SEDIMENTARY FACIES ANALYSIS
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
S P E C I A L P U B LI CA TI O N N U M B E R 22 I N T E R N A T I O NA L A S S OCIA T I O N
OF
THE
OF S ED I M E N T O L O GI S T S
Sedimentary Facies Analysis A TRIBUTE TO THE RESEARCH AND TEACHING OF HAROLD G. READING
EDITED BY A. GUY
b
Blackwell Science
PLINT
© 1995
The International Association
of Sedimentologists and published for them by Blackwell Science Ltd Editorial Offices: Osney Mead, Oxford OX2 OEL
DISTRIBUTORS
Marston Book Services Ltd PO Box 87 Oxford OX2 ODT
(Orders:
Tel: 01865 791155 Fax: 01865 791927
25 John Street, London WClN 2BL
Telex: 837515)
23 Ainslie Place, Edinburgh EH3 6AJ 238 Main Street, Cambridge Massachusetts 02142, USA 54 University Street, Carlton Victoria 3053, Australia
USA Blackwell Science, Inc. 238 Main Street Cambridge, MA 02142
(Orders:
Tel: 800 215-1000
Other Editorial Offices:
617 876-7000
Arnette Blackwell SA 1, rue de Lille, 75007 Paris France Blackwell Wissenschafts-Verlag GmbH KurfUrstendamm 57 10707 Berlin, Germany Feldgasse 13, A-1238 Wien Austria
Fax: 617 492-5263) Canada Oxford University Press 70 Wynford Drive Don Mills Ontario M3C 119
(Orders:
Tel: 416 441 2941)
Australia Blackwell Science Pty Ltd 54 University Street
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
Carlton, Victoria 3053 Tel: 03 347-0300
(Orders:
Fax: 03 349-3016)
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, except as permitted by the UK Copyright, Designs and Patents Act
A catalogue record for this title is available from the British Library ISBN 0-86542-898-0
1988, without the prior permission of the copyright owner. First published 1995
Library of Congress Cataloging-in-Publication Data Sedimentary facies analysis: a tribute to
Set by Setrite Typesetters, Hong Kong
the research and teaching
Printed and bound in Great Britain
of Harold G. Reading/
at the Alden Press Limited,
edited by A. Guy Plint.
Oxford and Northampton
p.
em.
(Special publication number 22 of the International Association of Sedimentologists) Includes bibliographical references and index. ISBN 0-86542-898-0 1. Facies (Geology) 2. Sedimentation and deposition. II. Reading, H.G.
[[[.
I. Plint, A. Guy.
Series: Special publication
of the International Association of Sedimentologists; no. 22. QE651.S43 1995 552' .5-dc20
94-30445 CIP
To Harold
Photograph courtesy of Tim Barrett
We offer this collection of papers as a token of our appreciation for your friendship, guidance and inspiration. In remembering your infectious enthusiasm, dedication and sometimes daunting expectations, we realize how deeply we were influenced by your philosophy and attitude; a gift that has, in no small measure, shaped the course of our professional ]jves. Your former students
Contents
IX
Preface
XI
Harold G. Reading
xm
Introduction
Clastic Facies Analysis 3
Alluvial palaeogeography of the Guaritas depositional sequence of southern Brazil Paulo S. G. Paim
17
Sedimentology of a transgressive, estuarine sand complex: the Lower Cretaceous Woburn Sands (Lower Greensand), southern England Howard D. Johnson and Bruce K. Levell
47
An incised valley in the Cardium Formation at Ricinus, Alberta: reinterpretation as an estuary fill Roger G. Walker
75
Gravelly shoreface and beachface deposits Bruce S. Hart and A. Guy Plint
101
The return of 'The Fan That Never Was': Westphalian turbidite systems in the Variscan Culm Basin: Bude Formation (southwest England) Robert V. Burne
137
Depositional controls on iron formation associations in Canada Philip Fralick and Timothy J. Barrett
157
Facies models in volcanic terrains: time's arrow versus time's cycle Geoffrey J. Orton
Tectonics and Sedimentation 197
Coarse-grained lacustrine fan-delta deposits (Pororari Group) of the northwestern South Island, New Zealand: evidence for Mid-Cretaceous rifting Malcolm G. Laird VII
Contents
vm
219
Sedimentation and tectonics of a synrift succession: Upper Jurassic alluvial fans and palaeokarst at the late Cimmerian unconformity, western Cameros Basin, northern Spain Nigel H. Platt
237
The use of geochemical data in determining the provenance and tectonic setting of ancient sedimetary successions: the Kalvag Melange, western Norwegian Caledonides Rodmar Ravnas and Harald Fumes
265
Differential subsidence and preservation potential of shallow-water Tertiary sequences, northern Gulf Coast Basin, USA Marc B. Edwards
Sequence and Seismic Stratigraphy in Facies Analysis 285
Seismic-stratigraphical analysis of large-scale ridge-trough sedimentary structures in the Late Miocene to Early Pliocene of the central North Sea Joe Cartwright
305
Millstone Grit cyclicity revisited, II: sequence stratigraphy and sedimentary responses to changes of relative sea-level Ole J. Martinsen, John D. Collinson and Brian K. Holdsworth
Facies Analysis in Reservoir Sedimentology 331
Productive Middle East clastic oil and gas reservoirs: their depositional settings and origins of their hydrocarbons Ziad R. Beydoun
355
The evolution of Oligo-Miocene fluvial sand-body geometries and the effect on hydro carbon trapping: Widuri field, west Java Sea Ray Young, W.E. Harmony and Thomas Budiyento
381
Index
Preface
This book stands out in the series of Special Publi
Bureau considered that this would be best done
cations of the International Association of Sedimen
through a Special Publication on a subject in line
tologists.
It
is
an
acknowledgement
of Harold
with Harold's work (obvious topics were clastics,
Reading's commitment to lAS, for whom he has
facies and depositional environments, sedimentation
been Publications Secretary, General Secretary and
and tectonics).
President successively, over the last 30 years.
It is therefore most appropriate that Guy Plint,
Harold has not only been source and inspiration
as Editor chosen for this special publication, has
of many of the lAS policies and activities over this
brought together a collection of original scientific
time, he has also been at the roots of 'facies sedimen
papers authored by Harold Reading's students, or
tology' as an art in itself, and as a major tool in the
students of theirs. To honour Harold Reading's own
broader field of geology.
scientific scope, the subject chosen is broad: sedi
More than providing his own personal contribution
mentary facies analysis. The contributions contained
to this branch of the earth sciences, Harold created a
in this Special Publication show to what extent facies
flourishing school of teaching and research. Harold's
sedimentology, as fostered by Harold Reading, is
approach has burgeoned from Parks Road, Oxford,
now established as a necessary basis to any under
to
standing of sedimentary rocks.
become
international
not
only
through
his
students, but also through 'his book'.
PETER HOMEWOOD
The Bureau of the lAS, taking up a suggestion by
/AS Publications Secretary
Robert Campbell of Blackwell Science, decided to put together a scientific tribute to Harold. The
IX
Harold G. Reading
Harold Reading was born in 1924 and, on leaving
at Rijswijk, arranged for Harold to investigate the
school, joined the Indian Army. This early experience
context of reported turbidites associated with English
left a lasting impression and undoubtedly contributed
Carboniferous deltaics in the Pennines and in south
to Harold's later concern for international cooper
west England.
ation. He went up to Oxford in 1948, initially to read
Cooperation with de Raaf and with Roger Walker,
Forestry, but his interests were diverted towards
one of Harold's earliest research students, developed
geology and he graduated in that subject in 1951.
a detailed appreciation of sedimentary structures
As an undergraduate, he visited North Norway to
and their role in understanding processes, and led
investigate the Late Precambrian and Cambrian
to the development of the style of facies analysis
stratigraphy of the Digermul Peninsula. This under
exemplified by the 1965 classic paper on the Carbon
graduate expedition not only shed significant new
iferous cycles of North Devon. Thereafter, Harold's
light on the stratigraphy of the area but also sowed
stable of research students grew rapidly as this volume
the seeds of a later rich sedimentological harvest.
amply testifies. Until his retirement, it was unusual
Three years at Durham under K.C. Dunham led to a
for him to have fewer than five or six doctoral
PhD with a project that involved mapping Carbon
students at any one time, this in addition to a full
iferous Y oredale cycles across an area of bleak
undergraduate teaching programme and responsi
Pennine moorland. Although the main thrust of the
bilities in college. This formidable work load was
study was stratigraphy and structure, the experience
carried out with great conscientiousness but Harold
of Carboniferous cyclicity was to set a further pointer
still had time to spare for external activities such
for the future.
as his involvement with lAS and JAPEC. During
On completion of his PhD Harold joined Royal
Harold's long career in Oxford, he only spent sus
Dutch Shell and immediately found himself in the
tained periods away on sabbatical on two occasions,
contrasting field conditions of Venezuela. This multi
the first in Leyden in the mid-1960s and the second
national, multidisciplinary environment developed
in Canada in 1972. The period in Holland led to
an appreciation of broader geological perspectives
close cooperation with structural geologists working
and the pragmatic, though rigorous, approach to
in the Cantabrians, an important extension of his
problem solving that has characterized Harold's
interests outside Britain.
career.
Of particular significance was a visit to
Harold's earliest students developed his early
Venezuela by Ph. Keunen who was, at that time,
interests, the Carboniferous deltas of Britain, and
actively promoting his pioneering work on turbidity
the tillites, shallow-marine and fluvial sediments of
currents
rigorous
northern Norway. Later, the Lower Palaeozoic of
approach to understanding depositional processes
Ireland and the Carboniferous of northern Spain
struck a chord with Harold, which was to be a
were added. As students were attracted to Oxford
and
their
deposits.
Kuenen's
cornerstone of his approach to sedimentology.
from around the world, the geographical spread
Harold returned to Oxford in 1957 as lecturer in
grew. However, geographical diversification was not
geology, a position that he held until retirement in
an end in itself but largely a result of Harold's
1991. When he took up his post, his teaching responsi
curiosity about wider controls on sedimentation,
bilities included mapping and palaeontology and
particularly the role of tectonics. He understood
stratigraphy. Sedimentology, as we know it, hardly
very early the implications for sedimentology of
existed. Harold first revived his interests in northern
Plate Tectonics, as exemplified by his pioneering
Norway through a further, largely stratigraphical
paper with Andrew Mitchell. Curiosity about new
expedition to Digermul. Perhaps more importantly,
geological ideas and the need to investigate their
he developed his interest in sedimentary process and
implications for sedimentology and vice versa has
environments through a relationship with Shell.
been a hallmark of Harold's geological thinking. By some standards, Harold has not been a prolific
Maurits de Raaf, then Head of Geological Research
XI
Harold G. Reading
XII
author, although his papers are always thoughtful
President have already been acknowledged by the
and stimulating. Published evidence of Harold's
Association
influence lies mainly in the rigour, originality and
Membership to Harold. It is worth remembering
appreciation of the wider geological perspective that
that it was in no small measure due to Harold's
itself
in
the
granting
of
Honorary
characterize many of the publications of his research
efforts that the Association changed from a largely
students and of second and third generation students.
European organization to one of real international
Harold edited one Special Publication of the lAS
stature. Harold's tireless efforts to meet and encour
on-strike-slip mobile belts, but his most valued publi
age sedimentologists of all ages and backgrounds
cation is the textbook Sedimentary Environments and
around the world and his endless patience and
Facies, initially written largely by Harold's former
diplomatic skill have been well rewarded in the
students and rigorously edited to reflect the high
healthy Association that we enjoy today. Harold
standards he espouses. The 3rd edition currently
has additionally been honoured by the Geological
occupies much of his 'retirement'.
Society of London with the award of the Lyell Fund
Although this book is essentially a celebration
and the Prestwich Medal and, most recently, by
of Harold's scientific influence, it is important,
SEPM with the award of its prestigious Twenhofel
especially in a Special Publication of the lAS, to
Medal.
acknowledge
his
enormous
contribution
to
the
development of sedimentology internationally. His unstinting efforts on behalf of the lAS, as Publi cations Secretary,
as General Secretary and as
JOHN CoLLINSON Shrewsbury, UK
Introduction
This volume is a very personal compilation. Unlike
from Harold's former graduate students and their
previous lAS Special Publications, it is not centred
students and co-workers, but to impose no constraint
on a specific geological theme, and for that I make
on topic, in order to illustrate the scope of Harold's
no apology. Instead, my intent was to illustrate, and
knowledge, interest and vision. In consequence, the
celebrate, the breadth of interest, energy and inspi
contents of this book are eclectic. The collection
ration that Harold Reading has brought to the field
of papers serves to highlight the power of facies
of sedimentary geology.
analysis, whether the rocks be volcanogenic, bio
Few would deny the depth of Harold's influence
genic, siliciclastic,
or even 'catastrophic'
on sedimentology, world-wide. In part, this is due to
olistoliths!),
his publications, in particular the enormously suc
method fostered by Harold.
cessful
Sedimentary
Environments
(mega
the scientific
Facies,
It is particularly appropriate that, amongst the
unquestionably the cornerstone for all those who
contributions, Ole Martinsen, John Collinson and
embark on sedimentary facies
and
and of course reflect
analysis! Equally
Brian
important of course, has been his pivotal role in the
Holdsworth
offer
new
interpretations
of
Namurian deltaic rocks in the northern Pennines,
foundation and development of the lAS, a contri
(upon which Harold cut his sedimentological teeth),
bution acknowledged recently with honourary mem
but which, judging from referees comments, still
bership of that Association.
provide fuel for heated debate! In similar vein,
His philosophy and attitude has of course travelled
Bob Burne presents a review and discussion of the
with his graduate students, drawn from 13 countries
depositional environment of the enigmatic Bude
on six continents. Because many of these students
Formation
returned home upon completion of their work in
1960s), but which is still subject to sharply divergent
Oxford, and others now work and teach outside
interpretations.
the UK, the approach Harold fostered during their
Roger Walker shows how important it is, both to
(which In
Harold a
studied
salutory
in
lesson
the to
early
us
all,
graduate days has continued to spread. (He may not
separate facts from interpretations, and to ques
know this, but in a geneological sense, Harold
tion one's cherished interpretation, when he boldly
is now a great-great grandfather to at least one
reinterprets as an incised valley fill, rocks he pro
young sedimentology student who doubtless is quite
claimed a turbidite channel deposit just nine years
unaware of the history of the supervisory influence
ago! As Editor of this volume, I am indebted to the
that has been passed down!)
following people whose thorough reviews served
Although initially conceived as a thematic volume with contributions to be invited from a panoply of
to clarify the papers, and who made my job that
leading
much easier: Gail M. Ashley, Timothy R. Astin, T.
sedimentologists,
two difficulties quickly
arose: first, just what was to be the theme? As
Christopher
Harold has been involved in so many areas of
Charlie S. Bristow, H. Edward Clifton, Thomas C.
Baldwin,
sedimentary geology, selection of any one topic
Connally,
simply served to highlight gross neglect of another.
Peter G. DeCelles, Frank G. Ethridge, Jill Eyers,
Edward
Janok
Cotter,
P.
Bhattacharya,
William
R.
Dupre,
Secondly, it rapidly became apparent that numerous
Stephen S. Flint, Edward C. Freshney, Robert L.
former students were anxious to pay their own
Gawthorpe, Roland Goldring, Anthony J. Hamblin,
personal tribute, and whose contributions could,
Alan P. Heward, Phillip R. Hill, Richard N. Hiscott,
alone, easily constitute a hefty volume! Of the 34
Richard S.
students whom Harold guided through doctoral
McCabe, Kathleen M. Marsaglia, Franco Massari,
theses between 1961 and 1994, 16 have authored, or
Gerrard V. Middleton, Robert A. Morton, George
co-authored papers in this volume.
Postma, William C. Ross, Alastair H. Ruffell, Bruce
In keeping with the sentiment of this festschrift, I
Hyde, Elana L.
Leithold,
Peter J.
W. Sellwood, Gary A. Smith, Roger G. Walker,
took the decision to limit contributions to those
James
xiii
D. L.
White,
John
A.
Winchester
and
Introduction
XIV
Jonathon Wonham, plus two people who chose to
clandestine spmt of this project,
remain anonymous.
essential intelligence on both Harold and his former
I am very grateful to Susan Sternberg, Edward Wates and Julie Elliott at Blackwell Science who provided guidance at critical phases in the prep aration of this book. I also thank Diana Relton (Earth Sciences,
Oxford) who entered into the
graduate students.
A. GuY PuNT London, Ontario
and provided
Clastic Facies Analysis
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
Spec. Pubis int. Ass. Sediment.
(1995) 22, 3-16
Alluvial palaeogeography of the Guaritas depositional sequence of southern Brazil P A U L O S. G. P AlM* Earth Sciences Department , Oxford University, Parks Road, Oxford OXI 3PR, UK
ABSTRACT
The Guaritas sequence is the uppermost stratigraphical level of the Camaqua Basin (southern Brazil) and comprises an alluvial, deltaic and aeolian continental facies association up to 800 m thick. Facies mapping of this unit has revealed a lateral association of tributary fans and trunk braided rivers developed under semi-arid conditions. Two main regions (lobes) of alluvial fan development can be discriminated and the source points of both coincide with synforms in the nearby basement. This depositional system presents a normal down fan facies change. An anomalous lateral change of facies within the trunk river system is interpreted as having been inherited from pre-existing alluvial fan deposits. The main alluvial facies comprise trough cross-stratified (74%) and horizontally bedded (7%) sandstones, massive (16%) and tabular cross-stratified (2%) orthoconglomerates, and massive mud stones (1%) . Vertical aggradation of three-dimensional subaqueous dunes, followed by an upper flow regime plane-bed phase, characterized the depositional events of the sandy areas of the alluvial system. Diffuse gravel sheets and minor longitudinal and transverse bars were the main geomorphological features of the gravelly alluvial reaches. Fine-grained sediments represent temporarily abandoned areas within the braided channel network.
INTRODUCTION
trending tectonic structure, in southern Brazil (Fig. 1}, and evolved during the latest phases of the Brasilia no orogenic cycle (strike-slip basins of Brito Neves & Cordani (1991}}. An extensional or transtensional event at the end of the Brasiliano orogenic cycle, and the consequent formation of intermontane basins, has been pro posed as the tectonic setting of the Camaqua Basin during the deposition of the Guaritas sequence (Fragoso-Cesar et al., 1984, 1992; Beckel, 1990, 1992). In the past decade, the Guaritas depositional sequence has received attention from several authors in terms of facies analysis and palaeoenvironmental interpretation (Becker & Fernandes, 1982; Fragoso Cesar et al., 1984; Jost, 1984; Lavina et a!., 1985; Beckel, 1990). Generally, these papers have indi cated continental sedimentation characterized by
The Guaritas depositional sequence constitutes the uppermost unit of the Camaqua Basin infilling and it is an unconformity-bounded stratigraphical unit: it overlies older deformed molasse strata (angular unconformity) and is covered by Permian sedimen tary rocks of the Parana Basin. The Guaritas sequence, about 800 m thick, is almost always flat-lying, although, near to regional faults some extensional reactivation has tilted the Guaritas deposits. The available radiometric dating, summarized in Soliani et al. (1984) and Fragoso Cesar et al. (1984), indicates a Cambro-Ordovician age for the deposition of the Guaritas sequence. The Camaqua Basin is located in a NE-SW * Permanent address: UNISINOS- Departamento de Geologia, Av. Unisinos 950, Sao Leopoldo RS, Caixa Postal 275, CEP 93022-000, Brazil.
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
3
P.S.G. Paim
4
Permo Triassic
T �
BRAZIL
>-
� L:...:!l
Upper Vendian to Ordovician
t-':(:�: ., .. . .•
� >:<: c::
-�
Guaritas depositional sequence older molasse sequences
Middle to Upper Proterozoic
1· ;;; � :f:f �
N
granites
1
meta volcanic/sedimentary rocks
Archean to Lower Proterozoic
Study area
Br153 .....--_
� B.:::J
Main roads
Scale (Km)
Major faults Ca9apava do Sui 2
6
-
-
0
6
3
12
18
Santana da Boa Vista
Fig. I. Location map and geological setting of the Camaqua Basin. Modified from DNPM/CPRM (1987).
alluvial fan and braided alluvial plain deposits, with associated aeolian and lacustrine sediments. A semi arid environment has been proposed for the overall Guaritas sequence. The most detailed study on the depositional sys-
terns of the Camaqua Basin was presented by Lavina et al. (1985). In this paper the alluvial facies were related to marginal alluvial fans (channel and debris flow deposits) associated with an axial braided alluvial plain. Gravelly longitudinal bars and sandy
5
Palaeogeography of the Guaritas sequence subaqueous dunes and transverse bars were the main morphological elements attributed to the alluvial palaeostreams. Petrological studies by De Ros et al. (1994) on samples from alluvial and aeolian facies of the Guaritas sequence indicate the presence of: (i) fresh feldspar and volcanic lithoclasts; (ii) aggregates of hematite; (iii) oxidized grains; (iv) caliche (concen tric interlayering of calcite and iron oxide); and (v) silcretes. These early diagenetic features reflect arid to semi-arid conditions during the deposition of the Guaritas sequence. Basin-wide facies mapping of the Guaritas se quence carried out by the author in 1988, reinforce previous interpretations indicating intermittent vol canic activity and an aeolian, alluvial and deltaic facies association (Fig. 2). Basin-scale changes of the alluvial facies charac teristics suggest that an objective delineation of dis tinct alluvial subenvironments is possible. These alluvial subenvironments, as well as a brief descrip tion and interpretation of the main alluvial facies, are the main subject of this paper. The alluvial deposits will be discussed in terms of their general features of texture, fabric, sedimentary structures and palaeocurrent pattern on a basin wide scale. Both the mean sedimentary facies charac teristics and the lateral facies changes within the alluvial system are described. The data base includes 403 outcrop descriptions distributed over an area of nearly 1600 km2 (see Fig. SA). This area was subdivided into 46 equal rec tangles (8 x 9 km) and mean values of several par ameters were calculated for each subdivision. The results of this approach are presented in Tables 1 & 2 and summarized in Figs 5 & 6. This approach involves comparison of values from different strati graphical levels. The consistent results (see Fig. 5) throughout the basin, with sampling at several stratigraphical levels (Fig. 2), suggest that the palaeoenvironments were more or less stationary throughout deposition of the Guaritas sequence. A detailed three-dimensional facies architecture analysis (architectural elements approach of Allen (1983) and Miall (1985)), aiming to build up a local alluvial model on a channel-fill scale, is part of my ongoing studies and will be the subject of another publication.
ALLUVIAL GENERAL
FACIES:
FEATURES
To simplify terminology the lithofacies classifi cation proposed by Miall (1977), as modified by Miall (1978), Rust (1978) and Bromley (1991), was adopted. Table 1 presents the main characteristics of each sedimentary facies described in the field. The terminology and classification scheme pro posed by the SEPM (Society of Economic Paleon tologists and Mineralogists) Bedforms and Bedding Structures Research Symposium (Ashley, 1990) for description of large-scale flow-transverse bedforms (excluding antidunes) was adopted. The alluvial deposits (Table 1) are sand dominated (facies S, 81% ) with a smaller amount of conglom erates (facies G, 18% ) and an insignificant amount of pelites (facies F, 1%). Facies S is composed mainly of medium- to coarse-grained sandstones (41% ), with a significant proportion of pebbly to very coarse-grained (25%) and fine- to very fine grained (15%) sandstones. Facies G is composed of pebbles (9%) and granules (8% ) and minor amounts of cobbles (1% ). A few boulders occur in the base of some conglom erate beds, mainly near the eastern border of the Camaqua Basin. The alluvial deposits are usually arranged in fining upward cycles bounded by fifth-order surfaces (sensu Miall, 1988). These cycles are 0.5-4 m thick and tens of metres in lateral extent (Fig. 3), both parallel to and perpendicular to palaeoflow, and can be classified as laterally extensive to sheet-like deposits following the classification of Friend et al. (1979). The proportion of the different textural classes within the fining upward cycles changes laterally with increasing gravel content toward both margins. Conglomerates (G)
Clast-supported conglomerates comprise around 18% of the alluvial facies and massive conglomerates are the most common (Table 1). Clast-supported conglomerates are a very common facies in the lowermost parts of the fining upward cycles. Massive clast-supported conglomerates (facies Gm, Table 1) are the main lithotype of facies G and normal grading, clast orientation and imbrication are their most conspicuous sedimentary features. Facies Gp is characterized by gravels (mainly pebbles) arranged in small- to large-scale, normally isolated, sets of tabular cross-stratification. This
P.S.G. Paim
6 A
10km B
\; ·>.1 � t:;;:.t1
Mainly alluvial facies
Mainly volcanic rocks
Q o
Mainly eolian facies Pre-Guaritas basement
E:f=3-g
Mainly deltaic facies
LA: j
Permo Triassic
Fig. 2. Three-dimensional view of Camaqua Basin and surrounding area (same region of Fig. 1): (A) topography and (B) sketch of the Guaritas sequence facies.
7
Palaeogeography of the Guaritas sequence
Table 1. Classification and relative percentage of the sedimentary lithofacies (lithofacies code adapted from Miall (1977, 1978) and Rust (1978)) Rock type Conglomerates (G)
Sandstones (S)
Mudstones (F)
Facies code
Description
Percentage
Gm
Massive or crudely bedded conglomerates (cobbles, pebbles and granules)
l6
Gp
Small- to large-scale tabular cross-stratified conglomerates (granules and pebbles)
Gt
Small- to large-scale trough cross-stratified conglomerates
Gms
Massive, matrix-supported conglomerates (boulders to granules dispersed in a muddy sand matrix)
St
Small- to large-scale trough cross-stratified sandstones (pebbly to very fine-grained)
Sh
Horizontally bedded sandstones (medium to very fine-grained)
Sp
Medium to pebbly sandstone with small- to large scale planar cross-stratification
Fm
Massive mudstones with mudcracks
Fl
Laminated to rippled very fine sandstone to siltstone
Table 2. Relative percentage of trough cross-stratification and horizontal lamination in each sandy textural class Texture Pebbly to very coarse grained
Facies St
Sedimentary structures
Percentage
Small scale Medium scale Large scale
10 49 41
Sh Coarse to medium grained
St
0 Small scale Medium scale Large scale
10
Sh Fine to very fine grained
St
Sh
10 43 37
Small scale Medium scale Large scale
15 37 28 20
facies commonly occurs associated with facies Gm (Fig. 3). Facies Gt is rare, finer grained than facies Gm and Gp and characterized by small- to large-scale trough cross-stratification (alternations of small pebbles and gravelly sands). This facies interfingers with facies Gm and grades into facies St (Fig. 4). Matrix-supported conglomerates (facies Gms) are
2
74 7
also rare and occur, locally, near the eastern border of the Camaqua Basin. The main characteristic of this facies is its chaotic arrangement of pebbles, cobbles and, less commonly, boulders floating in a muddy to sandy matrix. Sandstones
Trough cross-stratification (facies St, 74% ), in places disrupted and/or deformed by convolution, and horizontal bedding (facies Sh, 7% ) are the main features of the alluvial sandy deposits (Figs 3 & 4). Planar cross-stratification (facies Sp) is rare. Facies St is characterized by very fine- to very coarse-grained sandstones with trough cross-bedding (Table 1). The cross-strata are predominantly of medium to large scale in all textural classes, but the proportion of small-scale trough cross-stratification increases as sandstones become finer grained (Table 2). This facies is the most common in the fining upward cycles. Convolute bedding is common in trough cross stratified sandstones (facies St). Within a single cross-stratified set, all gradations may occur from oversteep foresets, recumbent folding to intense deformation and even complete destruction of the former bedding (facies Sm and Spo of Bromley, 1991). Deformation near the top of the cross-
8
P.S.G. Paim
Fig. 3. Main alluvial lithofacies: facies Gt, St, Sh, Spo and, in the uppermost part of the picture, Fl, Gm and Gp. Bar scale is 2 m long.
stratified set is commonly characterized by downcur rent oversteepening of the cross-strata (Figs 3 & 4), and the intensity of convolution increases down the slip-face. Horizontal bedding (facies Sh) does not occur associated with pebbly and very coarse-grained sand stones and comprises 10% of the sedimentary struc tures of medium- to coarse-grained sandstones and 20% of the fine- to very fine-grained sandstones (Table 2). This facies is often related to the upper most parts of the fining upward alluvial cycles (Figs 3 & 4). Planar-tabular cross-stratified sandstones (facies Sp) are not common in the Guaritas sequence alluv ial deposits (Table 1). They occur as small- to large scale sets in pebbly to medium-grained sandstones and are normally interlayered with facies St. Other facies
Massive mudstones are rare and commonly mud cracks are their most conspicuous feature (facies Fm). Very fine-grained sandstones and siltstones (facies Fl) are also, and can be either horizontal (Fig. 3) or, more rarely, cross-laminated (Table 1). Both usually occur in the uppermost parts of the fining upward alluvial cycles. Alluvial facies: summary of general features and interpretations
The textural aspects (Table 1) suggest that the alluv ial facies of the Guaritas sequence represent bedload stream deposits in which the bedload was predomi-
nantly sandy and the suspension load, if deposited, was almost completely eroded by subsequent flood events. This type of stream commonly has a braided pattern characterized by low sinuosity and highly mobile channels (Collinson, 1986). The sheet-like geometry of the fining upward cycles enclosed by fifth-order bounding surfaces suggests broad, shallow channels. In terms of the gravelly facies, the dominance of clast-supported conglomerates (Table 1) is indicative of gravel deposition by strong tractive flows, whereas the finer grained material (sand and mud) was still being carried in suspension (Rust & Koster, 1984). Thin beds of facies Gm associated with laterally extensive channels suggest the development of dif fuse gravel sheets (Hein & Walker, 1977) by very extensive and shallow sheet-floods (Collinson, 1986). Thicker deposits of facies Gm suggest deeper and less ephemeral flows (Rust, 1978) causing more extensive vertical aggradation of gravel bars with low depositional dips. These deposits commonly have been associated with the development of longi tudinal and/or diagonal gravelly bars (Smith, 1970; Rust, 1972, 1978; Miall, 1977, 1978; Rust & Koster, 1984; Collinson, 1986) under high water and sedi ment discharge (Hein & Walker, 1977). Conglomerates with planar cross-stratification (facies Gp) has been related to (i) two-dimensional dune migration (transverse and/or linguoid gravel bars of Hein & Walker (1977), Miall (1977) and Middleton & Trujillo (1984)) as well as to (ii) later modifications of longitudinal bars (Smith, 1970; Rust, 1978; Enyon & Walker, 1974) in modern alluvial gravelly reaches. The frequent occurrence of
;;,o
!:)
�
�
-§ � -.:;, �
"'
C)
§
;::. s
"' "'
.E Fig. 4. Detailed view of Fig. 3 (enlargement of its lower part): facies Gt, St, Sh, Spo and thin tabular beds of Gm. Bar scale is 2 m long.
"' "' ;:s '"' "'
'.0
10
P.S.G. Paim
isolated sets of facies Gp within deposits of facies Gm could be explained more easily by the second hypothesis. Trough cross-stratified conglomerates are rare (facies Gt) and have been associated with (i) three dimensional dune migration, as observed by Fahnestock & Bradley (1973) and Galay & Neill (1967), and (ii) channel scour-and-fill structures (Miall, 1977; Middleton & Trujillo, 1984). The same criteria previously used to interpret facies Gp can also be applied in this case: the solitary nature of this facies suggests the deposition of gravel in depressions around diffuse gravel sheets. Matrix-supported conglomerates (facies Gms) are also rare and represent mud- and debris-flow deposits commonly associated with an alluvial fan setting (Blackwelder, 1928; Bull, 1963; Hooke, 1967; Rust & Koster, 1984; Collinson, 1986; Blair & MacPherson, 1992). Sandy sediments constitute the majority of the Guaritas alluvial deposits (Table 1) and are exten sively dominated by facies St (Table 2). Trough cross-stratified sandstones have been related almost invariably to migration of three-dimensional dunes (e.g. Collinson, 1970; Williams, 1971; Harms et al., 1975; Miall, 1977; Rust, 1978). In braided alluvial settings these bedforms usually have been associated with in-channel deposition (Cant & Walker, 1976, 1978; Cant, 1978; Walker & Cant, 1984). Such repetitive sand deposits commonly are considered as flood-stage bedforms (Williams, 1971) and are larger in deep channels (Cant, 1978). The association of facies St with the lower and middle part of sheet-like fining-upward cycles suggests this facies could be related to flood stage in shallow channels. Subcritical climbing trough cross strata (facies St) indicate subaqueous dune aggra dation. Sporadic lateral accretion of these bedforms is indicated by inclined planes (first-order bounding surfaces of Miall (1988)) dipping perpendicular to the dune migration direction (Paim, 1994). The absence of third-order surfaces (except the rare occurrence of lateral accretion surfaces) associ ated with the subcritical climbing of the trough cross-bedded sets (facies St) suggests rapid depo sition of a sandy load, transported by traction plus suspension, without macroform (sensu Jackson, 1975) development. Deformation of trough cross-stratified sandstones is a very conspicuous feature of the Guaritas sandy alluvial facies. Recumbent folding in cross-bedded sandstones commonly has been attributed to shear
stress acting on a liquefied sand bed and caused by current drag (Allen & Banks, 1972; Doe & Dott, 1980; Owen, 1987) or by the movement of large bedforms over an unconsolidated substrate during high-flow stages (Plint, 1983). Horizontal bedding (facies Sh) occurs most often in the finest fraction of the sandy deposits (Table 2). This textural control, associated with the occurrence of parting lineation and scattered small pebbles and granules near the base of the horizontally bedded sets, indicates its origin as an upper flow regime bedform. Deposits with the same characteristics of facies Sh usually have been linked to an upper flow regime phase developed during flood stages on the channel floor (McKee et al., 1967; Williams, 1971; Miall, 1977) or under the influence of high-velocity and low depth flows on the top of sand-flats (Cant and Walker, 1978; Miall, 1977; Collinson, 1986). The common occurrence of this facies (Sh) on the upper most parts of the fining upward cycles supports an interpretation involving upper flow regime currents reworking the top of the previous alluvial deposits. Planar-tabular cross-stratified sandstones (facies Sp) are rare. Within alluvial settings this facies commonly has been related to slip-face advance of two-dimensional dunes (transverse -linguoid or lobate bars of Collinson (1970, 1986), Smith (1970), Williams (1971), Asquith & Cramer (1975), Miall (1977), Cant & Walker (1978) and Cant (1978); or sand waves and straight-crested megaripples of Smith (1970), Collinson (1986) and Miall (1978)). Smith (1970) related the origin of the two dimensional dunes to the development of 'deltas' in pre-existing channel-floor depressions, whereas Cant & Walker (1978) related them to flow expansion at channel junctions or places where the channels widen. Facies Fm and Fl are not common in the alluvial system of the Guaritas sequence (Table 1). Their rarity and generally lenticular geometry (Fig. 3) are suggestive of waning flood deposits settling on to temporarily abandoned areas of the braided system (Cant, 1978; Cant & Walker, 1978; Miall, 1978). In general, diffuse gravel sheets and longitudinal/ diagonal bars were the main geomorphological elements of the gravelly reaches, whereas sub aqueous three-dimensional dunes characterized the sandy portions of the Guaritas alluvial system. The predominance of vertical aggradation of dunes instead of downstream and/or lateral accretion of more stable sandy accumulations (e.g. sand-flats)
11
Palaeogeography of the Guaritas sequence suggests a highly variable hydrological character and predominance of the upper part of lower flow regime conditions within the channels. Debris-flow and sheet-flood deposits suggest the presence of alluvial fans within the alluvial system as well as flashy discharge due to sporadic, but torren tial, rainy seasons. The above interpretations together suggest an alluvial drainage developed under semi-arid con ditions (large discharge fluctuations) with alter nation of flood events and dry seasons. These conclusions are reinforced by the aeolian associ ation (Lavina et al., 198S) and by petrographical evidence related to early diagenetic processes (De Ros et a/., 1994).
ALLUVIAL
FACIES:
LATERAL CHANGES
The previous section describes the pattern of alluvial sedimentation in terms of mean regional values and, in this way, reflects the major features of the alluvial deposit. In the following section, spatial variation in some sedimentary features is described and, when possible, interpreted. To achieve this, the mean values, per unit area, of several sedimentary par ameters were calculated using the outcrop locations and grid presented in Fig. SA.
Palaeocurrent pattern
The pattern of sediment transport within the entire Camaqua Basin was calculated using the grid and outcrops shown in Fig. SA. In order to eliminate problems associated with the analysis of several types and scales of sedimentary features (Miall, 1977) mean vectors were calculated only from trough cross-stratification. By using only one rank of sedimentary features, difficulties related to vector magnitude were eliminated (Allen, 1963; Miall, 1974). In addition, dunes seem to be associ ated with high-stage flow and, consequently, should be good indicators of the true downstream direction (Miall, 1977). The distribution of the palaeocurrent vector means (Fig. SB) indicates two major dispersal compart ments within the alluvial system: 1 from the eastern border to the basin axis the sedimentary transport was almost perpendicular to the regional tectonic trend (a general mean vector of
282°, with a correlation coefficient of 0. 86), reflecting a sedimentary input towards the basin axis; 2 from the basin axis to the western border, palaeo currents were predominantly parallel to the struc tural trend (general mean vector of 211 with a correlation coefficient of 0.96). o,
Pattern of textural dispersion
The alluvial deposits of the Guaritas sequence are composed primarily of sandstones (mainly facies St and Sh), minor conglomerates (mainly facies Gm and Gp) and trace amounts of fine-grained sediments (facies Fm and Fl), as has been described in the previous section. In this paper three types of alluvial deposits are distinguished: sandy (:2: 70% sand stones); mixed (70-30% sandstone); and conglom eratic ( ::::: 30% sandstone). Figure SC shows the percentage of sandstone (rela tive to conglomerate) through the entire basin and illustrates a gradual decrease from sand dominated alluvial deposits along the basin axis (axial alluvial sedimentation), to mixed alluvial deposits toward both basin margins (marginal alluvial sedimen tation). Likewise, Fig. SD presents a plan view of the spatial changes of the percentages of coarser grained sediments (conglomerates plus pebbly and very coarse-grained sandstones) relative to finer grained sediments (coarse to very fine sandstones). A pattern quite similar to the former (Fig. SC) can be seen. Clearly, the facies St and Sh are gradually replaced by facies Gm towards both basin borders. In both cases (Figs SC & SD) the only exception to the general pattern of sediment distribution is a NW -SE trending intrusion of coarse material in the southeast region of the basin. Alluvial facies: interpretation of lateral changes
Figure 6 presents an interpretation of the alluvial palaeogeography of the Guaritas sequence based on the lateral variations of the textural and palaeocur rent data. This figure was constructed according to the following considerations. The palaeocurrents suggest the coexistence of two distinct alluvial subenvironments (Fig. SB) with almost orthogonal mean sedimentary transport pat terns (282° versus 211°). 1 The first dispersal system (282°), developed in the eastern part of the basin, is characterized by the highest palaeocurrent vector dispersion and by palaeo flow almost perpendicular to the tectonic trend of
12
P. S.G. Paim B
12 -
Mean vector and number of readings per area Boundary between tributary alluvial fan system and trunk braided river system
50
�
Mean vector and number of readings of both alluvial systems
\
5
A
3 46 ,514 / 44 ti 99
!
26 / 64 /
;;
101
I
--
32
-36
c
§
>90 80-89 70-79
D
1
60 41
\
100 ""'-89 14
"4
/
3 N
---
""'...... 26
H 30-39 ��29 � 0
Fig. 5. Lateral changes within the alluvial system: (A) grid and location of alluvial outcrops used to calculate palaeocurrent and textural mean values; (B) palaeocurrent mean values per area; (C) percentage of sandstone relative to conglomerate; and (D) percentage of coarser grained sediments (gravel plus pebbly to very coarse-grained sand) relative to finer grained sediments (coarse to very fine-grained sand plus mud).
the basin. This subenvironment is interpreted as a tributary alluvial fan (sensu Rust & Koster, 1984). 2 The second dispersal system (211°), represented by palaeoflow parallel to the basin axis, by low palaeocurrent vector disperson, and characteristic of
the western portion of the basin, is interpreted as a trunk braided river (sensu Rust & Koster, 1984). As Collinson (1986) stated, 'it is sometimes poss ible to identify individual fans by the establishment of a radial pattern of palaeocurrents over an area'.
13
Palaeogeograph y of the Guaritas sequence
E
c
Gm Gt St
Sh Spo
n
o -
Sm 1
-
2
-
3
-
Mudstones
Sandstones
Conglomerates
Sandy alluvial deposiiS
Mixed alluvial deposiiS Tributary fan streams Braided trunk river streams Reworked tributary fans
--A
Boundary between trunk rivers and tributary fans Plan view of the alluvial palaeogeography
B
Western margin trunk river facies association
C
Axial sandy trunk river facies associalion
(reworked alluvial fans)
D
Eastern margin proximal alluvial fan facies
E
Summary of the alluvial facies
association
Fig. 6. Alluvial palaeogeography and lateral facies changes: (A) plan view of the alluvial subenvironments; (B) and (D) vertical profiles of marginal facies association; (C) vertical profile of axial facies association; and (E) ideal vertical arrangement of the main facies. Facies code from Miall (1978); Bromley (1991).
Here, the tributary fan mean vectors (Fig. 5B) indicate the coalescence of two main fan lobes (Fig. 6A). The northern fan has a radius of 15 km, whereas the southern fan has a radius of 20 km. These dimen sions are comparable to the size of recent examples of semi-ariel and arid alluvial fans documented by Heward (1978). The point source of both lobes coincides with structural lows (synforms composed
of easily erodible metapelites) in the nearby base ment (Fig. 2). The southern lobe was more important than the others in terms of sedimentary input, as can be deduced both from it having penetrated furthest into the basin (Fig. 5B) and from the major intrusion of coarsest material from southeast to northwest (Figs 5C & 5D) in the southeast region of the basin. Comparison of the grain size distribution (Figs 5C
14
P. S.G. Paim
& 50) and palaeocurrent mean vectors (Fig. 5B) reveals some obvious relationships as well as some discrepancies. The gradual decrease of grain size, from both basin margins towards the basin axis (Figs 5C & 50) can be interpreted as a consequence of lateral alluv ial fan input. This matches the palaeoflow data of the eastern side of the basin, but not that on the western side (Fig. 5B). The textural, palaeocurrent and facies data sum marized in Fig. 6A demonstrate that the grain-size distribution within the trunk braided river system does not show a downcurrent fining, which is a very common characteristic of many braided alluvial environments (e.g. Smith, 1970; Miall, 1977, 1978; Rust, 1978; Collinson, 1986). Instead, the trunk rivers present a lateral change from sand-dominated deposits near the basin axis to mixed deposits towards the western basin border. Such a discrepancy can be related to a dominant alluvial input from the eastern border (tributary alluvial fans) causing the development of a trunk braided river system on the western side of the Camaqua Basin. The emplace ment of trunk rivers in the western region could cause a major remoulding of the alluvial fan deposits on the western border without erasing the down-fan fining. Some of these previous alluvial fan deposits could be preserved and thus could explain some atypical palaeocurrent readings made near the west ern margin, which point to a southeasterly directed sediment discharge. Transitions between purely sandy or gravelly reaches were the norm inside the alluvial system of the Guaritas sequence. The facies are usually arranged as fining upward cycles, bounded by fifth order bounding surfaces, with the major facies super imposed in the following order: Gm-Gp-St-Sh Fm. This characteristic vertical arrangement, reflecting the proportion of each facies (Table 1), is illustrated in a summary (idealized) vertical profile (Fig. 6E) incorporating the mean values of the principal facies observed throughout the basin. Comparison of the summary profile (Fig. 6E) with actual sections around the basin, and with the regional textural variation (Figs 5C & 50), facilitates the identification of some common marginal and axial facies associ ations, summarized in Figure 6: profiles B and D typify common marginal facies associations, whereas profile C represents the axial part of the basin.
CONCLUSIONS
The alluvial deposits of the Guaritas sequence reflect a lateral association of tributary alluvial fans and trunk braided rivers. The alluvial fans show a down stream decrease in mean grain size whereas the trunk rivers present no longitudinal variation in texture. Instead, the trunk rivers show a lateral grain-size change that is interpreted to have been inherited from a hypothetical alluvial fan system fed from the western margin of the basin. The alluvial fans are dominated by water laid deposits and comprise two main lobes. The source points of both alluvial fan lobes coincide with structural lows, suggesting control by basement topography. Semi-arid conditions during the alluvial depo sition are suggested by: sheet-flow and debris-flow deposits; petrological evidence, such as fresh feld spar and volcanic lithoclasts, interstitial hematite aggregates, caliche and silcretes; and the association with aeolian facies (although aeolian facies are com mon in early Palaeozoic sequences because of the absence of land vegetation regardless of climate). Diffuse gravel sheets and longitudinal bars were the main geomorphological elements in the alluvial gravelly reaches. Subaqueous dune aggradation, fol lowed by partial reworking of the deposit by upper flow regime currents, characterized the sandy reaches. An idealized channel-fill succession is typified, from base to top, by: (i) a horizontal to slightly undulatory erosional surface; (ii) gravel deposits, representing diffuse gravel sheets and longitudinal bars (Gm), locally with avalanche faces (Gp); and (iii) sandy deposits, consisting mainly of three dimensional subaqueous dunes (St), rare two dimensional dunes (Sp), and plane beds (Sh), on the top.
ACKNOWLEDGEMENTS
Thjs study was carried out during the tenure of a postgraduate scholarship awarded by the Research Council of the Brazilian Government (CNPq), and forms part of the author's D .Phil. thesis at the University of Oxford, England, written under the supervision of Dr H.G. Reading. Field-work costs were supported by CNPq (Grant 413321/ 88-6), Universidade do Vale do Rio dos Sinos (UNISINOS), and Company of Research of Mineral
Palaeogeography of the Guaritas sequence Resources (CPRM) of the Brazilian Government. The author wishes to thank H. G. Reading and H.C. Jenkyns for criticism and revision of an earlier version of the manuscript. Later reviews by G. Plint, G.V. Middleton, A.P. Hamblin and P.A. Allen have enabled me to make several very useful improve ments to the paper.
RE FERENCES
J.R.L. (1963) The classification of cross-stratified units, with notes on their origin. Sedimentology, 2, 93-114. ALLEN, J.R.L. (1983) Studies in fluviatile sedimentation: bars, bar-complexes and sandstone sheets (low-sinuosity braided streams) in the Brownstones (L. Devonian), Welsh borders. Sediment. Geol., 33, 237-293. ALLEN, J.R.L. & BANKS, N.L. (1972) An interpret atiOn and analys1s of recumbent-folded deformed cross bedding. Sedimentology, 19, 257-283. ASHLEY, G.M. (1990) Classification of large-scale sub aqueous bedforms: a new look at an old problem. 1. sediment. Petrol., 60(1), 160-172. AsQUITH, G.B. & CRAMER, S.L. (1975) Transverse braid bars in the Upper Triassic Trujillo sandstone of the Texas Panhandle. J. Geol., 83, 657-661. BECKEL, J. (1990) Metalogenia del Cu, Pb y Zn en ALLEN,
Ia Cuenca de Camaquii durante e/ Cicio Orogenico Brasiliano- Rio Grande del Sur (Brasil). Unpublished PhD thesis, Universidad de Salamanca,274 pp. BECKEL, J. (1992) Evolus;ao geotect6nica da Bacia do Camagua, RS- Proposta de classificas;ao dentro de um conceito mobilista. J Workshop s6bre as bacias molassicas brasilianas,
Boletim
Especial
Resumos
Expandidos,
pp. 1-5. Sao Leopoldo. BECKER, M. R. & FERNANDES, L.A.D. (1982) Caracter izas;ao faciol6gica de uma sequencia vulcano-sedimentar eo-Paleoz6ica na f6lha Passo do Tigre ( RS). Acta geo/. Leopold. , 13(6), 287-321. BLACKWELDER, E. (1928) Mudflow as a geologic agent in semi-arid mountains. Geol. Soc. Am. Bull., 39, 465-483. BLAIR, T.C. & McPHERSON, J.G. (1992) The Trollheim alluvial fan and facies model revisited. Geol. Soc. Am. Bull., 104, 762-769. BRITO NEVES, B.B. & CoRDANl, U.G. (1991) Tectonic evolution of South America during the Late Protero zoic. Precam. Res. , 53, 23-40. BROMLEY, M.H. (1991) Variation in fluvial style as revealed by architectural elements, Kayenta Formation, Mesa Creek, Colorado, USA: evidence for both ephem eral and perennial fluvial processes. In: The Three dimensional Facies Architecture of Terrigenous Clastic Sediments and its Implication for Hydrocarbon Discovery
(Eds Miall, A.D. & Tyler, N.). Soc. Econ. Paleontol. Mineral. Spec. Publication, Concepts Sedimentol. Paleontol. Tulsa, 3, 94-102. BuLL, W.B. (1963) Alluvial fan deposits in Western Fresno County, California. J. Geol., 71, 423-251. CANT, D.J. (1978) Development of a facies model for a_nd Recovery
15
sandy braided river sedimentation: comparison of the South Saskatchewan river and the Battery Point Forma tion. In: Fluvial Sedimentology (Ed. Miall, A.D.). Can. Soc. petrol. Geol., Calgary, Memoir 5, pp. 627-639. CANT, D.J. & WALKER, R.G. (1976) Development of a braided-fluvial facies model for the Devonian Battery Point, Quebec. Can. J. Earth Sci. , 13, 102-119. CANT, D.J. & WALKER, R.G. (1978) Fluvial process and facies sequence in the sandy braided South Saskatchewan river, Canada. Sedimentology, 25, 625-648. CoLLINSON, J.D. (1970) Bedforms of the Tana river, Norway. Geogr. Annaler, 52A, 31-55. COLLINSON, J.D. (1986) Alluvial sediments. In: Sedimen tary Environments and Facies (Ed. Reading, H.G.). pp. 20-62. Blackwell Scientific Publications, Oxford. DERos, L.F., MORAD,S. & PAlM,P.S.G. (1994) The role of detrital composition and climate on the diagenetic evolution of continental molasse: evidence from the Cambro-Ordovician Guaritas sequence, southern Brazil Sediment. Geo/., 92, 197-228. DNPM/CPRM (1987) Cachoeira do Sui- F6lha SH. 22-Y-A. Projeto mapas metalogeneticos e de previsao de recursos minerais, texto e mapas. Escala 1:250000. Brasflia. MME/DNPM/CPRM, 13 pp. DoE, T.W. & Dorr, R.H.,J R. (1980) Genetic significance of deformed cross bedding- with examples from the Navajo and Weber sandstones of Utah. J. sediment. Petrol., 50(3), 793-812. EYNON, G. & WALKER, R.G. (1974) Facies relationships in Pleistocene outwash gravels,Southern Ontario: a model for bar growth in braided rivers. Sedimentology, 21, 43-70. FAHNESTOCK, R.K. & BRADLEY, W.C. (1973) Knik and Mantanuska rivers, Alaska: a contrast in braiding. In: Fluvial Geomorphology (Ed. Morisawa, M.). State University of New York, Binghamton, Proc. 4th Geomorph. Symp., pp. 220-250. FRAGOSO-CESAR, A.R.S., LAVINA, E.L., PAlM, P.S.G. & FACC!Nl, U.F. (1984) A antefossa molassica brasiliana do cinturao Dom Feliciano no escudo do Rio Grande do Sui. Congress. Bras. Geol., 33 (Anais 7), 3272-3283. FRAGOSO-CESAR, A.R.S., MACHADO, R., SAYEG, H.S. & FAMBRJNJ, G.L. (1992) Bacias orogenicas do ciclo brasiliano no Rio Grande do Sui e Uruguai. J Work shop
s6bre
bacias
moltissicas
brasilianas,
Boletim
pp. 47-53. FRIEND, P.F., SLATER, M.J. & WILLIAMS, R.C. (1979) Vertical and lateral building of river sandstone bodies, Ebro Basin,Spain. J. geo/. Soc. London, 136,39-46. GALAY, V.J. & NEILL, C.R. (1967) Discussion of 'No menclature for bed forms in alluvial channels'. J. Hydr. Div. A mer. Soc. Civ. Eng. , 93, 130-133. HARMS, J.C.,SOUTHARD, J.B., SPEARING, D.R. & WALKER, R.G. (1975) Depositional Environments as Interpreted Especial Resumos Expandidos,
from Primary Sedimentary Structures and Stratification
Society of Economic Paleontologists and Mineralogists, Tulsa,Short Course 2. HElN, F.J. & WALKER, R.G. (1977) Bar evolution and development of stratification in the gravelly braided Kicking Horse River, British Columbia. Can. J. Earth Sci. , 14, 562-570. HEWARD, A.P. (1978) Alluvial fan sequence and mega sequence models: with examples from Westphalian Sequences.
16
P.S.G. Paim
D- Stephanian B coalfields, Northern Spain. In: Fluvial (Ed. Miall, A.D.). Can. Soc. petrol. Geol.,Calgary, Memoir 5,597-604. HooKE, R. L E B. (1967) Process and arid-region alluvial fans. J. Geol., 75, 438-460. JACKSON, R.G., II (1975) Hierarchical attributes and a unifying model of bed forms composed of cohesionless material and produced by shearing flow. Geol. Soc. Am. Bull., 86,1523-1533. losT,H. (1984) Sedimentacao e vulcanismo durante o ciclo Brasiliano no Rio Grande do Sui: Uma revisao. Congres. Bras. Geol., 33, 3421-3457. LAVINA, E.L., FACC!Nl, U.F., PAlM, P.S.G. & FRAGOSO CESAR, A.R.S. (1985) Ambientes de sedimenta�ao da Bacia do Camaqua, RS. Acta geol. Leopold., 21, 185-227. McKEE, E.D., CROSBY, E.J. & BERYHILL, H.L. (1967) Flood deposits, Bijou Creek, June (1965) J. sediment. Petrol. , 37, 829-851. MIALL, A.D. (1974) Palaeocurrent analysis of alluvial sedi ments: a discussion of directional variance and vector magnitude. J. sediment. Petrol. , 44(4), 1174-1185. MIALL, A.D. (1977) A review of the braided-river depo sitional environment. Earth Sci. Rev. , 13, 1-62. MIALL, A.D. (1978) Lithofacies types and vertical profile models in braided river deposits: a summary. In: Fluvial Sedimentology (Ed. Miall, A.D.). Can. Soc. petrol. Geol., Calgary, Memoir 5, 597-604. MIALL, A.D. (1985) Architectural-element analysis: a new method of facies analysis applied to fluvial deposits. Earth Sci. Rev. , 22, 261-308. MIALL, A.D. (1988) Architectural elements and bounding surfaces in fluvial deposits: anatomy of the Kayenta Formation (Lower Jurassic), Southwest Colorado. Sedi ment. Geol., 55,233-262. MIDDLETON, L.T. & TR UJ IL LO,A.P. (1984) Sedimentology and depositional setting of the Upper Proterozoic Scan lan Conglomerate, Central Arizona. In: Sedimentology of Gravels and Conglomerates (Eds Koster, E.H. & Sedimentology
Steel, R.J.). Can. Soc. petrol. Geol.,Calgary, Memoir 10, 189-201. OwEN, G. (1987) Deformation process in unconsolidated sands. In: Deformation of Sediments and Sedimentaty Rocks (Eds Jones, M.E. & Preston, R.M.F.). Geol. Soc. London, Spec. Pub!., No. 29, 11-24. Blackwell Scientific Publications, Oxford. PAlM, P.S.G. (1994) Depositional Systems and Palaeogeo graphical Evolution of the Camaqua and Santa Barbara Basins, Brazil. Unpub. D.Phil. thesis, University of Oxford,277 pp. PuNT, A. G. (1983) Sandy fluvial point bar sediments from the Middle Eocene of Dorset, England. In: Modern and Ancient Fluvial Systems (Eds Collinson, J.D. & Lewin,J.). Spec. Pub!. int. Ass. Sediment. No. 6, 1933. Blackwell Scientific Publications, Oxford. RusT,B.R. (1972) Structure and process in a braided river. Sedimentology, 18,221-245. RusT, B.R. (1978) Depositional model for braided alluv ium. In: Fluvial Sedimentology (Ed. Miall, A. D.). Can. Soc. petrol. Geol., Calgary, Memoir 5, 605-625. RusT, B.R. & KosTER, E.H. (1984) Coarse alluvial deposits. In: Facies Models (Ed. Walker,R.G.). Geosci. Can., Reprint Ser. 1, 71-89. SMITH, N.D. (1970) The braided stream depositional environment: comparison of the Platte River with some Silurian clastic rocks,North Central Appalachians. Geol. Soc. Am. Bull. , 81,2993-3014. SOLIANI, E., JR.,FRAGOSO-CESAR, A. R.S.,TEIXEIRA, W. & KAWASHITA, K. (1984) Panorama geocronol6gico da por�ao meridional do escudo Atlantica. Congress. Bras. Geol., 33, 2435-2449. WALKER, R.G. & CANT, D.J. (1984) Sandy fluvial systems. In: Facies Models (Ed. Walker, R.G.). Geosci. Can., Reprint Ser. 1,71-89 (2nd Edn). WILLIAMS, G.E. (1971) Flood deposits of the sand-bed ephemeral streams of central Australia. Sedimentology, 17, 1-40.
Spec. Pubis int. Ass. Sediment. (1995) 22, 17-46
Sedimentology of a transgressive, estuarine sand complex: the Lower Cretaceous Woburn Sands (Lower Greensand), southern England H O W A R D D . J O H N S O N* and B R U C E K . L E V E L L t
* Department of Geology, Imperial College of Science, Technology and Medicine, t Shell
Prince Consort Road, London SW7 2BP, UK; and Exploration and Production Ltd, Shell-Mex House, Strand, London WC2R ODX, UK
UK
ABSTRACT
A sedimentological investigation of the Lower Cretaceous Woburn Sands of southern England has been used to develop a depositional model for a transgressive estuarine (or embayment) sand complex. The Woburn Sands average 70 m in thickness, but are over 100 m thick in places, and infill a NE-SW trending trough 25-30 km wide, which cuts into the western end of the NW-SE trending London Brabant land mass. Initially, this trough was of limited extent to the northeast but opened out into a broader shallow sea to the south and southwest. Subsequently, during the course of the major Early Cretaceous (Aptian-Albian) transgression, this feature formed a seaway connecting two previously separate basins (the North Sea Basin to the north and the Weald-Wessex-Channel Basin to the south) . The Woburn Sands record this overall transgressive history i n the form of six main facies bodies, which occur in five erosionally based units (from bottom to top) : (i) Orange and Heterolithic Sands; (ii) Silver Sands; (iii) Silty Beds; (iv) Red Sands; and (vi) Transition Series. The lowermost deposits (the Orange and Heterolithic Sands; equivalent to the Lower W oburn Sands) display convincing evidence of tidal current deposition (e.g. bimodal-bipolar palaeocurrent patterns, herringbone cross-bedding, clay drapes and wavy- flaser bedding) . They are further character ized by large-scale, subhorizontal and low-angle erosion surfaces, which are interpreted as tidal channel bases and tidal shoal accretion surfaces, respectively. The overlying sand deposits (the Silver and Red Sands; equivalent to the Upper Woburn Sands) display similar evidence of tidal current activity but are distinguished by overall coarser grain sizes, better sorting, lack of clay layers and the abundance of large-scale cross-bedding. The large-scale structures in the Silver and Red Sands dip mainly towards the south (inferred ebb direction), whereas similar structures in the exposed Orange Sands dip mainly to the northwest (inferred flood direction) . The overall sequence is interpreted in terms of a transgressive tide-dominated estuary or embayment model. The Lower Woburn Sands (Orange and Heterolithic Sands) were deposited in m utually evasive ebb and flood tidal channels and intervening tidal shoals, probably in an inner estuarine environment . In contrast, the higher energy Silver and Red Sands were deposited in the outer reaches of an estuary or embayment where greater water depths allowed the build-up of large-scale bedforms. The southward increase in both cross-bed set size and sand-body thickness in the Red Sands probably reflects general southward deepening. The final element in this facies succession is the draping of the sand complex by slowly deposited fossiliferous marine beds (Transition Series and Basal Beds of the Gault), which are overstepped to the north by the shallow-marine muds of the Gault. The description, interpretation and depositional model outlined here may assist in the recognition and prediction of similar shallow marine sand bodies. This study demonstrates, for example, that thick, high reservoir-quality sands with favourable geometries for stratigraphical traps can accumulate in transgressive estuaries and embayments. The resulting sand complex could be expected to comprise several erosively bounded, lenticular units displaying rapid lateral thickness and facies variations and an upward increase in reservoir quality and sand continuity. This would contrast with the more tabular geometry and more gradual lateral thickness and facies variations of similar tide-dominated deposits developed in offshore/shelf environments. Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
17
18
H. D. Johnson and B.K. Levell INTRODUCTION
Shallow-m arine sands (i.e those deposited in w ater depths of 10-200 m and rangi n g from inshore/sub tidal to offshore/neritic environmen ts) h ave received less attention th an most other cl astic deposits and , despite recent adv ances, current facies models remain rel atively generalized. Recent studies demonstrate that sh allow-marine sands occur in a v ariety of settings and owe their variability to several factors, p articularly the complex relationships between fluctu ations in w ater depth , subsidence rates, morphology of the coastal zone, sediment supply and the hydraulic regime of the b asin, includin g the shoreline (e . g. Swift & Thorne, 1991) . Most studies of shallow-marine sand bodies distin guish those resulting from tide-domin ated processes and those formed m ainly by w ave- and storm-domin ated processes (Johnson & B aldwin , 1986; Dalrymple, 1992; W alker & Plint, 1992). In this context the Lower Cretaceous Woburn S ands of southern E n gl and h as been quoted repe atedly as a prime example of an ancient tide-domin ated shallow-marine sand complex, with the spectacul ar l arge-scale cross-beddin g interpreted as the deposits of tid al sand w aves (e . g. Ben tley , 1970; de R aaf & Boersma, 1 97 1; Walker, 1984; Buck , 1985 ; Dalrymple, 1992). However, there h as been less agreement on the specific type of tid al environment preserved in the Woburn S ands and on its overall genetic evolution. Because the resultin g product is a thick sand of high reservoir quality, i t w as felt th at a better understanding of the Woburn S ands would assist in the development of strati graphical models of tid al sand bodies. This could aid their recognition in the subsurface , allow comparison with the geo metry and intern al ch ar acteristics of other sh allow m arine s and bodies (e. g. Exum & H arms, 1968; McCubbin, 1969; Campbell, 1971; Spearin g, 1 975) and contribute to the development of depositional models for hydrocarbon exploration and production. This w as the b ackground to a field study conducted in 197 9 when both authors were employed by the Koninklijke/Shell Exploratie en Produktie L ab oratorium (KSEPL) in Rijswijk , The Netherlands. Subsequently, the results were presented at the 1 980 Annu al Conference of the AAPG (Johnson & Levell, 1980) and documented in an internal report in 1 982. This l atter report forms the b asis for this p aper, which is presented here for several re asons. First, the model presented here is different in several respects to those interpretations published
both before and since completion of our work. Secondly, the 1 980 Abstract is cle arly an i n adequate reference document, but is nevertheless quoted by workers studying these exposures. Thirdly, these exposures comprise sand quarries, which are constantly changing as a result of con tinuing s and extraction. Hence, new observations are frequently m ade and so the d at a acquired during the course of this study, despite the time l apse , are still worthy of fuller documentation. Publishin g the results of this study will also allow our interpretation to be more critically eval u ated and will form a more l astin g contribution to the analysis of this important geological unit. The m ain aim of this p aper, therefore , is to provide a sedimentological description and to argue a deposi tional model for the Lower Cretaceous (Apti an- Albi an) Woburn S ands at Lei ghton Buzzard, southern E n gl and (Fi gs 1 & 2). It is not our intention to comprehensively evalu ate our findin gs in the context of more recent rese arch undertaken on the Woburn S ands, p artly because much of this remains unpublished, p articularly a detailed an alysis of sedimen tary structures of Buck ( 1987) , a litho strati graphical study by Eyers ( 1992 a) and a recent sequence strati graphical analysis by Won h am ( 1993).
GEOLOGICAL FRAMEWORK Stratigraphic framework
The Woburn S ands h ave been shown throu gh field m appin g and borehole evidence to comprise a lens shaped sand complex up to c. 100 m thick (Fig. 3 ) , which in fills a rel atively n arrow (25 -3 0 km wide) N E - SW trendin g trough (Bristow, 1963; Wy att et a/., 1986). The trough, which m ay be p artly tec tonic in ori gin (Eyers, 1 991) , came into existence in the L ate J ur assic, but w as infilled only in the E arly Cre taceous. At Lei ghton Buzzard, which is on the western m argin of the trough, the upper Aptian to lower Alb i an Woburn S ands unconformably overlie Upper Jurassic cl ays and are overstepped northwards by the Albian G ault (Fig. 4). Ammonites and brachiopods found in the b asal phosphatic gr avels of the Woburn S ands are assi gned to the upper Apti an ( nutfieldiensis zone; Casey, 196 1) . Althou gh the Woburn S ands currently exposed contain an abundan t and diverse ichno-
Transgressive estuarine sand complex
19
0'
2'
- Aptian-Albian Lower Greensand outcrop distribution
0
50
100km
,
LHTON BUZZARD
I
l. Location map showing the distribution of Lower Cretaceous outcrops.
Fig.
fauna they do not contain a shelly fauna, possibly because of leaching of calcium carbonate. The over l ying beds be long to the Transition Series, which comprises a thin ( 1 -2 m) , complex and relatively poorly exposed succession of variable lithologies, including the Shenle y Limestone and various iron cemented beds referred to locall y as 'Carstone ' ( Fi g. 3). The Shenle y Limestone contains lower Albian (tardefurcata to mammillatum zone) fauna and represents a depositional hiatus with a complex depositional and diagenetic history (Eyers, 1992b). The condensed lower Albian Gau lt (dentatus zone) represents a northward overstep, which can be related to the ' 108 Ma' maximum floodin g surface of Haq et a!. ( 1987). The paucity of datable fauna within the Woburn Sands essentially precludes further correlation wi th events elsewhere in the Lower Greensand basin (Ruffell & Wach, 199 1 , i n press). Palaeogeographical setting and depositional environments
Duri n g
the
Early
Cretaceous
( Ryazanian
to
Barremian) the London- Brabant land mass formed an intermi ttent land barrier between a shallow marine southern North Sea Basin (the Boreal Sea) to the north, and a freshwater Wealden Basin to the south (Fi g. 5A) . This southern basin, to ge ther with the Channel, Southwestern Approaches, Celtic Sea and Bristol Channel Basins ( Ziegler, 1988, 1990) , formed a series of mainly separate and active, fault bounded basins in the southern British Isles , which underwent a c. 40-million-year period of alluvial sedimentation and the deposition of 'Weald en ' facies ( P. Allen, 198 1 ) . This period o f non-marine sedimentation was terminated by the major Aptian-Albian marine trans gression and the deposition of the Lower Greensand Group. This was mainly a consequence of continued sea-floor spreading and northward extension of the Atlantic Ocean and the Early Cretaceous eustatic sea-level rise. This resulted in progressive, northeastward marine inundation of the Southwestern Approaches, Channel and Weald Basins. Simultaneously, there was also marine northwestward-directed transgression through the Paris Basin into sou thern England and
20
H. D. Johnson and B.K. Levell
@
f!7' Jane's Pit
Heath and Reach
A
@
Old { Llnslade Road
0
Three main periods of marine transgression are recorded in the Lower Greensand of southern England, each associated with depos it ion of exten s ively cross-bedded t idal deposits (Bridges, 1982). The upper Apt ian to lower Alb ian Woburn Sands are associated w ith the transgressive breaching of the London-Brabant land ma�s (Figs SB & SC) . However, although the Woburn Sands clearly represent, in broad terms, a transgressive shallow marine sand deposit , more precise environmental interpretations have remained uncertain. The large scale cross-bedding, extensive bioturbation and evi dence of reversin g currents has led most authors to su ggest a t idal environment (Lamplugh, 1922; Schwarzacher, 1953; Bentley, 1970; de Raaf & Boersma, 197 1 ) , but these authors d isagree , or are non-committal, about specific t idal subenviron ments, with su ggestions ranging from open shelf to tidal fiat for different facies in the complex. From these possibilities, two main deposit ional models emerge for all or part of the Woburn Sands: 1 t idal shelf or seaway, such as the present-day Straits of Dover/En gl ish Channel (e. g. Brid ges, 1982 ) ; 2 tidal estuary o r embayment (e. g. Johnson & Levell, 1980) . These alternatives w ill be considered here in the light of our observations. Similarly, the uncerta inty as to whether the breakthrou gh across the London Brabant land mass (the 'Bedforsh ire Strait' of K irkaldy ( 1939)) occurred as a result of southward or northward extension of a coastal embayment will also be considered.
1km
Lithostratigraphical subdivision and relationships Fig.
2. Location of the sand pits used in this study and the
line of the cross-section illustrated in Fig. 4.
southerly expansion throughout the southern part of the North Sea Basin and adjacent areas (e . g. West Netherlands, Broad Fourteens and Lower Saxony Basins; Ze igler, 1990) . The precise t iming of full marine connection between these basins is uncertain, but most recent reconstructions show this to have occurred by the late Aptian or early Alb ian, at around the time of deposit ion of the Woburn Sands and the overlyin g Gault. Hence , on a regional scale, the depositional h istory of the Woburn Sands would appear to be related to marine transgression that resulted in the connection of two main intracratonic basins.
cThe Woburn Sands comprise up to six facies bodies, which are readily d istinguished on the basis of their l ithofac ies characteristics (mainly grain size, com posit ion , colour, clay content, sed imentary struc tures and bioturbation) . Most of these facies bodies are separated by major subhorizontal to low-an gle erosion surfaces, and five discontinuity-bounded units have been defined as follows (Table 1 & Fi g. 3 ) : 1 Transition Series (you n gest) ; 2 Red Sands;
Fig. 3. (Opposite.) Composite vertical section through the Woburn Sands.
LITHOSTRATIGRAPHIC NOMENCLATURE
-�--'·-·
-�------,
ENVIRONMENTAL SUMMARY
0- ·-:__-_c
"' :0
:;;: :;;
-' 3:
0
------
.. 0 ..
,., :; "' (.')
:;
:;
"'
(.')
- . "Dentatus "--
Cl
�����
Blanket shelf muds
t:=== = -:::::-=---
Carstone and Shanley Limestone
-
BURROW TYPES
u
-
10 Transition Series
u c
"'
rocn
(f) u
u
a:
�� 0
--- -- · "'"'''
\���-
Red Sands
''- -- ...Y.C
<1>
20�"'
en�
- _____/Slow deposition/reworked, transgressive e:i3> deposits
"����
"'
<1>
0:<1>
u
--
"'
�u
:=C
en"' (f)
Silly Beds
�\�� ''\�
� . ''''"
,,,,,
''-'-'-'->--."."''-':-
c
"'
5
.0"'
�-g �"'
Q)(f)
a. a.
=>
u c
"'
Silver Sands
(f)
:;;
.2 Ui
30-
"'
..: :;;
a. a. =>
"'
u c
"'
(/)
E
40 -
::l .0 0
Heterolithic Sands
3:
--
"'
u c "' (f) c
5
.0 0
3: :;;
-'
�� ------�-
-s==: �L
-
�-
�
.. ...
----� ---� !"- .,_.
-- -- ----
High energy, ebb dominated channel! shoal complex (estuary mouth or open marine, sea strait environment)
(:==.:r> =
==M /Low energy, estuary abandonment! transgressive deposits
J High energy, ebb dominated, estuary mouth channel/shoal complex (= ebb-tidal delta environment)
�-
c
"li
� -� ��
=
= =M
= = = =
Low-moderate energy, estuary shoal deposits
=
�
�--.1!....,
"' u c
50 -
"'
� ..}.\.._ �
(f) c 3:
-
e
en
�-
�
�� ��
..... � -:::'{J! � � � /// !))})
3:
0
�
·-··--�
Orange Sands 60-
.__
�
Moderate-high energy, flood dominated, channel-fill sands intercalated with tidal shoal deposits
�--��---""!.
? 70 -
Large-scale cross-bedding Trough cross-bedding Ripple cross-lamination Intraclasts
Basal transgressive deposits with phosphate nodules and reworked faunas
��[�
UPPER JURASSIC
J:?/-::-1 B. Q D
== =
. EJ c=J 1:-�;1 E;J '
Wavy bedding
� Strongly bioturbated
Flaser bedding
6:11 Shells and shell debris
Low-angle erosion surfaces
Q Plant debris
Concretions/nodules
-& Occasional burrows
H. D. Johnson and B.K. Levell
22
Table 1. Summary of the l ithofacies and reservoir characteristics of the main units within the Woburn Sands Interval
Lithology
Gault Clay
Grey fossiliferous claystones
Transition Series
Iron-cemented pebbly sands (basal beds); glauconitic & phosphatic fine-coarse, partly argillaceous Transgressive lag deposits. sands. In-situ lenses of richly fossiliferous limestone (Shenley Lmst.). Reworked clasts of iron-cemented sst. (Carstone) & Shenley Lmst. Rapid lateral lithological variations.
Muddy shelf.
Three main types of cross-bedding: I
Red Sands
Med.-v.coarse sand Mod.-poorly sorted. Ferruginous with up to 20% bv detrital iron oxide (red colouration) up to 2.5% bv heavy minerals 100% sand.
Silty Beds (/) Q z <( (/) z a:
::;:)
Ill 0 Silver
;:
Dep. environment
Physical sedimentary structures
Sands
"//!/1/1/1
0.-":
11
l
��T �1 %
III
;;
�:JtYti,;;tff
J
Giant cross-bedding with avalanche foresets & infilling, large scours ca 3m deep and 1OOm wide. Wedge-shaped sets (-2-3m thick) superimposed on low-angle (4-8°), S-dipping surfaces.
-�-am0.1-4m thick tabular & J trough cross-bedding.
Grey-green glauconitic & lignitic Coarse sands have flat bases, large rippled or flat surfaces & internally cross-bedded or horizontally laminated. clays, silts & f. sands. Minor crs., well sorted sand Fine sands occasionally show low-angle to horizontal lamination but mainly layers & lenses. bioturbated. -20-40% sand. Med.-v.crs. sand. Well-sorted. Quartz arenites. Minor carbonaceous debris. Locally Fa-cemented e.g. clay clasts on erosion surfaces). 100% sand.
��/-71"lr ��: +m Northern area:
(J.t,t,l,ffZt:;Z,'
1
'L1 1 n-:
Southern area:
Large-scale, low-angle (2-4°) surfaces separated m by o.5-2m thick, tabular cross-bed sets. 1-3m avalanche-type cross-bedding, partly filling urs. Complex low-angle s
���
Main structures (in order of decreasing importance) current ripple Numerous thin clay cross-lamination, trough cross-bedding, scour & fill structures & low-angle layers. Intraclasts of cross lamination. Herringbone cross-bed patterns. Abundant clay drapes clay & carbonac. produce wavy & !laser bedding. Large-scale low-angle surfaces (dipping debris. -90% sand. -40)
3
Fine to crs. mod. sorted sands. Scattered quartz granules, clay pebbles & clay drapes. Iron oxide cement in liesegang rings & around clay deposits. -95-100% sand.
Large-scale subhorizontal (1) & low-angle (2) erosion surfaces.
1
��
=
2
D
/////////////////////I
I
�
Large avalanche foresets fill deep scours (1-5m thick). Low angle surfaces separated by cross-laminated, cross-bedded & bioturbated sands. Flaser & wavy bedding/clay drapes.
S ilty Beds; Silver S ands; Oran ge S ands and the Heterolithic S ands ( Lower Woburn S ands or Brown S ands). This inform al scheme generally follows that of Ben tley ( 1970) and is read il y correl ated w ith the
4 5
Transgressive or local abandonment deposit.
Variety of large-scale cross-bedding:
Heterolithic Fine to v.fine mod. Sands sorted sands.
Orange Sands
High-energy, ebb-dominated complex (estuary/embayment mouth or open marine sea strait).
High-energy ebb-dominated embayment mouth channel-shoal complex (cf ebb-tidal deltas)
Moderate-energy, tidal shoal deposits within an inner estuarine/inner embayment ?margmal to tidal channel complex (=Oranae Sands). High-energy flood-dominated tidal channel-fill sands with intercalated moderate-to high-energy tidal shoal deposits.
term inology of prev ious workers (Table 1 & Fig. 3 ) , inclu d in g the schemes o f W yatt e t al. ( 1 986) and the more recent form al l ithostratigraphy of Sheph ard Thorn et a/. ( 1986) . The vertical and l ater al rel ationsh ips of these units are summar ized in Figs 3 & 4, respectively. The b ase
Transgressive estuarine sand complex
23
Table 1. (Continued. ) Fauna and biogenic sedimentary structures
�'it�fe':.�ent SEAL
Ammonites, belemnites, bivalves, brachiopods.
g
POOR to V-POOR -partly sealing due to cementing & argill. content
Abundant ammonites, bivalves, belemnites, astropods & oysters. Partly reworked & phosphatized. henley Lmst.= brachipods, echinoids & crustacea.
No preserved fauna (?leached) 2 main types of bioturbation:
Reservoir characteristics
Pal
Unimodal to the S-SSE GOOD Very locally reduced by minor Fe cementation. No shale layers.
Minor reversals II Funnel to v-shaped burrows due to vertical animal escape or sediment collapse/inti II. Large burrows (1O's & herringbone patterns mm wide, up to -1OOmm high) caused by large bivalve or crustacaean.
No distinct burrow types.
Unfossiliferous. Negligible bioturbation-rare single clay-lined burrows (Ophiomorpha) towards the top in some places (e.g. New Trees).
Unfossiliferous (?leached). Extensively bioturbated (ca. 10-50% of primary structures destroyed). Horizontal, slightly sinuous, clay-lined burrows are the most common type & occur mainly in cross-laminated sands. Occasional vertical to oblique burrows.
Unfossiliferous (?leached). Moderately to strongly bioturbated (30-50% of primary structures destroyed) & variety of burrow types: (i) narrow vertical tubes, (ii) sinuous subhorizontal burrows producing colour mottling. (iii) iron-cemented vertical to steeply inclined burrows, (iv) subhorizontal branching burrows Thalassinoides. (v) v-shaped burrows.
-?Om max. 1-2m
0->15m
I Intense, small-scale colour mottling (5mm diam.) -pale core & darker rim. Caused by horizontal burrows & resulting in negligible destratificatation.
No fauna observed. Strongly bioturbated throughout- fine sediments effectively destratified.
Thickness
(?up 1o10's m max.)
Bimodal-bipolar S-SW modes are dominant (directions of all major structures)
Bimodal-bipolar WSW mode dominant & ENE mode slightly subordinate
Bimodal-bipolar NW mode dominant with minor S-SSW mode
of the lowest unit, the Orange and Heterolithic Sands, was not seen but is thought to directly overlie the phosphati c gravels and sands of the fossiliferous basal beds re corded in abandoned pits ( Lamplugh, 1922). The relationship between these two lower most sands has also not been observed directly, with
-thin permeable sands probably laterally extensive (=Storm layers)
VERY GOOD vertically & laterally uniform
MODERATE to POOR -discontinuous shale layers
1-2m
Sheet-like/tabular
E-W lenticular geometry in N
Large-scale southward thickening wedge in S.
g
-drapes irre . surface of ilver Sand. -dissected cut-out by erosional base of the Red Beds.
2-15m
Tabular within study area.
<25m
Uncertain-restricted to E part of study area. Possibly interfinger to W with Orange Sands.
up to -50m
Uncertain, greater N-S continuity ct. E-W.
MODERATE to GOOD -distinct higher permeability zones within N-S trending channels.
Sheet-like
Lenticular
POOR/SEAL Isolated structures show dips to S
Geometry
either lateral interfingering or erosional contact both possible. However, facies similarities (discussed later) suggest that the Orange and Heterolithic Sands are probably lateral equivalents. The boundary between the top of the Woburn Sands and the overlying Gault is generally poorly
24
H. D. Johnson and B. K. Levell
D Gault Cia'{ D Transition Series [(g);j Red Sands (RS)
�j:j:j:j Silty Beds
- Silver Sands (SS) � Heterolithic Sands (HS) EO(,<�J Orange Sands (OS) North D Jurassic clays ..,.
0
2km
w
--
Approx. depth of present-day exposures
0 10 E: "
20 )§
0
30
10
·t <.) Q)
40.0 :: 0 50 a; .0
20 30
60 �
Q)
70::2:
40
80
50 60
Fig. 4. Cross-section through the Woburn Sands illustrating the vertical and lateral relationships between the main lithostratigraphical units (sec Fig. 2 for location) . The locations at which some of the main lithostratigraphical boundaries can be seen arc indicated by single vertical lines. Note the dashed line indicating the approximate depth of present-day exposures. Datum is the base of the cristatum subzone.
exposed and has not been studied here in any detail . However, this impo rtant and richly fossilife ro us inte rval has been studied e xtensively in the past by palaeontologists and biostratigraphe rs, who have meas ured many detailed vertical profiles (e.g. Lamplugh, 1 922; Wright & W right, 19 47; Casey, 19 61 ; Owen, 1 9 72). These data have been incor porated into Table 1 and Figs 3 & 4.
SEDIMENTOLOGICAL CHARACTERISTICS
This section outlines in detail the sedimentological cha racteristics of the six main facies types. The key points of description and interp retation are summarized in Table l .
Transgressive estuarine sand complex A
25
B
BOREAL BOREAL
SEA
SEA
:r:
,'
�
Main palaeocurrent directions 50 100km �---
5. Three schematic palaeogeographical maps illustrating the transgressive history of the Lower Cretaceous in the southern North Sea- English Channel area, and the evolution of the 'Bedfordshire Strait' which ultimately connected the Boreal Sea and the Wealden Basin. Aptian-Albian outcrop shown in black (a) Ryazanian-Valanginian; (b) Aptian: Woburn sands, Folkcstonc sands, Hythe and Sandgate beds; (c) Albian, Gault clay. (Based on Ziegler, 1988, 1990.) Fig.
'----'"'---..J100km
H. D . Johnson and B. K. Levell
26 Orange Sands
D_escription
Based on the available exposures at the time of our field-work in 1 9 79 (B ryant's Lane, Stone Lane and Sheepcott quarries; Fig. 2), these sands a re moder ately sorted, fine- to coa rse-grained and contain some quartz granules, clay pebbles and wood frag ments. The o range colour is due to widespread iron oxide, whi ch o ccurs as a cement, in Liesegang rings and in rims around clay (e.g. clay pebbles, drapes and burrow linings). Large-scale erosion surfaces within the unit have been divided a rbitra rily into two types:
NE
I Subhorizontal erosion surfaces a re essentially flat and extend up to 200 m. Lo cally they cut down in con cave-upward scours 4- 6 m deep ( Fig. 6) . The e rosion surfaces are spa ced 5 - 10 m apart verti cally and a re normally overlain by coa rse lags of granules and mud flakes. 2 Low-angle erosion surfaces o ccur within the units bounded by e rosion surfaces of type I and pass laterally and down-dip into the subhorizontal e rosion surfa ces ( Fig. 6) . They are spaced at intervals of a few decimetres to 1 m and separate inte rvals with a variable array of cross-bedding, cross-lamination and burrows ( Fig. 7) . The more deeply e rosive parts of the subhorizontal erosion surfaces a re overlain by 1 -5 m thick tabular
f<---- Flood-dominated tidal channel
Tidal bar -----+1
SW
metres 0
8 .
Transport toN
Photograph location
+----- 150metres--------•
6. An example of large-scale facies relationships in a flood tidal-channel complex in the Orange Sands. The base of the channel is a horizontal erosion surface lined with intraformational clay clasts. A series of low-angle (4-8°) erosion surfaces (right side of the photograph) mark the flanks of a tidal bar, which is characterized by small-scale cross-bedding and moderate to strong bioturbation . The inclinations of the low-angle erosion surfaces (inferred bar flank surfaces) increase laterally (to the left) and eventually pass into high-energy, channel-fill deposits displaying tabular avalanche foresets up to 4 m high. A second flood-tidal channel sequence is also exposed in the lower part of the section. The simplified field sketch (sec Fig. 3 for legend) shows the broader relations between the bioturbated tidal bar sands and the avalanche foresets of flood-dominated, tidal channel-fill sands (from Bryant's Lane pit) .
Fig.
Transgressive estuarine sand complex
27
7 . Physical and biogenic sedimentary structures i n the Orange Sands. (A) Tidal bar deposits comprising small- to moderate-scale cross-bedding (10-SOcm thick) separated by horizontal and low-angle erosion surfaces with thin intraformational mud-flake conglomerates, and occasional clay layers. The numerous low-angle reactivation surfaces give a characteristic wedge-shaped appearance to the cross-bed sets (see also ( D ) ) . ( B ) Close-up of the central part of (A) illustrating some details of the bioturbation. Note in particular the simple vertical burrows and a large V-shapcd burrow (lower centre of photo) . (C) Large-scale tabular cross-bedding infilling a flood tidal channel (upper half of photo) . Vertical burrows increase in density in the deeper part o f the channel and arc inclined perpendicular t o the forcscts. The underlying deposits display oppositely-dipping cross-bedding, horizontal and inclined erosion surfaces and moderate bioturbation. (D) Wedge-shaped cross-bedding with numerous reactivation surfaces, which arc occasionally overlain by clay drapes or oppositely-dipping cross-lamination. Fig.
or wedge-shaped c ross-bedding with NE-dipping avalanche foresets separated by large-scale , low angle erosion surfaces (Fig. 6) . These low-angle e rosion surfaces may flatten up-dip into more closely spaced, subhorizontal e rosion surfaces. This is ac companied by a change in sedimentary structures from 1-5-m-thick avalanche cross-bedding to 0.20. 7-m-thick sets of t rough cross-bedding, cu rrent ripple cross-lamination and flaser and wavy bedding. Palaeocurrent directions from all these deposits a re variable, but there is abundant evidence of reve rsals , especially in the smaller-scale structures. The larger structures show mainly northwest-flowing palaeo cu rrents but with clear, subordinate reve rsals (Fig. 8).
Burrowing has destroyed , on average, some 3050% of the p rimary structures and has been sub divided into five types: 1 Na rrow ( c. 2 mm) vertical tubes that form a branching network with sections 10 -20 mm long. These tubes have no clay lining and a re extremely fragile, being visible only on wind-sculpted faces. They resemble burrows p roduced by polycheate worms in modern sands of estuaries and tidal flats (Schafer, 1 9 72 ) . 2 Sinuous, subho rizontal burrows p roducing c. 5-mm-diameter colour mottling. These burrows occur mainly in the ripple-laminated sands. 3 Simple , ve rtical or steeply inclined tubes ( c. lO mm diameter and 50-200 mm long) with clay linings,
H. D. Johnson and B.K. Levell
28
N
I
N
I
Red Sands
n=89
Silver Sands
n=157
Interpretation
The interbedding of deposits wi th opposed palaeo current modes, evidence of rapid lateral variation in flow regime (as shown by the intercalation of large and small-scale structures), the range of burrow types , and the clay drapes, are all common character istics of high-energy, shallow-water tidal deposits (de Raaf & Boersma , 197 1 ; N io & Yang, 1 99 1 ) . The subhorizontal erosion surfaces, therefore, probably define the bases of tidal channels and, at least in the areas of deepest scour, these channels carried northwestward flowing water. The low-angle erosion surfaces are interpre ted as the accretionary flanks of in-channel bars on which curren t dominance was less pronounced and low-energy structures were preserved (e.g. Yang & N io, 1989 ) . There is no evidence that the preserved portions of these bars were e ither emergent or suffered severe wave activity . The Orange Sands thus represent a h igh energy, tidal channel complex w ith mutually evasive ebb and flood tidal currents. Heterolithic Sands
Description
Orange Sands
n=203
Heterolithic Sands (scale x2)
n=52
8. Palaeocurrent distributions based mainly on l arge scale cross-bedding.
Fig.
wh ich are frequently the sites of iron-oxide precipi tation (Fig. 7A & B) . Th is type projects normal to the bedding even when this is inclined (Fig. 7 C) , and is w idespread throughout the Orange Sands. 4 Complex, subhorizontal to inclined, branching burrow networks c. 10-40 mm in diameter and w ith enlarged, bulbous junctions (Fig. 7D). Iron cementation preserves these in three dimensions. This type most closely resembles the crustacean burrow system Ophiomorpha. 5 Nested cone-shaped burrows (V-shaped in two dimensions, Fig. 7 B) . These appear to represent the collapse of sedimentary lamination i n to 20-30-mm w ide horizontal tubes, but may also occur as iron cemented , V-shaped laminae. They are especially common in sands just above and below major erosion surfaces.
This unit comprises moderately sorted, fine to very fine grained sands w ith numerous clay layers, scattered clay flakes and woody detritus. The sands are largely grey, while the clay drapes and surround ing sands are some times rust-coloured due to iron oxides. The sands contain low-angle ( c. 4 °) erosion sur faces tens of metres long that closely resemble those of the Orange Sands, with the exception that they do not pass downwards into channel-fill facies and are often overlain by relatively continuous clay drapes. The main sedimentary s tructures are, in order of decreasing importance: current ripple cross lamination (Fig. 9A) , trough cross-bedding, scour and-fill s tructures and low-angle cross-lamination. Clay drapes occur on set boundaries and foresets w ith in all types of cross-stratification and sometimes produce wavy and flaser bedding (Fig. 9A & B). Although not measured in detail , clay drape d istr i bution is suggestive of tidal bundles, possibly w ith neap-spring tide cycles (e.g. Visser, 1980). Evidence of bidirectional currents is ubiquitous in all these s tructures (e.g. Fig. 9A) . The larger foresets commonly have superimposed smaller sets w ith reversed dips . Cross-stratification type varies over
Transgressive estuarine sand complex
29
Fig. 9. Physical sedimentary structures in the Heterolithic Sands. (A) Flaser bedding associated with small-scale, current ripple cross lamination. Herringbone patterns are occasionally developed but normally the southwest ebb direction is dominant. (B) Clay draped foresets separated by thick clay layers, which are internally disrupted by bioturbation.
short distances, both laterally and vertically, and there are no progressive vertical changes in either set thickness or grain size. Palaeocurren ts are bimodal-bipolar, with a dominant southwest directed mode (Fig. 8). Burrowing, which has destroyed around 10-50% of the primary sedimentary fabric, is dominated by sinuous, horizontal, clay-lined forms (Fig. 10). Burrows are most common in the flaser and wavy bedded subfacies and may be virtually absent in the larger decimetre-scale cross-bed sets. The pre dominance of horizon tal burrows results in little disturbance of the sedimentary structures. Less common types are vertical to oblique and rare , spiral clay-lined burrows (Fig. lOD).
Interpretation
The bimodal-bipolar palaeocurrent pattern , and the assemblage and variability of sedimentary structures in this unit suggest a shallow-marine tidal origin . The thickness o f this facies (up t o 25 m was recorded by Bentley ( 1970)) and the lack of features rep resenting emergence (rootlets, wave-reworked sur faces, desiccation cracks, etc . ) suggest deposition in a subtidal environment of moderate water depth and fluctuating flow conditions. The lack of distinct channel-fill facies and the ubiqui tous presence of low-angle inclined erosion surfaces wi th a constan t south-eastward dip suggests accretion on a broad subtidal shoal. Relatively low-energy curren ts are
30
H.D. Johnson and B.K. Levell
Fig. 10. Biogenic sedimentary structures i n the Heterolithic Sands. (A) Strong bioturbation i n current ripple cross laminated and partly ftaser bedded sands. The dominant biogenic structures are horizontal, clay-lined burrows. (B) Moderate bioturbation mainly by horizontal, clay-lined burrows with a single inclined burrow (southeast of lens cap) . Bidirectional, cross-lamination with occasional clay-ftasers is still well-preserved. (C) Plan view of the dominant burrow type in this facies comprising horizontal, sinuous, clay-lined burrows with back-fill laminae. (D) Isolated example of a vertical, clay-lined burrow resembling Ophiomorpha. Background facies is ripple laminated, ftaser bedded sand.
suggested by the small scale of cross-bedding and the predominance of r ipple cross-lamination , whereas the extensive clay drapes suggest relatively long periods of quiet water conditions. Silver Sands
Description
The Silver Sands consist of well-sorted , medium- to very coarse-grained quartz arenites (previously used as glass sands ) . Sooty and woody carbonaceous matter is locally abundant, but clay drapes are absent. These sands truncate all earlier deposits w ith a major planar to regionally concave -upwards erosion surface which is lined with granules and clay flakes. The lag deposit overlying this erosion surface is well cemented by iron oxides.
In the northwest of the area (around Heath and Reach, Fig. 2) several pits expose up to 20 m of S ilver Sands, and major low-angle (2-4° apparent dips) erosion surfaces w ith a constant southwestward dip can be p icked out throughout the u n it (Fig. llA ) . Between these planar to slightly undulose erosion surfaces tabular cross-bed sets from 0.5 to 2 m thick occur (Fig. 1 1B ) . The erosion surfaces terminate abruptly down-dip, resulting in thickening of some cross-bed sets, and formation of hanging set bound aries. No s ingle surface could be traced from the top of the 1 5 -m -thick unit to the base. There is only occasional evidence of reversing palaeo currents, such as at the base of the unit in Munday's H ill quarry, resulting in an overwhelmingly domi nant southwestward dip to all scales of forese t (Figs 8 & 1 1) . I n the extreme north o f the area o f the Silver
Transgressive estuarine sand complex
31
Fig. 11. Physical sedimentary structures i n the Silver Sands from the northern part of the study area. (A) A complete vertical section through the Silver Sands ( c. 15 m thick) illustrating a sequence of moderate- to l arge-scale cross-bedding (set thickness c. 0.5-2.0 m ) separated by undulatory horizontal to low-angle erosion surfaces (dipping to right ) . Cross bedding is mainly undirectional to the southwest (ebb-sets) . A prominent erosion surface lined with intraformational clay clasts and quartz granules forms a ledge j ust below the contact between the dark and light coloured sands. The dark sands represent ebb-dominated channel-fill sands within the upper part of the Orange Sands. (B) Closer view of the southwest directed, ebb-dominated Silver Sands (quarry face c. 10 m high ) . Note the undulatory nature of the erosion surfaces and corresponding irregularities in cross-bed set thickness (c. 0.5-2 m ) . There are no positive signs of reversing currents in these ebb-dominated sands. The ledges produced along the various erosion surfaces are not lined with clay; these sections are in c. 100% quartz sand.
Sands are somewhat thinner and more iron-stained. Cross-bedding is predominantly of the avalanche type in se ts 1 - 2 m thick . Three-dimensional ex posures show these sets to be the infills of spoon shaped scours with the appearance of micro-de ltas, rather than the deposits of migrating large-scale megaripp les (or sand waves) with flat bases. In the Shenley Hill area further to the east (Fig. 2), tabular sets 2-4 m thick comprise the bulk of the Silver Sands (Fig. 1 2 ) . These were relative ly poorly exposed during our fie ld -work but were photographed by Bentley ( 1970) and have been
re-exposed recently ( J . Eyers, 1994 pers. comm . ) . Internally these sets show complex lateral transitions from single avalanche se ts, which are a lways directed to the south or southwest (Fig. 8) , and which pass into thinner tabular or wedge-shaped sets of up- and down-slope dipping foresets, separated by low-angle or subhorizontal erosion surfaces (Fig. 13). These sets have only been seen in sections paralle l to the transport direction , and their strike configuration is unknown. Bentley ( 1970) claimed to be able to map the E -W trend of individua l sand-wave crests in this area.
V.> N
::t: \:::)
0' �
a
Cl
;: ., ;: "' t:l;)
;>;: !:""'
"' "' "' :::::
Fig. 12. Large-scale, sand-wave-type cross-bedding in the Silver Sands (Munday ' s Hill Quarry) . Individual sets are up to 4 m thick and are internally complex, ranging from avalanche foresets to low-angle surfaces (6-8°) separated by upslope- and downslope-dipping cross-bedding (see Fig. 13 for details) . Note also the sharp, planar contact between the Silver Sands and Silty Beds and the erosional contact between the Silver and Red Sands. The dark area along this contact represents a ridge of strong iron cementation ('Carstone rib ' ) . Finally the sequence is capped by the marine Gault Clay.
Transgressive estuarine sand complex
33
Fig. 13. Internal characteristics of the sand-wave-type cross-bedding in the Silver Sands in the Munday 's Hill- Double Arches area (large-scale structures illustrated in Fig. 12). (A) Tabular cross-bedding with simple, angle-of-repose foresets. Note small oppositely-dipping set (by lens cap) separating the main , thick sets and a prominent reactivation surface in the upper thick set (Double Arches Quarry) . (B) Relatively small tabular cross-beds, with each displaying an upward decrease in grain size (Munday's Hill Quarry) . (C) The low-angle surfaces dipping to the left represent an ebb directed (to southwest) sand-wave lee face . Internally these coarse to granule grade sands display upslope dipping cross bedding which represents flood-directed megaripples (Munday ' s Hill Quarry ) . (D) Herringbone patterns developed in relatively small sets of trough cross-bedding (Munday's Hill Quarry ) .
In the extreme west of the area, west of Heath and Reach (Sheepcott quarry; Fig. 2) , a heterolithic facies of decimetre-scale cross-bedded sands with persistent clay drapes overlies the Orange Sands. These beds, formally assigned to the Lower Woburn Sands, could be a lateral facies variant of the Silver Sands. In general, the Silver Sands are weakly bio turbated, apart from the upper few metres in some pits (e.g. New Trees) , where clay-lined Ophiomorpha-type burrows occur. Interpretation
The textural and mineralogical maturity of the Silver Sands, the uniformly large size of cross-bedding and
the lack or scarcity of clay drapes and burrows suggest deposition in a higher energy environment than both the Heterolithic and Orange Sands. None the less, the relatively infrequent evidence of revers ing palaeocurrents still suggests tidal deposition. The extensive subhorizontal erosion surface at the base of the unit and the evidence of a bedform complex up to 1 5 m thick, at least in the Heath and Reach area, suggest deposition in a high-energy, current-dominated bar system after a period of wide spread erosion. There is no clear channel-fill facies and it is possible that deposition occurred in an environment in which interbar depressions were created by the accretion of shoals rather than the active cutting of channels. In more detail, the cross-bedding indicates large ,
34
H. D. Johnson and B.K. Levell
possibly sand-wave-type bedforms with super imposed megaripples over the majority of the out crop area, and microdelta-like forms infilling scour hollows in the north. Taken together with a similar change in cross-bedding style in the younger Red Sands (see below ) , this change could be related to a southward increase in water depth across the study area (Fig. 2 ) , which allowed the construction of larger bedforms in the south of the area. In comparison with the shoal complex of the Heterolithic Sands, current energy was apparently much higher and there was less time for the settling of suspended clay. The textural and mineralogical maturity of the sands also indicates a more intensive phase of reworking prior to final deposition. Silty Beds
Description
The Silty Beds (Lamplugh & Walker, 1903; Lamplugh , 1922) comprises a thin interval (mainly 1 - 2 m ; maximum 3 .6 m) of grey-green, glauconite and carbonaceous-rich clays, silts and argillaceous fine-grained sands with rare thin beds of coarse- to very coarse-grained, well-sorted sands. The fine grained sediments are strongly (90- 100 % ) bio turbated (Fig. 14) and have in places suffered soft-sediment deformation. The coarse beds are 50- 1 00 mm thick , have sharp bases and flat or rippled tops, and are internally cross-laminated or plane-laminated. The unit sharply overlies a slightly undulating surface on top of the Silver Sands (Fig. 12). This undulose surface was considered by Lamplugh ( 1922) and Bentley ( 1970) to represent the preserved morphology of shoal or sand-wave crests in the underlying Silver Sands. The Silty Beds are sharply overlain by either the Transition Series or the Red Sands (Fig. 4). Interpretation
This interval represents a period of slow sedimen tation in a marine environment in which both current and wave energy were apparently too weak to cause significant erosion of the unconsolidated sands on the sea-floor. The coarser grained sand layers are interpreted as lags, resulting from reworking and winnowing of the Silver Sands (cf. Levell, 1980). The Silty Beds are interpreted to represent a phase of low energy and low sediment influx (an estuarine abandonment phase) relative to the high-energy,
14. Silty Beds, overlying the Silver Sands with a sharp, planar contact (next to trowel ) . The mottled texture in the Silty Beds is the .result of extreme bioturbation. A low-angle laminated sand bed with an erosional base (storm layer) is present in the upper part of the sequence (Bryant's Lane Quarry) .
Fig.
subtidal sand shoal and channel complexes described from the underlying deposits. The precise depositional environment of this unit is debatable. The lack of marine fauna and abundant carbonaceous material could indicate restricted marine conditions, possibly within a deeper, pro tected part of an estuary ( Lamplugh, 1922). The facies could also be interpreted as tidal flat deposits (Eyers, 1992b) , although the absence of tidal channels and the characteristic fining upward sequence capped by in situ marsh deposits (e.g. Evans, 1965) fails to support this. Alternatively, the abrupt cessation of high-energy conditions marked by the sharp base of this unit, and the presence of reworked and glauconite-rich horizons within it, is more characteristic of a sudden deepening, rather
Transgressive estuarine sand complex
than shallowing in water depth. Hence , a more offshore shallow marine environmen t is a further possibility . The Silty Beds probably represen t a relatively long time interval, which marks the end of the underlying Orange- Heteroli thic and Silver Sands depositional succession. This i n terpreta tion implies that the overlying Red Sands form a genetically separate sand body. Red Sands .
Description
The unit comprises moderately to poorly sorted, medium- to very coarse-grained sands (occasionally gravelly), which are distinguished from all older units of the Woburn Sands by their high conte n t (up to 20% bulk volume) of detri tal iron oxide in the form of rust-coloured ooliths of goeth i te and angular ir9nstone chips. It is also rich in heavy minerals ( Bentley, ( 1970) reports up to 2 . 5 % bulk volume) . I n addition, diagene tic iron oxide occurs a s a local cement around mud clasts and on certain foresets. The Red Sands erosionally truncate all earlier deposits. In the Shenley Hill area they were shown by Bentley ( 1970) to occupy E- W trending shallow scours up to 6 m deep and 100 m wide. South of this area the Silty Beds and the Silver Sands are progressively cu t out by erosion beneath the sou th ward thickening Red Sands until, south of Leighton Buzzard, the sand pits expose only this last unit (Fig. 4), which is presumed to overlie Orange and/ or Heterolithic Sands. The erosion surface at the base of the Red Sands is i n terpreted as a sequence boundary by Ruffell and Wach (in press). Three main types of cross-bedding characterize the Red Sands. Type I consists of thick (up to 5 . 5 m) sets of avalanche cross-bedding that overlie prominent erosional scours up to 3 m deep and more than 100 m wide . The overall set geome try thus resembles giant scale trough cross-bedding. The uppermost se t of the unit is overlain by a coset of successively smaller se ts with a particularly i n triguing geometry (Fig. 15). The thick set of avalanche foresets at the base dip at 25 -30° and abut sharply against a basal erosion surface locally lined with granules and pebbles. Major reactivation surfaces periodically separate bundles of foreset laminae with an average thickness of 130 mm (range 50-220 mm) . Each of
35
these tabular or wedge-shaped bundles is in turn subdivided by some times i n tersecting minor dis continuity planes that define the individual foresets. The major reactivation surfaces are convex-up, flattening toward the top of the set, where they are separated by cosets or sol i tary sets of decimetre scale cross-bedding. At the top of the sets, where the reactivation surfaces flatten out, there are some up-dip thinning sets 0.05 - 0.2 m thick of oppositely dipping cross-bedding. This geometry closely resembles, on a large scale, the ebb and flood shields described from several modern tidal environments (e.g. Klein , 1970) . Type ll is characterized by wedge-shaped cross-sets (0.3- 2 m thick) with sou th- or southwest-dipping foresets, which are superimposed on 2-4-m-thick sets of low-angle southward-dipping erosion surfaces (dips 4-8°) . Individual cross-sets taper either up or down the inclined surfaces. Small sets (less than 0.5 m thick) occasionally thicken down-dip into single large-scale ( c. 2-3 m thick) avalanche sets. Type II cross-bedding passes laterally i n to type I in both down- and up-palaeocurren t direction . Type III comprises 0. 1 -0.4-m-thick sets o f trough and tabular cross-bedding bounded by subhorizontal erosion surfaces, and forms the bulk of the Red Sands in the channelized facies of the northern area.
Additional information on the complex and varied geometry of these sedimen tary structures has been obtained recently from ground-penetrating radar studies ( Bristow, in press) . The Red Sands are extensively bioturbated (Fig. 16). The most widespread burrow type is a c. 5-mm diameter colour mottling in which pale-coloured burrow fills contrast with the darker iron-rich sands surrounding them (Fig. 16A) . This structure results from horizontal burrowing and has caused only minimal disruption of the primary stratification. Such bioturbation is apparently similar to that termed 'cryptobioturbation' by Howard & Frey ( 1975) . This structure was found by them in the G eorgia (USA) estuaries and a ttributed by them to amphipod crustaceans. The colour mottling could be due either to a slight grain-size fractionation ( the iron pellets fall mostly in the fine- to very fine-grained sand classes ( Bentley, 1970) or to some oxidation/ reduction difference caused by organic slime. The second and most distinctive burrow type is the nested inverted cone or funnel-shaped struc-
36
H. D. Johnson and B. K. Levell
Fig. 15. Large-scale (c. 3-4 m thick ) , ebb-directed (to south) , avalanche foresets and associated top set deposits.,.(A) Foreset packets are separated by major reactivation surfaces. The latter are ascribed to stronger than normal flood-currents because they are overlain by packets of oppositely dipping, wedge-shaped cross-beds. The large-scale structure is interpreted as an ebb-dominated bedform which filled an erosional hollow. (B) Close-up of central part of (a) showing oppositely-dipping (to northeast), flood-directed sets, possibly analogous to flood-shield bedforms. Locality: Pratts' Lane Quarry.
ture (Fig. 1 6 B) . The V-shaped laminae (in two dimensions) of these structures could have formed either by sediment collapse i n to the underlying horizontal tubes, which are sometimes associated with these structures, or could have formed as an escape structure by upward movement of an animal (e.g. large pelecypod or crustacean). The escape s tructure interpretation may be supported by the association of this sort of burrow with large erosion surfaces overlain by migrating bedforms. Perhaps the animals preferred the relatively slow deposition areas of channel floors but were none the less able to escape by efficient vertical burrowing when buried beneath advancing bedforms.
A third burrow type consists of horizontal 40-mm diameter subhorizontal tubes with miniscus-shaped spreiten. The size, shape and back-filling spreiten resemble the 'press structures' produced by £chino cardium cordatum ( Reineck & Singh , 1973). If these burrows are echinoid burrows, they would indicate fully marine conditions. Interpretation
The Red Sands accumulated in a high-energy marine or marginal marine environment dominated by southward flowing, but sometimes reversing, cur rents (Fig. 8) . There is no evidence of distinct
Transgressive estuarine sand complex
37
Fig. 16. Biogenic sedimentary structures in the Red Sands. (A) Typical mottled appearance of the Red Sands resulting from horizontal burrows. Note the alternation of non-bioturbated laminae and the retention of structure within the bioturbated layers. Locality: Pratt's Lane. (B) Two V-shaped burrows occurring within a large-scale cross bedded sand body at the contact between two different sediment types (coarser, massive sands above) . The lower cross-bedded sands also display the mottled texture detailed in (a) . Locality: Grovebury South.
channel-fill or shoal deposits, and the whole uni t seems to represent rather a blanke t-like field of megaripples and megarippled sand waves (cf. J . R.L. Allen, 1980) , which increased systematically in height southwards. The sedimentary s tructures closely resemble tidal sand waves found in modern subtidal environments (e.g. Berne et a!. , 199 1 ) . There i s no evidence of emergence o r s ignificant wave activity , but the b ioturbation suggests similar marine or marginal marine cond itions to the under lying units. The lack of clay drapes indicates a more or less constantly active water column .
Transition Series
Description
The Transition Series comprises a d istinctive thin (c. 1 -2 m thick) , and laterally extensive deposit at the junction between the previously described units and the overlying Gault. I t has been studied inten sively by palaeontologists because of its rich fauna (for more extensive descriptions see Lamplugh ( 1922 ) , Wright & Wright ( 1947 ) , Hancock ( 1958) , Casey ( 1961 ) and Owen ( 1972)) . Lithologically the Transition Series is laterally variable , consisting mainly of partly cemented pebbly
38
H. D. Johnson and B.K. Levell
sands (at the base) and argillaceous, glauconitic and phosphatic fine- to coarse-grained sands and sandy clays. A distinctive and unusual limestone, the Shenley Limestone, occurs as discontinuous in situ bands, up to c. 0.6 m thick , and as reworked frag ments in the lower part of this unit. This limestone, which sharply overlies the underlying deposits, contains a unique fauna, including rich and dis tinctive brachiopods and a unique assemblage of echinoids and Crustacea. The whole Transition Series is highly fossiliferous, with abundant ammon ites, bivalves, belemnites, gastropods and oysters. There is abundant evidence of reworking and dis continuous sedimentation in the form of reworked faunas and several condensed horizons. The Tran sition Series forms a laterally extensive drape over the underlying Woburn Sands and extends laterally over Jurassic sediments (Owen, 1972) . Interpretation
The Transition Series represents a composite trans gressive , shallow-marine lag, which forms a type of abandonment deposit on top of the underlying, mainly high-energy, subtidal deposits. The rework ing of limestone beds, concretions and iron cemented sandstones ( ' Carstone'), mixing of faunal zones and condensed intervals indicate slow depo sition and strong current activity. The abundant and highly varied fauna supports a shallow marine, probably offshore, depositional environment, with the coarser grained layers representing winnowed lag deposits.
PALAEOGEOGRAPHICAL EVOLUTION
Reconstruction of the palaeogeographical evolution of the five erosionally bounded units depends not only on the process interpretations given above but also on the overall palaeogeographical setting and the stratigraphical relationships between the units. The first point to establish is the direction of the transgression . Several factors lead us to propose a northward transgression . 1 The affinities of the fauna described by Casey ( 196 1 ) from the lower part of the complex are with faunas from the southern Wealden Basin, rather than from the northern Southern North Sea Basin. 2 In tidal embayments it is common for ebb-directed sets to be preferentially preserved, either because ebb flows are often enhanced by freshwater run-off
or because these flows reach maximum velocities at times of lower water, and are thus preferentially preserved in the topographically lower parts of the deposit. The predominance of southwestward directed palaeocurrents leads to the proposal of open sea in the southwest. 3 The increase in set size southwards in the Red Sands (and possibly in the Silver Sands) , coupled with a southward increase in the thickness of the Red Sands, could be explained by greater water depth in the south. Water depth is, of course , of especial importance in the preservation of sets up to 5 . 5 m thick. In modern tidal basins (e.g. estuaries and embayments) the largest bedforms frequently occur in the deeper areas (see Fig. 17). 4 The geometry of the deposit is one of southward dipping facies units (Fig. 4) . This is most readily explained by a northward transgression . Two factors lead us to suggest that the Red Sands should be considered separately from the older sub divisions of the Woburn Sands. 1 They rest erosively on the Silty Beds, which appears to be a slowly deposited unit representing the cessation of sand supply to the underlying Silver Sands shoal system. 2 Their distinctive petrography, rich in detrital iron oxide fragments with goethite ooliths, suggests a change in the source material. The most likely possibility is the erosion of presently unexposed iron-rich intervals derived from further updip to the north , which may include the iron-stained Orange Sands, the carstones of the Silver Sands or some other, older or laterally equivalent formation(s) (e.g. the early Cretaceous Carstone of the Norfolk area, Casey & Gallois, 1 973). The only reported occurrence of similar goethite ooliths is from an early Cretaceous oolith-rich sand unit further north (Deepdale Pit), which may be equivalent to the Silver Sands (Eyers, 1992a) . The deposition of the Woburn Sands began with the partly erosional and partly tectonic creation of a trough trending NE-SW through the Bedfordshire area. The basal deposits are gravels and phosphates with rolled, derived faunas and abundant wood material, suggesting transgressive coastal erosion (whether open-marine or on the margins of a brackish embayment is unknown). The overlying Orange and Heterolithic Sands represent subtidal channel and shoal deposits, respectively. The relationship between these two units could be either lateral interfingering or a sharp erosional contact (Fig. 17). l n the first case
Transgressive estuarine sand complex
39
�� Tidal channel fills I . I High-energy tidal shoals C:=J Moderate-energy tidal shoals
}
}
Orange Sands Heterolithic Sands
b) Model 1
� "'A :, ::-:=--. ·:-::::_� -:-::=· ::- :: -_-.-_. -.::-:.: ::���-::-.� �;:.;_ -::-::-.·-. .�- �=._ � :s_ �:::. _.:.:.:::= � �;=-; ;�;� CHANNEL
. __ .•.•.. •
::::
HIGH ENERGY CHANNE HIGH ENERGY L SHOAL SHO L - -. . . ::::::::::-::=·. .� : - .:0 - _ - =-== = . '\ . ;- _
MODERATE ENERGY SHOAL
== -==;::;===;;:::_:::: ----== _ :: :; :
B
0
?
c) Model 2
A
5m B
17. (a) Distribution of the Orange and Heterolithic Sands (Lower Woburn Sands or Brown Sands) below the Silver Sands erosion surface, and (b) two possible models depicting possible lateral relationships (model 1 preferred) .
Fig.
40
H. D. Johnson and B.K. Levell
(Fig. 17b) , the Heterolithic Sands would be situated towards the centre of the estuary or embayment, with the Orange Sands being a marginal , flood dominated , channel complex. Such a palaeo geography with marginal channels hugging the edges of an embayment or estuary would be consistent with the generally rotary circulation models for estuaries, where tidal currents are deflected by Coriolis force to the margins (e.g. Schubel , 197 1 ) . Nevertheless, a problem remains with this inter pretation in that it is difficult to understand why the Heterolithic Sands shoal complex accreted steadily to a thickness of 25 m , without being dissected by a migrating channel system. However, the similarity between the Heterolithic and Orange Sands leads us to infer a laterally interfingering relationship (Fig. 17b). Two features place these two units in the relatively landward and protected parts of the embayment or estuary, rather than on the open shelf. 1 The clear presence of distinct channel and shoal complexes is characteristic of the estuary proper, rather than the seaward extensions. Furthermore the flood dominance of the Orange Sands channels indicates mutually evasive ebb and flood channels. 2 The well-developed clay drapes indicate periods of slack water in a generally current-dominated regime. Whether these are diurnal drapes or not, periods of tidal slack water are more likely to occur in systems with rectilinear tides. Such tides are more common in the inner and more restricted parts of estuaries, whereas in estuary mouths and on the open shelf, tides tend to be rotary (Terwindt, 1973). The major erosion surface at the base of the Silver Sands indicates a shift in depositional setting, and is explained most readily by the lateral shift of a major, wide, estuary mouth channel system. The overlying Silver Sands, with the lack of a distinct channel-fill facies and the large-scale bedform complexes up to 15 m high , suggest, in the context of a continuing northward transgression, deposition in estuary mouth subtidal shoals. Modern analogues might be the outer reaches of Chesapeake Bay ( Ludwick 1970, 1974; Colman eta/. , 1988), or Delaware Bay ( Knebel et al . , 1988) , or the seaward portions of ebb-tidal deltas (Greer, 1975 ; Hayes, 1975 ) . A s mentioned above , the Silty Beds represent the cessation of sand supply to the Silver Sands shoal, followed by a period of slow deposition. Reasons for the abandonment of this shoal could be autocyclic (e.g. a shift in the shoal pattern related to the movement of main tidal channels) or allocyclic (e.g.
sea-level rise leading to the area becoming an open marine shelf rather than an estuary mouth. However, alternative interpretations consider this interval to represent tidal flat deposits and hence a relative shallowing of sea-level of the facies succession has been proposed (Eyers, 1992b) . In gross facies terms (e.g. the style of cross bedding, grain size and lack of clay drapes) the Red Sands closely resemble the Silver Sands. However, there is a major difference in their petrography. The Red Sands could, therefore , represent a similar outer estuarine shoal complex but formed at a time when the up-dip tidal channels were eroding different material. They would therefore represent the shift of tidal channels to debouch again in the Leighton Buzzard Area. An alternative and more dramatic interpretation would relate the Red Sands to the final breaching of the London-Brabant land mass and thus to continued coastal erosion at the margin of the transgressing embayment. As discussed above, the erosion surface at the base of the Red Sands has been interpreted as evidence for a sea-level low stand (sequence boundary) punctuating the overall late Aptian transgression. This explanation would presumably equate the depositional settings of the Silver Sands and Red Sands, with the Silty Beds representing either the most distal (offshore) or most proximal (tidal flat) depositional settings. On the grounds of the facies differences we have described , and of simplicity, we prefer to view the succession as one of continued overall transgression and attribute the baSal erosion surface of the Red Sands to changes in tidal dynamics, possibly associ ated with the breaching of the Lori don - Brabant land mass. The Transition Series marks the final abandon ment of the high-energy, tide-dominated sanely shoal complexes. Sedimentation was slow, discontinuous and interspersed with periods of physical reworking. Subsequent basin deepening resulted in the wide spread deposition of shallow-marine/shelf muds (the Gault) .
DEPOSITIONAL MODEL FOR A TRANSGRESSI V E ESTUARINE - EMBAYMENT SYSTEM
A generalized depositional model can be constructed for a transgressive estuarine -embayment system based on observations from both the Woburn Sands and from known facies and bedform distributions
Transgressive estuarine sand complex
within modern meso- and macrotidal estuaries and embayments (Fig. 18). This model also can be compared with estuarine facies models developed mainly from a synthesis of modern environments and from conceptual considerations (Dalrymple et al., 1 99 1 ) . The essential elements o f the inner estuarine environment are as follows. 1 Discrete , mutually evasive, ebb and flood tidal channels and interchannel tidal shoals. 2 Tidal channel sands will be relatively coarse-grained and characterized by large-scale cross-bedding. 3 Tidal shoals will consist of more heterolithic sands, generally with smaller scale structures, stronger bio turbation and more frequent and laterally extensive clay layers than the higher energy channel-fill deposits.
b)
41
4
Reservoir quality and heterogeneity will be strongly influenced by the contrast between channel and shoal deposits. In general, the channel fills may be expected to form discrete zones of higher per meability (Fig. 18b), whereas the tidal shoals will comprise variable but generally lower quality and more heterogeneous reservoirs ( c.f. Orange and Heterolithic Sands). The essential features of the outer estuarine environ ment are as follows. 1 Minor lithological differentiation between tidal channel and shoal deposits. 2 Well-sorted sands with fewer clay intercalations and less bioturbation compared with the inner estuarine sands. 3 Large-scale cross-bedding, partly reflecting large estuary mouth bars and sand waves, and also
Tidal shoals
'
3
a)
Basic Characteristics
Woburn S. Equiv.
Fossiliferous clays
Gault Clay
Reworked, fossiliferous, partly cemented sands
TRANSGRESSIVE LAG
Transition Series
Well-sorted sands. Large-scale, complex cross-bedding. Ebb-dominated palaeocurrent directions. Mainly minor bioturbuation.
OUTER ESTUARINE/ EMBAYMENT TIDAL SHOALS
Silver Sands and Red Sands
Pre-transgressive deposits
� l
Upper Jurassic Clays
18. A model for a transgressive estuarine- embayment depositional system (see Fig. 3 for legend). (a) Idealized vertical section through a transgressive estuarine-embayment complex based on the Woburn Sands. (b) Block diagram illustrating some of the essential differences in sand body characteristics between the inner estuarine/embaymer.t and outer estuarine/embayment environments. (c) Idealized section through a transgressive estuarine- embayment indicating the potential stratigraphic trap geometry.
Fig.
H. D. Johnson and B.K. Levell
42
the infilling 0f deep ebb-dominated channels and interbar troughs. 4 Possibly an overall seaward i ncrease in the size of cross-bedding in response to deeper-water conditions. 5 Higher quality and more homogeneous reservoir characteristics compared with the inner estuarine sands as a result of both stronger tidal currents and increased wave activity. Shale layers are likely to be very thin and discontinuous, or absent. The vertical sequence through this type of trans gressive sand complex would tend to have lower reservoir quality sands in the lower part and higher reservoir quality sands towards the top (Fig. 1 8a) . The transgressive nature of the sequence would lead ultimately to the blanketing of the overall lenticular sand-body complex by shelf muds and could thereby, form a stratigraphical trap (Fig. 1 8c) .
DISCU S S ION
The difficulty of postulating a generalized facies model for this type of succession is that these deposits are influenced by a particularly wide range of physical processes, both autocyclic and allocyclic; they may receive sediment supplied from different sources (landward and/or seaward; Schubel, 197 1 ) and their preservation potential i s linked closely to the rate and sense of relative sea-level fluctuations. These variables form the basis of Dalrymple et al. 's ( 199 1 ) classification of estuarine facies, in which they argue that these deposits are diagnostic of transgressive conditions. In this scheme the Woburn Sands would be classified as a tide-dominated estuary in which fluvial processes were negligible . Compar able modern environments in terms of physical processes, facies and sedimentary structures would include the southwest Netherlands area (e.g. Haringvliet and Oosterschelde Estuaries) (Oomkens & Terwindt, 1960; Terwindt, 1973; van den Berg et al. , 1980; Yang & Nio, 1989 ) , Severn Estuary/ Bristol Channel area, (Hamilton, 1979; Harris & Collins, 1985 ; J . R . L. Allen, 1 990) , Ossabaw Sound (Greer, 1975 ) , Jade Estuary/German B ight ( Reineck, 1963; Reineck & Singh, 1973) and Broad Sound, Australia (Cook & Mayo, 1977). The stratigraphical relationships recorded in the Woburn Sands and summarized in the depositonal model allows comparison with other stratigraphical models associated with tide-dominated estuaries, incised valleys and transgressive systems tracts , and
allows consideration of their sequence stratigraphical implications (e.g. van Wagoner et at. , 1990) . The major third-order lowstand and maximum flooding events would be recorded respectively by the basal lag deposit of the Woburn Sands and by the condensed horizons in the Lower Gault Clay. We have seen no direct (i.e. preserved) evidence of fluvial incision at the base of the Woburn Sands, which comprises a typical shallow-marine lag deposit. The non-preservation of predicted lowstand fluvial deposits among shallow marine sandstones successions is commonly argued to be the result of either (i) minor lowstand fluvial deposition, due to a predominance of fluvial bypassing during incision, and/or (ii) extensive shallow marine reworking during subsequent relative sea-level rise. In many cases this combination results in the apparent merging of lowstand and initial marine flooding sur faces, and removal of all lowstand fluvial deposits (e.g. Bergman & Walker, 1987; Plint & Walker, 1987; Plint & Norris, 1 99 1 ; Thorne & Swift, 199 1 ; Walker & Bergman, 1992). The greatest amount of such reworking is likely to occur in tide-dominated settings, particularly in confined areas such as estuaries or embayments, due to high-energy tidal currents and their ability to incise deeply into the substrate. Holocene incised valleys on the inner continental shelf of southeastern USA , for example, show a hierarchy of erosional events that reflect not only earlier fluvial incision but later deep tidal scour (c. 15 -30 m ) , particularly at the confluence of tidal streams, in tidal channels associated with tidal flats, in tidal inlets and in channels associated with ebb tidal delta shoals (Oertel et a/. , 199 1 ) . Preservation potential o f transgressive estuarine successions is relatively high due to their relatively protected position within palaeovalleys (Swift et a/. , 1980; Demarest & Kraft , 1987). The degree of preservation is related most closely to the depth of incision of the basal erosion surface , or sequence bounding fluvial incision, plus any additional down·· cutting by tidal scour associated with the initial flooding surface and , to a lesser extent, the degree of erosion associated with the ravinement surface (Dalrymple, 1992). The latter is equivalent to the ravinement/shoreface erosion surface of wave dominated shorelines , but in tide-dominated systems it is characterized by tidal current erosion (e.g. the bedload parting zones of tidal seas; Stride, 1982 ) . Complete preservation of an idealized tide dominated transgressive stratigraphy, which seems to be extremely rare, would predict tidal shelf sand
Transgressive estuarine sand complex
deposits (e.g. tidal sand ridges and tidal sheet sands with sand waves and associated deposits; Stride, 1982; Nummedal & Swift, 1987; Dalrymple, 1992) overlying the ravinement surface . All the main sand bodies in the Woburn Sands appear to occur below the transgressive marine erosion surface and hence are assigned to the estuarine part of the sequence. Shelf sand wave deposits may occur further to the south and southeast within the Folkestone Beds (Narayan, 1 97 1 ; J . R. L . Allen, 1 982; Bridges, 1982). The main difference between tidal shelf sands and tidal estuarine sands is that the latter will be domi nated by complex interfingering of tidal-channel, tidal-shoal and, in places, tidal-flat deposits (cf. Maguregui & Tyler, 1 99 1 ) . Thus in the Woburn Sands, concave-upward erosion surfaces and rela tively rapid lateral thickness changes are common. In more basinward settings individual sand bodies are likely to have tabular geometries, with individual sand and shale layers having much higher continuity (cf. Surlyk and Noe-Nygaard, 1 99 1 ) . Distinguishing between these different types of shallow-marine sand body is particularly important in evaluating the nature of such deposits in the subsurface because of their different geometries and lateral extent (e.g. Banerjee, 199 1 ; Brownridge & Moslow, 199 1 ; Jenette et a!. , 1992).
ACKNOW L E DGEMENTS
We would like to extend our sincerest thanks to Harold Reading for giving us our initial introduction to the Woburn Sands and for making us aware of its sedimentological significance. The basis for this work, its methodology and evaluation stems directly from the lessons learnt by the authors under Harold's previous supervision. We are extremely grateful for this and for his ongoing interest and influence in our work. The authors would also like to thank the Konin klijke/Shell Exploratie and Produktie Laboratorium, Riswijk, The Netherlands, for supporting both the project and its publication. We also acknowledge Shell Research B . V. and Shell Internationale Pet roleum Maatschappij B . V . for their permission to publish this paper. We are also indebted to Guy Plint for his patience and encouragement during the preparation of this paper. The assistance and openness of the two referees, Jill Eyers and Alistair Ruffell , is also gratefully acknowledged. Finally, Ian Glenister is thanked for redrafting the figures.
43 REFERENCES
J . R . L . ( 1 980) Sand waves: a model of origin and internal structures. Sediment. Geol. , 26, 281 - 328. ALLEN, J . R . L . ( 1 982) Mud drapes in sand wave deposits: physical model with application to the Folkestone Beds (Early Cretaceous, southeast England ) . Proc. R. Soc. London, Ser. A, 306, 291 -345 . ALLEN, J . R. L . ( 1990) The Severn Estuary in southwest Britain: its retreat under marine transgression, and fine sediment regime. Sediment. Geol. , 66, 13-28. ALLEN, P . ( 1981 ) Wealden research - ways ahead. Proc. Geol. Assoc. , 100, 529-564. BANERJEE, I. ( 1991 ) Tidal sand sheets of the Late Albian Joli Fou-Kiowa-Skull Creek marine transgression, Western Interior Seaway of North America. In: Clastic Tidal Sedimentology (Eds Smith, D .G . , Reinson, G . E . , Zaitlin, B .A . & Rahmani, R . A . ) . Can. Soc. petrol. Geol. , Calgary, Memoir 16, 335-347. BENTLEY, A . ( 1 970) Sedimentation studies in the Lower Greensand; a report on the sedimentology of the Aptian and lower Albian strata near Leighton Buzzard, Bedfordshire. Unpublished BSc thesis, University of Oxford, 137 pp. BERGMAN, K.M. & WA LKER , R.G. ( 1987) The importance of sea-level fluctuations in the formation of linear con glomerate bodies; Carrot Creek Member of Cardium Formation, Cretaceous Western Interior Seaway, Alberta, Canada. J. sediment. Petrol. , 57, 651 - 665 . BERNE , S . , DuRAND, J. & W EB ER , 0. ( 1991) Architecture of modern subtidal dunes (sand waves ) , Bay of Bourgneuf, France. I n : The Three-Dimensional Facies Architecture of Terrigenous Clastic Sediments and its Implication for Hydrocarbon Discovery and Recovery (Eds Miall, A . D . & Tyler, N . ) . Soc. econ. Paleontol. Mineral. , Concepts Sediment. Paleont . , Tulsa, 3 , 245 -260. BRIDGES, P . H . ( 1982) Ancient offshore tidal deposits. I n : Offshore Tidal Sands: Processes and Deposits (Ed. Stride, A . H . ) , pp. 172- 192. Chapman & Hall , London. BRJSTOW, C . R . ( 1 963) The stratigraphy and structure of the Upper Jurassic and Lower Cretaceous Rocks in the area between Aylesbury (Bucks) and Leighton Buzzard (Beds) , Unpublished PhD thesis, University of London. BRISTOW, C . S . ( 1995) Internal geometry of ancient tidal bedforms revealed using ground-penetrating radar. In: Tidal Signatures in Modern and Ancient Sediments. (Eds Fleiruning, B .W. & Bartholoma). Spec. Pubis. int. Assoc. Sediment. , No. 24. Blackwell Science, Oxford (in press) . BROWNRIDGE, S . & MOSLOW, T.F. ( 1991 ) Tidal estuary and marine facies of the Glauconitic Member, Drayton Valley, central Alberta. I n : Clastic Tidal Sedimentology (Eds Smith, D . G . , Reinson , G . E . , Zaitlin, B .A . & Rahmani, R . A . ) . Can. Soc. petrol. Geol . , Calgary, Memoir 16, 107- 122. BucK, S. ( 1985) Sand-flow cross strata in tidal sands of the Lower Greensand (early Cretaceous ) , southern England. J. sediment. Petrol. , 55, 895-906. BucK, S . ( 1987) Facies and sedimentary structures in the Folkestone Beds (Lower Greensand, Early Cretaceous) and equi valent strata in southern England. Unpublished PhD thesis, University of Reading. CAMPBELL, C.V. ( 1971) Depositional model - Upper ALLEN ,
44
H. D. Johnson and B. K. Levell
Cretaceous Gallup beach shoreline, Ship Rock area, northwestern New Mexico. J. sediment. Petrol. , 41, 395-409. CASEY, R. ( 1961 ) The stratigraphical palaeontology of the Lower Greensand. Palaeontology, 3, 486-62 1 . CASEY, R . & GALLOIS, R.W. ( 1973) The Sandringham Sands of Norfolk. Proc. Yorks. geol. Soc. , 40, 1 -22. COLMAN , S . M . , BERQUIST, C . R . , J R . & HOBBS, C.H. , I I I ( 1988) Structure, age and origin o f the bay-mouth shoal deposits, Chesapeake Bay , Virginia. Mar. Geol. , 83, 95 - 1 13 . CooK, P . J . & MAYO, W . ( 1977) Sedimentology and Holocene history of a tropical estuary (Broad Sound, Queensland) . Aust. Bur. Miner. Resow·. , Geol. Geophys. Bull. , 170, 206 pp. DALRYMPLE, R.W. ( 1992) Tidal depositional systems. In: Facies Models: Response to Sea Level Change (Eds Walker, R.G. & James, N . P . ) , pp. 195 - 218. Geological Association of Canada, St Johns. DALRYMPLE, R.W. , ZAITLIN , B . A . & B OYD , R . ( 1991 ) Estuarine facies models: conceptual basis and strati graphic implications. J. sediment. Petrol. , 62, 1 130- 1 146. DEMAREST, J .M. , II & KRAFT, J . C. ( 1 987) Stratigraphic record of Quaternary sea levels: implications for more ancient strata. In: Sea-Level Fluctuation and Coastal Evolution (Eds Nummedal, D . , Pilkey, O . H . & Howard , J . D . ) . Spec. Pub!. Soc. econ. Paleont. Miner . , Tulsa, N o . 4 1 , 223-229. DE RAAF, J . F . M . & BoERSMA, J . R . ( 197 1 ) Tidal deposits and their sedimentary structures (seven examples from Western Europe) . Geol. Mijnbouw, 50, 479 - 50 1 . EvANS, G. ( 1 965) Intertidal flat sediments and their environments of deposition in the Wash. Q. J. geol. Soc. London, 121, 209-245. Exu M , F. A. & HARMS, J .C. ( 1 968) Comparison of marine bar with valley-fill stratigraphic traps, western Nebraska . Bull. Am. Assoc. Petrol. Geol. , 52, 1 85 1 - 1868. EYERS, J. ( 199 1 ) The influence of tectonics on early Cretaceous sedimentation in Beclfordshire, England. J. geol. Soc. London , 148, 405 -414. EYERS, J . ( 1 992a) Lithostratigraphy of the Lower Greensand and Gault Clay (Lower Cretaceous) of the Bedfordshire Province, England. Unpublished PhD thesis, Open University. EYERS, J. ( 1992b) Sedimentology and palaeoenvironment of the Shenley Limestone (Albian, Lower Cretaceous): an unusual shallow-water carbonate. Proc. Geol. Assoc. , 103, 293-302. GREER, S . A . ( 1975) Estuaries of the Georgia coast USA: III. Sandbody geometry and sedimentary facies at the estuary/marine transition zone, Ossabaw Sound Georgia; a stratigraphic model. Senckenbergiana Mar. , 7, 105 - 135 . HAMILTON, D. ( 1979) The high-energy, sand and mud regime of the severn estuary, S . W . Britain. In: Tidal Power and Estuary Management (Eds Severn, R .T. , Dineley, D . L . & Hawker, L . E . ) , pp. 162- 1 72. Trans atlantic Arts Incorporated, Albuquerque, Colston Paper 30. HANCOCK, J . M . ( 1 958) The Lower Cretaceous near Leighton Buzzard. I n : The London Region Geologists Association Guide, No. 30, pp . 36-40. HAQ, B . - U . , HARDENBOL, J. & VAIL, P . R. ( 1987) Chron-
ology of fluctuating sea-levels since the Triassic (250 million years ago to the present) . Science, 235, 1 156- 1 167. HARRIS, P.T. & COLLINS, M . B . ( 1985) Bedform distribution and sediment transport paths in the Bristol Channel and Severn Estuary. Mar. Geol. , 62, 153- 166. HAYES, M .O. ( 1 975) Forms of sand accumulation in estuaries. I n : Estuarine Research , Vol . II, Geology & Engineering (eel . Cronin, L . E . ) , pp. 3 -22. Academic Press, London. HowARD, J . D . & FREY, R.W. ( 1 975) Regional animal sediment characteristics of Georgia estuaries. Senckenbergiana Mar. , 7, 33- 103. JENElTE, D .C. , JONES, C.R. , VAN WAGONER, J .C. & LARSEN , J . E . ( 1992) High resolution sequence strati graphy of the Upper Cretaceous Tocito Sandstone: the relationship between incised valleys and hydrocarbon accumulation, San Juan Basin, New Mexico. I n : Sequence Stratigraphy Applications to Shelf Sandstone Reservoirs: Outcrop to Subswface Examples (Eels van Wagoner, J.C. , Jones, C.R . , Taylor, D . R . , Nummedal, D . , Jenette, D .C. & Riley, G.W. ) . American Associ ation of Petroleum Geologists Field Conference, September 199 1 . JOHNSON , H . D . & BALDWIN , C.T. ( 1986) Shallow silici clastic seas. In: Sedimentary Environments and Facies (Ed. Reading, H . G . ) , pp. 229-282. Blackwell Scientific Publications, Oxford . JoHNSON, H . D . & LEVELL, B . K . ( 1 980) Sedimentology of a Lower Cretaceous subtidal sand complex, Woburn Sands, southern England. Bull. Am. Assoc. petrol. Geol. , 64, 728-729. KIRKALDY, J .F. ( 1939) The history of the Lower Cretaceous period in England. Proc. Geol. Assoc. , 50, 379-413. KLEI N , G . DE VRIES ( 1970) Depositional and dispersal dynamics of intertidal sand bodies. J. sediment. Petrol. , 40, 1 095 - 1 127 . KNEBEL, H . J . , FLETCHER, C . H. , III & KRAFr, J . C . ( 1 988) Late Wisconsinan - Holocene paleogeography of Dela ware Bay: a large coastal plain estuary. Mar. Geol. , 83, 1 15- 133. LAMPLUGH, G.W. ( 1922) On the j unction of Gault and Lower Greensand near Leighton B uzzard (Beclfordshire) . Q. J. geol. Soc. London, 78, 1 -80. LAMPLUG H , G.W. & WALKER, J.F. ( 1903) On a fossiliferous band at the top of the Lower Greensand near Leighton Buzzard. Q. J. geol. Soc. London , 59, 234-265 . LEVELL, B . K. ( 1 980) Evidence for currents associated with waves in Late Precambrian shelf deposits from Finnmark, North Norway. Sedimentology, 27, 153- 166. LuDWICK, J .C. ( 1 970) Sand waves and tidal channels in the entrance to Chesapeake Bay, Virginia. Va. J. Sci. , 21 ' 178- 184. LUDWICK, J .C. ( 1 974) Tidal currents and zig-zag sand shoals in a wide estuary entrance. Geo/. Soc. Am. Bull. , 85, 717-726. MAGUEREGUI, J . & TYLER, N. ( 1991 ) Evolution of Middle Eocene tide-dominated deltaic sandstones, Lagunillas field, Maracaibo Basin, Western Venezuala. In: The Three-Dimensional Facies Architecture of Terrigenous Clastic Sediments and Its Implication for Hydrocarbon Discovery and Recovery (Eels Miall, A . D . & Tyler, N . ) .
Transgressive estuarine sand complex Soc. econ. Paleontol Mineral . , Concepts Sediment. Paleont., Tulsa, 3 , 233-244. McCuBBIN, D . G . ( 1 969) Cretaceous strike-valley sandstone reservoirs, Northwestern New Mexico. Bull. Am. Assoc. petrol. Ceo!. , 53, 2 1 14-2140. NARAYA N , J. ( 197 1 ) Sedimentary structures in the Lower Greensand of the Weald, England, and Bas-Boulonnais, France . Sediment. Geol. , 6, 73 - 109. N10, S.-D. & YANG, C .-S. ( 1991 ) Diagnostic attributes of clastic tidal deposits: a review. I n : Clastic Tidal Sedimentology (Eds Smith, D . G . , Reinson, G . E . , Zaitlin, B . A . and Rahmani , R . A . ) . Can . Soc. petro l . Geo l . , Calgary, Memoir 16, 3-28. NuMMEDAL, D . & SwiFr, D .J.P. ( 1987) Transgressive stratigraphy at sequence-bounding unconformities: some principles derived from Holocene and Creataceous examples. In: Sea-Level Fluctuation and Coastal Evol ution (Eds Nummedal, D . , Pilkey, O . H . & Howard, J . D . ) . Spec. Pub!. Soc. econ. Paleont. Mineral, Tulsa, 4 1 , 241 - 260. OERTEL, G . F . , HENRY, V . J . & FOYLE, A . M . ( 1 99 1 ) Impli cations of tide-dominated lagoonal processes on the preservation of buried channels on a sediment-starved continental shelf. In: Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (Eds Swift, D . J . P . , Oertel, G.F. , Tillman, R . W . & Thorne, J . A . ) . Spec. Pubis i n t . Assoc. Sediment., No. 1 4 , 379-393. Blackwell Scientific Publications, Oxford . OoMKENS, E . & TERWINDT, J . H . J . ( 1960) Inshore estuarine sediments in the Haringvliet, Netherlands. Ceo/. Mijnbouw, 39, 701 -710. OwEN , H.G. ( 1 972) The Gault and its junction with the Woburn sands in the Leighton Buzzard Area, Bedfordshire and Buckinghamshire. ?roc. Geol. Assoc. , 83, 287-312. PuNT, A.G. & WALKER, R . G . ( 1 987) Morphology and origin of an erosion surface cut into the Bad Heart Formation during major sea-level change, Santonian of West-Central Alberta, Canada. J. sedim. Petrol. , 57, 639-650. PuNT, A . G . & NORRIS, B . ( 1 99 1 ) Anatomy of a ramp margin sequence: facies successions, palaeogeography, and sediment dispersal patterns in the Muskiki and Marshybank formations, Alberta Foreland Basin. Bull. Can. Pet. Geol. , 39, 18-42. REINECK, H . E . ( 1963) Seclimentgefuge in Bereich cler Sucllichen Norclsee. Abh. Senckenb. Natwforsch. Ges, 505, 1 - 138. REINECK, H . E . & SING H , I . B . ( 1973) Depositional Sedi mentary Environments with Reference to Terrigenous Clastics. Springer-Verlag, New York, 431 pp. RuFFELL, A . H . & WACH, G . D . ( 199 1 ) Sequence strati graphic analysis of the Aptian- Albian Lower Greensand in southern England. Mar. Petrol. Ceo/. , 8, 341 -353 . RuFFELL, A . H . & WACH, G . D . (in press) Comparison of depositional sequences in the arenaceous beds of the Aptian-Albian boundary (Cretaceous) in southern and eastern England. In: The Mesozoic and Cenozoic Sequence Stratigraphy of European Basins (Eels Graciansky, P.C. & Jacquin, T . ) . Soc. econ . Paleont. Mineral. Tulsa. ScHAFER, W. ( 1972) Ecology and Palaeoecology of Marine Environments , p. 568. Oliver & Boyd, Edinburgh, (Also:
45
Aktua-palaeontologie nach Studien in der Nordsee, p. 666. W. Kramer, Frankfurt. ) . ScHUBEL, J . R . ( 1971 ) Sources o f sediments to estuaries. In: The Estuarine Environment, Estuaries and Estuarine Sedimentation (Eel. Schubel, J . R . ) , pp . V, 1 - 19. American Geological Institute, Washington. SCHWARZACHER, W. ( 1953) Cross-bedding and grain size in the Lower Cretaceous sands of E. Anglia. Geol. Mag. , 90, 322-330. SHEPHARD-THORN , E . R . , HARRIS, P . M . , HIGHLEY, D . E . & THORNTON, M . H . ( 1 986) An Outline Study of the Lower Greensand of Parts of South-east England. Report for the Department of the Environment, British Geological Survey, Keyworth. SPEARING, D.R. ( 1975) Shannon Sandstone, Wyoming. In: Depositonal Environments as Interpreted from Primary Sedimentary Structures and Stratification Sequences. Society of Economic Paleontologists and Mineralogists, Short Course No. 2, pp. 104- 1 14. STRIDE, A . H . ( 1982) Offshore Tidal Sands: Process and Deposits. Chapman & Hall, London, 213 pp. SuRLYK, F. & NoE-NYGAARD, N . ( 1991 ) Sand bank and clune facies architecture of a wide intracratonic seaway: Late Jurassic- Early Cretaceous Raukelv Formation, Jameson Land, East Greenland. I n : The Three Dimensional Facies Architecture of Terrigenous Clastic Sediments and Its Implication for Hydrocarbon Discovery and Recove1y (Eels Miall, A . D . & Tyler, N . ) . Soc. econ. Paleontol. Mineral . , Concepts Sediment. Paleont . , Tulsa, 3 , 261 - 276. SwiFT, D . J . P . & THOR N E , J .A . ( 1 99 1 ) Sedimentation on continental margins, 1: a general model for shelf sedimentation. I n : Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (Eels Swift, D . J . P . , Oertel, G.F. , Tillman, R.W. & Thorne, J . A . ) . Spec. Pubis i n t . Assoc. Sediment . , No . 1 4 , p p . 3 - 3 1 . Blackwell Scientific Publications, Oxford. SWIFT, D . J . P . , MOIR, R . & FREELAN D , G . L. ( 1980) Quaternary rivers on the New Jersey shelf: relation of seafloor to buried valleys. Geology, 8, 276-280. TERWINDT, J . H . J . ( 1 973) Sand movement in the in- and offshore tidal area of the southwestern part of The Netherlands. Geol. Mag. , 52, 69-77. THORNE, J . A . & Swwr, D . J . P. ( 1991) Sedimentation on continental margins, VI: a regime model for depositional sequences, their component systems tracts, and bounding surfaces. I n : ShelfSand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (Eds Swift, D . J . P . , Oertel , G . F . , Tillman, R . W . & Thorne, J . A . ) . Spec Pubis int. Assoc. Seclimento l . , No. 14, pp. 189-255. Blackwell Scientific Publications, Oxford. VAN DEN BERG, J . M . , GoESTEN , M . J . B . G . & S M U LDERS, F. ( 1980) Dynamics and sequential analysis of a mesoticlal shoal and intershoal channel complex in the eastern Scheidt (southwestern Netherlands). Sedimem. Geol. , 26, 263 -279. VAN WAGONER, J .C . , MITC H U M , R . M . , CAMPIO N , K . M . & RAHMA N IAN , V . D. ( 1 990) Siliciclastic Sequence Stra tigraphy in Well Logs, Cores, and Outcrops: Concepts for High-resolution Correlation of Time and Facies. American Association of Petroleum Geologists, Tulsa, Methods in Exploration, 7, 55 pp. ViSSER, M . J . ( 1 980) Neap-spring cycles reflected in
46
H. D. Johnson and B. K. Levell
Holocene subtidal large-scale bedform deposits: a preliminary note. Geology, 8, 543-546. WALKER, R.G. ( 1 984) Shelf and shallow marine sands. I n : Facies Models ( E d . Walker, R . G . ) . Geosci. Can . , Reprint Ser. 1 , 1 4 1 - 170. WALKER, R.G. & BERGMAN , K .M. ( 1992) Late Cretaceous clastics: cores and outcrops. In: Field Trip Guidebook: Trip No. 6. Canadian Society of Petroleum Geologists/ American Association of Petroleum Geologists, Calgary, 77 pp. WALKER, R . G . & P uNT, A . G . ( 1992) Wave- and storm dominated shallow marine systems. In: Facies Models: Response to Sea Level Change (Ed. Walker, R.G. & James, N . P . ) , pp. 2 19-238. Geological Association of Canada, St Johns. WoNHAM, J . P . ( 1993) Sedimentology and sequence stra tigraphy of tidal sandstone bodies: implications for reservoir characterisation. Unpublished PhD thesis, University of Liverpool.
C.W. & WRIGHT, E . ( 1947) The stratigraphy of the Albian Beds at Leighton Buzzard. Geol. Mag. , 84, 1 6 1 - 168. WvArr, R.J . , LAKE, R . D . , MoORLOCK, B . S . P . & SHEPHARD THOR N , E . R . ( 1986) The Leighton Buzzard Area: Geo logical Report on 1: I 0 000 Sheets SP82NE, Southeast; SP92NW, Southwest; SP93NW, Northeast, Southwest, Southeast. British Geological Survey, Keyworth . YANG, C.-S. & N i o , S.-D. ( 1989) An ebb-tide delta depo sitional model - a comparison between the modern Eastern Schelde tidal basin (southwest Netherlands) and the Lower Eocene Roda Sandstone in the southern Pyrenees (Spain ) . Sediment. Geol. , 64, 175- 196. ZIEGLER, P . A . ( 1 988) Evolution of the Arctic- North Atlantic and the Western Tethys. Mem . Am. Assoc. petrol. Geo l . , 43, 198 pp. ZIEGLER, P.A. ( 1990) Geological Atlas of Western and Central Europe. Shell Internationale Petroleum Maatschaappij B . V . , The Hague, 239 pp.
WRJGHT,
Spec. Pubis int. Ass. Sediment. (1995) 22, 47-74
An incised valley in the Cardium Formation at Ricinus, Alberta: reinterpretation as an estuary fill R O G ER G . W A L K E R Department of Geology, McMaster University, Hamilton, Ontario L8S 4Ml, Canada
ABSTRACT
The Ricinus incised valley (Turonian-Coniacian Cardium Formation) is at least 50 km long, 10 km wide and up to about 30 m deep. It trends NW-SE, parallel to tectonic trends and shoreline trends in the Cardium Formation. I t cuts into shallow-marine mudstones and sandstones (Raven River Allomember of the Cardium), and is truncated by a transgressive surface of erosion (E5). The valley fill forms a small oil reservoir in the subsurface, but the field is disrupted by thrust faults, and the western margin is poorly constrained because of thrusting and Jack of wells. The valley fill consists of interbedded sandstones and mudstones, with both brackish and fully marine trace fossil assemblages. Some of the sandstones are structureless, with individual beds apparently over 4 m in thickness; others are parallel-laminated with beds up to 3 m thick. Cross-bedding in sets thicker than 5 em is extremely rare. The mudstones contain some thin silty and sandy laminations, but are commonly extensively bioturbated with a fully marine trace fauna (robust Helminthopsis, Zoophycos, Planolites, Anconichnus, Skolithos, Ophiomorpha, Thalassinoides, Teichichnus, Asterosoma, Terebellina and Rosselia). The mudstones can be traced as a sheet about 4 m thick along the entire length of the valley; they appear to pinch out against the eastern wall of the valley. Chert-pebble conglomerates are scattered throughout some wells, but tend to be concentrated near the northern end of the valley. Here, they arc associated with sandstones and with fully marine mudstones. The valley was formerly interpreted to have been cut and filled by turbidity currents, but this interpretation is difficult to reconcile with the depositional environments of the rest of the Cardium Alloformation. It is suggested here that the valley was cut by a river during a falling stage of relative sea level, and was filled transgrcssivcly as an estuary. There is absolutely no evidence for tidal currents, and little indication of waves. The very thick and structureless sandstones suggest pulses of sediment swept into the estuary from the bay head, possibly as density underflows. There is little or no evidence for the reworking of these deposits within the estuary. There arc no central basin mudstones (no turbidity maximum), although this could be due to incomplete preservation of the entire length of the estuary. The geometry of the estuary and the facies distribution of the fill suggest that the open sea Jay to the north. Chert pebbles moving in the longshore drift system were introduced into the estuary from the north by storms; they were not introduced from the fluvial (southern ) end of the estuary. The orientation of the incised valley, parallel to depositional strike for over 50 km, remains a maj or problem.
INTRODUCTION
storm-dominated shallow-marine sandstones (Walker & Eyles, 1 988) , and transgressive incised shoreface deposits ( Bergman & Walker, 1987 , 1988; Pattison & Walker, 1 992). The main purpose of this paper is to re-examine the allostratigraphical (NACSN , 1983) position of the sand body (Figs 1 & 2 ) , and to reinterpret the incised valley fill as estuari ne. Ricinus is a small oil field, discovered in 1969.
The Ricinus sand body in the Cardium Formation of Alberta (Figs 1 & 2) is over 50 km long, at least 10 km wide, and up to about 30 m thick (Walker, 1985 ) . I originally interpreted it as a channel fill, with channel cutting and filling by turbidity currents. This interpretation has always been difficult to reconcile with the sedimentology of the other facies in the Cardium Formation, which consist largely of
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
47
48
R.G. Walker
9
10
8
7
37
RICINUS 0
Wells
.
Cores
OFF-FIELD 36
o
Wells
"
Cores
60' ALBERTA
35
054'
0
3
/�K�\�9' \ o
I
so
�
1 5:
4
Fig. I. Map showing locations of
6
EDMONTON
w--.:_
E5 fields
�
E4 fields
33
�
� \, U '\.
RICIN
c::
�
116'
sw
32
CALGARY . - 1'
114'
6 miles
5
-10 KM
c.20 km
NE
Cardium fields associated with the E4 (fields outlined) and E5 (black) bounding discontinuities. Note location of Ricinus close to and partly straddling the edge of the deformed belt (thrust symbol, inset lower left). The main map shows locations of wells and cores. All of the wells immediately east of Ricinus are shown, but only a few Caroline wells have been used. Note also the locations of cross-section A-A' through to G-G'. Number at right indicates Townships, and number along top indicates Ranges west of the 5th Meridian (114°).
Fig. 2. Stratigraphy of the Cardium Alloformation in the Ricinus area, drawn roughly to scale vertically. Bounding discontinuities E2, E3, E6 and E6.5 cannot be recognized here. Sandier upward successions A-D are assinged to the Hornbeck Allomembcr (note that succession D is commonly truncated by E4). E5 RSE (regressive surface of erosion) defines the base of the Ricinus Allomember, and E5 TSE (transgressive surface of erosion) defines the top. There is some evidence that E5 TSE cuts downward progressively southwestward. IT and RT refer to initial and resumed transgressions, as discussed in the text.
Cardium Formation incised valley
Initial oil-in-place was estimated as 2.88 x 106m3 (18. 1 million bbl) , with initial recoverable oil esti mated at 2.27 x 106m3 ( 14.3 million bbl) (Geological Survey of Canada, 198 1 a , b). However, the signifi cance of the field goes well beyond its minor econ omic potential. It is the only incised valley known in the Cardium, which is surprising in a formation dominated by eight regionally extensive erosion surfaces, each controlled by major fluctuations of relative sea-level. The incised valley interpretation is complicated by the fact that the valley is straight, at least 50 km long, and is parallel to regional strike (Fig. 1 ) . The new interpretations are based upon all o f the data available as of April 1993 , as illustrated in seven new cross-sections presented here (see Figs 3-9 and Appendix 1 ) . There are 172 on-field wells, of which 71 are cored (Fig. 1 ) . All measurements are given in metric units, except where specific reference is made to wells where the well logs and cored intervals were originally recorded in feet.
INTRODUCTION TO THE CARDIUM ALLOFORMA TION
The Cardium Alloformation (NACSN , 1983) is Upper Cretaceous (Turonian - Coniacian), and occurs in the Alberta foreland basin. This basin is part of the Western Interior Seaway, which was open from the Boreal Ocean to the Gulf of Mexico during the Turonian . Sediment was supplied from the active cordillera in the west. The Cardium is about 100 m thick, and is dominated by open-marine mudstones and storm influenced sandstones with hummocky cross stratification (HCS). There is one prograding shoreface succession (Kakwa Allomember; Plint & Walker, 1987) overlain by non-marine deposits (Musreau Allomember), and several examples of incised transgressive shoreface deposits. The relationships of these various deposits have been documented by mapping eight bounding discon tinuities, at least five of which can be traced basin wide (Piint et a/. , 1986; Wadsworth & Walker, 199 1 ; Walker & Eyles, 199 1). These were originally termed erosion surfaces and lettered El- E6, E6. 5 and E7. The regional trend of the various shoreface sand bodies is consistently NW-SE (Fig. 1 ) , parallel to the active cordillera to the west of the foreland basin.
49
Structural problems at Ricinus
The exact morphologies of the Ricinus valley and sand body are difficult to reconstruct because Ricinus lies at the eastern limit of the Foothills deformed belt (Figs 1 , 10 & 1 1 ) . The sand body is broken by several SW-NE trending faults, interpreted to have formed in response to different distances of thrusting toward the northeast. The fault positions have been located mainly by offsets in the northeastern margin of the field, which can be fairly well constrained by the close juxtaposition of wells with 'Ricinus' well log responses versus 'off-field' log responses (Figs 1 & 3-7). Within Ricinus, the sand body is repeated by thrusting in many wells (see Fig. 5 ) . In a few wells the dips exceed 45° and are locally vertical. The former turbidite interpretation
The interpretation of Ricinus as a channel cut and filled by turbidity currents (Walker, 1985) was based on the presence of very thick structure less sandstones and the presence of many beds characterized by a Tabc and Toe internal structure (Bouma, 1962). In the early 1980s, the possibility of turbidity currents in the Western Interior Seaway was suggested by the need to explain how a very large number of sharp based HCS beds had been emplaced well below fair weather wave base. Thinking was also influenced by the occurrence of well-defined turbidites overlain by HCS beds in the Jurassic Passage Beds in the Banff area. Here , both the turbidites and HCS beds have identical palaeocurrent orientations (Hamblin & Walker, 1979), suggesting emplacement of the HCS beds by turbidity currents, but above storm wave base. Turbidite ideas were first applied to the Cardium before the presence and significance of sea-level changes in the Cardium had been documented (Piint et a/., 1986), and before much of the important recent work on estuarine deposits had been pub lished (Dalrymple et a/. , 1992). The main purpose of this paper is to suggest an alternative to turbidity currents for cutting and filling the incised valley at Ricinus. Allostratigraphy
The most useful subdivision of the Cardium is allostratigraphical (NACSN , 1983) rather than lithostratigraphical (Fig. 2). An allostratigraphical unit 'is a mappable stratiform body of sedimentary
50
R.G. Walker
rock that is defined and identified on the basis of its bounding discontinuities ... [one may thus define] as single units discontinuity-bounded deposits charac terized by lithic heterogeneity' (NACSN , 1983 , p. 865). I regard the Cardium Alloformation as a formally defined stratigraphical unit, and have appended some thoughts on allostratigraphy in Appendix 2. The Cardium Alloformation has been subdivided into 14 allomembers, with eight well defined bounding discontinuities (E1 through to E7; Plint et al. , 1986) . Throughout this paper, informal use of the term Cardium (without a modifier) implies the Cardium Alloformation. The term Cardium Formation will be used only to designate the litho stratigraphical unit.
regionally extensive sheet-like sandier upward successions of the Hornbeck Allomember (Fig. 2). These successions are labelled A through to D in Figs 2-8 and collectively form the datum in these cross-sections. Succession C is cored in wells 7-9-36-8W5 and 10-17-34-7W5 (Fig. 4; all wells lie west of the 5th Meridian , and the W5 will be omitted from all subsequent well numbers cited). Succession C consists of bioturbated sandy mudstones that gradually become sandier upward; the top consists of a 1-cm-thick layer of chert grains about 0.5 mm (7-9-36-8) or 3 mm (10-17-34-7) in diameter. This sandier upward succession, capped by a gritty layer, is similar to those described by Walker & Eyles ( 1 988) from the Raven River Allomember; suc cessions A, B and D have not been cored but are probably similar to succession C. The sand bodies that form reservoirs at Garrington, Crossfield and Caroline (E4 fields in Fig. 1; Pattison & Walker, 1992) are assigned to the Burnstick Allomember. They rest in NW-SE trending steps on the E4 erosion surface (Fig. 2) , and are inter preted as transgressive shorefaces. Each incised shoreface is underlain by a surface of initial trans gression (IT in Fig. 2) , and is truncated by a surface of resumed transgression (RT in Fig. 2). An RT surface can pass southwestward into an IT surface if and when the transgression pauses and another still stand shoreface becomes incised. An example can be seen in well 7-33-36- 10, section A (Fig. 3 ) , where
OUTLINE OF CARDIUM STRATIGRAPHY AND SEDIMENTOLOGY
The base of the Cardium Alloformation is taken at bounding discontinuity El (Figs 2 -6). In the Ricinus area E1 has never been cored, and it is probably a correlative conformity (see Appendix 2) rather than a surface with demonstrable erosion . Its log identi fication is compatible with the E 11og identifier used by Plint et al. ( 1 986) and Wadsworth & Walker ( 199 1 ) . Erosion surfaces E2 and E3 cannot be ident ified in this area. Surface E4 has demonstrable erosion (Figs 3, 5 & 6), and cuts into a series of
7-33-36-10
7-26-36-10
7-36-36-10
1-31-36-9
8-32-36-9
11-33-36-9
10-34-36-9
11-36-36-9
7-32-36-8
@
® �
���
�
--
-
--
0
�
�Es�
�
�
�TSE�
�
---'
F �
� �
?-
�
�
�-
-
_t
_
Fig. 3.
-EI-
Gamma-ray and resistivity well-log cross-section A-A', hung collectively on a series of more -or-less parallel markers above and below the Cardium. Depths are marked at 50-m intervals and cored intervals are shown as black bars. Wells 7-33-36-10, 7-26-36-10, 7-36-36-10 and 7-32-36-8 have 'off-field' well-log signatures.
51
Cardium Formation incised valley 7-36-36-10 15-23-36-10
® �
5-10-36-9
4-18-36-8
10-18-36-8
-
�� z 0
11-18-36-9
7-19-36-8
7-9-36-8
®
-
�
--- �/
-
----
-TSE �co�
?-
::;;� 0:: 0 LL
__
'3
/
___________
,
BLACK
�E4�
::;; ::;) 0 0:: "' u
'
f\_SE___
�E4�
_,
fc
\ -E. I-
j
Fig. 4.
Cross-section B-B', hung communally on the sandier upward successions A-C, and the top of the black blanket (dashed line). See caption of Fig. 3 for other conventions, and note draping of upper markers over the Ricinus sand body. The shoulder on the resistivity and gamma-ray logs typical of E7 is not present in 7-9-36-8, nor is the narrower waist in the well-log profile immediately below E7. The waist appears to be cut out by E7 toward 7-9-36-8, and E7 is probably only about 2 m above E5 (see Figs 12-14). Wells 7-36-36-10, 15-23-36-10, 11-18-36-9, 7-19-36-8 and 7-9-36-8 have 'off-field' well log signatures.
©
1-5-34-8
6-9-34-8
14-10-34-8
10-10-34-8
6-14-34-8
6-24-34-8
11-20-34-7
10-21-34-7
©
F?
--
-------
-- EI -
Fig. 5.
-
§ a>
----
g
"' "'
g---"'
Cross-section C-C', hung on sandier upward successions A-C. Other conventions explained in the caption of Fig. 3. Wells 1-5-34-8 and 10-21-34-7 have 'off-field' well-log signatures. F indicates a fault repeat of the top of the Ricinus Allomember in 10-10-34-8, and F? indicates possible faulting. RRl is the first sandier upward succession in the Raven River Allomember.
52
R.G. Walker
1 0-8-33-7
4-9-33-7
3-10-33-7
10-12-33-7
7-10-33-7
6-24-33-7
11-25-33-7
®
®
� �
g " N
z 0
--
0 0 N
�:::!:
0::: 0 LJ... 0 ...J ...J <[
?
E4---
BLANKET
rc � TB 15 0::: <[ u
?
�N
--
0
+A
t
g N
-
EI-
-
Fig. 6.
Cross-section D-D', hung on sandier upward successions A-C. Other conventions explained in the caption of Fig. 3. Note the erosional truncation of markers between ES TSE and E7 by the E7 bounding discontinuity between wells 7-10-33-7, 10-12-33-7 and 6-24-33-7. Wells 10-12-33-7, 6-24-33-7 and 11-25-33-7 have 'off-field' signatures.
11-18-33- 7
3-19-33-7
8-19-33-7
6-20-33-7
4-21-33-7
6-23-33-7
©
----- ----
E4 formed during an initial transgression , and sand was deposited on the surface during a stillstand in transgression . This sand was truncated as the E4 transgression continued, leaving a 4-m-thick sand body. However, over most of the study area, E4 is probably overlain only by a thin lag. The Raven River Allomember extends from the top of this lag to the ES bounding discontinuity (Fig. 2). The lower part consists of black mudstones (the 'black blanket' of Walker, 1983) characterized by parallel and featureless gamma-ray and resistivity logs (Figs 2 & 4-8). In logs, the top of the black
Fig. 7.
Cross-section E-E', hung on the B sandier upward succession. Note truncation of the top of the 'black blanket' and top of RR1 by bounding discontinuity ES RSE, and the apparent pinchout of facies D between wells 3-19-33-7 and 8-19-33-7 against the wall of the incision. Facies D and the sandstones above (3-19-33-7) appear to be eroded by ES TSE toward well 11-18-33-7.
blanket is taken at the first deflections of the gamma ray and resistivity logs (the dashed line in Figs 4-8). Here, the black blanket grades into bioturbated siltstones and sandstones, and HCS sandstones , organized into sandier upward successions. The first of these successions is identified as RR-1 (Figs 5 & 7). The successions are eroded to various depths by ES. Erosion surface E5 has received much attention (Fig. 2; Walker & Eyles, 199 1 ) , partly because of its economic importance in locating the shoreface gravels in the Carrot Creek area (Bergman &
53
Cardium Formation incised valley
7-16-34-8
16-16-34-8
®
'-"'"""E7"'"""-
9-16-34-8
®
�,L
Cross-section F-F', hung comunally on sandier upward successions A-C. Note thinning and disappearance of facies C and/ or D from well 16-16-34-8 eastward toward well 10-22-34-8, and possible erosion of these facies by the thick sandstone in well 7-16-34-8 (possible erosion surface shown by question marks). Note also the truncation of markers between ES TSE and E7 by the E7 bounding discontinuity.
3-19-33-7
6-24-34-8
!15
Fig. 8.
�
10-22-34-8
�ISE"""'
�
BLANKET
16-30-34-8
1 2 -19-35-8
ll-30-35-8
16-30-35-8
3-6-36-8
®
3-6-36-8
7-18-36-8
10-18-36-8
16-13-36-9
6-24-36-9
13-26-36-9
1 2-34-36-9
2-4-37-9
) -�
!15 �0 M
Fig. 9.
South (G) to north (G') cross-section showing the continuity of facies C and/or D along the length of the field. These muddy facies can be recognized by their distinctive indentation in the gamma-ray and resistivity well-logs. The mudstones occur in the uppermost part of the core in 12-34-36-9 (Fig. 26), and also appear to occupy the same position in wells 7-18-36-8, 10-18-36-8 and 13-26-36-9. In these wells, the Ricinus Allomember is unusually thin, suggesting that sandstone originally above facies C and/or D has been eroded at the ES TSE bounding discontinuity.
54
R.G. Walker
Walker, 1987, 1988) , and also for determining the remnant erosional topo·graphy on top of major Cardium reservoirs at Pembina, Willesden Green and Ferrier (shown black and identified as E5 fields in Fig. 1 ; Walker & Eyles, 1988, 199 1 ) . In this paper, the base of the Ricinus sand body is correlated with E5 - specifically with a falling stage of relative sea-level (E5 RSE- regressive surface of erosion) . Over most of the Cardium basin, the subaerial E5 surface was modified during transgression , and is designated E5 TSE (transgressive surface of erosion) in Figs 3-9. Erosion beneath E5 RSE is illustrated by the truncation of log markers, particularly in Figs 5 - 8 . The extent o f erosion a t the E 5 R S E surface can also be shown by isopaching the E4 to E5 RSE interval. If it is assumed that E4 is relatively planar in the Ricinus area, the shape of the Ricinus incision can be seen in Fig. 10. The new term Ricinus Allomember (Fig. 2) is introduced here for the sand 10
9
8
7 37
body that rests on E5 RSE and is truncated by E5 TSE. The extent of this sand body, and the type wells are the same as for the Ricinus Member as defined by Walker ( 1985, p. 1 98 1 ) . Discontinuity E 6 i s a n erosion surface with little topographical relief and it cannot be identified reliably in this area. Although E6.5 has more relief than E6 it also is difficult to identify in the study area. Discontinuity E7 consistently forms a small shoulder on the gamma-ray and resistivity logs, and can be identified easily. The topography of the E7 surface has been mapped by Wadsworth & Walker ( 199 1 ) . They showed a deep 'valley' or 'low' on the erosion surface j ust east of Ricinus, and part of this can be seen in Figs 4 & 6. In Fig. 4, E7 is not present in the 'expected' position in 7-9-36-8 (Fig. 12) . The lag associated with E5 is prominent in core (Fig. 13), and is overlain by transgressive mudstones. A second lag, with chert pebbles up to 1 . 1 em diameter, occurs about 2 m above E5 (Fig. 14) ; it shows on the gamma-ray and resistivity logs and is identified as E7 (?) in Fig. 4. A similar relationship is shown in Figs 6-8, where log markers in the E5 TSE to E7 interval are truncated by E7. Log marker E7, which is nor mally well developed and easy to identify, is absent
10
36
9
8
7 37
35
36 34 6
35 33
34
Isopach, E4 to E5 RSE (base of Ricinus
6 32
incised valley), metres Fig. 10.
Isopach map of the interval between E4 and E5. On field, the map shows the E4-E5 RSE interval, and off field, the E4-E5 TSE interval. Isopachs show height above E4; thus the 4 m isopach shows the deepest part of the valley. Note that the edge of the valley is approximated by the 28 m isopach. The line F shows one of the faults at Ricinus (Fig. 1); north of this fault, all of the data points have been shifted southwestward to line up the axis of the field as defined by the well density - see Appendix C.
lsop1ch, E5 RSE to E5 TSE, metres
33
32
Fig. 11.
Isopach map of the E5 RSE to E5 TSE interval (the thickness of the Ricinus sand body). See Fig. 10 and Appendix C for explanation of line F.
55
Cardium Formation incised valley
Fig. 12.
Core from well 7-9-36-8, which lies just east of the margin of Ricinus. B =bottom, T =top. Note siderite horizon (below label) and chert pebbles (above label) associated with ES, and the interval of laminated mudstones (the 'laminated blanket' of Walker, 1983) between ES and E7. Note siderite horizon (under label) and chert pebbles (above label) associated with E7, and the much blacker mudstones without silty laminations above E7. The separation of ES and E7 over Ricinus is 40-48 m, but is about 2m in this core. Thus there is about 40 m of erosional relief on E7 immediately to the east of Ricinus (Wadsworth & Walker, 1991).
in the easternmost wells in these cross-sections, suggesting that E7 and E5 are co-planar (or that E7 has actually cut into and removed E5). The top of the Cardium Alloformation is taken at the top of the lag that rests on E7.
Fig. 13.
Detail of siderite horizon, chert pebbles, and eroded sandstone clast (large triangular shape) at the ES TSE horizon, well 7-9-36-8, 8634 ft. See Fig. 12 for stratigraphical context of this photograph, but note that core has been rotated in Fig. 13 to show different features. Core is 3 in. (7.6 cm) wide.
GEOMETRY OF THE RICINUS INCISED
VALLEY
AND SAND BODY
The geometry of the Ricinus sand body was estab lished initially from many cross-sections - seven are presented here (Figs 3 - 9). After it had been deter-
mined that bounding discontinuities E7 , E5 TSE , E5 RSE , E4 and El could be traced and identified throughout the study area, maps of various strati graphical intervals were made using SURFER 4.0 (Golden Software Inc. , Golden, Colorado) (see Appendix 3). ·
56
R.G. Walker
which is the base of the valley. The eastern edge of the valley is approximated by the 28 m isopach. The valley trends roughly NW-SE, with a straight and well-defined eastern margin that has at least 24 m of relief in T34 R8-9 (Fig. 10). This eastern valley wall has a gradient of about 0.007° westward into the valley. The eastern margin remains a straight and well-defined feature from T35 R8-9 to the southern most fault in T32 R7, but northward, the valley seems to broaden and become less well defined in T36 R9. Erosion at the eastern valley wall is evident in the cross-sections; Figs 3 - 8 all show truncation of log markers by E5 RSE (particularly in Figs 5, 7 and 8) , or convergence of E5 RSE toward the top of the 'black blanket' (Fig. 6) , or convergence toward E4 (Fig. 5). The eastern margin is characterized by a rapid change from on-field to off-field log signatures, allowing a boundary to be drawn to within 600 m (e.g. in land section 2 1 , T35 R8 and land sections 17-20, T36 R8; Fig. 1 ) . The western margin i s very poorly defined because there is little well control; many of the well logs suggest extensive thrusting and are impossible to use in the isopach maps. The maximum depth of the valley is in the centre of the field (T34 R8 -9), and the valley becomes shallower both northwestward and southeastward. The northern and southern ends of the valley are faulted. Wells on line with the Ricinus sand body but outside the faulted ends have 'off-field' well-log signatures. Sand body geometry
Fig. 14. Detail of siderite, bioturbation and chert pebbles
at the E7 horizon, well 7-9-36-8, 8639 ft. See Fig. 12 for stratigraphical context of this photograph. Core is 3 in. (7. 6 cm) wide.
Incised valley geometry
An isopach map of the E4 to E5 interval is shown in Fig. 10. On-field, the map shows the E4 to E5 RSE interval, and off-field, it shows E4 to E5 TSE. Erosion surface E4 is not exactly planar (Figs 3 -8) , but regionally it may be taken to approximate a plane. If so, the E4 to E5 isopach can be interpreted as showing the on-field topography of E5 RSE,
The sand body geometry was defined by isopaching the E5 TSE to E5 RSE interval (Fig. 1 1) , with the assumption that E5 TSE (the upper bounding dis continuity) approximates a planar surface. The sand body isopach map closely resembles the map of the E4 to E5 RSE interval (the morphology of the base of the valley) , with the sand body thinning northwestward and southeastward from a maximum thickness greater than 24 m in T34 R8. In the northwest, the isopach of the sandstone seems to define a valley more clearly than the E4 to E5 RSE map. The prominent eastward deflections of the 8 and 4 m isopachs in T34 R7 -8 are due to the 1 1-m-thick sandstone in well 1 1-20-34-7; this sand stone has apparently been thrust unusually far to the northeast (Figs 2 & 5). With this exception, the eastern edge of the sand body is remarkably straight. The sand body is characterized by a very sharp
Cardium Formation incised valley
base, and blocky to blocky and serrated gamma-ray and resistivity well-log signatures (Figs 3 - 8) ; these will be termed 'on-field' signatures. They contrast with off-field wells (e .g. Fig. 3, wells 7-33-36- 10, 7-26-36- 10 and 7-36-36- 10) , which are characterized by a gradual sandier upward signature between E4 and E5 and a capping sand body that is only a metre or two thick.
57
Facies A: structureless sandstones
This facies consists of sandstones that lack any physical or biological structures (Figs 1 5 - 17), or that are dominantly structureless but contain parallel lamination and/or ripple cross-lamination in the uppermost part of the bed . Grain size is mostly in the fine-grained sand range, with a few beds of very
FACIES OF THE RICINUS ALLOMEMBER
The Ricinus sand body contains a suite of facies that are quite unlike those found in any other part of the Cardium. Five facies were recognized originally (Walker, 1985 ) , and this scheme is modified and simplified here.
Fig. 15.
Structureless sandstones with no apparent bedding surfaces and no mudstone partings. B =bottom, T =top. Core 3 in. (7 .6 em) wide, each core sleeve 2.5 ft (about 75cm) long. Well 8 -10-34-8, 8109-8124 ft.
Fig. 16.
Structureless sandstone. The absence of structure appears to be primary - there is no evidence of the destruction of former structure by dewatering or bioturbation. Core 3 in. (7. 6 cm) wide. Well 8-10-34-8, 8115 ft.
58
R.G. Walker Structureless sandstones: percent of core
100 ••• ••
... .
u
6
. .
Structureless sandstones:
1m)
average bed thickness 5
4
3
2 . .
.
.. .
.
.
.
•
.
..
•
•
.
... . .
.
0 33 Southeast
34
34
35
36
36
37
Northwest
Fig. 17.
Distribution of facies A structureless sandstones from T33 northward to T37.
fine-grained sand and even fewer of medium-grained sand. X-radiographs have not been made of these sandstones, but it is believed that the absence of sedimentary structures is real. Structures are readily observed in other sandstone facies of similar grain size (Figs 18- 20) . It seems reasonable to infer that where the core shows absolutely no indication of structure (Figs 15 - in the four core sleeves on the right - & 16) , there is probably nothing there. The percentage of facies A beds in core varies from zero to 100, but there is no trend from south to north (Fig. 17). Because of the intense fracturing, particularly in the southern part of Ricinus, it is difficult to estimate individual bed thicknesses. However, several cores contain sandstones that are unfractured, contain no internal grain-size breaks, have no mudstone partings, and have no layers of eroded mud clasts (Fig. 15 - note that there is one mudstone parting near the top of the fourth sleeve from the left). Possibly some of these criteria for multiple beds have been lost in places where core is
Fig. 18.
Sandstone showing crude parallel stratification at the base, well-developed parallel lamination in the middle, and ripple cross-lamination at the top. Core 3 in. wide (scale in em). Well 11-18-33-7, 8541 ft.
broken or cut. Nevertheless, there appear to be at least four single beds thicker than 4 m (in wells 1 1-1 8-33-7 , 7-36-33-8, 2-1-34-8, 14-27-36-9), with two more reaching thicknesses of 6.3 (7-28-34-8) and 6.7 m ( 16- 10-34-8). There is an overall tendency for the thicker beds to be concentrated in the southern part of Ricinus; the thinner ( < 1 m) beds occur
Cardium Formation incised valley
59
Fig. 19.
Almost every piece of core in these boxes shows parallel lamination. The individual pieces of core are cut, obscuring possible bed contacts. There are no angular divergences, intersections or truncations of lamination. B =bottom, T =top. Core is 3 in. ( 7. 6 cm) wide, and each core sleeve is 2. 5 ft ( about 75em) long. Well 14-2-34-8, 7726-7740 ft.
throughout the sand body but are more common in the north, particularly in Townships 36 and 37 (Fig. 17). These thinner beds have sharp bases marked by a jumble of eroded mud clasts and/or a concentration of coarse grains and granules of chert. Grading occurs in about half of the beds, either because of an upward disappearance of the coarse grains and granules (coarse tail grading) , or because of a progressive upward fining throughout the bed (dis tribution grading) . Internally, the beds are domi nantly structureless, but some show the development of crude parallel lamination toward the top. About one-third of the beds have a single layer of ripple cross-lamination at the top. The most common association of structures is Tbe (Fig. 18), with a few beds showing the succession Tal>e· These beds are much more common in the northern part of the sand
Fig. 20.
Parallel lamination, commonly emphasized by finely comminuted organic material. Well 14-2-34-8, 7708ft.
body; the average thickness is about 50 em (range 10- 135 cm , n= 153 beds). Trace fossils are very rare in this facies, and are restricted to Ophiomorpha. Facies B: parallel-laminated sandstone
These sandstones are characterized by parallel lamination (Figs 19 & 20); the stratification is characterized by subtle changes in grain size and by very thin carbonaceous laminae (Fig. 20) . I ndividual beds appear to be as thick as 2 m (maximum 6 m) in the southern part of the sand body, and thin north ward to an average of about 60 em . The percentage of parallel-laminated sandstones in the cores shows a very weak tendency to decrease northward (Fig.
R.G. Walker
60
Grain sizes are consistently in the fine-grained sand range. Trace fossils are rare , and are restricted to Ophiomorpha.
21).
100 80
Parallel-laminated sandstones percent of core
60 Facies C: interbedded bioturbated sandstones and mudstones
This facies consists of alternations of sandstones (thickness typically·1 -3 em) and mudstones (thick nesses a few millimetres to about 1 em) (Figs 22 & 23). I t occurs throughout the sand body (Fig. 9), but is thickest (10-15 m) in the central area (T34 R8) (Fig. 11). Contacts with the other sandstone facies (structureless, parallel-laminated) tend to be gra dational , and granules and pebbles of chert are rare (compare with facies D, below) . The sandstones are characterized by small-scale ripple cross-lamination. Where ripple forms can be seen, they are commonly asymmetrical, suggesting current ripples rather than wave ripples. The facies is not intensely bioturbated; the commonest trace fossil is Planolites (Fig. 23), with some examples of Ophiomorpha, Arenicolites , Conichnus and Rosselia (Fig. 22 - the Rosselia in sleeve 12, counting from the right, was figured by McEachern & Pemberton , 1992, p. 63, A). Facies D : bioturbated mudstones
This facies is less sandy and more thoroughly bio turbated than facies C (Figs 24 - 26). It contains a larger suite of trace fossils; the commonest traces are simple, horizontal sand-filled tubes about 1 em diameter (Planolites), with Helminthopsis (Fig. 27). Zoophycos (Fig. 27), Anconichnus, Skolithos, Ophiomorpha,
Thalassinoides ,
Teichichnus, Aste
and Rosselia. Scattered chert granules and pebbles are com monly associated with the bioturbated sandy mud stones (Fig. 26), and they occur throughout the thickness of the facies in many wells. The coarser
rosoma, Terebellina
Fig. 22.
•
40
20
.
.
0
. .
.
..
·-
.
I I I ·-···· ··-····-·-·--· ·-·.· -•
6
Parallel-laminated sandstones average bed thickness (m)
5 4 3 •
2
..
•
.
•
•
·. ·.
.
;,
..
••
0 �---.----.--,,--r33 34
Southeast
34
35
36
36 37
Northwest
Fig. 21.
Distribution of facies B parallel -laminated sandstones from T33 northward to T37.
clasts rarely occur in distinct beds - they are more commonly bioturbated into the sandy mud stone. The clasts occur throughout the field, from the southernmost (11-21-34-8) to northernmost ( 12-34-36-9) occurrences of the facies. Where the lower contact of this facies is seen in core, it is always sharp, and marked by the incoming of chert granules and pebbles (Figs 24, 25 & 28), together with rip-up clasts of mudstone and siderite. Good examples occur in wells 11-21-34-8 (1 mm chert grains; Fig. 24), 16-30-34-8 (3 mm grains) , 2-34-36-9 (with chert pebbles up to 1 . 4cm), 12-34-36-9 (1.3 em pebbles; Fig. 25) and 2-4-37-9
( Opposite. ) Well 12-19-35-8 boxed with the base of the well at the lower right; B =bottom, T =top. Note the sharp base of the Ricinus sandstone (BASE R), overlain by parallel-laminated sandstone that grade into facies C, the interbedded bioturbated sandstones and mudstones (sleeves 11-14 from right). Toward the top of this facies, it becomes interbedded with parallel-laminated sandstones (sleeves 14 and 15), and grades up into structureless sandstones (top five sleeves). Core is 3 in. (7.6 cm) wide, and each core sleeve is 2.5 ft (about 75cm) long. Well 12-9 -35-8, 8990-9048 ft. Note Rosselia (ROSS) in sleeve 12 - this example was figured by McEachern & Pemberton, 1992. During core photography, handwritten labels with way-up arrows, well locations and the words 'Ricinus' and 'bottom' were pinned on to the core boxes. Thus 'Ricinus bottom' refers to each photograph of three or four core boxes, not to the bottom of the Ricinus Allomember. The base of the Ricinus Allomember is marked by the printed letters BASE R.
62
R.G. Walker
and pebble lags mark the very sharp contact with underlying facies. These lags are normally only one to a few centimetres thick, with chert granules and pebbles up to 4 cm diameter, and abundant sideritized mud clasts. These basal lags occur from the southern to northern parts of the sand body, but there is no systematic distribution of clast sizes. Conglomerates also mark the top of the Ricinus Allomember, where they occur as lags on the E5 TSE surface (Fig. 25) . This surface truncates the Ricinus Allomember, and the conglomerates actually belong to the Carrot Creek Allomember. These conglomerates have been cored only in three wells (7- 19-35-8, 12- 18-36-8 and 12-34-36-9) ; they are 1 .2-2.5 m thick and have chert pebbles up to 4 cm diameter (in 12- 18-36-8). In T36 R9 (particularly in land sections 24 -27) conglomerates are abundant from bottom to top of the Ricinus sand body (Figs 26 & 29). The coarser clasts occur at the bases of graded sandstone beds, or form conglomerate beds up to about 1 m thick, with chert pebbles 1 -2 em diameter. A few pebbles reach 3 - 4 cm, with a maximum of 7.5 cm diameter (in 3-27-36-9). These bedded conglomerates are structureless to crudely stratified, and are inter bedded with structureless or parallel-laminated sand stones, and with thin mudstone partings (Fig. 29). The scattered granules and pebbles associated with the bioturbated mudstone facies have been described above. They are typically mixed into the mudstones by bioturbation, or may fill distinct bur rows (Fig. 30) .
LATERAL AND VERTICAL FACIES DISTRIBUTIONS
Fig. 23. Interbedded bioturbated sandstones and
mudstones, showing abundant Planolites (P), and a possible example of Zoophycos (Z) at the bottom of the photograph. Note ripple cross -lamination within the sandstones. Well 12-19-35-8, 9016 ft.
( 1 .7 em pebbles; Fig. 28). Thus the coarser clasts at this sharp contact become generally fi n er grained southward. Facies E: conglomerates
Conglomerates and scattered pebbles occur through out the Ricinus Allomember. At the base, granule
Sandstones
The Ricinus Allomember trends NW-SE, with a maximum thickness of about 25 m in T34 R8 (Fig. 1 1 ) . It thins against the eastern wall of the valley (Fig. 1 1) but the western margin is poorly constrained by data and disrupted by extensive thrusting. Within the main sand body, the sandstone facies (structure·· less, and parallel-laminated) show no preferred vertical trends. Some wells are almost exclusively structureless, others almost exclusively parallel·· laminated. Most wells show interbedding of these two facies. There is a weak tendency for the average bed thickness in both facies to decrease northwest ward (Figs 17 & 2 1 ) , but there is almost no change
Cardium Formation incised valley
63
Fig. 24.
This core is boxed top-to-left; B =bottom, T =top. The parallel-laminated sandstone facies is overlain sharply by the bioturbated mudstone facies. The sharp contact is marked by the incoming of chert granules (CH). The bioturbated mudstones contain Zoophycos (Z; it first appears near the base of the facies), Helminthopsis (H), Rosselia (R), Thalassinoides (TH) and Teichichnus (TE). Chert granules and pebbles are dispersed throughout the bioturbated mudstone facies, but become abundant and coarser toward the top (in the sandier core sleeve on the left). Core is 3 in. (7.6 cm) wide, and each sleeve is 2.5 ft (about 75 em) long. Well 11-21-34-8, 7022-7052 ft. See comment about handwritten label 'Ricinus bottom' in caption of Fig. 22.
along the field in the percentage of each facies in core (Figs 17 & 2 1 ). Mudstones
Because mudstones form an important part of many estuary fills, the position of Facies C and D within Ricinus was investigated in detail. These facies occur throughout the field (Fig. 9) , with a maximum thicknesses in T34 R8-9. An isopach map (not included here) of the interval between the base of the incised valley and the base of facies C or D shows that in the valley axis these facies occur about 12 m above the base. However, toward the north eastern margin , the height above the base decreases progressively to about 4 m (Figs 7 & 8). The isopachs are parallel to the trend of the valley wall. This
relationship can be interpreted to imply partial valley filling (more sand in the centre, less toward the margin) before the muddier facies C and D were deposited as a sheet that pinched out toward the valley wall. The wells that contain facies C or D appear to have no preferred locations within the field. This may be due to thrusting, particularly on the western side. Alternatively, there is some evidence that the distribution of facies C and D is masked by erosion of E5 TSE into the Ricinus sand body. Erosion of facies C and D is strongly suggested in Fig. 7 , between wells 3-19-33-7 and 1 1-18-33-7, and in Fig. 9 (wells 7-1 8-36-8 and 10- 18-36-8, and at the northern end in wells 13-26-36-9, 12-34-36-9 and 2-4-37-9). The patchy distribution of facies C and D may also be due to original depositional pinchout against the
64
R.G. Walker
Cardium Formation incised valley
eastern valley wall, as suggested in Fig. 7 (between wells 3-19-33-7 and 9-19-33-7) and Fig. 8 (between wells 16- 16-34-8 and 10-22-34-8) . Also , it is possible that some deposits of facies C and D were eroded during subsequent filling of the Ricinus valley, and before the development of ES TSE . This is suggested in Fig. 8, where facies C and/or D in well 9-16-34-8 might possibly be scoured by the sandstone with a blocky well-log response in 7-16-34-8. Although relatively few well-logs display the deeply indented pattern shown in Fig. 9 (indicative of facies C and/or D ) , and despite the limited core control , it seems likely that these facies originally extended from the northern to southern parts of the field as a sheet. This sheet averages about 4 m in thickness, and seems to pinch out against the eastern wall of the valley. The western margin is unknown. The mudstones of facies D, with their fully marine trace fauna, are apparently limited to the northern and central parts of the field (2-4-37-9 southward to 1 1-21 -34-8) . They have very sharp, and probably erosive bases in wells 1 1-21-34-8, 16-30-34-8, 2-34-36-9 and 12-34-36-9.
COMPARISON WITH OTHER CARDIUM ALLOMEMBERS
The major differences between the Ricinus ABo member and other Cardium allomembers are reviewed here. Because all of the facies at Ricinus are apparently marine, I will not make any com parisons with the non-marine Musreau Allomember. There are many asymmetrical scoured surfaces in the Cardium. The scours are linear, oriented NW-SE, and have gentle dips to the southwest and steeper dips to the northeast. Surfaces E4 and ES are typical (Walker & Eyles, 199 1 ; Pattison & Walker, 1992) . Ricinus is also a linear scour oriented NW-SE, but incomplete data for the southwestern side preclude any analysis of its symmetry.
Fig. 25.
65
The facies overlying the valley-floor scour at Ricinus are completely different from those over lying E4 and ES elsewhere. At Ricinus, the typical facies are thick, structureless and parallel-laminated sandstones. Elsewhere, the E4 and ES incised scours are overlain by up to 20 m of conglomerate (Carrot Creek Allomember; Bergman & Walker, 1987, 1988 ) , or coarse and pebbly sandstones with sets of cross-bedding 1 0 cm or more thick (Burnstick Allo member; Pattison & Walker, 1992). The coarse facies in the Carrot Creek and Burnstick Allo members are interpreted as transgressive shoreface deposits. The structureless and parallel-laminated sandstones at Ricinus are found nowhere else in the Cardium. Typical open-marine Cardium deposits consist of mudstones, and wave-rippled and HCS sand stones interbedded with bioturbated mudstones. There are no definite HCS sandstones at Ricinus. Common and abundant ichnospecies in the open marine mudstones include Chondrites, Gyrochorte, Palaeophycus, Planolites, Rhizocorallium , Thalas
and Zoophycos (Pemberton & Frey, 1984, p. 282) . This is a different suite from that found in the mudstones at Ricinus, where Rosselia is much more abundant, Palaeophycus is rare and Chondrites was not observed. Finally, the one prograding shoreface in the Cardium , the Kakwa Allomember, is dominated by swaley cross-stratification (Plint & Walker, 1987 ) . Swaley cross-stratification has not been identified at Ricinus.
sinoides
INTERPRETATION OF RICINUS
Could Ricinus be an incised shoreface?
The orientation of the Ricinus sand body suggests the possibility that it is an incised shoreface , similar to those at Garrington , Crossfield and Caroline
( Opposite. ) Sharp base of Ricinus sand body (BASE R) on underlying mudstones of the Raven River Allomember. Note lag j ust above base (L). The sandstones are parallel-laminated and structureless, and are very abruptly overlain by the bioturbated mudstones facies in the upper half of the figure, at CH. Note sudden incoming of chert pebbles (CH) up to 1 em in diameter. The top of the Ricinus sand body is marked by erosion surface E5 TSE (E5 in photograph), which is characterized by eroded sandstone clasts at the base (j ust above E5 label in photograph). The lag on E5 is about 15em thick, and is overlain by transgressive laminated mudstones (the 'laminated blanket' of Walker, 1983; LB in photograph). B =bottom, T =top. Core is 4 in. (10. 2em) in diameter, and each core sleeve is 2ft (about 60cm) long. Well 12-34-36-9, 9254-9307 ft. See comment about handwritten label 'Ricinus bottom' in caption of Fig. 22. Core is boxed top to-right and is continuous from upper to lower panels in the figure.
66
R.G. Walker
Fig. 26. Bioturbated mudstone facies with abundant scattered chert granules and pebbles (small white specks throughout
photograph), and a prominent conglomerate bed (CGL). B =bottom, T = top. Core is 3 in. (7.6 cm) wide and each core sleeve is 2.5 ft (75 em) long. Well 3-32-34-8, 9040-9066 ft. See comment about handwritten label 'Ricinus bottom' in caption of Fig. 22.
(associated with E4; Pattison & Walker, 1992), and at Carrot Creek (associated with E5; Bergman & Walker, 1987, 1988) . However, the abundance of thick beds of structureless sandstone at Ricinus is most atypical of a shoreface, where the sand is normally being constantly reworked by longshore currents, rip currents and waves. The Carrot Creek shorefaces are characterized by conglomerates, and the Burnstick shorefaces by cross-stratified coarse and granule-sandstones, and conglomerates. The prograding Kakwa shoreface is dominated by swaley cross-stratification. None of these features is present at Ricinus. Mudstones 4-5 m thick are not characteristic of shorefaces, although it could be argued that the mudstones at Ricinus formed during a minor trans gression of the postulated shoreface . However, their geometry is incompatable with this hypothesis. At
Ricinus, the mudstones appear to pinch out against the bounding discontinuity on the eastern side. If this discontinuity were interpreted as the initial trans gressive surface underlying a Ricinus shoreface, the mudstones would downlap on to this surface . This would be an almost impossible situation for a minor transgression of a shoreface ; the mudstones ought to pinch out against the western wall and continue seaward parallel to the eastern bounding discon tinuity. All of these arguments, taken together, suggest that Ricinus is not any form of incised shore face, despite the orientation of the sand body. Origin of the valley
The comparisons of Ricinus with other parts of the Cardium, along with discussions of a possible shore face origin , suggest that Ricinus is (i) not an incised
Cardium Formation incised valley
67
Fig. 27.
Zoophycos (Z) and robust Helminthopsis (H; flattened black marks) in well l l -21-34-8, 7035 ft. Core is 3 in. (7.6 cm) wide.
transgressive shoreface deposit, (ii) not a prograding shoreface , and (iii) not an open-marine HCS sand body like those in the Raven River Allomember (Walker & Eyles, 1988) . The basal incision (E5 RSE) appears to define a NW-SE trending valley at least 24 m deep, with a marine fill. I now reject my earlier suggestion that the valley was cut by turbidity currents (Walker, 1985). The Cardium contains mainly shallow-marine to shore line deposits, influenced by about eight major fluc tuations of relative sea-level . In this overall setting, the most likely origin for the Ricinus valley is fluvial , with incision during a falling stage of sea-level. The valley was subsequently filled in an estuarine setting. The basal deposits of the Ricinus sand body consist of structureless and parallel-laminated sandstones, commonly with a lag of chert pebbles, siderite
Fig. 28.
Sharp and erosive basal contact of bioturbated mudstone facies D on underlying sandstones in well 2-4-379, 9326 ft. Chert pebbles are up to 1. 7 em diameter, and siderite intraclasts are larger than 2 em. Note possible Zoophycos (Z) only a few centimetres above base of the bioturbated mudstone facies. Core 3 in. (7 .6 em) diameter.
pebbles and ripped-up mudstone clasts. There is no preserved angle-of-repose cross-bedding in sets thicker than about 5 em, nor any other facies indica tive of fluvial deposition. It therefore appears that any fluvial sediment in the valley was completely reworked by waves during initial transgression of the valley. At the same time , the valley walls were probably wave scoured and the valley widened. The
68
R.G. Walker
Fig. 29. Interbedded parallel-laminated sandstones (facies B, centre of figure), bioturbated sandstones and mudstones (facies C), and conglomerates (facies E, in 2nd sleeve) in well 3-27-36-9, 8885-8914 ft. Chert pebbles are up to about 2 em diameter. B =bottom, T =top. Core is 3 in. (7 .6 em) wide, and each core sleeve is 2.5 ft (75 em) long. See comment about handwritten label 'Ricinus bottom' in caption of Fig. 22.
interpretation of the base of the valley as a regressive surface of erosion (ES RSE) applies to the main phase of valley erosion; if the floor and walls of the valley have indeed been modified during trans gression , the surface seen today is strictly a ravine ment surface, or surface of initial transgression (IT) . P alaeogeography
There are no unquestionable data that indicate flow direction in the valley, and the problem of palaeo geography is compounded by the incomplete preser vation of the valley; both ends are faulted. The following indicators, taken together, suggest that the sea lay to the north. First, the valley appears to widen northward (Fig. 10), and the eastern wall becomes less steep. Second, the bed thicknesses of
the structureless and parallel-laminated sandstones decrease northward (Figs 17 & 2 1 ) . Third, conglom erates are best developed at the northern end of the valley (T36 R9, land sections 24-27) , but they are not cross-bedded fluvial conglomerates. They are interbedded with mudstones and sandstones with a marine trace-fossil fauna, suggesting that the pebbles were being moved by waves in the longshore drift system, and introduced into the marine end of the estuary by storms. There are far fewer granules and pebbles associated with the thick structureless sand stones at the southern end of the valley. Finally, the mudstones of facies D have a sharp base with a conglomeratic lag. In the four wells with data avail able , the clast size in the lag decreases progressively southward , from 1 .7 em to about 1 mm. This also suggests marine transgression from north to south, and hence an open sea to the north.
Cardium Formation incised valley
Fig. 30.
Chert granules and pebbles scattered within the bioturbated mudstone facies. Note granules filling distinct burrows (centre), and Zoophycos (Z) in upper part of photograph. Well 3-32-3411, 9050ft. Core is 3 in. (7.6 cm) wide.
Deposition in the Ricinus estuary
Deposition in the Ricinus estuary is not closely comparable with that of any modern estuaries, nor does it conform to any existing estuary models ( Dalrymple et al. , 1992; Reinson , 1992). The absence of angle-of-repose cross-bedding in sets thicker than about 5 em indicates the absence of two- and three dimensional dunes in the estuary. Tide-dominated estuaries (e.g. the Gironde (Allen , 199 1 ) and the Bay of Fundy ( Dalrymple et al . , 1990)) and river dominated estuaries (e.g. the Mgeni (Cooper, 1993))
69
are both dominated by dune bedforms, and presum ably would be dominated by angle-of-repose cross bedding if these estuaries were preserved in the geological record. Such cross-bedding is absent at Ricinus. Wave-dominated estuaries are characterized by a tripartite facies distribution consisting of (i) an upstream river-dominated bayhead delta, (ii) a central basin (turbidity maximum) where mudstones are deposited, and (iii) a downstream, marine dominated area where sand is introduced into the estuary from the sea. This tripartite subdivision is not seen at Ricinus, although this might be due to incomplete preservation of the entire length of the estuary. There is no suggestion of a bayhead delta prograding northward, or marine sands prograding southward; in both cases, sandier upward facies successions would be expected (but are not present) as the sands built on to central basin mudstones. Wave-formed or wave-influenced sedimentary structures (symmetrical ripples, hummocky cross stratification) are rare at Ricinus. As in many modern estuaries, this is probably due to partial closing of the estuary at the seaward end, and limited pen etration of large waves from the sea. Wave base in the Ricinus estuary may therefore have been only a few metres below sea level. The sandstones at Ricinus commonly contain evidence of rapid deposition from strong currents, with no subsequent reworking. Rapid deposition is suggested by the thick, structureless sandstones; some of the flows may have deposited beds thicker than 4 m, mostly at the southern end of the valley (T33 and T34). Other flows continued along the length of the valley to deposit structureless sand stones 50- 150 em thick in T36 and T37 (Fig. 17). The common occurrence of parallel lamination suggests that flows were rapid. The presence of Tahc and Tbe beds (Fig. 18) indicates that some of the rapid flows waned during deposition. Flow distance from the southernmost fault in T32 and T33 to the northern end of the valley (T37) is more than 45 km. The beds that contain only parallel lamination also imply rapid flows, with single depositional events forming beds up to 3 m thick in the south, and about 1 m thick in the north. The preservation of structure less, parallel-laminated, and Tabc and Tbc beds indi cates that there was no wave reworking after the original deposition of the sand. However, the estua1y was not necessarily deep because these facies have been preserved; their preservation may simply reflect the proposed shallow depth to wave base.
70
R.G. Walker
Because there is no evidence for fluvial or tidal facies in the estuary, and no evidence for wave reworking, I suggest that the structureless, parallel laminated, Tabc and Tbc beds were emplaced by density underflows generated at a bayhead delta (not observed) in the southern end of the estuary. Similar underflows appear to have been generated in the estuary of the Congo (Zaire) River (Heezen et al. , 1964) , and finer grained underflows originated from the delta of the Colorado River in Lake Mead and flowed the entire length of the lake (about lOO km; Grover & Howard, 1938). The mudstones of facies C and D appear to occur as a sheet about 4-5 m thick that spreads from end to end of the valley. In many places, the structureless and parallel-laminated sandstones grade vertically into the interbedded bioturbated sandstones and mudstones of facies C (Fig. 22) . However, the mud stones of facies D are characterized by a very sharp base, and a basal layer of chert pebbles and mudstone clasts (Figs 24, 25 & 29) . The trace-fossil fauna of these mudstones is fully marine. I suggest that after deposition of the lower sandstones, the estuary was rapidly flooded, with wave scouring of the substrate, cut-off of sediment input from the fluvial (southern) end of the estuary, introduction of pebbles by storms from the open (northern) end of the estuary, and quiet deposition of mud. Wave scouring in the estuary could have increased during flooding ( deepening ) if the barrier at the seaward end was removed. Deposition of mud was interrupted periodically by storms, introducing sand and pebbles at various stratigraphical horizons within the mudstones. The relationship between the mudstones of facies C and D is not clear. The log response suggests that both facies tend to occur at the same stratigraphical horizon (Fig. 9); however, facies C appears to have gradational contacts with facies A and B, whereas facies D has a distinctly erosional relationship with underlying facies. None of the cores shows facies C and D occurring together, and there is not enough data to determine their possible lateral relationships. After deposition of the sheet of mudstones, sand was again introduced into the estuary, and the same facies occur both above and below the mudstones. Because of the patchy distribution of the mudstones, there is a strong possibility that a falling stage of relative sea-level caused reintroduction of sand into the estuary and local scouring of the mudstones. This is tentatively suggested in Fig. 8. The depth of scouring would be at least 15 m if facies C and D in
well 16-1 6-34-8 (Fig. 8) have been cut out in well 7-16-34-8 (Fig. 8). Examination of sandstone cores adjacent to preserved mudstones showed no lag horizons or facies changes that might indicate scouring. Estuarine deposition was terminated by a final rise of relative sea-level , and the estuary deposits were truncated by ES TSE. There appears to be some topographical relief on this surface, with deeper scouring down to the west (Figs 3 - 5 & 7). The ES TSE surface is typically overlain by a thin lag of chert pebbles (Figs 13, 25 & 29) , and then by transgressive mudstones with thin silty laminations the 'laminated blanket' of Walker, 1983.
VALLEY ORIENTATION
The Ricinus valley trends parallel to the incised shorefaces in the Burnstick (Pattison & Walker, 1992) and Carrot Creek (Bergman & Walker, 1987, 1988) Allomembers (Fig. 1 ) ; an estuary would be expected to be perpendicular. The valley is incised into black mudstones and open marine HCS sand stones and bioturbated mudstones of the Raven River Allomember, and it is the only known incised valley in the Cardium (both in subsurface and out crop). The orientation of the Ricinus valley seems to require that the depositional strike in the basin changed from NW -SE (incised transgressive shore faces associated with E4) , to N E - SW (perpendicular to the Ricinus valley), and back to NW-SE (incised transgressive shorefaces associated with ES). One possibility that has been considered (and rejected) is that the Ricinus valley is a strike-parallel tributary of a main valley that trends SW- NE. The marine end of Ricinus appears to be at the northern end, so the postulated main valley would most likely be north of Ricinus. No such valley is known. It also seems unlikely that the gravel that occurs throughout the section at the northern end of Ricinus could have been emplaced into the main valley, and thence into the Ricinus valley via a right angle turn. The orientation of Ricinus remains enigmatic!
CONCLUSIONS
1 The sand body at Ricinus is defined by bounding discontinuities below and above. The lower discon tinuity defines an incised valley at least 50 km long, 24 m deep and 10 km wide. The valley trends
71
Cardium Formation incised valley
N W- S E , parallel t o regional tectonic trends and parallel to regional shoreline trends in the Cardium. 2 The valley is filled with structureless and parallel laminated sandstones. There are also mudstones that form a layer 4-5 m thick that appears to extend as a sheet along the entire length of the valley. The sheet probably pinches out against the eastern wall of the valley. The mudstones contain a fully marine trace-fossil fauna. 3 The geometry of the valley and distribution of facies suggest that the open sea lay to the north. 4 There is no evidence of fluvial or tidal processes , and wave reworking of the sandstones is rare. At least part of the fill (the structure less sandstones and the Tahc and Tbe beds) appears to have been emplaced by density currents that maintained sand in suspension, and then deposited it very rapidly. The sand was not reworked after its initial deposition. This suggests that the valley was domi nated by input of large amounts of sand from the fluvial end. This sand appears to have slumped, perhaps on a bayhead delta, and then travelled along the estuary as a density underflow. 5 A phase of marine transgression terminated sand input and led to the deposition of a valley-wide layer of marine mudstone. A subsequent drop of relative sea-level led to renewed sand emplacement, prob ably with local scouring of the mudstones. 6 The valley is interpreted as a lowstand incision formed by a river. The fill consists of a variety of facies, but broadly can be termed estuarine. 7 The orientation of a major valley 50 km long, parallel to depositional strike, remains enigmatic.
ACKNOWLEDGEMENTS
I began work at Ricinus in 1982 whilst on sabbatical leave at Amoco Canada. I thank the company for their support. Work on this paper commenced in 1993 , and I thank David James and Wascana Energy I nc. for access to well logs. The research has been financially supported by the Natural Sciences and Engineering Research Council of Canada, in the form of Operating and Strategic Grants. I particu larly thank James McEachern for his help in ident ifying and explaining the trace fossils. And finally, none of the work would have been possible without the help of Harold Reading; it was Harold who first emphasized to me in 1961 the difference between gradational and abrupt facies contacts, following the thinking of Johannes Walther and anticipating many
of the better ideas now incorporated into allostra tigraphy and sequence stratigraphy.
AP PENDIX
1
Cross-sections
The choice of a datum for the cross-sections is difficult. Possible datums above Ricinus tend to drape over the sand body. The 'Cardium zone marker' (E7 in the cross-sections) commonly used by industry and some other workers is a very bad choice because it is a mappable erosion surface with more than 40 m of relief (Wadsworth & Walker, 1 99 1 ; Figs 4 & 6-8). In this paper I have chosen a group of resistivity signatures A - D in Figs 3-8 that appear to define four sandier upward successions immediately below E4. The top of the 'black blanket' (Figs 4-8) is almost parallel to the sandier upward successions. Together, horizons A - D can be taken to approximate an original , almost fiat, sea-floor, into which the Ricinus valley is incised. All 71 cores have been measured and photo graphed. All of the off-field well-logs west of Ricinus have been used where the Cardium is present and recognizable , but some cannot be used because the rocks west of Ricinus are too deformed. Enough wells north and south of the valley (Fig. 1) have been used to define the ends of the valley, and most of the wells immediately east of the valley (between Ricinus and Caroline) have been used. Only a small proportion of Caroline wells are shown on the map (Fig. 1 ) . The cross-sections show gamma-ray and resistivity well-logs, with cored intervals shown as black bars. Depths are mostly in metres, with depth markers shown at 50-m intervals, but a few older wells are shown in feet (numbers greater than 6000, with depth markers at 100-ft intervals) . In the northern and central parts of the field , a few cores show dipping beds. However, the generally parallel nature of E 1 , successions A - D , and the top of the black blanket, shows that thicknesses in most wells are not seriously distorted by variations in dip. The southern part of the field appears to be more extensively thrusted, with more fracturing of the sandstones in core. Steeply dipping beds in section D can be identified in cores, and in those parts of the well-logs where thicknesses appear to be expanded. I have assumed that in section D (Fig. 6) the eastern most well ( 1 1-25-33-7) is undeformed because the
72
R.G. Walker
thicknesses of various stratigraphical intervals are very similar to those in the other cross-sections. Other wells in section D have been drilled into dipping rocks, and I have therefore reduced the well logs so that the thicknesses of units A - D below E4 are roughly constant across the section.
AP P ENDIX
2
Allostratigraphy
Allostratigraphy first appeared m the North American Stratigraphic Code in 1983 (NACSN , 1983 ) . It does not appear to have been used in the Western Interior Seaway of North America until 1987 (Plint et a!. , 1987, p. 366) , with reference to the unconformity-bounded members of the Cardium Formation proposed by Plint et al. ( 1 986) . An allostratigraphical unit 'is a mappable stratiform body of sedimentary rock that is defined and identified on the basis of its bounding discontinuities' (NACSN, 1983 , p . 865). I suggest (following the definitions of sequence stratigraphy) that this definition be extended to read ' . . . bounding discontinuities and their correlative conformities' . The bounding dis continuities were not defined in the stratigraphical code, but clearly include erosion surfaces (regressive and transgressive surfaces of erosion) , condensed horizons of reduced or zero deposition (commonly maximum flooding surfaces), and firmgrounds and hardgrounds. Allostratigraphy is descriptive, and does not involve any of the interpretive aspects of sequence stratigraphy. It can commonly be applied on a smaller scale than sequence stratigraphy (where a sequence is a 'relatively conformable succession of genetically related strata bounded by unconformities or their correlative conformities'; van Wagoner et al. , 1990, p. 22) . An unconformity was defined as 'a surface separating younger from older strata along which there is evidence of subaerial-erosional trunc ation, or subaerial exposure, with a significant hiatus indicated' (van Wagoner et al. , 1990, p. 22) . In the Cardium , there i s absolutely n o 'evidence of subaerial-erosional truncation' on surfaces E5, E6, E6.5 and E7, although the surfaces are interpreted to have initially formed subaerially. The 'evidence' was entirely removed during erosive shoreface transgression . Sequence stratigraphers are now recognizing the need to subdivide many stratigraphical units on a finer scale, thus giving rise to discussions of 'high-
resolution sequence stratigraphy'. This subdiscipline is converging on allostratigraphy, but unfortunately retains the concept that units are bounded by uncon formities, and that the strata bounded by the uncon formities are 'genetically related' . I n the Cardium (Fig. 2) , the only type 1 sequence boundary, with valley incision, is E5 RSE at Ricinus. If I ignore Ricinus for a moment, it is possible to define a 'sequence' between E4 and E5, by arguing that both are unconformities, and that both developed initially during subaerial exposure (E4, see Pattison & Walker, 1992; E5, see Walker & Eyles, 199 1 ) . The implication is that the rocks between E4 and E5 are 'genetically related' as a sequence. This sequence consists of coarse sand stones and conglomerates deposited in a trans gressing shoreface (Burnstick member, Fig. 2), overlain by black mudstones (the 'black blanket', Fig. 2) and very fine- to fine-grained hummocky cross-stratified sandstones in the upper part of the Raven River Allomember. There is presumably a maximum flooding surface immediately below or j ust within the black blanket. I suggest that on sedimentological grounds, there is absolutely no 'genetic relationship' between the transgressive con glomerates and the overlying (highstand) black mudstones and HCS sandstones. Water depths were very different, as were the wave climates, sediment transport processes and grain sizes of available sedi ment. The salinities may also have been different. Thus sequence stratigraphy is not the most appro priate technique for descriptive subdivision of a unit like the Cardium. Allostratigraphy does not involve problems of defining terms such as unconformity , genetically related, and evidence of subaerial erosional truncation, and allows the Cardium to be subdivided using bounding discontinuities (however they formed). The most useful bounding discon tinuities are the transgressive surfaces of erosion (E4, E5 TSE, etc. ) , and the flooding surfaces that j uxtapose relatively offshore quietly deposited mud stones on top of nearshore conglomerates (Fig. 25). The allostratigraphical units so defined correspond to transgressive systems tracts (the Burnstick Allomember and its correlative conglomerate lag on E4) and highstand systems tracts (the Raven River Allomember) ; these systems tracts do contain 'genetically related' strata. The Ricinus sand body is also defined by two bounding discontinuities; E5 RSE and E5 TSE. It is therefore defined in this paper as the Ricinus Allomember of the Cardium Alloformation. It is
73
Cardium Formation incised valley
genetically distinct both from the highstand deposits of the Raven River, and the transgressive deposits of the Carrot Creek Allomembers. The scale of allostratigraphical subdivision, like that of lithostratigraphical subdivision , is left to the discretion of the individual worker. In the Cardium , the transgressive surfaces of erosion are complicated by the fact that in many places there appear to be stillstands of sea-level during the overall trans gression. Incised shorefaces are cut during these stillstands. The shoreface is floored by the initial surface of transgression (IT), and is truncated by a surface of resumed transgression ( RT) (Fig. 2, Burnstick Allomember ) . However, the RT surface will pass laterally into another IT surface if there is another stillstand during the transgression (Fig. 2). Thus the two Burnstick sand bodies shown in Fig. 2 are defined by different bounding discontinuities, and hence would qualify as different allomembers. Such fine subdivision seems unnecessary at present; the erosional envelope beneath the sand bodies can be regarded as one bounding discontinuity, as can the contact between the RT surfaces and the over lying transgressive mudstones. Thus the Burnstick Allomember (and the Carrot Creek Allomember) consist of a series of separate sand bodies at slightly different stratigraphical horizons, but all developed during one overall phase of transgression. My final comment on allostratigraphy concerns names. The stratigraphical code does not address this problem - it simply states that 'the principles and procedures for naming allostratigraphical units are the same as those for naming of lithostratigraphi cal units' (NACSN, 1983 , p. 867). The Cardium Formation is a formally defined lithostratigraphical unit. Some workers have suggested that if an allo stratigraphical scheme is to be used, the term Cardium Alloformation is forbidden because Cardium has already been used. It is clear that if existing names are banned, the number of names will double if existing units are subdivided allo stratigraphically. This is ridiculous. I do not believe that the term Cardium Alloformation is in any way confusing. It is defined as those rocks between E l and the top of the transgressive lag o n E7, and encompasses roughly the same rocks as the Cardium Formation. It has the advantage of emphasizing a natural break in sedimentation, whereas the base of the Cardium Formation is defined at the incoming of sandstones at some arbitrary point within the first sandier upward succession. There is room for both a Cardium Formation and a Cardium Alloformation.
Alloformations and allomembers will commonly be heterogeneous, encompassing both horizontal and vertical facies changes. Thus the concept of formally defining a 'type section' needs to be re-evaluated. Also, any formal definition of an allo stratigraphical unit must include descriptions of the bounding discontinuities, so that they can be recog nized by other workers.
AP PENDIX
3
P reparation of computer maps
It has not been possible to restore all of the thrusts that displace sandstones within the field. Conse quently, some of the wells penetrate sandstones that have been thrust northeastward, and some thickness may be slightly exaggerated if the well penetrates dipping strata. The effect of exaggerated thickness is reduced in the mapping because SU RFER calculates each grid-point using data from the 10 closest wells using a kriging method. Thus the thickness distor tions are probably not great, but the maps should be stretched southwestward in the reader's imagination to allow for the foreshortening due to thrusting. It can be seen in Fig. 1 that the main sand body as defined by drilling density does not seem to be displaced laterally by the faults that define the southernmost three compartments. In my data base, the locations of the wells in the northern compart ment have been shifted about 4 km southwestward to bring them on line with the rest of the field; off field wells to the southwest were also shifted, but off-field wells to the northeast were not relocated.
REFERENCES
ALLEN, G.P. (1991) Sedimentary processes and facies in the Gironde estuary: a recent model for macrotidal estuarine systems. In: Clastic Tidal Sedimentology (Eds Smith, D.G . , Reinson. G.E. , Zaitlin, B . A . & Rahmani, R.A.). Can. Soc. petrol. Geol . , Calgary, Memoir 16, 29- 40. BERGMAN, K . M. & WA L KER , R.G. (1987) The importance of sea level fluctuations in the formation of linear con glomerate bodies: Carrot Creek Member, Cretaceous Western Interior Seaway, Alberta, Canada. J. sediment. Petrol. , 57, 65 1 - 665. BERG MA N , K . M. & WALKER, R.G. (1988) Formation of Cardium erosion surface E5, and associated deposition of conglomerate; Carrot Creek field, Cretaceous Western Interior Seaway, Alberta. In: Sequences, Stratigraphy, Sedimentology: Surface and Subswface (Eds James, D.P.
74
R. G. Walker
& Leckie, D.A.). Can. Soc. petrol. Geol . , Calgary, Memoir 15, 15-24. B o u M A , A.H. ( 1 982) Sedimentology of some Flysch Deposits. Elsevier, Amsterdam, 168 pp. CooPER, J .A.G. ( 1993) Sedimentation in a river dominated estuary. Sedimentology, 40, 979- 1 0 1 7 . DALRYMPLE, R.W., KNIGHT, R.J . , ZAITLIN , B.A. & MIDDLETON , G. V . ( 1 990) Dynamics and facies model of a macrotidal sand-bar complex, Cobequid Bay - Salmon River · estuary, Bay of Fundy, Canada. Sedimentology, 37, 577- 6 1 2 . DALRYMPLE, R.W. , ZAITLIN , B.A. & BoYD, R. ( 1992) Estuarine facies models: conceptual basis and strati graphic implications. J. sediment. Petrol. , 62, 1 1 30- 1 146. GEOLOGICAL S U RVEY OF CANADA ( 198 1a) Gas Pools of Western Canada. Geol. Surv. Canada, Map 1558A, Ottawa. GEOLOGICAL S u RVEY OF CANADA ( 1981b) Oil Pools of Western Canada. Geol. Surv. Canada, Map 1558B, Ottawa. GROVER, N.C. & HowARD, C.S. ( 1938) The passage of turbid water through Lake Mead. Trans Am. Soc. civ. Eng. , 103, 720- 790. HAMBLIN, A.P. & WALKER, R.G. ( 1979) Storm-dominated shallow marine deposits: the Fernie-Kootenay (Jurassic) transition, southern Rocky Mountains. Can. J. Earth Sci. , 16, 1673- 1690. HEEZEN, B.C., MENZIES, R.J . , SCHNEIDER, E. D . , EWING, W.M. & GRANELLI, N.C.L. ( 1964) Congo submarine fan. Bull. Am. Assoc. petrol. Geol. , 48, 1 126- 1 149. McEACHER N , J.A. & PEMBERTO N , S. G. ( 1992) lchnological aspects of Cretaceous shoreface successions and shore face variability in the Western I nterior Seaway of North America. In: Applications of Ichnology to Petroleum Exploration; a Core Workshop (Ed. Pemberton, S. G. ) Society of Economic Paleontologists and Mineralogists, Tulsa, Core Workshop 1 7 , 57-84. NACSN ( N ORTH AMERICAN COMMISSION ON STRATIGRAPHIC NoMENCLATURE ) ( 1 983) North American stratigraphic code. Bull. Am. Assoc. petrol. Geol. , 67, 84 1 -875. PATTISO N , S. A.J. & WALKER, R. G. ( 1992) D eposition and interpretation of long, narrow sand bodies underlain by a basinwide erosion surface: Cardium Formation, Cretaceous Western Interior Seaway, Alberta, Canada. J. sediment. Petrol. , 62, 292-309. PEMBERTO N , S. G. & FREY, R.W. ( 1984) I chnology of storm influenced shallow marine sequence: Cardium For-
mation (Upper Cretaceous) at Seebe, Alberta. In: The Mesozoic of Middle North America (Eds Stott, D.F. & Glass, D.J.) Can. Soc. petrol. Geol . , Calgary, Memoir 9, 281 -304. PuNT, A.G. & WALKER, R. G. ( 1987) Cardium Formation 8. Facies and environments of the Cardium shoreline and coastal plain in the Kakwa field and adjacent areas, northwestern Alberta. Bull. Can. petrol. Geol. , 35, 48-64. PUNT, A. G., WALKER, R.G. & BERGMAN, K.M. ( 1986) Cardium Formation 6. Stratigraphic framework of the Cardium in subsurface. Bull. Can. petrol. Geol. , 33, 2 1 3-225 . PUNT, A. G., WALKER, R.G. & BERGMA N , K.M. ( 1 987) Cardium Formation 6. Stratigraphic framework of the Cardium in subsurface: Reply. Bull. Can. petrol. Geol. , 35, 365-374. REINSON, G.E. ( 1 992) Transgressive barrier island and estuarine systems. In: Facies Models - Response to Sea Level Change (Eds Walker, R.G. & J ames, N.P. ) . pp. 179- 194. Geological Association o f Canada. VAN WAGONER, J.C., MITCHUM, R. M. , CAMPION, K.M. & RAHMANIAN , V.D. ( 1990) Siliciclastic Sequence Stra tigraphy in Well Logs, Cores, and Outcrops. American Association of Petroleum Geologists, Tulsa, Methods in Exploration Series, 7 , 55 pp. WADSWORTH, J.A. & WALKER, R.G. ( 1 99 1 ) Morphology and origin of erosion surfaces in the Cardium Formation (Upper Cretaceous, Western Interior Seaway, Alberta) and their implications for rapid sea level fluctuations. Can. J. Earth Sci. , 28 , 1507- 1520. WALKER, R. G. ( 1983) Sedimentology and stratigraphy in the Caroline-Garrington area, Bull. Can. petrol. Geol. , 3 1 , 2 1 3 -230. WALKER, R.G. ( 1 985) Cardium Formation at Ricinus Field, Alberta: a channel cut and filled by turbidity currents in the Cretaceous Western Interior Seaway. Bull. Am. Assoc. petrol. Geol. , 69, 1963- 198 1 . WALKER, R.G. & EYLES, C.H. ( 1988) Geometry and facies of stacked shallow marine sandier upward sequences dissected by erosion surface: Cardium Formation, Willesden Green, Alberta. Bull. Am. Assoc. petrol. Geol. , 72, 1469- 1494. WALKER, R. G. & EYLES, C.H. ( 199 1 ) Topography and significance of a basinwide sequence-bounding erosion surface in the Cretaceous Cardium Formation, Alberta, Canada. J. sediment. Petrol. , 61, 473-496.
Spec. Pubis in!. Ass. Sediment. (1995) 22, 75-99
Gravelly shoreface and beachface deposits B R U C E S . H A R T* L and A . G UY P Ll N T"! *Geological Survey of Canada, Box 6000, Sidney, British Columbia VBL 4B2, Canada; and "I Department of Earth Sciences, University of Western Ontario, London, Ontario N6A 5B7, Canada
ABSTRACT
Modern gravel-dominated shoreface deposits are poorly documented, especially from progradational settings. This makes the interpretation of possible ancient analogues very difficult. In the absence of unique diagnostic features, ancient conglomeratic shoreface deposits may be identified on the basis of: (i) the inferred relationship between sedimentary process and deposit, based on sedimentary structures, fabric and texture of the deposit, palaeocurrent data, etc . ; (ii) vertical and lateral successions of facies; (iii) gross morphology of a deposit; and (iv) stratigraphical setting. Much more is known about gravelly beachface deposits because modern analogues are easily accessible. Several distinct 'end-member' types of gravelly shorcface/beachface deposits can be recognized: (i) river mouth bar deposits, which show evidence of rapid deposition, scouring, reworking by waves and (possibly) post-depositional failure; (ii) mixed sand-gravel systems in which sedimentary structures within the sandstones allow determination of depositional environment; (iii) swash-aligned gravelly systems dominated by ncar-perpendicular wave approaches, which show shore-normal facies zonation; (iv) driji-aligned gravelly systems , which show evidence of pronounced longshore transport generated by oblique wave incidence; and (v) transgressive lag deposits. Observations and interpretations of ancient conglomeratic shorefore deposits raise process-related questions that have yet to be addressed by studies of modern equivalents.
INTRODUCTION
depositional environments must rely on interpret ation of the processes which gave rise to the observed lithofacies , and relating these interpreted processes to stratigraphical context, partially analogous modern environments, and the three-dimensional morphology of the deposit. Progradational gravel-dominated shoreface and beachface systems are not common, and are rarely documented. Three 'end-member' settings for grav elly shorefaces may be identified. The first involves areas of active erosion of coastal bedrock outcrops (e.g. Zenkovitch , 1967; Ogren & Waag, 1986). The second setting is in high-latitude areas that were subject to Pleistocene glaciation (e.g. Forbes & Taylor, 1987 ) . There, gravel can be derived from tills, outwash deposits, or other coarse-grained sub- to proglacial deposits. Thirdly, progradational
It is commonly accepted that one of the most fruitful approaches to facies analysis involves comparison of modern sedimentary environments and potential ancient analogues. This approach, based on the uniformitarian principle that the present is the key to the past, forms the basis of most texts dealing with facies models and its roots can be traced back for at least a century to Johannes Walther. Walther advocated an 'actualistic' approach to the examin ation of ancient sedimentary deposits, but knew that ancient facies might exist which have no direct modern analogues (Middleton, 1973) . In such cases, as explained by Reading (1986), reconstruction of 1 Present address: Department of Geosciences, Penn sylvania State University, U niversity Park, Pennsylvania 168 02, USA.
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
75
76
B.S. Hart and A . G. Plint
gravelly coastlines may develop where the shoreline is adjacent to uplands yielding abundant coarse grained sediment. Most modern examples of such coastlines are associated with fan-deltas. In fan delta settings, gravelly shorefaces develop in a 'tran sition zone' between alluvial plain and offshore settings (e .g. Wescott & Ethridge, 1980) . Holocene sea-level rise has inhibited widespread develop ment of progradational systems, and therefore most studies of modern gravelly coasts have involved transgressive gravelly shorelines along bedrock coasts and in glaciated areas. Given the relative scarcity of gravel dominated coastal deposits, and the logistical difficulties involved in their investigation, it is no surprise that modern gravel-dominated shorefaces represent a type of environment that has received very little attention from sedimentologists. Modern gravel beachfaces have been the object of considerable study (Bluck , 1967; Carr, 1969; Dobkins & Fol k , 1970; Carter & Orford , 1984; Forbes & Taylor, 1987 ) , but there have been very few studies of the movement of pebble-sized material on the shoreface (Neate, 1967; Mathews, 1980; Gillie, 1983) and the nature of sedimentary structures present on modern gravelly shorefaces is poorly documented. Studies of gravel on modern inner shelves (e.g. Gillie, 1983; Forbes & Boyd, 1987) are helpful in this regard , and gravels have also been reported from modern sand dominated shorefaces (e . g . Hunter et a/., 1979; Shipp, 1984; Ross & Long, 1989 ) . Ancient conglomeratic shoreface/inner shelf deposits have been recognized in outcrop (e.g. Clifton, 1973, 198 1 ; Leckie & Walker, 1982; Leithold & Bourgeois, 1984; Leckie, 1988; Massari & Parea, 1988; Cheel & Leckie, 1992) but in most of these cases, pebble-size material forms a small pro portion of a sand-dominated sedimentary deposit. Kleinspehn et a!. ( 1984), Massari & Parea ( 1988) and Massari et a!. ( 1 986) described progradational shoreface and beachface gravels associated with fan delta settings. Conglomerates have been described using subsur face data (Bergman & Walker, 1987; Arnott, 199 1 ) and interpreted a s shoreface deposits o n the basis of stratigraphical position and the geometry of the conglomerate bodies as a whole. Ancient con glomeratic beachface deposits are clearly recogniz able, both in outcrop and core (e.g. Nemec & Steel 1984; Plint & Walker 1987; Postma & Nemec, 1990) . This paper will review the sedimentology of gravel dominated shoreface and beachface deposits. M uch
of the data presented was drawn from studies of conglomerates in the Upper Cretaceous Cardium Formation of western Canada (Piint & Hart, 1988; Hart & Plint, 1989, 1 99 1 ; Hart, 1990, 199 1 ) . In the Cardium Formation, sedimentological studies are complemented by extensive stratigraphical studies (e.g. Plint et a!., 1986; Bergman & Walker, 1987; Hart, 1990; Hart & Plint , 1993a ) . This work has demonstrated that the chert-pebble conglomerates in the Cardium were deposited many tens to hun dreds of kilometres basinward of any potential source -they are therefore not interpretable as alluvial fan or fan-delta deposits.
SEDIMENT TRANSPORT PROCESSES
There is considerable disagreement between authors and between disciplines (e.g. geologists and coastal geomorphologists and engineers) about the termin ology used to describe coastal zonation. The shore face is defined here as the broadly concave-upward zone where sediment is transported by wave pro cesses on a regular basis (i.e. between low-tide level and 'fair-weather wave base'; Fig. 1). The upper portion of the shoreface is the region of breaking waves (surf ) or 'nearshore zone' of coastal geomor phologists (see Komar, 1976; Dubois, 1992 ) . The lower shoreface extends seaward from the breaker zone to the depth at which shoaling surface gravity waves no longer entrain sediment (sometimes des cribed as the 'transition zone' between the nearshore and offshore; Komar, 1976; Dubois, 1992). In the stratigraphical record , the vertical transition from interbedded sandstones and mudstones to thick sand stones (and/or conglomerates) in ancient prograding shoreline successions is typically defined as the tran sition from inner shelf to shoreface deposition (e.g. Plint & Walker, 1987 ) . The width and maximum water depth of the shoreface and nearshore are a complex function of sea-floor slope, wave climate and tides, all of which vary between sites and over the course of time at any given site (Wright & Short, 1984) . The beachface comprises that region between low-tide level and the landward limit of wave action (Fig. 1 ) . There exists abundant literature concerning shoreface and beachface processes (e.g. Zenkovitch, 1967; Komar, 1976; Niedoroda et a/., 1985; Carter, 1988) , and an extended discussion will not be presented here. Numerical analysis of the relationship between flow parameters and the movement of gravel in
Gravelly shorefaces
77
SHOREFACE
�....:::--
Swash ------------------- limit
--
����-----�--�Depth at breaker zone
Fairweather
Mean low tide
wave base
Fig. l. Schematic diagram showing coastal zone terminology used in this paper.
shallow marine settings has not advanced to the point where predictive estimates of sediment trans port can be made confidently. This shortcoming can be related both to the logistical problems of working in the high-energy shoreface zone and the problems inherent in understanding the movement of pebble size materials in general (see below). Bulk textural properties of a deposit, clast form and fabric all appear to play important roles, which are not (yet) fully quantifiable. Williams et al. ( 1989 ) , for exam ple, used sophisticated instrument packages to study the movement of marine gravels, yet felt compelled to state that 'caution is required if attempting to deduce bedload transport rates for marine gravels with significantly different grain size distribution or in flow regimes radically different from the present' . Given these uncertainties, only a qualitative over view of the subject will be presented in this section. Shoreface processes
In a general way, there are three main forces respon sible for sediment transport on the shoreface. These are: (i) wave motion; (ii) currents induced by shoaling waves; and (iii) other currents. During wave shoaling, progressive distortion of sinusoidal deep-water waves results in the forward motion under the wave crest becoming shorter but faster, and the backward motion under the trough becoming longer and slower. This time-velocity asymmetry has important consequences for sedi ment. As described by Zenkovitch ( 1967) and Bruun ( 1988) , the shoreface profile is a function of sediment grain size and wave type (height, wavelength, period, form ) . For any particle resting on the shoreface there are forces acting to move the particle upslope (stronger but shorter landward motion under the crest) and downslope (slower but lon2:er seaward
motion under the trough, plus gravity) . For a given grain size, slope and wave conditions, there will exist a point at which the upslope and downslope forces are balanced, and forward motion will equal the backward motion . Seaward of this point, sediment will migrate offshore, whereas further landward, sediment will migrate onshore. This effect is grain size dependent: for the same wave conditions and bottom slope, fine-grained sediments may move off shore whereas coarse-grained sediments move onshore (Zenkovitch, 1967) . When a mix of grain · sizes is available on the sea-floor, a graded profile can be developed with large particles moving onshore and smaller particles offshore (Zenkovitch, 1967 , p. 1 15 ) . Shoreface profiles are a function o f several factors, including wave height, period and sediment grain size. Although gravelly systems appear to respond less quickly to changes in wave regime than their sandy counterparts (Carter, 1988), Hart & Plint (1989) concluded that gravel on the shoreface at Chesil Beach moved offshore during storms then onshore during fair-weather conditions. Gillie ( 1983) described both onshore and offshore movement of gravel on the inner shelf in response to differing wave conditions. Kidson & Carr ( 1959) used radioactively 'tagged' pebbles and found that at some critical (but undefined) distance offshore, pebble movement was very limited (even during storms) and exchange of gravel between the shoreface and shelf did not occur. I ntuitively, this 'critical' distance must be related to the depth at which oscillatory wave motion can initiate movement, and is therefore a function of sediment size, submarine slope and wave climate. Longshore currents (sensu stricto) are generated over the upper shoreface by the dissipation of the longshore component of energy in breaking waves. The upper shoreface, owing to the high energy levels
78
B.S. Hart and A. G. Plint
and potential for unidirectional longshore currents, is considered to be a major sediment transport zone for coarse-grained sediments (e.g. Carter & Orford, 199 1 ) . Rip currents are another type of current induced by breaking waves. Channels can be excavated by these seaward flows (Davidson-Arnott & Greenwood, 1976; Gruszcyriski et a!., 1993) and pebbles and even boulders can be transported off shore (Zenkovitch, 1967; Gruszczynski et a!., 1993 ) . In sandy settings, coarser material accumulates as lags in the base of the rip channels (Davidson-Arnott & Greenwood, 1976; Hunter et a!., 1979). Rip cur rents are not typically well developed on steep, 'reflective' shorelines such as those associated with most modern gravelly shorelines (Carter, 1988) . However, it must be remembered that these are generally transgressive systems and progradational shorelines may have different profiles. Tidal, wind-induced and fluvial currents may also transport sediment on the shoreface . In wave dominated settings, perhaps the most important role of tides is to change water depth and hence the geographical limits of the surf and swash zones (Ross & Long, 1 989) . Currents generated by coastal set-up (with onshore winds) flow offshore along the sea-floor in the shoreface zone, and thus enhance offshore sediment transport during storms (Niedoroda et at., 1985). Close to river mouths, interaction of fluvial and shoreface processes is to be expected. Gravity-driven mass transport processes may occur in 'mouth bar' settings (Kleinspehn et a!., 1984) as rapid deposition leads to oversteepening and resultant instability.
Finally, in some settings, seaweed can enhance both landward and seaward transport (Woodborne et a!., 1989 , and references therein ) , although the geological significance of this process is hard to assess. Pebbles and shell material attached to hold fasts are commonly observed along beaches adjacent to seaweed beds (Fig. 2 ) . Seaweed communities are best developed along rocky or gravelly coasts in cool waters of high latitudes (Carter, 1988). Once trans ported out of its initial habitat, the plant dies and decays, leaving no lasting trace of the holdfast on the clast. Woodborne et a!. ( 1 989) reported kelp rafting of a 30-kg boulder, and kelp rafted pebbles in 350 m water, 120 km offshore, and also reported that about 15% of the kelp along the South African coast is torn from the bottom during storms each year. If only 1% of the algae carried clasts, pebble transport by this process could be 'significant' over geological time (Woodborne et a!., 1989 ) . Beachface processes
Beachface processes along gravelly coasts are better documented than those below low-tide level. The roughness of gravel beachfaces can reduce the velo city and distance travelled by the swash, relative to sandy systems (Carter, 1988). The high porosity and permeability and low water table of gravelly beaches also reduces the run-up distance of swash, and reduces or eliminates backwash by allowing water to freely percolate into _the sediment. When the water table is high, for example during falling tides or storms, surface clasts on the beachface are more susceptible to movement by wave swash and back-
Fig. 2. Clasts attached to seaweed holdfasts on a transgressive shoreline, Island View Beach, British Columbia. Lens cap for scale.
Gravelly shorefaces wash . Forbes & Taylor ( 1 987) reported that beach morphology is sensitive to the amount of sand in the system. Oblique swash motion on the beachface , generated by oblique wave approach , transports clasts alongshore, with grains returning directly sea ward during the backwash. This alongshore 'zig-zig' motion should not be confused with longshore cur rents in the surf zone, even though both processes transport clasts along the shoreline. Overwash may be important during storms on some gravel barrier systems (Carter & Orford , 198 1 , 1984) . Pebble-filled streams and drainage rills (developed by groundwater flow during falling tide) may cut
Fig.
3. Streams cutting across beachface. (A) Pebble-lined stream cutting across sandy beachface, China Beach, British Columbia. Note steep, locally overhanging channels walls. View toward back beach area. Note 'pulse' of gravel in background being transported seaward through the channel. (B ) Large drainage rills incised into Chesil Beach, England.
79
across sandy beaches (Clifton et al., 1973; Fig. 3A) . The permeability of gravel-dominated systems tends to limit the development of such small channels, although they may nevertheless develop (Fig. 3 B ) . Bluck ( 1 967) suggested that particles can b e trans ported through the sediment body in certain portions of coarse-grained beachfaces. Bluck suggested that the backwash percolating through the gravel appeared to 'flush' finer particles seaward. How ever, unless the beach consisted of an open, clast supported framework of pebbles and cobbles, the sieving process is unlikely to operate.
80
B.S. Hart and A . G. Plint GRAIN SIZE AND SHAPE
As with sandy deposits, the textural characteristics of gravels reflect the processes responsible for their deposition. Unfortunately, gravels do not readily lend themselves to the same types of routine size analyses as sands, and therefore much less is known about their grain-size distribution. On the other hand , the size of gravels makes them much more amenable to analysis of clast shape, usually expressed as form indices derived from measurement of long ( L), intermediate (!) and short ( S) axes. As noted in the preceding section, accurate predictions of entrainment velocity are very difficult to quantify as a result of sorting and clast shape effects. Textural trends
The threshold conditions for movement of a given clast depend not only on flow parameters but also on clast shape, fabric of the deposit and the size of the grain relative to that of the host deposit (Komar & Li , 1986) . Although the largest clasts are sometimes used to evaluate flow competence , such grains on the surface of a deposit project higher into the flow and are thus subjected to higher shear stress than the bulk of the sediment. Komar ( 1 987) suggested that flow competence is therefore best evaluated by relating the bed shear stress and the ratio of the diameter of the largest clast to the median diameter of the deposit as a whole . Maejima ( 1 982) found, however, that mean grain size and D10 show very similar trends across gravelly beaches, lending some justification to the use of D10 (D10 =intermediate diameter of the 10 largest clasts ) . Williams & Caldwell ( 1 988) suggested that the short axis of clasts was the most 'hydraulically sensitive' length on gravel beaches. Further investigation appears warranted. In the previous section it was noted that graded shoreface profiles are expected to develop under equilibrium conditions, with larger clasts moving onshore. A vertical section through a progradational shoreface deposit might therefore display an upward coarsening trend . To date, the only attempt to quan tify such a trend appears to be the work presented by Hart & Plint ( 199 1 ) on the Bay Tree outcrop of the Cardium Formation. Representative measured sections through two, well-exposed examples of grav elly shoreface deposits in the Cardium Formation are presented in Fig. 4. Vertical trends in L (long axis length) , I (intermediate axis length) , S (short
axis length) and D10 from two measured sections at the Bay Tree locality are shown in Fig. 5. Regression lines from D10 in section BT2, and L and I from both sections have positive slopes, indicative of an up-section increase in grain size, and high correlation coefficients. These results suggest that upward coarsening trends (not necessarily 'recognized' by D10) may develop on some gravelly shorefaces, although the lack of a distinguishable trend for the S axis suggests that this dimension may not have been hydrodynamically significant at Bay Tree. Dupre et al. ( 1980) , in a study of a Pleistocene, mixed sandy and gravelly shoreface system, reported that the coarsest clasts were to be found within deposits of the upper surf zone. A comparable up section increase in the thickness, abundance and lateral continuity of conglomerate beds in sandstone dominated shoreface deposits can be seen in some sections of the Cardium Formation (Fig. 6). Some studies of beachface deposits indicate that the coarsest sediments are found on the uppermost (landward) part of the beachface (Biuck, 1967; Kirk, 1980; Massari & Parea, 1988) , suggesting that upward-coarsening trends should develop in pro gradational systems. Maejima ( 1982 ) , however, reported that the coarsest sediment was present in a zone of berm accretion on the beachface. In mixed sand-gravel systems, the plunge point (at the toe of the beach during fair-weather conditions) may also be a position where coarse sediment accumulates (Dupre et al., 1980; Kirk, 1980; Short, 1984). Well sorted and stratified very coarse sandstones and granule conglomerates are present as beachface deposits in the Cardium Formation (Piint & Walker, 1987; Arnott, 199 1 ) . Gravelly shorefaces and beaches may reveal along shore trends in grain size generally taken to represent downdrift fining. Two examples from outcrops of the Cardium Formation show that such trends may not always be clear (Fig. 7 ) . Pattison & Walker ( 1992) used along-strike grain-size variations to infer sediment dispersal direction in pebbly sandstone shoreface deposits of the Cardium Formation in the subsurface. This type of trend was inferred to develop in response to preferential longshore trans port of finer grades, with smaller clasts 'out running' coarser sizes. Although such processes are docu mented from modern settings, Carter ( 1988) has emphasized that downdrift coarsening may also develop under some circumstances. Gravels and conglomerates typically consist of at least two grain-size classes of material, gravel clasts
Gravelly shorefaces
81
BT7
BT6 BT2
E
0
Ravinement Surface (Hart & Plint 1993a)
---r-t:l;;;;::--..:.o
A
MNA·IO
MNA·9
, __
0
B
Fig.
.-�Channelized base
/
4. Measured sections of Cardium Formation. (A) Shoreface conglomerates at Bay Tree (Alberta), represented by seven measured sections along cliff face oriented nearly parallel to depositional strike. Total distance between BT7 and BT4 about 300m. Nearly constant thickness along strike suggests tabular geometry. Beach face deposits are preserved at top of BT7. (B) Shoreface/mouth-bar conglomerates at Elephant Ridge (British Columbia) . Sections measured along diff face at slight angle to depositional strike. Total distance along strike about 200m. Note suggestion of channel form, with conglomerates scouring into u nderlying swaley cross-stratified shoreface sandstones. Sections hung on coarse-grained wave rippled surface inferred to be a nearly planar marine flooding surface (Hart, 1990) . Facies codes: A, clast-supported pebble conglomerate; B, imbricated clast-supported conglomerate; C, graded conglomerates; D, decimetre-thick cross-bedded conglomerates; E, matrix-supported conglomerate; F, interstratified conglomerate and sandstone (beachface deposits); G, amalgamated conglomerates; H, matrix-supported pebbly granule conglomerate; I, poorly sorted pebbly sandstones with localized shallow scour-fill conglomerates; 16, swaley cross-stratified sandstones; 17, cross-bedded sandstones, locally pebbly.
82
�• • • • I � iJ !IL�I � (A) BT2
100 80 60 0 40 � 2 o
0�
-----
�
�
.g
I
I
B.S. Hart and A. G. Plint
•• • ••
(B) BT7
•
I
100
: jJ�� • r: .. � �••• • I = r: I
"' c 0
·o;
100 80 E 60 40 20 I 0 o 25
. .
X
�
•
. ..
100 80
�
•
.
_,
.
.
..
I
50
0
75
100
L_ ..1. I __JI _..1. ----=L " O _25- 50_ 7 5 _ 100
Non-dimensional height above base
and sand and/or mud matrix. The proportion of each defines whether the deposit is open framework (lacking matrix completely) , clast supported (clasts in contact with spaces infilled by matrix) or matrix supported (clasts 'floating' in matrix). These differ ences can have significant economic importance, as they directly affect reservoir properties such as porosity and permeability (e.g. Clarke, 1979; Cant & Ethier, 1984; Arnott, 1991). There is a consensus that shoreface and beachface conglomerates tend to be better sorted than fluvial conglomerates, and that sand and gravel tend to be better segregated in deposits of shallow marine environments (e.g. Emery, 1955; Clifton, 1973). It has been suggested that rapid deposition of sand and gravel in fluvial deposits may produce pebbly sands, whereas prolonged reworking of a mixture of sand and gravel by waves will tend to segregate the two constituents (Clifton, 1973). Well-sorted, clast supported conglomerates are common in shoreface conglomerates at the Bay Tree locality and elsewhere in the Cardium Formation (Figs 4 & SA & B ) . Arnott (1991) examined a linear conglomerate body of the Cardium in the subsurface. Although dominantly clast-supported, matrix-supported con glomerates were also present but tended to be con centrated preferentially along the landward margin
Fig.
5. Trends in non-dimensional L, S, I, and 010 versus non dimensional height above base from two sections at Bay Tree , Alberta. See text for explanation. Relative section locations shown in Fig. 4A.
of the body. The matrix-supported conglomerates were interpreted to have been deposited in the vicinity of fluvial distributary mouths. Although these results and others suggest that texture (in particular sorting) is a valuable tool for discriminating between fluvial and littoral conglom erates, texture alone cannot be used to distinguish between fluvial and shoreface conglomerates (Nemec & Steel, 1984). Matrix- or clast-support of conglomerates should not be considered diagnostic of, respectively, fluvial and littoral deposits. For example, Howell & Link (1979) found that, in the Eocene of California, non-marine conglomerates tended to be clast-supported (in places open frame-· work), whereas wave-influenced deposits include both clast- and matrix-supported conglomerates. Leithold & Bourgeois (1984), N ielsen et al. (1988) and Massari & Parea (1988) have also reported matrix-supported conglomerates interpreted to have been deposited in a shoreface setting. Studies of fluvial conglomerates suggest that they too are typi·· cally clast supported (Harms et al., 1982). Low·· wave-energy shorelines may lack sufficient energy to effect wave segregation of sand- and pebble-size material. Finally, the nature of the sediment supplied to the shoreline (ratio of sand to gravel, fluctuations of this ratio through time) and sediment transport
CJ ;:; '<:: "'
� "' ;::.. 0
� .,
8
Fig. 6. Up-section increase in thickness of shallow, pebbly scour fills in swaley cross-stratified shoreface sandstones from a single outcrop of Cardium Formation at Cutbank Lake, Alberta. Photographs (A)-(D) represent a vertical interval of about 6 m . Conglomerates are interpreted as rip-current and longshore trough deposits. Pocket knife (arrowed) for scale.
00 w
84
B.S. Hart and A. G. Plint
N
Vl c 0 ·.::; "'
>
Q;
Vl .c 0 .....
0
Q;
.c
E
::J z
BT4 A
9
I
q C? I
9
s
MNA-10
N
20 30 40 50 60 70
010 (mm)
MNA-9
infiltrate? Comparison of the size of clasts inferred to have been rolled on the bed with the size of the sand in the matrix is sometimes used as a starting point (Harms et al., 1982) . The application of this approach to shoreface conglomerates is difficult because: (i) shallow marine settings are typified by combined flows (especially during significant sedi ment transport events such as storms) and criteria for the threshold of movement under such conditions are not easily established (especially for gravels), and (ii) this approach is probably not valid for high sediment concentrations or transport rates such as might typify the surf zone. Textural evidence can sometimes provide clues as to the origin of the matrix. For example, in the Cardium Formation at Bay Tree, a sandy matrix is common in clast-supported shoreface conglomerates but is typically absent from adjacent finer (granule or fine pebble) conglomerates (Fig. SA) . This may suggest that the sand was able to infiltrate the larger pore throats of the coarser deposits but not through the more restricted openings between smaller clasts.
· MNA-8
MNA-6
Clast shape MNA-4
MNA-3
s
MNA-2
10 B
20
30
40
010 (mm)
50
Fig. 7. Longshore trends in D10 in shoreface conglomerates of Cardium Formation at: (A) Bay Tree and (B) Elephant Ridge (Mount Niles), showing a weak, southeasterly fining trend, which, based on regional palaeocurrent data (Hart, 1990; Hart et a/., 1990) is the dominant longshore transport direction. Each 'observation' at Bay Tree corresponds to a D10 value calculated from individual beds in each section. The D10 values at Elephant Ridge represent averaged values for each measured section. The location of sampling points is shown in Fig. 4A & B .
mode (see next section) must also be considered. Carter & Orford (1991) emphasized that episodic supply can induce textural changes that are inde pendent of process regime. The origin of a sandy matrix in a conglomerate can be problematic. Was the sand transported and deposited with the gravel , or is the sand a later
Many studies have suggested that form indices can be useful for distinguishing between 'beach' and 'fluvial' pebbles (Cailleux, 1945; Dobkins & Folk , 1970). The study by Dobkins & Folk ( 1970) suggested that the 'maximum projection sphericity' and 'oblate- prolate index' were useful discrimi- nators for distinguishing between beach and river clasts and several studies have since investigated the usefulness of these criteria (partial summary in Hart 199 1 ) . Hart (199 1 ) , using pebble shape data from Bay Tree and Chesil Beach , concluded that no form parameter derived to date can be used to unambigu ously distinguish shoreface or even beachface gravels and conglomerates from fluvial deposits. Thus, although pebble-shape segregation does develop in fluvial and beachface environments, and pebble zonation does develop on individual beaches (Orford, 1975), shape criteria alone cannot be con sidered diagnostic of depositional environment . Although Howard ( 1 992) was able to show statistical differences between shape indices of ancient fluvial and beach gravels, considerable overlap was present between fields and none of the 'cri tical values' derived from studies of modern gravels provided a successful means of differentiating beachface conglomerates.
Gravelly shorefaces
85
Fig.
8. Clast-supported pebbly shoreface conglomerate. (A) Bay Tree outcrop; note open framework portions. (B ) Core ( 10-18-63-5W6) note crude stratification and possible imbrication to right.
B EDFORMS AND SEDIMENTARY STRUCTURES
In most ancient deposits, sedimentary structures provide the key information for palaeoenvironmen tal reconstruction. Knowledge of the types of bed forms and morphological elements in analogous modern environments is therefore critical . Unfortu nately little is known about the variety of bedforms that may form on gravelly shorefaces. Shoreface
Gravel wave ripples (Fig. 9A & B ) , formed by oscillatory water motion are excellent indicators of subaqueous deposition at depths above storm wave base but can be formed at considerable depths, well beyond the limit of the shoreface. Leckie (1988) reviewed existing literature on the generation and environmental significance of 'coarse-grained' rip ples. Although most examples appear to be from shelf settings, gravel wave ripples have been reported from modern (Shipp , 1984; Hart & Plint, 1 989} and ancient (Leithold & Bourgeois, 1984; Massari & Parea, 1988} gravelly shorefaces. Crestlines tend to
be nearly shore-parallel due to wave refraction (Fig. lOA}, and both onshore and offshore migration (suggested by asymmetric profiles) have been reported. Gravel wave ripple form sets may be preserved , especially when the bedform is draped by finer-grained sediment. Cross-bedding may be produced by migration of the bedform under the influence of asymmetric wave motion or wave motion with weak superimposed unidirectional currents. Bi- and polymodal fabrics with modes oriented in onshore and offshore directions may indicate the presence of shore-parallel symmetrical gravel wave ripples (ct. Leckie & Walker, 1982) even if form sets are not observed (Hart & Plint, 1989 ) . Alternatively, migrating gravel wave ripples may produce fabrics with dominant dips in the direction of ripple migration (Hart & Plint, 1989; Fig. lOB) . In the absence of preserved form sets, it may be difficult to distinguish landward dips generated by onshore migration of gravel wave ripples from those produced in gravels by strong unidirectional offshore flows (cf. Cheel & Leckie, 1992) . Nearshore bars of various types are also formed by wave processes. These bars develop on gently dipping shorelines which tend to be sand-dominated
86
B.S. Hart and A. G. Plint
Fig.
9. Coarse-grained wave ripples developed in poorly sorted coarse sandy conglomerate; Cardium Formation on Elephant Ridge. (A) Symmetrical crestlines and (B) interference ripples. Scale bar in centimetres.
(Carter, 1988). Large, broadly convex upward 'macroforms' apparently representing nearshore bars have been reported from shoreface successions comprising sandstones, pebbly sandstones and con glomerates by Leithold & Bourgeois (1984), Massari & Parea (1988) and Massari eta!. (1986). Such forms are found in the upper shoreface (Davidson-Arnott & Greenwood , 1976) and may or may not migrate. Migratory bars centimetres to decimetres in height may move onshore and (particularly in tidal settings) weld themselves to the beachface, typically produc ing onshore oriented tabular to sigmoidal cross-beds
capped by planar stratified sandstones or pebbly sandstones (Fig. llA & B ) . Troughs associated with longshore bars have been recognized (Leithold & Bourgeois, 1984; Massari & Parea, 1988) as broadly lenticular conglomerates with concave upward bases cut in sandstones. Such troughs should be elongate and nearly shore-parallel. Ross & Long (1989) cored a modern nearshore bar system in a mixed gravel and sand setting and found that gravel was not restricted to the troughs, but could also be found on the bar crest. The presence of longshore bars and troughs from ancient gravelly
Gravelly shorefaces
87
8
A
Mean x-bed orientation
Mean crestline orientation
Crestline orientation
Fig.
10. Summary of palaeocurrent information from Elephant Ridge section of Cardium Formation (Fig. 4B) . (A) Coarse-grained wave ripple crestlines (a) and decimetre-scale cross-bedded conglomerates (b) . Wave ripples indicate NW-SE trending shoreline, in agreement with other data from the Cardium Formation in ths area (Hart, 1990) . (B) Pebble fabric from a near-symmetrical coarse-grained wave ripple at Elephant Ridge suggesting onshore migration of the wave ripple form.
shoreface deposits suggests that steep, 'reflective' profiles (typical of modern transgressive gravelly shorefaces) were not developed at the time of depo sition (Wright & Short, 1984). Broad, pebble-lined scours that are oriented per pendicular, or nearly so, to the palaeoshoreline in shoreface sandstones may be interpreted as rip-current channels (e.g. Fig. 6C; Leithold & Bourgeois, 1984; Massari & Parea, 1988; Gruszcyzynski et a/., 1993 ) . Shell debris may form a significant component of such lags, facilitating the recognition of depositional environment. Cross-bedded sandstones and conglomerates are common in shoreface deposits, and may be oriented onshore, offshore or alongshore (Dupre et al., 1980; Leithold & Bourgeois, 1984; Massari & Parea, 1988). These structures may be the product of asymmetrical wave motion, and rip, longshore, tidal and (close to river mouths) fluvial currents . At the Bay Tree section of the Cardium Formation, cross-bedded conglomerates several decimeters thick (with cross bedding defined by size or textural variations) are present. This large-scale cross-bedding (Fig. 12A) was apparently produced by the migration of dune like forms at least 60 em high. Palaeocurrent evi dence indicates that these bedforms and similar cross-sets in other outcrops of the Cardium migrated in a shore-parallel direction (Figs lOA, 12B & 13A ) . Such bedforms have not been documented from modern gravelly shorefaces, although B ujalesky & Gonzalez-Bonorino (199 1 ) reported asymmetric 'washed-out (gravel) dunes' up to 10 em high with
wavelengths typically about 5 m and crests oriented at high angles to the shoreline from the lower beach face (intertidal) . These bedforms appeared to have been produced originally by longshore currents at high tide. At Bay Tree (Fig. 4A) over 50% of the conglomer ate is organized into crudely bedded, amalgamated units, 1-3 m thick. Centimetre to decimetre thick, massive clast-supported pebble conglomerate beds are the primary constituent of the amalgamated units, but open framework , imbricated, matrix supported conglomerates and thin sandstone beds (generally < lO cm thick) are also present. Shallow scours, depositional thinning and lateral facies changes limit the lateral extent of individual beds to a few metres. Despite the internal scouring, the overall appearance of the conglomerate, revealed especially by the sandstone beds, is of a pronounced horizontal stratification in the strike sections at Bay Tree (Fig. 14) . Presumably, 'clinostratification', such as that illustrated by Massari & Parea ( 1 988), would be observable in shore-normal sections. Most of these shoreface conglomerates can be interpreted as the products of gravel bedload sheets such as described from gravelly fluvial systems (Hein & Walker, 1977; Whiting et a/., 1988). The lateral dimensions of these 'bedforms' were probably of the order of several metres, explaining the lateral impersistence of individual conglomerate beds in the amalgamated deposits. The possible existence of this type of sediment transport has been suggested by Mathews ( 1980), who found that tracers on a
88
B.S. Hart and A. G. ?lint
Fig.
1 1. Migratory bars of upper shoreface and beachface. (A) Poorly sorted pebbly sand , Island View Beach, British Columbia ( B ) Sigmoidal, landward-directed sandstone cross-set in upper portion of Cardium Formation shoreface conglomerates at Bay Tree, Alberta.
Gravelly shorefaces
89
Fig.
12. Decimetre-scale cross bedding in Cardium Formation shoreface conglomerates. (A) Bay Tree (scale= 15 em) and (B) Elephant Ridge (pocket knife for scale) . In both cases, migration direction was alongshore, to the southeast.
modern gravelly foreshore moved alongshore on the lower beachface (i.e. in the surf zone at high tide) as discrete 'slugs' of sediment representing volumes of 1-10m\ rather than as a uniform gravel sheet. These 'slugs' moved 10- SOm alongshore during periods of high wave energy, but were reworked during fair-weather conditions. One may speculate that, during strong longshore flows, the bedload sheets may have grown in height and developed slipfaces, such as has been observed in modern gravelly streams (Whiting et at., 1988) . Because of the amalgamated (and hence indistinguishable) con tacts between individual beds ('flow units') , the use
of quantitative techniques such as plots of maximum clast size versus bed thickness or Markov chain analyses become impractical. Pebble fabrics from the shoreface portion of the conglomerate at Bay Tree are typically polymodal (Fig. 13B ) . In most cases, the principal mode is towards the south or southeast (alongshore) . Other modes are observed at about 90° and 180° to the principal mode in nearly all cases. All of these samples represent crudely stratified to cross-bedded clast-supported pebble conglomerates generally from the middle to lower portions of the section. The dominant dip direction is approximately that of
(A�
90
�
B.S. Hart and A. G. Plint (B)
(C)
(E)
(F)
�
�
� (G)
~
(Hl
(I)
� �
Fig.
13. Palaeocurrent data from Bay Tree section. (A) Cross-bedded sandstones and conglomerates. (B- I ) Pebble fabrics from conglomerates. Pebble fabrics are for sites labelled in Fig. 4A: (B) BTI-11 (beachface conglomerates); (C) BTI- 15 (plunge-step conglomerates); (D) BT4-2 (?wave rippled conglomerate) ; (E) BTI-5; (F) BT3-7; (G) BT3- 1 2; (H) BT5-9; ( I ) BT4- 13. See text for further description.
the dominant cross-bed orientation in the sand stones. This suggests a downcurrent dip of clasts lying on the lee side of gravel bedforms transitional
between bedload sheets and dunes. Massari & Parea (1988) also reported the presence of polymodal fabrics from shoreface conglomerates but did not elaborate on the potential significance of their obser vations. They also indicated that rose diagrams of clast long axes from upper shoreface deposits could be isotropic, polymodal, bimodal or unimodal. Clifton (1973) suggested that pebble segregation and the lenticularity of conglomerates, although not necessarily diagnostic, were useful criteria for distinguishing between wave-worked and fluvial gravels. Wave-worked gravels tend to be more 'lat erally regular' than lenticular, owing to the uniform ity with which waves act over large areas. He also suggested that winnowing of a deposit by waves could segregate sand- and gravel-sized material . Sub sequent studies (Leithold & Bourgeois 1984; Massari & Parea 1988; Hart & Plint 1989) have tended to confirm Clifton's textural and stratigraphical obser vations. It should be reiterated, however, that rip current deposits and longshore trough conglomerates may both have lenticular geometries, providing that they were buried below the influence of wave rework ing. Lenticular, channelized conglomerates may also be present in river mouth settings (Kleinspehn et al., 1984; Fig. 4B). Planar cross-bedded conglomerates several deci metres thick at the top of some sections at Bay Tree (Fig. 15) probably record the progradation of a gravel plunge step at the toe of the beach (low tide level) -a feature common on coarse-grained beaches (Forbes & Taylor, 1987; Massari & Parea,
Fig.
14. Pronou nced horizontal stratification in shoreface conglomerates at Bay Tree. Stratification emphasized by sandstone interbeds.
Gravelly shorefaces
91
Fig. 15. Planar tabular conglomerate cross-set at Bay Tree, i nferred to have been produced by the offshore progradation of a plunge step at the base of the beachface. From top of section BT2 ( see Fig. 4A) .
1988; Hart & Plint, 1989) . Somewhat larger steps are reported from some gravelly shorefaces (Kirk, 1980; Carter, 1988; Massari & Parea, 1988; Postma & Nemec, 1990) . A pebble fabric from the plunge step deposits at Bay Tree shows offshore dipping pebbles (Fig. 13B) . B ased on (unpublished) obser vations at Chesil Beach, progradation of the plunge step seems most likely to occur either following high tide or after a storm surge when the water table in the beach is still high. At such times, backwash dislodges pebbles from the beachface and transports them seaward. These clasts then avalanche over the plunge step causing it to prograde. Body and trace fossils provide the clearest evidence of marine deposition (Clifton, 1973; Kleinspehn et a/., 1984; Leithold & Bourgeois, 1984) . Unfortunately, neither are common in high energy, conglomerate-dominated shoreface deposits (Leithold & Bourgeois, 1984; Massari & Parea, 1988; Hart & Plint, 1989 ) . Trace fossils are rare in conglomeratic shoreface deposits of the Cardium Formation, although traces are present locally in thin interbedded sandstones (Fig. 16) . Although gravels are generally inhospitable to infaunal macro benthos, we have observed (using SCUBA) crab burrows in gravels on the Chesil Beach shoreface. However, the preservation potential of these bur rows seems very low owing to the lack of a contrast ing sediment fill. In general , the abundance of trace fossils may be related directly to the percentage of
Fig. 16. Burrowing ( Diplocraterion) in sandstone within shoreface conglomerates, Cardium Formation (core from 16-19-5 1 - 1 0W6, scale bar= 5 em).
92
B.S. Hart and A. G. Plint
sand, partly because this forms a more hospitable substrate, and partly because sand is better able to preserve traces of burrowing. Beachface
The structure of beachface gravels and conglomer ates is well established (Maejima, 1982; Nemec & Steel, 1984; Forbes & Taylor, 1987; Massari & Parea, 1988; Hart & Plint, 1 99 1 ) . Well-stratified, laterally continuous gravels and interbedded sands that dip gently (typically S0-20°) in a seaward direction are
considered diagnostic of gravelly beaches (Fig. 17 A & B ) . In examples from the Cardium Formation, individual gravel ·laminae/beds can be normally or inversely graded. Sand tends to be well-segregated from gravel, at least in high-energy systems , and commonly forms 'sand runs' in the intertidal zone (Fig. 18A) . Beach face dips are generally inversely related to the per centage of sand in the system and directly related to grain size. Shape sorting and seaward-dipping pebble imbrication (Fig. 18B) may be well developed. Landward-dipping stratification and imbrication may
Fig.
17. Beachface deposits. (A) Gently dipping interstratified beachface conglomerates and sandstones, Bay Tree. Note low angle truncation surface. (B) Gently dipping stratification in beachface conglomerates, Cardium Formation, Cutpick Hill, Alberta.
Gravelly shorefaces
93
Fig.
18. Modern beachface gravels at Chesil Beach, England. (A) Good segregation of sand- and pebble-sized constituents on beachface. (B) Offshore (to right) imbrication of beachface gravels. Shape sorting is not well developed at Chesil Beach.
be present in cases, and are apparently the result of berm accretion (Maejima, 1982; Forbes & Taylor, 1987 ) . Maejima ( 1 982) noted that broad , concave upward scour surfaces may represent the effects of beach cusp formation. Massari & Parea ( 1988) found that beachface conglomerates were internally trunc ated in places by low-angle erosion surfaces inter preted to represent storm-wave planation of the beachface (Fig. 17A). Enigmatic 'high-angle scours' were described by Leithold & Bourgeois ( 1 984) and Bourgeois & Leithold ( 1 984) from gravelly sandstone 'shoreface' deposits; similar structures have been observed in
outcrops of the Cardium Formation (Fig. 19). The scours may have steep walls (in places stepped or undercut) that scour into hummocky/swaley cross stratified and planar laminated sandstone. Multiple phases of fill can be inferred for some units, and sandstone 'intraclasts' , woody debris and shell frag ments are observed in places. Leithold & Bourgeois ( 1 984) suggested very rapid cut-and-fill (nearly instantaneous) at shallow depths just seaward of the breaker zone. An alternative interpretation is that the scours are the product of small streams incised into the beachface. Intuitively, it seems unlikely that high-angle scours can be
94
B.S. Hart and A. G. Plint
Fig.
19 Conglomerate filling �teep sided scour cut into planar laminated, fine-grained sandstone ?beachface deposits of Cardium Formation along Murray River, British Columbia. The scour may have been cut subaerially by a small stream crossing the beach, or may be the product of subtidal scour by rip-currents (Chiocci & Clifton, 1991) . See text for discussion.
maintained in non-cohesive sandy substrates in a subaqueous setting, especially in the presence of intense wave activity. However, Gruszcyzynski (pers. comm . , 1993) reported seeing (using SCUBA) vertical-sided, rip-current channels cut into sandy sediment, and speculated that bacterial or algal binding may have rendered the sediment cohesive. Alternatively, high-angle walls (locally overhanging) can be cut in loose sand, and maintained in subaerial settings by capillary forces (Fig. 3A) . Cohesion may be promoted by salt cementation (in a spray zone j ust above the limit of the swash) and by organic films. The example shown in Fig. 19 is from planar stratified sandstone, interpreted as beachface , which directly overlies swaley cross-stratified (shoreface) sandstone, the stratigraphical position of which, and relation to relative sea-level changes, is well con strained (Hart & Plint, 1993a) . Facies descriptions of wash over deposits have been provided by Carter & Orford ( 198 1 ) and Forbes & Taylor (1987) . Massari & Parea ( 1988) have described washover deposits from Messinian and Pleistocene gravelly. shorelines.
FACIES SUCCESSIONS
Stochastic variations in storminess, sediment supply, relative sea-level and other factors probably account for much more of the variation in sedimentary struc tures and sediment textures than is commonly appreciated . The morphology of gravelly shorefaces
and beachfaces can be very sensitive to textural factors such as the amount of sand in the· system. The effects can be non-linear: process affects mor phology which in turn affects process. In this respect , it may be overly simplistic to attribute facies vari ations at any given site to short-term changes in energy level (e.g. storm -non-storm) , sea-level , etc. Modern gravelly beachfaces show much intersite variation and , in places, much temporal variation at any given site. Facies models based on 'snapshots' of any particular beachface or shoreface are therefore clearly of limited applicability. It is equally clear that sedimentary facies and textural or pebble shape criteria derived from the beachface must not be applied when attempting to identify shoreface conglomerates. Several previous authors have presented sum maries of criteria for distinguishing between fluvial and shoreface/beachface gravels (Clifton, 1973; Leckie & Walker, 1982; Ethridge & Wescott, 1984; Nemec & Steel, 1984; Reddering & Illenberger, 1988). Based on these studies, and our own studies of shoreface and beachface gravels and conglomer ates, an attempt is made in Table 1 to outline the expected character of preservable gravelly shoreface and beachface deposits as a function of several vari ables. The factors that appear to be of greatest significance are: (i) change in relative sea-level (direction and magnitude) , (ii) textural character- istics of the sediment, including proportion of sandl in the system , (iii) proximity to sediment source (river mouth) ; and (iv) orientation of the shoreline
95
Gravelly shorefaces Table 1.
Summary of expected preservable characteristics of gravelly shoreface and beachface deposits Geometry and stratigraphical position
Transgressive
Processes
Facies
Thin lags continuous with flooding surfaces Topographically 'trapped' shoreface deposits
Barrier rollover Rafting
Thin lags Wave ripples and storm beds
Lobate (plan) Lenticular (cross-section) Correlate with other progradational deposits
Rivers/waves/tides Slope failures
Lenticular beds Mixed palaeocurrents Mix of fluvial and marine facies
Mixed sand and gravel
Tabular (cross-section) ?Shore-parallel elongation Correlate with other progradational deposits
Waves/tides - shoreface Swash/backwash - beachface
Longshore bars/troughs Rip-current scours Palaeocurrents reflect processes Segregation of sand and gravel increases with wave energy
Gravel - drift aligned
Tabular (cross-section) Shore-parallel elongation Correlate with other progradational deposits
Longshore currents dominant
Weak shore-normal zonation Bedload sheets/dunes Palaeocurrents shore parallel
Gravel - swash aligned
Tabular (cross-section) ?Beach ridges Shore-parallel elongation Correlate with other progradational deposits
Onshore/offshore motions
Pronounced shore normal zonation Wave ripples Onshore/offshore palaeocurrents
Regressive River mouth
with respect to dominant wave-approach direction. It is expected that in most instances transgressive gravelly shorelines will leave little permanent record of their existence. Thin pebbly lags representing abandonment of isolated clasts during shoreface retreat, and possibly organic rafting seaward of the shoreface (see Leithold, 1989; Woodborne et a!., 1989; Liu & Gastaldo, 1992), are the expected litho logical expression . Thicker successions might be preserved where the shoreface has 'backed up' against pre-existing relief ('bevels', cliffs, etc . ; see Bergman & Walker, 1987, 1988; Hart & Plint, 1993b) . Stratigraphically, these deposits will correlate with marine flooding surfaces (Plint et a!. , 1986) . Progradational packages are likely to represent regional shoreline regression, although they may also develop in areas of high sediment supply during relative sea level rise (e.g. modern fan-deltas) . Ideally the shoreface conglomerates will overlie shelf deposits and , in turn, be overlain by beachface conglomerates. The latter tend to be quite distinctive
both in outcrop and in core and the relative strati graphical position can be used to help infer a shore face origin for the underlying deposits. Non-marine facies may or may not be present above the beach facies. Roughly tabular geometries are expected for indi vidual progradational packages, except close to river mouths where lenticular geometries may occur (Fig. 4B) . These river-mouth deposits will reflect the interaction of fluvial, tidal and wave-induced cur rents. Rapid deposition in the transition zone will generate poorly sorted deposits and , potentially, instability related gravity flows (Kleinspehn et a!., 1984 ) . Channelized and burrowed deposits may be found in close proximity. Away from river mouths, the influence of marine processes will become more apparent. In 'mixed' settings with appreciable amounts of sand, segre gation of sand and gravel constituents should be proportional to the wave energy level of the environ ment. As a general rule , the potential for finding
96
B.S. Hart and A. G. Plint
marine trace fossils increases with the amount of sand in the deposit. Longshore bar systems and rip current deposits may be distinguishable (Leithold & Bourgeois, 1984; Massari & Parea, 1988; Gruszcynski et al., 1993). Studies of modern gravelly barriers have led to a distinction between swash-aligned and drift-aligned barrier systems, each with distinctive process morphology characteristics (Carter & Orford, 199 1 ) . Swash-aligned barriers are oriented nearly perpen dicular to the direction of wave approach and are dominated by onshore-offshore motions. Drift aligned barriers are those oriented at an angle to the direction of wave approach, and are dominated by longshore transport. Shore-parallel facies zonation is best developed in swash-aligned beaches and are 'smeared' in drift-aligned systems (Carter & Orford, 199 1 ) . Gravelly shoreface deposits tend to appear massive, with crudely developed stratification and sedimentary structures revealed by subtle textural variations (Bergman & Walker, 1987 , 1988; Hart & Plint, 1989, 199 1 ; Plint & Hart, 1988). By careful study of palaeocurrent information, both in the conglomerates and in interbedded sandstones, it should be possible to recognize the dominant direc tion of wave approach (Hart & Plint, 1989 ) .
SUMMARY
At present, sufficient criteria exist to permit the identification of gravelly shoreface deposits, at least in outcrop. However, few truly 'diagnostic' criteria exist, with faunal or trace faunal evidence being possible exceptions. The inadequacy of previously proposed diagnostic criteria (such as shape para meters) reflects the range of processes that can affect the character of shoreface conglomerates. In outcrop, the stratification style, sedimentary structures and palaeocurrent information can be powerful indicators of depositional process. In the subsurface , conglomerate body morphology, orien tation and stratigraphical relation to other units may aid interpretation. In both outcrop and subsurface , detailed facies analysis should make it possible to distinguish subtypes of gravelly shoreface deposits by clarifying the nature and relative importance of specific depositional processes (e.g. wave-induced and gravity driven currents, etc . ) . More studies are needed t o document the range of sedimentary processes and structures present on modern gravelly shorefaces. To do so will require
the development of innovative sampling and moni toring techniques. Difficulties encountered with establishing the threshold of grain movement and bedload transport rates under combined flows typical of sandy shoreface settings are doubly present on gravelly shorefaces, where clast morphology, bulk textural parameters and fabric play important roles. The establishment of phase diagrams for bed con figurations remains problematic even in sandy shore face settings, and extremely little documentation is available for gravelly shoreface settings. Empirical investigations of ancient deposits may provide, for some time to come, the best insight into the processes operative on modern gravelly shorefaces.
ACKNOWLEDGEMENTS
The work on shoreface conglomerates of the Car dium Formation which forms the nucleus of this paper was completed during the senior author's doctoral studies at the University of Western Ontario. Funding for that study was made possible through research grants (to A . G . P . ) from the Natu ral Sciences and Engineering Research Council, the Department of Energy, Mines and Resources and the University of Western Ontario. Additional logis tical support was provided by Canadian Hunter Ltd . , Esso Resources Ltd . , Home Oil Ltd. and Unocal Canada Ltd. To all these agencies and companies, we are very grateful . Helpful comments by reviewers E. Leithold and F. Massari are greatly appreciated. We thank the Canadian Society of Petroleum Geol ogists for permission to reproduce Figs 4A, SA, 9A and 12A, originally published in Plint & Hart ( 1988). This paper was written during the tenure of a Visiting Fellowship by the senior author at the Geological Survey of Canada's Pacific Geoscience Centre.
REFERENCES
ARNOTT, R.W.C. ( 1991) The Carrot Creek 'K' Pool, Car
dium Formation, Alberta - a conglomeratic reservoir related to a wave-reworked distributary mouth bar com plex. Bull. Can. petrol. Geol. , 39, 43-53. BERGMAN, K . M . & WA L KER , R . G . ( 1987) The importance of sea level fluctuations in the formation of linear con glomerate bodies; Carrot Creek Member of Cardium Formation , Cretaceous Western Interior Seaway, Alberta, Canada. J. sediment. Petrol. , 57, 651 - 665 . BERGMAN, K . M . & WALKER, R.G. ( 1 988) Formation of Cardium erosion surface E5 , and associated deposition of conglomerate: Carrot Creek Field, Cretaceous West-
Gravelly shorefaces ern Interior Seaway, Alberta. I n : Sequences, Strati graphy, Sedimentology: Surface and Subswface (Eds James, D . P . & Leckie, D . A . ) . Can . Soc. petrol. Geol . , Calgary, Memoir 1 5 , 15-24. BLUCK, B . J . ( 1967) Sedimentation of beach gravels: examples from South Wales. J. sediment. Petrol. , 37, 128- 156. BOURGEOIS, J. & LEITH OLD, E.L. ( 1984) Wave-worked conglomerates - depositional processes and criteria for recognition. I n : Sedimentology of Gravels and Conglom erates (Eds Koster E . H . & Steel, R . J . ) . Can . Soc. petrol. Geol . , Calgary, Memoir 10, 331 - 343. B RU U N , P . ( 1988) The Bruun Rule of erosion by sea level rise: a discussion on large-scale two- and three dimensional usages. J. coast. Res. , 4, 627-648. B UJALESKY, G . C . & GONZALEZ-BONORINO, G. ( 1991) Gravel spit stabilized by unusual (?) high-energy wave climate in bay side, Tierra del Fuego. In: Coastal Sedi ments '91, American Society of Civil Engineers, Seattle, pp. 960-974. CAILLEUX, A. ( 1945) Distinction des galets marins et flu viatiles. Bull. Soc. geol. France, 5, 325-404. CANT, D . J . & ETHIER, V . G . ( 1984) Lithology-dependent diagenetic control of reservoir properties of conglomer ates, Falher Member, Elmworth Field, Alberta. Bull. Am. Assoc. petrol. Ceo!. , 68, 1044- 1 054. CARR, A.P. ( 1969) Size grading along a pebble beach: Chesil Beach, England. J. sediment. Petrol. , 39, 297-311. CARTER, R.W.G. ( 1988) Coastal Environments. Academic Press, New York, 617 pp. CARTER, R . W . G . & ORFORD, J . D . ( 1981) Overwash pro cesses along a gravel beach in southeast Ireland. Earth Sutf Process. Landf , 6, 413-426. CARTER, R . W . G . & ORFORD, J .D . ( 1984) Coarse clastic barrier beaches: a discussion of the distinctive dynamic and morphosedimentary characteristics. Mar. Ceo!. , 60, 377- 389. CARTER, R.W.G. & ORFORD, J . D . ( 1991) The sedimentary organization and behaviour of drift-aligned gravel bar riers. l n : Coastal Sediments '91, American Society of Civil Engineers, Seattle, pp. 934- 948. CHEEL, R.J . & LECKIE, D .A. ( 1992) Coarse-grained storm beds of the Upper Cretaceous Chungo Member (Wapiabi Formation), southern Alberta, Canada. J. sediment. Petrol., 62, 933-945. CHIOCCI, F.L. & Cuno N , H . E . ( 1991) Gravel-filled gutter casts in nearshore facies - indicators of ancient shore line trend. In: From Shoreline to A byss: Contributions in Marine Geology in Honor of Francis Parker Shepard. (Ed. Osborne, R . H . ) Soc. econ. Paleont. Miner. Spec. Pub!. No. 46, pp. 67-76. CLARKE, R . H. ( 1979) Reservoir properties of conglomer ates and conglomeratic sandstones. Bull. Am. Assoc. petrol. Ceo!. , 63, 799-803. CuFrON , H . E . ( 1973) Pebble segregation and bed lenticu larity in wave-worked versus alluvial gravel. Sedimen tology, 20, 173- 187. CunoN, H . E . ( 1981) Progradational sequences in Miocene shoreline deposits, southeastern Caliente Range, California. J. sediment. Petrol. , 5 1 , 165 - 184. CLIFTON , H . E . , PHILLIPS, R.L. & HUNTER, R . E . ( 1973) Depositional structures and processes in the mouths of
97
small coastal streams, southwestern Oregon . I n : Coastal Geomorphology (Ed. Coates, D . R . ) , pp. 115- 140. State U niversity of New York at Binghamton . DAVIDSON-ARNOlT, R . G . D . & GREENWOOD, B . ( 1976) Facies relationships on a barred coast, Kouchibouguac Bay, New Brunswick, Canada. I n : Beach and Nearshore Sedimentation (Eds Davis, R.A. & Ethington , R . L . ) . Spec. Pub ! . Soc. econ. Paleont. Mineral, Tulsa, 24, 149- 168. DoBK.JNS, J . E . & FOLK, R . L . ( 1970) Shape development on Tahiti-Nui. J. sediment Petrol., 40, 1167- 1203. D uBOIS, R . N . ( 1992) A re-evaluation of Bruun's Rule and supporting evidence. J. coast. Res. , 8, 618 - 628. DUPRE, W . R . , CLIFTON, H . E . & HUNTER, R.A. ( 1980) Modern sedimentary facies of the open Pacific Coast and Pleistocene analogs from Monterey Bay, California. I n : Quaternary Depositional Environments of the Pacific Coast (Eds Field, M . E . et a!.), pp. 105 - 120. Soc. econ. Paleont. Mineral. Pacific Section, Tulsa. EMERY , K . O . ( 1955) Grain size of marine beach gravels. J. Ceo!. , 63, 39-49. ETHRIDGE , F.G. & WESCOTT, W.A. ( 1984) Tectonic setting and recognition of hydrocarbon reservoir potential of fan-delta deposits. I n : Sedimentology of Gravels and Conglomerates (Eds Koster, E . H . & Steel, R.J . ) . Can. Soc. petrol. Geol . , Calgary, Memoir 10, 2 17 -235 . FORBES, D . L . & BOYD, R . ( 1989) Gravel ripples o n the inner Scotian Shelf. J. sediment. Petrol., 57, 46-54. FoRBES, D . L. & TAYLOR, R . B . ( 1987) Coarse-grained beach sedimentation under paraglacial conditions, Canadian Atlantic coast. I n : Glaciated Coasts ( Eds Fitzgerald , D . & Rosen P . ) , pp. 5 1-86. Academic Press, New York. GILLIE, R . D . ( 1983) Field measurements of gravel move ment by wave action on the inner continentia! shelf. I n : Proceedings of the Canadian Coastal Conference (Ed. Holden, B . J . ) , pp. 73-89. National Research Council of Canada, Ottawa. GRUSZCZYNSKI, M . , RUDOWSKI, S., SEMIL, J . , SLOMINSKI, J. & ZROBEK, J. ( 1993) Rip currents as a geological tool. Sedimentology, 49, 2 17 -236. HARMS, J . C . , SOUTHARD, J . B . & WALKER, R . G . ( 1982) Structures and Sequences in Clastic Rocks. Society of Economic Palaeontologists and Mineralogists, Tulsa, Short Course 9, 200 pp. HART, B .S. ( 1990) The sedimentology and stratigraphy of the Upper Cretaceous Cardium Formation in northwestern Alberta and adjacent British Columbia. Unpublished PhD thesis, University of Western Ontario, London, Ontario, 505 pp. HART, B.S. ( 1991) A study of pebble shape from gravelly shoreface deposits. Sediment Ceo!. , 73, 185 - 189. HART, B . S . & Pu NT , A . G . ( 1989) Gravelly shoreface deposits: a comparison of modern and ancient facies sequences. Sedimentology, 36, 55 1-557 . HART, B . S . & Pu NT , A . G . ( 1991) Conglomeratic shore face deposits from the Cretaceous Cardium Formation, Alberta, Canada. l n : Coastal Sediments '91, A merican Society of Civil Engineers, Seattle, pp. 949-959. HART, B.S. & Pu N T , A . G . ( 1993a) Origin of an erosion surface in shoreface sandstones of the Kakwa Member (Upper Cretaceous Cardium Formation, Canada ) : importance for reconstruction o f stratal geometry and depositional history. I n : Sequence Stratigraphy and Facies
·
98
B.S. Hart and A. G. Plint
Associations (Eds Posamentier, H.W. , Summerhayes, C . P . , Haq, B . U . & Allen, G . P . ) , Spec. Pubis int. Ass. Sediment No . 18, 45 1 - 467. Blackwell Scientific Publi cations, Oxford. HART, B.S. & Pu NT , A . G . ( 1 993b) Tectonic influence on deposition and erosion in a ramp setting: Upper Cretaceous Cardium Formation, Alberta foreland basin. Bull. Am. Assoc. petrol. Ceo/. , 77, 2092-2 107. HART, B . S . , VANTFOORT, R . M . & Pu NT , A.G. ( 1 990) Dis cussion on 'Is there evidence for geostrophic currents preserved in the sedimentary record of inner to middle shelf deposits?' J. sediment. Petrol. , 60, 633 -635. HEI N , F.J. & WALKER, R . G . ( 1 977) Bar evolution and development of stratification in the gravelly, braided, Kicking Horse River, British Columbia. Can. J. Earth Sci. , 14, 562-570. HowARD, J . L. ( 1 992) An evaluation of shape indices as palaeoenvironmental indicators using quartzite and metavolcanic clasts in Upper Cretaceous to Palaeogene beach, river and submarine fan deposits. Sedimentology, 39, 47 1 -486. HowELL, D . G . & LINK, M . H . ( 1 979) Eocene conglomerate
sedimentology and basin analysis, San Diego and the southern California borderland. J. sediment. Petrol. , 49, 5 1 7 -540. HUNTER, R . E . , CLIFTON , H . E . & PHILLIPS, R . L . ( 1 979)
Depositional processes, sedimentary structures and predicted vertical sequences in barred nearshore sys tems, southern Oregon coast. J. sediment. Petrol. , 49, 7 1 1 -726.
KmsoN, C. & CARR, A . P . ( 1 959) The movement of shingle over the seabed close inshore. Geogr. J. , 125 , 380-389. KIRK, R . M . ( 1 980) Mixed sand and gravel beaches: mor phology, processes and sediments. Prog. phys. Geogr. , 4, 189 - 2 10. KLEINSPEH N , K.L. , STEEL, R .J . , JOHANNESSEN , E . & N ET LAND, A. ( 1 984) Conglomeratic fan-delta sequences,
Late Carboniferous-Early Permian, Western Spits bergen. I n : Sedimentology of Gravels and Conglomerates (Eds Koster, E . H . & Steel, R.J . ) . Can . Soc. petrol. Geo l . , Calgary, Memoir 1 0 , 279- 294. KOMAR, P . O . ( 1 976) Beach Processes and Sedimentation . Prentice-Hall, Englewood Cliffs, N J , 429 pp. KOMAR, P . O . ( 1 987) Selective gravel entrainment and the empirical evaluation of flow competence . Sedimentology, 34, 1 165- 1 176. KOMAR, P.O. & L1, Z. ( 1 986) Pivoting analysis of the
selective entrainment of sediments by shape and size with application to gravel threshold . Sedimentology, 33, 425-436. LECKIE, D . A . ( 1 988) Wave-formed, coarse-grained ripples
and their relationship to hummocky cross-stratification. J. sediment. Petrol. , 58 , 607 :..._ 6 22. LECKIE, D . A . & WALKER, R . G . ( 1 982) Storm- and tide dominated shorelines in Cretaceous Moosebar-Lower Gates interval: outcrop equivalents of Deep Basin gas trap in Western Canada. Bull. Am. Assoc. petrol. Geol. , 66, 138- 157. LEITHOLD, E.L. ( 1989) Depositional processes on an
ancient and modern muddy shelf, northern California. Sedimentology, 36, 179-202. LEITH OLD, E . L . & BouRGEOIS, J. ( 1 984) Characteristics of coarse-grained sequences deposited in nearshore, wave-
dominated environments - examples from the Miocene of south-west Oregon. Sedimentology, 3 1 , 749-775. L1u , Y. & GASTALDO, R . A . ( 1992) Characteristics and provenance of log-transported gravels in a Carboniferous channel deposit. J. sediment. Petrol. , 62, 1072 - 1083. MAEJIMA, W. ( 1982) Texture and stratification of gravelly beach sediments, Enju Beach, Kii Peninsula, Japan . J. Geosci. Osaka City Univ. , 25, 35-5 1 . MASSARI, F. & PAREA, G.C. ( 1988) Progradational gravel beach sequences in a moderate- to high-energy, micro tidal marine environment. Sedimentology, 35 , 88 1 - 9 1 3 . MASSARI, F . , PAREA, G . C . , RAINONE, M . L . , VEDOVATO, L. & YIVALDA, P. ( 1 986) Elementi di sedimentologia delle paleospiagge Pleistoceniche marchigiane. I n : Alti Riunione Gruppo Sedimentologia CNR, Ancona, 5 - 7 giugno 1986, p p . 8 1 - 103. MATHEWS, E . R. ( 1 980) Observations of beach gravel trans port, Wellington Harbour entrance, New Zealand. N.Z. J. Geol. Geophys. , 23 , 209-222. MIDDLETON , G . V . ( 1 973) Johannes Walther's law of the correlation of facies. Geol. Soc. Am. Bull. , 84, 979-988. NEATE, D.J . M . ( 1 967) Underwater pebble grading of Chesil Bank . Proc. Ceo/. Assoc. , 78, 419-426. NEMEC, W . & STEEL, R.J. ( 1 984) Alluvial and coastal conglomerates: their significant features and some com ments on gravelly mass-flow deposits. In: Sedimentology of Gravels and Conglomerates (Eds Koster, E . H . & Steel, R.J . ) . Can. Soc. petrol. Geol . , Calgary, Memoir 10, 1 - 3 1 . NIEDORODA, A . W . , Swwr, O . J .P. & HOPKINS, T.S. ( 1 985)
The shoreface. I n : Coastal Sedimentary Environments (Ed. Davis, R . A . , Jr. ) , pp. 533 - 624, Springer-Verlag, New York. NIELSEN, L . H . , JOHANNESSEN, P.N. & SURLYK, F. ( 1 988) A Late Pleistocene coarse-grained spit-platform sequence in northern Jylland, Denmark. Sedimentology, 35 , 9 1 5 - 937. OGREN , D . E . & WAAG, C.J. ( 1 986) Orientation of cobble and boulder beach clasts. Sediment. Geol. , 47, 69-76. ORFORD, J . 0 . ( 1975) Discrimination of particle zonation on a pebble beach. Sedimentology, 22, 441 -463. PArnsoN , S . A . J . & WALKER, R . G . ( 1 992) Deposition and
interpretation of long, narrow sandbodies underlain by a basinwide erosion surface: Cardium Formation, Cre taceous Western Interior Seaway, Alberta, Canada. J. sediment. Petrol. , 42, 292-309. PuNT, A . G . & HART, B .S . ( 1 988) Field Guide to the Upper Cretaceous Dunvegan (Cenomanian) and Cardium (Turonian) Formations in the Dawson Creek - Fort St. John area, British Columbia. Canadian Society of Pet-· roleum Geologists' Field Guide to 'Sequences, Stratigra-· phy, Sedimentology: Surface and Subsurface Technical Meeting', 14- 1 6 September, Calgary, Alberta, 5 1 pp. PuNT, A.G. & WALKER, R . G . ( 1 987) Cardium Formation 8. Facies and environments of the Cardium shoreline and coastal plain in the Kakwa Field and adjacent areas, northwestern Alberta. Bull. Can. petrol. Geol. , 35, 48-64.
PUNT, A . G . , WALKER, R . G . & BERGMAN, K . M . ( 1 986) Cardium Formation 6. Stratigraphic framework of the Cardium in subsurface. Bull. Can. petrol. Geol. , 34, 2 1 3 -225. PoSTMA , G. & NEMEC, W . ( 1 990) Regressive and trans-·
Gravelly shorefaces gressive sequences in a raised Holocene gravelly beach, southwestern Crete. Sedimentology, 37, 907-920. READING, H . G . ( 1986) I ntroduction . I n : Sedimentary En vironments and Facies (Ed. Reading, H .G ) , pp. 1 -3 . Blackwell Scientific Publications, Oxford. REDDERING, J . S . V . & ILLENBERGER, W . K . ( 1988) Dis cussion on palaeoenvironmental significance of clast shape in the Nardouw Formation, Cape Supergroup. S. Afi·. J. Ceo!. , 9 1 , 554-555 . Ross, N. & LONG, B . F . ( 1989) Evolution morpho sedimentaire de Ia barre de deferlement: un example dans le golfe du Saint-Laurent, Quebec. Geogr. phys. Quat. , 43, 377-388. SHIPP, R.C. ( 1984) Bedforms and depositional sedimentary structures of a barred nearshore system, eastern Long Island, New York. Mar. Ceo!. , 60, 235-259. SHORT, A . D . ( 1984) Beach and nearshore facies: southeast Australia. Mar. Ceo!. , 60, 261 -282. WESCOTr, W.A. & ETHRJDGE, F.G. ( 1 980) Fan-delta sedi mentology and tectonic setting - Yallahs fan delta, .
99
southeast Jamaica. Bull. Am. Assoc. petrol. Ceo!. , 64, 374-399. WHITING , P .J . , DIETRICH, W . E . , LEOPOLD, L . B . , DRAKE, T . G . & SHREVE, R . L . (1988) Bedload sheets in hetero geneous sediment. Geology, 16, 105- 108. WILLIAMS, A .T. & CALDWELL, N . E . ( 1988) Particle size and shape in pebble-beach sedimentation. Mar. Ceo!. , 82, 199-215. WILLIAMS, J .J . , THORN E , P.O. & H EATHERSHAW, A.D. ( 1 989) Comparisons between acoustic measurements and predictions of bedload transport of marine gravels. Sedimentology, 36, 973-979. WooDBORNE, M . W . , RoGERS, J. & JARMA N , N. ( 1989) The geological significance of kelp-rafted rock along the west coast of South Africa. Ceo-Mar. Lett. , 9, 109 - 1 18 . WRJGHT, L . D . & SHORT, A . D . (1984) Morphodynamic variability of surf zones and beaches: a synthesis. Mar. Ceo!. , 56, 93- 1 18. ZEN KOVITCH , V.P. ( 1967) Processes of Coastal Develop ment. Oliver and Boyd, London, 738 pp.
Spec. Pubis int. Ass. Sediment. (1995) 22, 101-135
The return of 'The Fan That Never Was': Westphalian turbidite systems in the Variscan Culm Basin: Bude Formation (southwest England) R OB E R T V . B U R N E Australian Geological Survey Organisation, P O B ox 378, Canberra, ACT 2601, Australia
ABSTRACT
The Westphalian Bude Formation of southwest England was one of the first successions to be interpreted as a subsea fan deposit. This interpretation has been questioned recently, and the alternative depositional environment of a wave-affected shelf has been proposed. The tectonic setting, palaeo geographical relationships and facies succession of the formation have been reassessed in an attempt to resolve this controversy. The formation was deposited in the synorogenic Culm Basin. It contains thick sandstone units, but is in other respects similar to the underlying turbidite-bearing Crackington Formation (Namurian-Westphalian) . There is no depositional continuity with the contemporaneous paralic and deltaic deposits of the Bideford Group that occur in a tectonically distinct succession near Westward Ho! Four facies are identified in the Bude Formation; the black shale facies, the muddy siltstone facies, the interbedded sandstone-shale facies and the sandstone-dominated facies. Both thickening upward and thinning upward facies successions occur at some levels, but generally the facies succession shows less predictable, although not random, alternations between the four facies. The suggestion that the Bude Formation represents the deposits of a wave-influenced shelf does not stand up to scrutiny. There is no substantial evidence for shallow water environments in the Formation . Evidence for storm-generated hummocky cross-stratification and combined-flow wave ripples is equivocal. The ichnofacies have no depth connotation and there is no evidence for sedimentological continuity with undoubted deltaic successions. By contrast there is overwhelming evidence for deposition of turbidite systems in an isolated fresh- or brackish-water basin. The black shale facies represent the deposits of fine-grained turbidity currents in an anaerobic or dysaerobic basin. The muddy siltstone facies represent the deposits of aerobic environments characterized by gentle currents, possibly fan-levee environments. The interbedded sandstone-shale facies consist of a variety of turbidites and were deposited in either levee, lobe or interchannel environments. The sandstone-dominated facies were deposited in fan-valley environments, and include channel-fill successions. It is concluded that the Bude Formation was deposited as subsea fans on the northern, inactive margin of a land-locked, foreland basin.
INTRODUCTION
The Westphalian Bude Formation (Owen, 1934; King, 1966, 1967, 1971; Freshney et al., 1979) was deposited in the synorogenic Culm Basin of south west England (Sedgewick & Murchison, 1840; Ussher, 1892; Thomas, 1988) (Figs 1 & 2). The formation crops out along the coast of north Cornwall and west Devon (Fig. 2) and it had become the custom to equate it with a contemporaneous succession, the Bideford Group, which crops out on
the coast near Westward Ho! in north Devon (Figs 2 & 3). Reading (1963) drew attention to the funda mental difference of opinion about the deposition of these successions. Owen (1950) and Prentice (1960a, b) considered that they were examples of paralic 'Coal Measures', but Ashwin ( 1957, 1958) interpreted them as turbidites. Reading ( 1963, p. 69) contrasted the evidence in the rocks around Westward Ho! for 'successive advances into a basin
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
101
102
R. V. Burne A 52 '
----
8
5 1'
ST
50' 6'
D [] �
Permian- Tertiary Upper Carboniferous Devonian - Lower Carboniferous
3'
• I+++I
Lizard and Start complexes
----&....- Thrust fault -- Normal fault
Granite
�
BCFZ
Bristol Channel fracture zone
SLFZ
Sticklepath - Lustleigh fracture zone
Strike- slip fault
Fig. 1. Location of the Culm Basin. (A) Location of southwest England with respect to the British Isles. (B) Location of the Culm Basin with respect to the other principal tectonic elements of southwest England (after Hartley & Warr, 1990). Suggested Silesian position of southwest England with respect to France produced by pre-deformation restoration along the Bristol Channel-Bray Fault (after Holder & Leveridge, 1986), all other structures shown in present-day positions.
of a low coastal plain, possibly fronted by deltas, alternating with periods of reduced supply when basin conditions were re-established' with the fact that 'No cyclic pattern of sedimentation has been recognized, nor has any unequivocal evidence of shallow water or coastal plain deposition been observed' in the Bude Formation. This implied that 'whilst coastal plains reached into north Devon in Lower Westphalian times, there is no evidence that they extended into Cornwall'.
Subsequent workers confirmed the existence of paralic cycles in the Bideford Group, (comprising the Northam and Abbotsham Formations), around Westward Ho! (Walker, 1964a, b; de Raaf et al. , 1965; Money; 1966; Elliott, 1976). They have also confirmed Reading's statement that (1963, p. 69) 'The precise depositional environment of the Bude Sandstones is not easy to establish'. Sedimentologists had generally recognized that the Bude Formation was, in part, a turbidite succession, but did not
103
Bude Formation subsea fans '40
4"20'
'30
'20
Bideford
'30
Bay
Hartland Point
51"00'
'20
Bude
Bay
'10 50'50'
BUDE
·
.
p;.
.
'00
•
Post- Silesian strata
WESTWARD HO! BIDEFORD FAULT BLOCK
m t-=-� �
Greencliff Beds Bideford group & Westward Ho! Formation
CULM BASIN
D [·.:.·: ·_.:J
Bude Formation
]
--Fault
SLFZ
SILESIAN
�
Crackington Formation 1
Sticklepath - Lustleigh fracture zone
Town
D
Village
30
National grid number tor Great Britain
Pre- Silesian strata
Fig, 2, Geological map of north Cornwall and west Devon. (Modified after Burne & Moore, 1971; Thomas, 1988.)
appear to be a 'true flysch' (Goldring & Seilacher, 1971). It contained some unusual characteristics, such as massive sandstones, which de Raaf (pers. comm., 1966) had informally termed 'Budes' and interpreted as some form of unusually fine-grained 'fluxoturbidite' (Unrug, 1963, 1965). A sedimentological facies analysis of the middle Bude Formation (Fig. 4) was undertaken to identify features definitive, rather than indicative, of depo sitional processes and environments (Burne, 1969a, b; 1970, 1971, 1973, 1976; Burne & Moore, 1971). It was concluded that the Bude Formation had been deposited in a land-locked basin, akin to the present Black Sea or Caspian Sea, that had been isolated by Hercynian continental collision. The
basin was generally filled with fresh or brackish water, although rare marine incursions occurred, probably the result of glacio-eustatic oscillations. The mature, fine-grained sandstones were derived from a land mass of older sedimentary rocks to the north and were deposited as turbidites from bottom hugging underflows. The turbidites included con ventional Bouma sequences, traction carpet deposits with current-reworked tops, irregularly laminated sandstones, and 'slurried' beds produced by the impact of dense turbidity currents on a soft substrate. The major 'slumped' beds had the same origin as the 'slurried' beds, but some sandstone dykes and volcanoes were formed by load-induced post depositional water-escape. The depositional
104
R. V. Burne R
HARTLAND TO BOSCASTLE
a WESTWARD HOI
z "' :J "' I 0.. f rJ) UJ
Amaliae Marine Band
s:
Usteri Marine Band Marine
• Q D � �
Greencliff Beds
Bideford Group
Westward Ho! Formation
Bude Formation
Crackington Formation
INDEX SHALES
ccs
Clovelly Court Shale
DPS
Deer Park Shale
ESS
Embury Shale
GAS
Gull Rock Shale
HQS
Hartland Quay Shale
SMS
Sandy Mouth Shale
SPS
Saturday's Pit Shale
TCS
Toms Cove Shale
WGS
Warren Gutter Shale
Brigantian •oom
environment was that of a submarine fan, with open fan, fan-levee, active fan-channel, and inactive fan channel facies being distinguished. Despite the apparently random nature of much of the succession, it was recognized that predictable associations occurred in parts (Burne, 1969b, Fig. 20) (Fig. 5), including thickening upward successions, interpreted as being due to the advance of a fan-channel/levee complex over an open fan, and thinning upward successions, interpreted as being the result of fan channel fill and abandonment.
Fig. 3. Stratigraphical correlation between the Culm Basin and Westward Ho! Silesian sequences. (Modified from King, 1967 ; Thomas, 1988.)
Melvin (1976, 1986) reached similar conclusions from a study of higher levels in the Bude Formation (Fig. 4). He suggested that deposition was from turbidity currents in a relatively shallow basin and noted that sediment coarser than fine-grained sand was absent, and the organization of the bundles of sandstones was complex. He suggested that sediment was supplied from a delta to the north by resedimen·· tation of delta-front sands, and by slumping of delta·· front muds and silts to form a lower prodelta turbidite fan. Melvin could not identify well-developed
105
Bude Formation subsea fans
B § � D G []] � �.
UJ 0 ::;) "'
Nodular shale Shale Siltstone Mudstone Slumped bed Thick-bedded and massive sandstone characteristics of BudeFormation Shales with thin sandstone Medium to thinly bedded sandstones with subordinate shales and siltstones
INDEX SHALES
600
WGS Warren Gutter Shale 'Anthracoceras' aegiranum horizon
SMS
Sandy Mouth Shale
SPS
Saturday's Pit Shale (key shale W)
TCS
Tom's Cove Shale (key shale Q)
LS
Longpeak Shale
HQS
Hartland Quay Shale
GRS
Gull Rock Shale
Gastrioceras amaliae horizon
Gastrioceras /isteri horizon
ES
Embury Shale Gastrioceras subcrenatum
400
200
horizon
Fig. 4. Representative lithological section of the Crackington Formation and Bude Formation based on exposures between Duckpool (SS 201 1 15) and Embury Beach (SS 214 195) (modified after Freshney et a/., 1979). King (1967) originally defined the base of the Bude Formation at a level 265 m below the base of the Tom's Cove Shale and the top of the formation at a level 150 m above the same datum, a section here referred to as the middle Bude Formation.
thickening upwards and thinning upward successions as described by Burne (1969b) from lower in the Bude Formation. However, he presented a facies model of various fan, fan-levee and fan-channel relationships that accounted for the facies suc cessions observed in measured sections. Apart from different degrees of organization in the succession, there is general correspondence between the fan interpretation suggested by Burne (1969b) and that proposed by Melvin (1986). Higgs ( 1987) strongly criticized Melvin's inter pretation in a vigorous discussion provocatively titled 'The Fan That Never Was?'. Higgs (1983, 1984, 1986a, b; 1991) tentatively identified wave influenced structures in the Bude Formation, and pursued the significance of this interpretation to its logical conclusion, that the 'Bude Formation was deposited largely, if not entirely, above storm wave base, as indicated by the fact that wave-influenced
structures occur throughout the succession, spaced no more than a few metres apart.' This paper attempts to resolve the controversy that has resulted from these contrasting inter pretations by reconsidering the tectonic setting, palaeogeographic relationships and facies analysis of the Bude Formation.
THE TECTONIC SETTING OF THE CULM BASIN
The convergence and collision of the Laurussian and Gondwanan continental masses between the Devonian and the Early Permian (Burne, 1969b, 1973; Ziegler, 1990) generated a deformation front that migrated northwards with time (Besly, 1988) to form the Variscan orogenic belt. During the Middle Devonian and Early Carboniferous, volcanicity and
R. V. Burne
106
COMPOSITE SEQUENCE
Possible interpretation
Mudstone
Open fan or inter
dominated
-channel sediment
sequence
1-:: . 'Thinning upward'
--- ---
/
-�::--......_
�-=--
::-�--
:� ----------- ----:-:-,
�
sequence
---==- � -==-
Sandstone dominated
l
Sediment in confines of partly active channel
Sediment in confines ----�-�-
of active large
sequence
channels
-
'Thickening upward'
.. ·.:::.:.:_"·
Sediment of channel levee
sequence
Mudstone
Open fan or inter
dominated
channel sediment
sequence
16/04117
I
Bed scale Approx. 1m (bu1 variable)
Bude Formation subsea fans
subsidence accompanied a period of extension in southwest England. Regional inversion of this Devonian to Early Carboniferous basin occurred in the early Namurian. Loading of its northern margin by the resulting thrust nappe development (Hecht, 1992) formed the Culm Basin that lay north of the developing Variscan Orogen and south of the cratonic Wales-Brabant Massif (Hartley & Warr, 1990) (Fig. 1). Evidence for a thin skinned thrust and-nappe regime implies that the orogen did not attain 'Alpine' proportions (Isaac et a/. , 1982) . Hartley & Warr (1990) suggest that the Culm Basin maintained its present geographical relationship with South Wales throughout the Variscan deformation. However, others suggest that the northern boundary of the basin corresponded to an area west of present day Le Havre, France (Fig. 1) (Holder & Leveridge, 1986), and subsequent transcurrent faulting along the Bristol Channel and Bray Fault has displaced the basin to its present position, south of the Bristol Channel. In the early Namurian the Culm Basin was prob ably part of a continuous deep-water flysch basin, the Rhenohercynian Basin, that extended from Cornwall to Upper Silesia (Ziegler, 1990) . Sedimen tation began to exceed subsidence in parts of this basin during the Namurian, and the Culm Basin became one of a number of basins isolated during the last stage of continental collision (Franke & Engel, 1988). In the Culm Basin evidence for this change may be seen in the turbidites of the Crackington Formation. Marine goniatites, which had been scattered throughout the shales lower in the succession, became restricted in the uppermost Namurian to marine bands deposited by marine incursions into an otherwise brackish-water basin (Freshney et a/., 1979). By the early Westphalian most Rhenohercynian basins had shallowed and established paralic conditions providing the classic localities for coal-measure cyclothems (Reading,
107
197 1) . The occurrence of correlatable marine bands throughout these basins indicates that they were separated by very low topographical relief, with the marine bands resulting from short-lived marine transgressions that entered the basins from the Moscow Depression (Ziegler, 1990). The trans gressions probably reflect high sea-levels in a series of short-term glacio-eustatic fluctuations, the last of which occurred in Westphalian D time. It was only in the Culm Basin that deep-water conditions persisted into the Westphalian (Ziegler, 1990). Deposition took place on the northern, non orogenic flank of the basin, as shown by the lack of palaeocurrents from the south (Fig. 6) (Ashwin, 1957; Burne, 1969b; Freshney eta/. , 1979; Melvin, 1986; Higgs, 1991) and by the compositionally mature, fine-grained nature of the sediments (Fig. 7) (Burne, 1969b; Freshney eta/. ; 1979; Melvin, 1986; Haslam & Scrivener, 199 1). Namurian processes deposited thin-bedded turbidites from currents flow ing eastward along the axis of the basin, forming the Crackington Formation (Freshney et a/. , 1979) . Deposition of the Bude Formation began in the Early Westphalian. Although 50% of the beds in the Bude Formation resemble those of the underlying Crackington Formation, it is distinguished by its content of massive sandstones (Fig. 4). The currents that deposited the Bude Formation generally flowed towards the south and southwest, although easterly flowing currents occurred at some horizons (Fig. 6) (Burne, 1969b, Freshney et a/. , 1979, Melvin, 1986). No post-Bude-Formation sediments are preserved in the Culm Basin.
TECTONICALLY
JUXTAPOSED
SUCCESSIONS IN THE CULM BASIN
Freshney & Taylor ( 1972) proposed a simple stra-
Fig. 5. (Opposite.) Composite 'sequence' (i.e. succession in present usage) for the Bude Formation reproduced from Burne ( 1969b, Fig. 20) showing the original interpretation, including the first recognition of 'thickening' and 'thinning' upward 'sequences' in subsea fan deposits. The composite succession includes only one representative of the various bed types· found in each part of the succession , and the scale shown relates to the thickness of these individual beds. Complete successions are actually 10-60 m thick. In the usage of the present paper the mudstone-dominated sequence corresponds to the black shale facies, now interpreted as fine-grained turbidites in an anaerobic or dysaerobic basin; the thickening upward sequence corresponds to part of the interbedded sandstone- shale facies, now interpreted as turbidites deposited in either levee , lobe or interchannel environments; the sandstone-dominated sequence corresponds to the sandstone-dominated facies, now interpreted as the deposits of fan-valley environments, including channel-fill successions; and the thinning upward sequence corresponds to part of the interbedded sandstone-shale facies now interpreted as the deposits of aerobic environments characterized by gentle currents, possibly fan levees.
R. V. Burne
108 A
Tool Marks
Directions
Senses Only
8
Current Scours
Directions
C
Senses Only
�
Cross Laminations
N
I
10 % Directions
N
=
15
20
Number of readings
Fig. 6. Structurally corrected palaeocurrents recorded from sole marks and cross laminations in the middle Bude Formation (data of Burne, 1969b).
tigraphy for the Culm Basin (Figs 3 & 4) in which all the dominantly turbidite sandstone-mudstone successions of Namurian and basal Westphalian rocks were classified as Crackington Formation, whereas the Bude Formation contained the younger
Westphalian deposits which 'while they still contain sequences of turbidites, have within them substantial numbers of massive sandstones and other associated facies characteristic of the rocks described by King (1966)' (Freshney & Taylor, 1972, p. 467). An ar bitrary junction between the two formations was set at the top of the Hartland Quay Shale (Freshney et al., 1979), a marker bed containing goniatites of the Gastrioceras amaliae marine band. The Bude Formation includes the youngest Westphalian rocks exposed in southwest England, the Warren Gutter Shale, which contain a fauna correlated with the 'Anthracoceras' aegiranum marine band (Freshney et al., 1979). The sedimentological contrast between the Crackington and Bude Formations and the con temporaneous paralic sediments exposed near Westward Ho! led Reading (1965) to suggest that the Culm Basin contained a number of sedimen tologically distinct though synchronous strati graphical successions separated by low-angle thrust faults. Burne & Moore (1971) concluded that two distinct stratigraphical successions could be recognized; a northern, essentially deltaic succession (the Westward Ho! Formation, Northam Formation, Abbotsham Formation and the Greencliff Beds), and a contrasting southern succession, consisting entirely of basin facies, comprising the Crackington Formation and the Bude Formation of Freshney et al. (1979) (Figs 2 & 3). The two successions are separated by a major E- W trending normal fault, with downthrow several hundred metres to the north (Freshney & Taylor, 1972). This fault is exposed on the coast (Burne & Moore, 1971, pp. 293-294) and can be traced inland eastward to the Mole Valley (National Grid Reference SS 675 260), 3 km east of South Molton (Fig. 1). The existence of tectonically juxtaposed suc·· cessions in the Culm Trough is also indicated by contrasting levels of organic maturation shown by vitrinite reflectance measurements from the two suc cessions (Cornford et al., 1987). Samples from the
Fig. 7. (Opposite.) Petrology and grain-size distributions of the various bed types in the Bude Formation. (A) Comparison of sandstone composition of the Bude Formation (data of Burne, 1969b; Freshney et al., 1979; Melvin, 1986); the Crackington Formation (data of Melvin, 1986); and the present-day Mississippi subsea fan (data of Roberts & Thayer, 1985). (B) Matrix composition of various Bude Formation bed types. Note high matrix content of 'slumped' and 'slurried'' beds, low matrix content of laminated and structureless sandstones, and intermediate matrix content of graded silty sandstones (data of Burne, 1969b) . (C) Plot of mean grain size () against sorting ( oG ) of Bude Formation sandstone grain·· size distributions, with and without matrix included (data of Burne, 1969b). Fields recognized by Melvin (1986) for the Bude Formation and Crackington Formation sandstones are outlined (note that Melvin measured o1 ) .
109
Bude Formation subsea fans QUARTZ
A
� D E2J]
FELDSPAR
8
Bude Formation sandstones Crackington Formation sandstones Mississippi Fan sands
ROCK FRAGMENTS QUARTZ
m; LIJ � E:;l
D
FELDSPAR & ROCK FRAGMENTS
c
Cross laminated beds Structureless & irregularly laminated sandstones Graded silty sandstones "Slurried" & "slumped" beds
MATRIX
3.0
2.0
MELVIN (1986) Bude Formation data
0G 1.0
}
MELVIN (1986) Crackington Formation data
0
Matrix Excluded
6.
Matrix Included
BURNE (1969b) Bude Formation data
110
R. V. Burne
section between Westward Ho! and Abbotsham form a statistically different group compared with those from stratigraphically equivalent rocks from the area between Hartland Quay (National Grid Reference SS 223 248) and Crackington Haven (SX 143 968). This suggests that the two areas have suffered differ ent thermal histories, i.e. different burial histories. 'These two areas must either now be in juxtaposition due to post burial tectonic events (e.g. thrusting) or they are not stratigraphic equivalents, the Bideford Formation (i.e. Bideford Group) being younger and hence less deeply buried than the Bude and Crackington Formations' (Cornford et al., 1987, p. 463). The stratigraphical equivalence of the sec tions was established by Edmonds et at. (1975, 1979) who identified the Gastrioceras amaliae marine band 550 m above the base of the Bideford Group. The Hartland Quay Shale, although of a contrasting facies, also contains this horizon and has been defined by Freshney et al. (1979) as underlying the base of the Bude Formation (Fig. 3). Edmonds et at. (1979) suggested that depositional continuity existed between these successions and concluded that the Northam and Abbotsham Formations (for which they proposed the name Bideford Formation, but for which the term Bideford Group is here retained) lie conformably between the Crackington Formation and the Bude Formation. Freshney et at. (1979) and Melvin (1986) both based their interpretation of the palaeogeography and facies of the Bude Formation on implied continuity between the deltas of the Bideford group and the depositional environments of the Bude Formation. According to Freshney (pers. comm., 1994) 'there is no physical evidence of thrusts in the Bideford area, but it is possible that strike-slip, east-west trending faults may define areas of contrasting thermal and burial history'. Although Freshney (pers. comm., 1994) points out that 'wedge-bedded sandstones characteristic of the Bideford Group can be traced eastward beyond Umberleigh (National Grid Reference SS 610 237) and appear to pass further east into a number of massive Bude-type sandstones', it has yet to be established whether this transition represents depo sitional continuity or tectonic juxtaposition. How ever, Cornford et at. (1987) concluded that the fluvial-deltaic Bideford Group is allochthonous and therefore the palaeogeography of the basin should not be constrained by the apparent associ ation between the Bideford Group and the deeper water Bude Formation. This conclusion supports the interpretation of Burne & Moore (1971) that
the succession in the area of Westward Ho! and Bideford is tectonically juxtaposed to the Culm Basin succession (Figs 1 & 2).
FACIES ANALYSIS OF THE BUDE FORMATION
Four facies are recognized in the Bude Formation, although these are, to a large extent, intergra dational. The term facies is used here to denote the assemblage of beds deposited within a depositional system. A depositional system comprises a sedimen tary environment together with the processes that may operate within it. Beds are formed by individual depositional processes. Facies descriptions
Black shale facies
Black shales (Fig. 8) are the most persistent facies in both the Bude Formation and the underlying Crackington Formation and form valuable lithofacies for stratigraphical correlation (King, 1967; Freshney
Fig. 8. Black shale facies showing regular millimetre-scale graded mudstone laminae with interbedded grey, muddy siltstone laminations, some with thin ripple cross laminations. Scale in centimetres (SS 1985 1313).
Bude Formation subsea fans
& Taylor, 1972). The facies may attain a thickness of 20 m (Fig. 4), and is dominated by uniform, dark mudstone composed almost entirely of sharp-based, graded laminae 5-10 mm thick. Thin sharp-based sets of ripple cross-laminated silt occur at the lower boundary of some of these laminations. Sole marks on these silts comprise prod marks or, rarely, obstacle scours around burrow entrances. Isolated thicker beds can occur in this facies. The clay minerals of the shales consist dominantly of illite (Burne, 1969b). Undeformed shales also contain kaolinite, whereas in deformed shales this is replaced by chlorite, and grey shales contain more kaolinite (Freshney et al., 1979). Siderite bands and ankerite nodules are found in some shales. The grading of the laminae is due to an upward increase in the content of organic carbon. Finely comminuted plant debris is common on parting surfaces. Freshney et al. (1979) record jarosite-natrojarosite efflor escences in sulphurous shales. Higgs (1991) found that carbon/sulphur (CIS) ratios (Berner & Raiswell, 1983, 1984) in four out of five black-shale samples were less than 10. The shales contain little evidence of biological activity apart from the common occurrence of Planolites (Fig. 9) (King, 1967). Tops of interbedded sandstones may contain Diplocraterion parallelum, an'd, more rarely Arenicolites, Teichichnus, Phyeodes
Fig. 9. Replacements of Planolites burrows preserved in a siderite vein in the black shale facies. A 4-cm-long live limpet for scale (SS 1983 0313).
111
and Skolithos (Higgs, 1991). Freshney & Taylor (1972) found that in the lower parts of the Crack ington Formation a marine fauna of goniatites occurs throughout the shales, whereas from the Namurian G1 zone the goniatites occur only in isolated bands within some black shales. The Goniatites, which are truly marine, occur in bands between layers of nodules that contain fossils of fish that were tolerant of low-salinity environments. Goniatites have been found only at two horizons in the Bude Formation (Fig. 4), the Sandy Mouth Shale and the Warren Gutter Shale. In the Warren Gutter Shale, goniatites are present only as spat (Freshney et at. , 1979). Goniatites have not been found in the middle Bude Formation, although fish remains and coprolites occur at two horizons (King, 1967). Cornuboniscus budensis and Elonichthys aitkeni, both palaeoniscid bony fishes, the acanthodian Acanthodes wardi and an eocarid crustacean Crangopsis huxleyi occur in the Saturday's Pit Shale, whereas the Coelocanth Rhabdoderma elegans occurs in the Tom's Cove Shale (Fig. 4). Muddy siltstone facies
This facies consists of grey to dark grey laminated muddy siltstones and thin, sharp-based silty sand stones that show parallel laminations and ripple drift cross-lamination. In places the facies is disrupted by intrastratal folding and small-scale synsedimentary faults. Scouring occurs locally (Fig. 10). The muddy siltstones are composed of carbonaceous material, clay minerals, and silt-sized quartz particles. The grading of the 1-10-mm-scale laminae reflects a decrease in the size of the quartz grains and an increase in the proportion of clay (Fig. 11). The silstones are composed of quartz, illite, kaolinite, chlorite, siderite and minor feldspar (Merriman, quoted in Higgs, 1991). Higgs (1991) found that organic carbon varied from 1.8% to 2.9%, plant fragments were conspicuous, burrows are absent, and C/S ratios (Berner & Raiswell, 1983, 1984) range between 22.8 and 39.0. King (1967) described Kouphnichnium (King, 1965; Goldring & Seilacher, 1971), i.e. xiphosurid feeding trails from this facies. The animals moved over silt-covered surfaces, prodding down into lower silty laminae. Current sole marks were formed before the trackways were emplaced, and the tracks are confined to one level within the graded laminated units, indicating that little sediment was deposited during the time in which the track was being made
112
R. V. Burne
Fig. 11. Photomicrograph of graded muddy siltstone lamination in the muddy siltstone facies. Carbonaceous material appears dark (SS 1 995 0400).
Fig. 10. Muddy siltstone facies showing alternations of laminated muddy siltstone and ripple cross-laminated units. Note irregular scouring and deposition of cross laminated beds at the level of the hammer (circled). Section youngs to the right (SS 201 077).
(Goldring & Seilacher, 1971). Other tracks initially regarded by King (1965) as xiphosurid mating-traces (Fig. 12), were reinterpreted by Higgs (1988) as Undichna or trails left by fish dragging their fins along the soft bottom during feeding. Interbedded sandstone-shale facies
This facies comprises interbedded 'event' beds (Einsele & Seilacher, 1982) and black shales. The relative abundance of various types of event beds in this facies shows no consistency of grouping. At some horizons beds of one of the following types dominate the succession. In other cases a predictable succession of event beds occurs, whereas elsewhere the facies consists of a less predictable succession of beds. The following types of event beds can be distinguished.
Fine-grained graded beds range from 1 to 12 em in thickness. Beds have sharp bases, in places bearing sole marks, and comprise laterally continuous, normally graded layers of siltstone and muddy silt stone (Fig. 13). The sediment is poorly sorted with median grain size 3.5
Bude Formation subsea fans
Fig. 12. Traces in the muddy siltstone facies originally interpreted by King (1965) as traces of xiphosurid mating behaviour, but reinterpreted by Higgs (1 988) as trails left by the fins of bottom-feeding fish (Undichna). Structures have a wave length of 9 em (SS 2017 0751).
Fig. 13. Grey muddy siltstone laminations (base of section) overlain by fine-grained graded beds, then passing up into laminated beds showing 'type 3' ripple drift of Walker (1 963) (SS 201 074).
113
parallel-sided beds 5-25 cm thick. They are gen erally laterally continuous, but may be scoured or may themselves expand into pre-existing scours. The sole marks are predominantly linear current ridge moulds, flute or groove moulds. The lower part of each bed is structureless and grading is obvious only in the topmost part, where aggregates of plant material and fragments of black mudstone may be concentrated (Fig. 18). Median grain size at the base of the unit is around 2.5 ¢, decreasing to about 3.0¢ at the top. The long axes of grains (Fig. 19) in the structureless unit show slight pre ferred orientation in the same direction as that of current flow indicated by sole marks on the base of the bed, and, although the bed as a whole is poorly sorted, the grain-sorting at any one level is moderate to good (excluding matrix oG 0.6-0.75), although matrix content may be up to 50% (Fig. 7). The structureless silty sandstone unit is overlain by a 5-20-mm-thick unit of uniform parallel-laminated silty sandstone of slightly finer grain size and then by a similar thickness of current ripple cross-laminated silty sandstone with either single sets or ripple drift. Water-escape structures or convolution are generally not found in these beds. The cross-laminated unit is overlain by a unit of alternating light and dark parallel-laminated silts, which is, in turn overlain by structureless grey silt. Structureless sandstones are sharp-based and between 10 cm and 1 m thick (Figs 20 & 21). The most common sole marks are flute moulds and linear
114
R. V. Burne
Fig. 14. Sharp based, laminated bed showing irregular ripple-drift with stoss side erosion and convolutions due to water-escape. Note graded top of the 12-cm-thick bed (SS 1999 0493) .
Fig. 15. View to the right of Fig. 14 showing a sharp erosive top to the same bed, and an absence of convolutions (SS 1999 0493) .
current-ridge moulds. Most of the bed is ungraded and most beds are completely structureless, although some contain faint traces of lamination, often dis rupted into dish structures (Stauffer, 1967). They are composed of moderately well-sorted fine-grained sand (excluding matrix aG 0.85-0.9; median grain size 2-2 4 <jl; matrix content 14-25%) (Fig. 7), with grains orientated either parallel to the current direction indicated by sole marks, or bimodally about this direction (Fig. 19). The beds may have .
sharp tops or may grade up, through an abrupt and irregular contact, into either irregular lamination comparable to hummocky cross-stratification, current ripple cross-lamination (Fig. 22) or, very rarely, dune-shaped cross-sets. The beds are gen erally laterally continuous, but are prone to amal gamation. They also can be scoured or can thicken in channels. They may also be confined to channels. 'Slurried' beds (Wood & Smith, 1959) consist of sharp-based units of a poorly sorted slurry of sand,
Bude Formation subsea fans
Fig. 16. Sandstone surface bearing straight-crested ripples (SS 2014 0779).
silt and mud (excluding matrix oa 0.9; including matrix oa 2.5; median grain size 3.5 ¢, matrix content 56-70%) (Fig. 7). They are generally between 2 and 20 em thick (Figs 23 & 24). The sharp base of the bed may bear sole marks, most commonly groove or prod and bounce moulds. The lower parts of the beds show grain orientation in the same sense indicated by sole marks (Fig. 19). The beds may be ungraded or graded only in the topmost part, but they commonly show a two-part division, with a basal massive, indistinctly laminated or, rarely, cross-laminat.ed silty sandstone, and an upper, more poorly sorted 'slurried' unit often containing large contorted fragments of mudstone. In places beds have 'frozen' in the act of peeling up fragments of mudstone from the substratum (Fig. 25). Sandstone or siltstone pseudo-nodules (Fig. 18), interpreted as load balls (Kuenen, 1965), and water-escape tubes may be found in the upper part of the 'slurried' unit.
115
Fig. 17. Sandstone surface bearing cuspate-crested ripples (SS 2014 0774).
Fig. 18. Interbedded sandstone-shale facies showing graded silty sandstones (A) separated by a 'slurried' bed (B). Note load balls falling from laminated unit on top of the 'slurried' bed, and laminated top of the underlying graded silty sandstone (SS 2018 0740).
116 A
R. V. Burne
N 41 Spec. 48
B
C
Irregularly Laminated Sandstones
0
Spec. 14
Graded Silty Sandstones
N 31 5
N 30 2
Spec. 49
Spec. 2A
Structureless Sandstones
N 31 8 Spec. 2C
N 295 Spec. 2B N 3 83
N 40
Spec. 73
Spec.
6 60
1
D
E
"Slumped" Bed
"Slurried" Bed
N
I
1
N 2 87 Spec. 4
N 27 5 Spec. 30
N
N 274 Spec. 52 b
N 3 02 Spec. 31 10
15
I
20
%
Fig. 19. The orientations of long axes of elongate grains N 2
69
Spec. 58
N
=
Number of grains
Spec.
=
Specimen number (BURNE 1969b)
10
15
20
%
This unit may be overlain by a thin unit of laminated sediment consisting of either current ripple cross lamination or the cross-laminated deposits of small sand-volcanoes. 'Slumped' beds are 4-20-m-thick beds of disrupted sediment (Figs 26-29), similar in character to the thinner 'slurried' beds (Figs 7 & 19). The name is well established (e.g. Freshney et al., 1979), but is misleading because the beds are probably not
within representative specimens of various types of sandstone bed (data from Burne, 1969b). Measurements were made from orientated thin-sections sawn parallel to stratification. Palaeocurrent directions from sole marks or cross-laminations associated with the beds are indicated by an arrow on the rose diagram.
true slump deposits. They are of variable extent. Some appear to have a sheet-like form correlatable between sections (King, 1967), whereas others are known from only one section (Edmonds eta!. , 1979; Freshney et a!., 1979) and may pass both laterally and downward into undeformed material. The beds are composed of sediment with three separate origins; exotic muddy sandstone (excluding matrix o0 1.2; including matrix o0 2.5; median grain size
Bude Formation subsea fans
117
Fig. 20. Structureless sandstones within the interbedded sandstone- shale facies showing traces of irregular lamination, possibly due to the amalgamation of individual units. Length of extended tape is 135 em (SS 1998 061 1 ) .
3.5
Thick sandstones with few visible structures are diagnostic of the Bude Formation (Fig. 30). Some sandstones are truly structureless, but in others structure seems to have been obscured by the lack of contrast in grain-size in the beds. Irregular bedding, cross-bedding, parallel lamination, convolute lami nation and parting lineation can all be discerned in places. Some sandstones consist of an amalgamated series of laterally continuous beds of structureless sandstones similar to those described above. Also found in this facies are irregularly laminated sandstones, which are relatively coarse-grained (median grain size of 2-2.9
than load moulds are not common, although moulds of transverse mud ripples have been observed. They may be laterally continuous, or they may wedge out laterally. The beds are characterized by various irregular and often interfering non-parallel lami nations, although the internal structures are some times difficult to distinguish due to lack of grain size contrast. In places, uniform cross-stratification in scalloped based sets up to 20 em deep and 30 em across occur throughout the bed (Fig. 31). The beds are composed of well-sorted medium sand, and represent the coarsest sands observed in the Bude Formation. The grain fabric shows a preferred orientation at right angles to the dip of the lamination (Fig. 19). Small 1-3-m-deep scours occur near the top of some sandstones (Fig. 32). These scours are often filled by overlying beds draping into them. The fill is similar to the sediment into which the scour is cut. Deeper (> 4 m), fiat-bottomed channels also occur in this facies. The fill is confined by the channel margins and comprises only a few units of either massive sandstone or irregularly laminated sand-
118
R. V. Burne
Fig. 21. Structureless sandstones showing irregular structures possibly due to amalgamation. Length of extended tape is 70 cm (SS 2014 0775).
Fig. 22. Structureless sandstone with upper part reworked into irregular laminations (SS 199 038) .
Fig. 23. 'Slurried' bed showing three part division of graded base, centre with homogenized shale fragments, and thin laminated top with load deformation (SS 201 1 0764) .
Bude Formation subsea fans
119
Fig. 24. 'Slurried' bed with contorted fresh shale fragments in upper part of bed (SS 2014 0775 ) .
Fig. 25. Lower part o f 'slurried' bed showing incorporation o f underlying sediments into the bed as How 'froze' (SS 201 077) .
stone. The basal unit of the channel fill can contain large mudstone clasts, suggesting that the processes that deposited these units were directly connected with the erosion of the channel. In places there is evidence that the deposition of sediment was originally localized at one side of the channel, possibly as some form of point bar. Successive channel cuts can be observed within one complex.
An example of a channel complex occurs in the massive sandstone facies between King's (1967) key shales P and Q, in repeated sections between Northcott Mouth (National Grid Reference SS 202 085) and Crooklets Beach (National Grid Reference SS 202 085) (Figs 33 & 34). Mapeo & Andrews (1991) and Tanner (1992) have provided structural interpretations for these features, however, the
120
R. V. Burne
Fig. 26. 'Slumped' bed about 80 m above the Sandy Mouth Shale with sand volcano (arrowed) on surface. Hammer for scale (SS 2015 1026).
Fig. 27. Sand volcano shown arrowed in Fig. 26. Note central crater and Hank lineations (SS 2015 1026).
evidence for sedimentologically distinct channel fills and channel erosion indicate that the discordances are true channel cuts rather than being the result of early thrusting, although these bed boundaries would have provided planes for preferential movement during deformation. Facies interpretation
Black shale facies
The preservation of well-laminated mudstones of
the black shale facies indicates deposition under tranquil conditions in a basin where sediment was introduced by discrete and gentle currents, at least some of which were bottom-hugging. The rare occurrences of pelagic goniatites and bivalves can be correlated with marine bands in other Westphalian Basins and probably represent brief connections with open marine conditions during eustatic high .. stands. There is no other indication of marine conditions in the succession. The preservation of nektonic fish remains along with their faecal material in well-laminated strata lacking remains of macro··
Bude Formation subsea fans
121
Fig. 28. The Black Rock 'slumped' bed exposed north of Upton Cross. Note load balls foundering from the top of the bed and 1 . 8 m long sandstone clast (arrowed) (SS 200 053).
Fig. 29. Sandstone clast shown arrowed in Fig. 28. Note that the deformation is reminiscent of that found in load balls, and when the sandstone clast is restored to its former geometry an original lenticular cross-section is revealed, which , together with the irregular internal laminations, suggest that the feature may originally have formed as a sand volcano on the top of the 'slumped' bed (SS 2000 0529).
benthos indicates an anaerobic environment (Savrda & Bottjer, 1991), although the restricted ichnofacies dominated by Planolites, suggests dysaerobic con ditions (Sageman et a!. , 1991). The oxygen depletion may have been caused by the increased oxygen demand of 'surplus' organic matter (Wetzel, 1991)
brought in by the fine-grained turbidity currents in the form of finely comminuted terrestrial plant matter. Evidence for anaerobic and dysaerobic conditions is supported by the CIS ratios for the shales (Berner & Raiswell, 1983, 1984) reported by Higgs (1991). The bed-top burrows reflect
122
R. V. Burne
Fig. 3 1 . Vertical section through climbing cross-stratified sets in irregularly laminated sandstones. Arrow shows younging direction (SS 2003 0818).
Muddy siltstones facies
Fig. 30. Thick, irregularly bedded sandstones of the sandstone-dominated facies. Figure for scale, section youngs to the right (SS 281 080) .
opportunistic colonization. They do not extend above the top of the sand bed, indicating that the animals responsible died when mud-deposition resumed (Higgs , 1991) , a characteristic of turbidite successions (Seilacher, 1982) . The laminations either result from the settling of dispersed suspended material that was received in discrete pulses, possibly due to seasonal flooding of distant rivers (Burne, 1969b; Melvin, 1976; Higgs, 199 1) , or as high concentration sediment clouds (fine-grained tur bidity currents) (Burne, 1969b; Melvin, 1976; Glenn & Kelts, 1991). Melvin ( 1986) has pointed out that these laminations could be described as T3_7 and T6_7 mud turbidites using the terminology of Stow & Shanmugam (1980).
The regular lamination of the muddy siltstone facies, absence of body fossils, presence of bottom-feeding traces, and high CIS ratios indicate that this facies was probably formed below storm wave base in an oxygenated and possibly freshwater environment. Higgs ( 1991) concluded that xiphosurid tracks are known only from rocks that appear, on other evi dence, to have been deposited in waters no deeper than a few tens of metres. However, Goldring & Sielacher ( 1971) point out that arthropods inhabit practically all marine environments, and that deeper parts of sedimentary basins where silt and clay laminae are deposited from low-velocity currents or distal turbidity flows provide the highest fossilization potential for xiphosurid trackways. It is concluded that this facies was deposited by a series of relatively gentle but persistent bottom-hugging currents (Burne, 1969b; Goldring & Seilacher, 197 1; Melvin, 1986) , the deposits of which Melvin (1986) has compared to T0 and T023 fine-grained turbidites o:f Stow & Shanmugam (1980). Interbedded sandstone-shale facies
All the bed types encountered in the interbedded
Bude Formation subsea fans
Fig. 32. Small scour in sandstone-dominated facies (outlined) with fill of structureless sandstone beds (SS 2014 0771 ) .
sandstone- shale facies originated from short-lived bottom-hugging density currents with a higher content of suspended matter than the surrounding water, i.e. turbidity currents in the sense of Sundborg (1956, p. 60). This conclusion is supported by the high matrix content of all the beds (Fig. 7). The beds are all composed of fine-grained sediment (virtually no sediment particles larger than 1 ¢) that could be carried in turbulent suspension by turbidity currents of reasonable velocity (Lowe, 1982) and would favour the development of autosuspension (Bagnold, 1962 ) . The following detailed interpretations may be made for each of the bed types in this facies. Fine-grained graded beds are interpreted as tur-
123
Fig. 33. 'Thickening upward' succession of black shale facies, interbedded sandstone-shale facies, and sandstone dominated facies, the latter forming a channel fill. Structureless sandstone beds have accreted against the channel margin (accretion surfaces outlined) . Cliffs are 60 m high, section youngs to the right (SS 2014 0630) .
bidites in which the structureless graded beds result when rapid deposition of sediment outpaces the ability of a decelerating current to make stable bed forms (Burne, 1969b; cf. Melvin, 1986). The result is a graded unit in which grains do not have an oppor tunity to be preferentially orientated and are depo sited in a reasonably stable packing position. Laminated beds are interpreted as laminated tur bidites, equivalent to the Bouma (1962) Toed, Ted or Tct types. The rate of deceleration of the depositing current was low enough for the deposited sediment to construct stable bedforms beneath the waning flow (Burne, 1969b; Melvin, 1986). Graded silty sandstones are interpreted as
A
>-'
ss 20000817
�
::>;:)
:<::::
B
O:l
!::
ss 20020774 SS 201 60n2
D [:. :: : ·.] � �==j r:t//1 • � ..
� 1 m[
Sm
Structureless sandstones
Graded silty-sandstones
Cross-bedded sandstones
Grey muddy-siltstone
& laminated beds
'Slurried' beds & fine-grained graded beds
Black shales
Duplex
Fig_ 34_ Cross-sections of a channel complex in sandstone-dominated facies between key shale P and key shale Q (Tom's Cove Shale) (King, 1967). The complex consists of an earlier channel filled with beds of structureless sandstone, and a later channel filled with cross-stratified irregularly laminated sandstone. (A) Exposure in section 39 of King ( 1967), North of Earthquake Cliff. Note shale clasts at base of channel fill, apparent lateral accretion of structureless sandstones, and final cross bedded channel-fill. (B) The same complex 500 m south in section 36 of King ( 1 967) .
3 "'
Bude Formation subsea fans
turbidites showing the almost complete Bouma se quence Tabcde (Burne, 1969b; Melvin, 1986) , although the 'd' division is rarely well developed (Bouma, 1962). In the rapid deposition of the 'a' division, grains do not have an opportunity to be preferentially orientated, although they are deposited in a reason ably stable packing position. Deposition of the massive units of the structureless sandstones may have taken place from the bedload of a current flowing during the strongly fluctuating hydrodynamic conditions and vigorous burst-sweep cycles of large turbidity currents (Hiscott, 1994). Deposition would be either as successive sheets of sediment (under conditions of plane bed with movement) or as more irregular lenses of sediment (under conditions of antidune flow). Subsequently the escape of pore-water, trapped in the sediment during deposition, and the compaction of the unit may have destroyed this primary structure. Alterna tively the structureless parts of these beds may have developed if the tangential stress applied by the depositing current was high enough to maintain the dispersive stress of a bedload, the volume of which was being greatly increased by sediment falling out of suspension. This process of transport causes pro nounced orientation in the grain fabric and is the equivalent of the traction carpet of Dzulynski & Sanders ( 1962), but is different from the process of grain flow (Sanders, 1965 ; Stauffer, 1967, p. 502) in which the tangential stress is applied by gravity to sediment resting on a slope. Traction carpets are essentially current-impelled grain flows (Middleton & Hampton, 1976) or cohesionless debris flows (Postma, 1986). As the current waned, the tangential stress applied to the bedload would decrease to a point where the dispersive stress in the traction carpet could no longer be maintained. At this point the traction carpet would abruptly cease to move. The reworked nature of the tops of these beds results from the tail of the turbidity current con tinuing to shear the surface of the sediment deposited by the body of the current, and reflects the probable irregularities of that surface. , The grain size distributions of the 'slurried' beds and 'slumped' beds are very similar to that of modern mudflows (Baker, 1965) except that there are no very coarse grains. All the features of the units considered here can be explained by deposition from turbidity currents carrying a very high concentration of fine-grained sediment in suspension. Bagnold ( 1968) has demonstrated that such currents expend little energy and sediment transport might therefore
125
continue infinitely but for the fact that the gradual loss of water from the current reduces the current's volume to a point where the concentration of the suspended sediment suppresses fluid turbulence, and the current would essentially evolve into a cohesive debris flow, or mudflow, without passing through an intervening tractional stage (Postma, 1986; Postma et al. , 1988). This dense current flowing over a soft muddy bottom would tend to peel up the substrate and incorporate it into the flow in the manner described by Dzulynski & Radomski (1966) . The mud would be contorted by the flow, and would tend to disaggregate. However, it would also increase the concentration of transported solids in the flow, thereby encouraging, 'freezing' of the flow. Postma et at. (1988) concluded that a rapid transformation of flow behaviour is necessary for 'soft' intraclasts to be preserved. When the critical concentration of solids is reached the flow would stop abruptly and leave a quick bed. The coarser sediment. of the unit would not be disturbed by this and would preserve any original preferred orientation (Fig. 19). Any sediment deposited on the surface of the unit at this stage would tend to founder into it because of the lack of strength of the bed (Figs 18 & 29) . Excess water would be expelled during the compaction of the bed, forming water-escape structures, pipes of matrix-free sand, and constructing sand volcanoes on the surface of the bed (Fig. 27) . This sequence of post depositional events has been reproduced experimentally by Nichols et al. (1994). The surface of the bed may continue to be sheared by a residual current, with the result that a thin unit of current ripple cross-lamination forms on the upper surface of the bed. Sand volcanoes on the surface of 'slurried' beds may show an asymmetry in the same sense as that indicated on the sole mark of th.e bed as a result of eruption into a current (Burne, 1970) . Sandstone-dominated facies
The comparatively coarse, moderately sorted nature of the beds of the sandstone-dominated facies suggests that they have been deposited from the bedload of a transporting current. The structureless sandstones of this facies have a similar origin to those of the interbedded sandstone shale facies, but are commonly amalgamated to form thick sand stones. The irregularly laminated sandstones are of two types. In the majority of beds the irregular lamination cannot be related to any bedform in the lower flow regime, but strongly resemble those
126
R. V. Burne
produced by Middleton (1965, Fig. 3) in an exper imental study of deposition from a current forming antidunes. ln other sandstones, better organized cross-stratification can be discerned, which can be related to migrating dunes. In both these sandstone types the structures are persistent throughout the bed and there is little evidence of grading, or waning flow conditions. Burne (1969b) considered that they were deposited by single events characterized by discrete, short-lived, but relatively powerful currents of almost constant velocity.
FACIES SUCCESSION
Description
The long-term evolution of a sedimentary succession at any given place involves temporal transitions between environments, each of which is character ized by different depositional systems producing distinct facies. The result is a facies succession. Reading ( 1971) noted the lack of predictability of Bude Formation facies successions when compared with other cyclic Silesian successions, such as the synchronous paralic or deltaic facies succession of the Bideford Group (de Raaf et al. , 1965). The facies succession of the Bude Formation (Figs 35 & 36) shows alternation between end-member con ditions of tranquil mudstone sedimentation and the rapid emplacement of channelled massive sandstones with a complete gradation of intermediate environ ments. Burne ( 1969b) identified a composite suc cession in the Bude Formation consisting of a succession of black shale facies, a thickening upward succession of interbedded sandstone-shale facies, a channelled sandstone dominated facies, and then a thinning upward succession of interbedded sand stone-shale facies and silty mudstone facies that passes abruptly up into another black shale facies (Fig. 5). This succession was derived largely from relationships observed in the section between the key shale 0 and the top of the Tom's Cove Shale Member of King (1967) (Figs 35 & 36). However,
much of the Bude Formation does not conform to well-developed thickening-up or thinning-up suc cessions (Melvin, 1986). The organization of these parts of the Bude Formation is less predictable, although it is not random (Fig. 4). Interpretation
The case against a storm-influenced shelf
Higgs (1991) concluded that The Bude Formation was deposited largely, if not entirely, above storm wave base, as indicated by the fact that wave influenced structures occur throughout the suc cession, spaced no more than a few metres apart.' Melvin (1986) thought that the presence of Planolites, xiphosurid trackways and possible hummocky cross stratification in the succession meant that the Bude Formation had formed as a subsea fan 'within relatively shallow (shelf as opposed to abyssal) water depths' (p. 26). Higgs (1987) queried how a fan could form at such shallow depths 'where the normal configuration is a wave-cut shelf ' (p. 378). Melvin (1987, p. 382) replied that it is 'hard to imagine how 1300 m could accumulate by the passive aggradation of storm deposits on a lacustrine shelf, wherein facies changes are determined by subtle salinity variations associated with a yo-yo effect of eustatic sea level changes'. This refers to Higgs's (1986b, 1991) interpretation that each alternation of 1-10m-thick sandstone with shales of similar thickness within the Bude Formation resulted from fluctuations in depth and salinity brought about by intermittent over-topping of the sill of a silled basin. According to Higgs (1991), the sandstones were deposited from river-generated storm flows, diverted and modified by wave activity. The river-fed turbidity current and the wave-induced oscillatory current never varied to the point where the wave processes dominated distally, as would be expected in the case of a localized river underflow operating through a tempest (Brenchley, 1985). Higgs (1991) interpreted the ripple cross-lamination developed in the upper parts of these sandstone beds as 'quasi-symmetrical
Fig. 35. (Opposite.) Lateral variation in facies successions between key shale 0 and the top of key shale Q (Torn's Cove Shale). Key shales and section numbers of King ( 1967). The individual beds of the interbedded sandstone-shale facies are shown in a generalized way to demonstrate the variability of this facie. Note the tendency for thickening upward sequences to follow the black shale facies, and for the muddy siltstone facies to occur beneath the black shale facies. Note the position of the channel complex of Fig. 34 in sections 39 and 36. (Locations: Earthquake Cliff, SS 201 079; Upton Cross, SS 200 049; Lower Longbeak, SS 198 032; Widemouth Sand, SS 199 023).
Bude Formation subsea fans
127 Section 5 (Widemouth Sand)
D.. B § 0
Irregular ly laminated or structure less sandstones
Section 9 (South of Lower Long beak)
"Slurried beds" or fine grained graded beds Cross laminated beds
Graded silty sandstones
� t:::::=::j
Muddy siltstones
D
Black shales
Section 1 8 1 (Upton Cross) 1
I
�m�
I Section 36 (South of earthquake) I I · >: : · . ·:. : - : · I · -: : : · : I .
[
TOM'S
COVE
(Key
SHALE Q)
shale
Section 39 (North of earthquake)
.. .·
·. · · . : .
· · ::'
.. ·
shale
.
.
P)
0)
.
128
R. V. Burne
Fig. 36. A 60-m-high cliff exposure of the facies succession of section 36 (SS 201 077) as logged in Fig. 35 . Section youngs to
the right. Arrows mark the three 'thickening upward' successions from key shale 0 (left), key shale P (centre) and Tom's Cove Shale or key shale Q (right) . Highlighted in the foreground is the channel-cut shown in the upper part of Fig. 34B. This is filled with cross-stratified, irregularly laminated sandstones comparable to those shown in Fig. 31 from the laterally equivalent channel fill shown in the upper part of Fig. 34A.
ripples' intermediate between current ripples and wave ripples. However, the statistics quoted by Higgs ( 1991) in support of a wave-generated origin for these ripples are not definitive (Reineck & Wunderlich, 1968) and they could equally represent non-equilibrium bedforms developed under waning flow conditions (Baas, 1994). Unequivocal oscil lation ripples have not been reported from the Bude Formation (Goldring & Seilacher, 1971), neither have 'clean' sandstones characteristic of wave sorting (Fig. 7; Seilacher, 1982). The 'quasi-symmetrical ripples', described by Higgs (1983, 1984, 1987, 1991) as storm-wave influenced structures are found in beds that are quite different from the wave-generated lithotypes described by de Raaf et al. (1977), and I consider that these structures have no depth-related connotation and are more logically interpreted as the product of reworking by the tail of a turbidity current of sediment previously deposited by the
body of the same turbidity current (Middleton & Hampton, 1976). Larue & Provine (1988) illustrate similar irregular cross-laminated sets from turbidites on Barbados, which they attribute to deposition as vacillatory turbidites emplaced by multiple surging flows. Higgs ( 1991) cited the occurrence of multi directional tool marks and mud-draped scours as further evidence of wave action in the Bude Formation. Diametrically opposed sole marks on the same surface are a characteristic of some tem pestites (Seilacher, 1982), but do not occur in the Bude Formation. The multidirectional sole marks that do occur are similar to those that Middleton & Hampton ( 1976) have attributed to divergence of flow directions within a strongly lobate turbidity current head. Mud-draped scours are not definitive of wave action because Normark & Piper ( 1991) have recorded them from submarine fan-channels.
Bude Formation subsea fans
Rare occurrences of hummocky cross-stratification in the Bude Formation were cited by Higgs (1991) as evidence of storm-wave activity reworking bottom sediments. Einsele & Seilacher (1991) claim that hummocky cross stratification is not found in tur bidites, but Prave & Duke ( 1990) have concluded that small-scale hummocky cross-stratification does not indicate a particular flow condition or depo sitional environment. They interpret examples similar to those described by Higgs (1983, 1984, 1991) to have been formed by antidunes generated by standing waves at the interface between the depositing body and tail of a turbidity current and an overlying low-density layer. Similar structures in turbidites also have been interpreted as antidune structures by Walker ( 1967), Skipper (1971), Hand et al. ( 1972) and Yagishita ( 1994) . Alternatively, Larue & Provine (1988) have interpreted hummocky cross-stratification found in turbidites as being due to deposition of laminated sand on a surface modified by fluidization and current scour - a situation comparable to that inferred for some of the Bude Formation examples. Mutti (1977, fig. 16) figures thin parallel laminae in a basin-plain-facies turbidite that closely resemble the structures interpreted by Higgs ( 1983, 1984, 1991) as hummocky cross stratification. In the absence of associated unequi vocal evidence for shelf sedimentation, such as that cited by Monaco (1992), the rare hummocky cross stratification recorded from the Bude Formation is interpreted to represent the result of either upper flow-regime structures or traction across a quick bed during deposition from a turbidity current. Brenchley et at. (1993) have presented an in terpretation for the Lower Ordovician Bell Island Group of Newfoundland that is remarkably similar to that presented by Higgs (1991) for the Bude Formation, but the facies succession is quite differ ent. In contrast to the Bude Formation, the Bell Island Group shows more predictable facies suc cessions, contains no sole marks, but wave ripples, hummocky cross-stratification, herring-bone cross stratification, gutter casts and bioturbation are all common. The case for turbidite systems
The substantial thickness of the Bude Formation, the relative continuity of single beds, and the absence of a shallow-water benthic fauna all support a turbidite-basin interpretation rather than that of a storm-affected shelf (Einsele & Seilacher, 1991).
129
The lack of predictable vertical associations in the Bude Formation is of itself evidence against a shallow water or paralic origin for the succession, because a 'large proportion of turbidite successions cannot be assigned to any form of regular cyclicity' (Piper & Stow, 1991, p. 371). Hiscott (1981) suggested that cyclical turbidite successions, identified solely on visual criteria, could be explained by chance occur rences within unordered successions of turbidite beds. However, Burne ( 1969b) recognized thicken ing upward and thinning upward successions, which he related to subsea fan environments. Similar successions have since been described from both ancient and modern turbidite fan environments. For example, Mutti & Ricci-Lucchi ( 1972) and Ricci Lucchi ( 1975) described turbidite successions from the northern Appenines of Italy that contain sub marine fan successions or 'second-order cycles'. These either followed one another directly or were separated by 2-100 m of monotonous succession interpreted as the deposits of the open fan or basin plain. They contained thickening upward cycles that represented prograding lobes (60% of these cycles were non-channelled), and thinning upward cycles that represented a channel-fill succession (80% of these were channelled). It is concluded that the Bude Formation can also be interpreted as the deposits of a turbidite system or subsea fan (defined without reference to shape). Thus the sandstone- dominated facies may represent the deposits of an accretionary fan-valley system (O'Connell et al. , 1991). The fan-valley is on a scale much greater than that of the outcrop, but within this overall valley 'thalweg' channels occur with a scale observable in outcrop. The facies has the complex organization typical of either the fan channel facies (Mutti, 1977), characterized both by thick-bedded channelized units and thinner bedded units, or of the channel-mouth facies of Mutti (1977), in which beds show lensing and wedging, and massive units are overlain abruptly by dune or ripple cross stratification. In the silty mudstone facies the evidence of frequent though gentle turbidity currents suggests deposition either in an elevated position within a fan-valley complex or on a channel levee. This interpretation is supported by the evidence of apparent flow instability (Piper & Stow, 1991). Normark & Piper (1991) described levee deposits as containing either discontinuous sands with wavy bedding and climbing ripples or abundant thin sand silt laminae. Further support for this interpretation
R . V.
no
is provided by Mutti ( 1977), who found that thin, rippled sandstone units that diverged or expanded in thickness were characteristic of a channel-margin facies. There are three possible environments in which the interbedded sandstone- shale facies could form (Mutti, 1977; Normark & Piper, 1991) . In channel levee and interchannel environments, turbidite deposits will reflect the degree to which sequential channelled flows overtop the levee. They will show variability in both vertical succession and thick ness. Deposition on a lobe at the end of a leveed fan-valley system will show less vertical or lateral variability, but preserve evidence of continuity of aggradation. Deposits on an unchannelized open fan will tend to be composed of laterally continuous beds of rather uniform character, although reflecting currents of different sizes. The character of the interbedded sandstone- shale facies is generally not uniform and it is concluded that it represents either the deposits of levee and interchannel environments or of lobes that accumulated at the end of leveed fan-valleys. The black shale facies is composed of fine-grained rhythmic couplets typical of profundal lake deposits (Glenn & Kelts, 199 1 ) and interpreted as the deposits of open fan or basin sedimentation.
CONCLUSIONS
It is concluded that the Bude Formation was depo sited in a land-locked synorogenic foreland basin. Deposition occurred on the northern, inactive margin of this basin. Despite the generally unpredict able nature of much of the succession, periods of black shale sedimentation alternated with periods of turbidite deposition in which subsea fans were constructed. The conclusion of Melvin ( 1986) that parts of these fans did, on occasion, rise above storm wave base cannot be absolutely excluded, but it seems more likely that the environment was a pro fundal lake or inland-sea basin (Glenn & Kelts, 1991) in some ways comparable with the East African Lakes (Cohen, 1990; Scholz & Rosendahl, 1990; Baltzer, 1991; Scott et al., 1991) or Lake Baikal (Hutchinson et al. , 1992). Earlier alternative interpretations of the Bude Formation (Owen, 1950; Prentice, 1962; King, 1965, 1967; Edmonds et al. , 1968; Freshney et al., 1972, 1979) seem to be based, in part, on a lack of appreci ation of the extensive variation of structures now
Burne
known to be possible in turbidite systems (Skipper & Middleton, 1975 ; Middleton & Hampton, 1976; Postma, 1986; Postma et al., 1988; Normark & Piper, 1991; Ghibaudo, 1992) . Reading (1987 , p . 8 ) has reminded us that: 'So often a new fashion or model is developed by the subordination of alternative hypotheses or possibilities'. The model developed by Higgs (1991) for a storm-influenced shelf is based on the equi vocal evidence of hummocky cross-stratification and 'quasi-symmetrical' ripples (Higgs, 1983, 1984, 1987, 1988, 1991). It is confirmed neither by facies associations nor by stratification characteristics of associated sediments, and, although Melvin ( 1 987) considered that the presence of wave ripples in the Bude Formation is not in dispute, no convincing illustration of them has been published to date. The problems regarding the depositional environment of the Bude Formation all disappear when it is realized that there is no other substantial evi dence for shallow-water environments in the Bude Formation. The ichnofacies have no depth conno tation (Goldring & Seilacher, 197 1 ; Sageman et al., 1991; Goldring, 1993). Whalley & Lloyd ( 1986) note that the thickness of the Bude Formation is exceptional for the depositional environment pro posed by Higgs (1983, 1984), and Higgs ( 199 1 ) himself found it 'curious that such a thick 1300 m shallow water succession is totally devoid of near shore or emergent features'. AU the beds of the Bude Formation have features consistent with having been emplaced by turbidity currents and there is no evidence that sediment flowed into the basin by any other mechanism. What evidence is there for the form of this turbidite system? Although some have argued for a prodelta setting (Melvin, 1986) and others for a base of slope ramp (Hartley, 1991) , the deposits are more compatible with a basinal rather than a slope environment. The only constraint on the depth of this basin is the requirement that the basin should have a marginal slope sufficient to ensure either the maintenance of hyperpycnal flow or the 'ignition' (Parker, 1982) of turbidity currents (Normark & Piper, 199 1 ). Although the black shalefacies contains evidence of changes in basin salinity, marine incursion, and degree of bottom oxygenation, none of these appear to have had any direct influence on either the supply of sand to the basin or the gen eration of turbidity currents. The supply of sediment to the basin was probably greatest during periods of relatively low basin water level and consequently
Bude Formation subsea fans
very few powerful turbidity currents were initiated during periods of high water level. The compositions of the various turbidites indicate that they all have the same provenance, but there are marked differences in the amount of matrix in the various beds (Fig. 7). A possible explanation for these different depositional characteristics from similarly sourced turbidity currents is to view them in terms of high- and low-density flows. Low-density flows would exhibit a proximal to distal progression from full Bouma sequences Tabcde through to Tde sequences by sedimentation from fully turbulent flows. High-density flows would deposit bedload, perhaps from traction carpets, in more proximal situations, whereas distally concentrated auto suspensions would flow until grain concentration increased to a limiting point for turbulence and the flow would be rapidly arrested by transformation into a debris flow. It is possible that the spilling of part of the turbidity current over channel levees would suddenly dissipate energy in the channelled flow, causing bedload deposition, possibly as a result of traction carpet collapse, whereas the overbank part of the flow would carry suspended sediment only and would also dissipate energy until it increased its concentration to that of a high-density flow, eventually 'freezing' to form a 'slurried' or 'slumped' bed. These beds of contorted sediment have con ventionally been regarded as having originated either as slumps or debris flows (Freshney et al., 1972; Melvin, 1986; Hartley, 1991) and their interpretation as mudflows or cohesive debris flows , which are the distal deposits of high-concentration turbidity currents, is a novel concept with considerable palaeogeographical significance. Thjck and extensive 'slumped' beds interbedded with black shales may represent large muddy flows triggered by changes in water level in the basin. The thicker, more restricted 'slumped' beds probably represent the distal deposits of channelled flows that deposited massive sands in more proximal environments. These units may be compared with the chaotic silt beds recorded by Nelson et al. (1992) from the outer lobes of the Mississippi fan. The determination of whether a fan model is appropriate for the Bude Formation depends on the recognition and distinction of fan-lobe, fan-channel/ mouth-bar and interchannel deposits. Some parallels can be drawn with the fan successions described by Mutti & Ricci-Lucchi (1972) and Ricci-Lucchi (1975), though these were formed on active tectonic margins. Although the Culm Basin was synorogenic,
131
the Bude Formation was deposited on the inactive margin of the basin (Burne, 1969b; Haslam & Scrivener, 1991). The sandstones are actually richer in quartz (Burne, 1969b; Freshney et al. , 1979; Melvin, 1986) than are the sands of a modern passive margin fan, the Mississippi fan (Roberts & Thayer, 1985) (Fig. 7). Shanmugam & Moiola (1988) con cluded that interpretations derived for active-margin fans should be applied with caution to passive margin fans because of differences in spatial distri bution of turbidite facies and their associations (cf. fig. 8 of Normark & Piper, 1991). Shanmugam & Moiola ( 1988) found that characteristics of mature passive-margin taos included: low sand-to-mud ratios in their sediment supply; hyperpycnal flow is import ant in fan evolution; flows are uniform and of low velocity; and very large muddy flows may be triggered by sea-level changes. The result is a fan characterized by steady growth of narrow levees, progradation of channel-levee systems, and uniform aggradation of lobes - all features consistent with the facies suc cession of the Bude Formation.
ACKNOWLEDGEMENTS
Harold Reading introduced me to the challenges of the Bude Formation, and supervised my doctoral research. A.F. King generously made correlations of the Bude Formation available to me prior to their publication. E.C. Freshney, W. D. Gill, L.R. Moore, J.F.M. de Raaf, E.R. Oxburgh, D. Tappin, E. K. Walton, G. Whitnall, A. Wood and G. Young are thanked for discussions of various aspects of the study. Norma Burne made invaluable contributions to this research. The Chadd family of Bude provided hospitality, assistance and friendship. R.J. Korsch, J.F. Lindsay, R. Norris and T.S. Loutit commented on an early draft of this paper. R. G. Walker provided encouragement over a long period, and, together with E.C. Freshney and A.G. Plint, vigorously reviewed the manuscript. I gratefully acknowledge the support and assistance of the Australian Geo logical Survey in many aspects of the preparation of this manuscript. Gail Hill of the Cartographic Services Unit (AGSO) drafted the diagrams.
REFERENCES
AsHWIN, D . P . ( 1957) The structure and sedimentation of
132
R. V. Burne
the Culm sediments between Boscastle and Bideford, north Devon. PhD thesis, University of London. AsHWtN , D . P . ( 1 958) The coastal outcrop of the Culm Measures in south-west England. In: Abstracts of the Proceedings of the Conference of Geologists and Ceo morphologists in the southwest of England, Exeter, 1 958 (Eds Simpson, S. & Robson, J . ) , pp. 2 - 3 . Royal Geo logical Society of Cornwall, Penzance. BAAS, J . H . ( 1 994) A flume study on the development and equilibrium morphology of current ripples in very fine sand. Sedimentology, 41, 185-209. BAGNOLD, R.A. ( 1962) Auto-suspension of transported sediment; turbidity currents. Proc. R. Soc. London, Ser. A , 265, 315-319. BAGNOLD, R.A. ( 1 968) Deposition in the process of hydraulic transport. Sedimentology , 1 0 , 45 -56. BAKER, P.E. ( 1965) The geology of the Mt Misery Volcano, St Kitts. PhD thesis, University of Oxford. BALTZER, F. ( 1991) Late Pleistocene and Recent detrital sedimentation in the deep parts of northern Lake Tanganyika (East African Rift) . In: Lacustrine Facies Analysis (Eds Anad6n, P . , Cabrera, L l . & Kelts, K . ) , Spec. Pubis int. Ass. Sediment . , No . 13, p p . 147 - 173 . Blackwell Scientific Publications, Oxford. BERNER, R.A. & RA ISWELL , R. ( 1983) Burial of organic carbon and pyrite sulfur in sediments over Phanerozoic time: a new theory. Geochim. Cosmochim. Acta, 47, 855-862. BERNER, R.A. & RAISWELL, R. ( 1 984) CIS method for distinguishing freshwater from marine sedimentary rocks. Geology, 12, 365-368. BESLY, B . M . (1988) Late Carboniferous sedimentation in northwest Europe: an introduction. In: Sedimentation in a Synorogenic Basin Complex: the Upper Carboniferous of Northwest Europe (Eds Besly, B.M. & Kelling, G . ) , pp. 1 - 7 . Blackie, Glasgow. BouMA, A.H. ( 1 962) Sedimentology of some Flysch Deposits. A Graphic Approach to Facies Interpretation. Elsevier, Amsterdam, 168 pp. BRENCHLEY, P.J. (1985) Storm influenced sandstone beds. Modern Geol. , 9, 369-396. BRENCHLEY, P.J . , PICKERILL, R.K. & STROMBERG , S.G. ( 1 993) The role of wave reworking on the architecture of storm sandstone facies, Bell Island Group (Lower Ordovician), eastern Newfoundland. Sedimentology, 40, 359-382. BuRNE, R.V. ( 1969a) Observations on the deposition of the Bude Sandstones. Proc. Ussher Soc. , 2, 109 - 1 10. BuRNE, R.V. ( 1969b) Sedimentological studies of the Bude Formation. PhD thesis, University of Oxford. BURNE, R.V. ( 1 970) The origin and significance of sand volcanoes in the Bude Formation (Cornwall). Sedimentology, 15, 2 1 1 - 228. BuRNE, R.V. ( 1971 ) Sedimentological unit definition , hypothesis testing, and multi-step dependence analysis for lithological sequences. Program with Abstracts, VIII International Association of Sedimentologists, Congress, August 31-September 3 Heidelberg, p. 16. BuRNE, R.V. ( 1973) Palaeogeography of south west England and Hercynian continental collision. Nature (Phys. Sci.) , 241, 129 - 13 1 . BuRNE, R.V. (1976) Submarine fans and their hydrocarbon potential. Abstracts, 25th International Geological Congress, August 1 7-24, Sydney, Vol. 2, pp. 340-34 1 .
BURNE, R.V. & MooRE, L.J. ( 1971 ) The Upper Carbon- iferous rocks of Devon and North Cornwall. Proc. Ussher Soc. , 2, 288-298. COHEN, A.S. (1990) Tectono-stratigraphic model for sedi- mentation in Lake Tanganyika, Africa. In: Lacustrine Basin Exploration: Case Studies and Modern Analogs (Ed. Katz, B . J . ) , Mem. Am . Assoc. petrol. Geol. , Tulsa, 50, 137- 150. CORNFORD, C . , YARNELL, L. & M URCHISON , D.G. ( 1987) Initial vitrinite reflectance results from the Carboniferous of north Devon and north Cornwall. Proc. Ussher Soc. , 6, 461 -467. DE RAAF, J.F.M . , BOERSMA, J.R. & VAN GELDER, A . ( 1977) Wave-generated structures and sequences from a shallow marine succession, Lower Carboniferou s , County Cork , Ireland. Sedimentology, 24, 45 1 -483. DE RAAF, J .F.M . , READING, H. G. & WAL KER , R.G. ( 1 965) Cyclic sedimentation in the lower Westphalian of north Devon, England. Sedimentology , 4, 1 - 52 . DZULYNSKI, S. & RADOM SKI , A. ( 1 966) Experiments on bedding disturbances produced by the impact of heavy suspensions upon horizontal sedimentary layers. BulL Acad. Pol. Sc. Ser. Sci. Geol Geogr., 14, 227-230. DZULYNSKI, S. & SANDERS , J . E . ( 1962) Current marks on firm mud bottoms. Trans. Conn. A cad. Arts Sci., 42, 57-96. EDMONDS, E . A . , McKEow N , M . C . & WILLIAMS , M. ( 1 975) British Regional Geology, South-West England. HMSO, London. EDMONDS, E . A . , WRIGHT, J . E . , BEER, K . E . , HAwKES , J . R . , WILLIAMS , M . , FRESHNEY, E . C . , et al. ( 1968) Geology of the country around Okehampton. Memoir of the Geological Survey of Great Britain for the One-inch Geological Sheet 324, New Series. Institute of Geological Sciences, HMSO, London, 256 pp. EDMONDS, E . A . , WtLLlAMS, B . J . & TAY LOR , R.T. ( 1979) Geology of Bideford and Lundy Island. Memoir of the Geological Survey of Great Britain for 1 : 50 000 Geological Sheet 292. New Series, with sheets 275, 2 76, 291, and part of sheet 308. Institute of Geological Sciences, HMSO, London, 143 pp. EtNSELE, G. & SEtLACHER, A. ( 1982) Cyclic and Event Stratification. Springer-Verlag, Berlin , 536 pp. EtNSELE, G. & SEtLACHER, A. ( 1991) Distinction of tempestites and turbidites. In: Cycles and Events in Stratigraphy (Eds Einsele, G . , Rieken, W. & Seilacher, A . ) , pp. 377-382. Springer-Verlag, Berlin. ELLIOlT, T. (1976) Upper Carboniferous sedimentary cycles produced by river-dominated elongate deltas . J. geol. Soc. London, 132, 199-208. FRANKE, W. & E NGEL , W. ( 1 988) Tectonic settings of synorogenic sedimentation in the Variscan Belt of Europe. In: Sedimentation in a Synorogenic Basin Complex: the Upper Carboniferous of Northwest Europe (Ed. Besly, B.M. & Kelling, G . ) , pp. 8-23. Blackie . FRESHNEY, E.C. & TAY LOR , R.T. ( 1972) The Upper Carboniferous stratigraphy of north Cornwall and west Devon. Proc. Ussher Soc., 2, 464-47 1 . FRESHNEY, E.C. , McKEOWN , M . C . & W ILLIAMS , M. ( 1 972) Geology of the Coast between Tintagel and Bude. Memoirs of the Geological Survey of Great Britain - Explanation of part of One-inch Geological Sheet 322, New Series. Institute of Geological Sciences, HMSO, London, 92 pp.
Bude Formation subsea fans FRESHNEY, E . C . , EDMONDS, E .A . , TAYLOR, R.T. & WILLIAMS, B . J . (1979) Geology of the country around Bude and Bradworthy. Memoir of the Geological Survey of Great Britain for I : 50 000 Geological Sheets 307 and 308. Institute of Geological Sciences, HMSO, London, 62 pp. GHtBAUDO, G . ( 1 992) Subaqueous sediment gravity flow deposits: practical criteria for their field description and classification. Sedimentology, 39, 423-454. GLEN N , C.R. & KELTS, K. ( 1 991) Sedimentary rhythms in lake deposits. In: Cycles and Events in Stratigraphy (Eds Einsele , G . , Rieken, W. & Seilacher, A . ) , pp. 188-22 1 . Springer-Verlag, Berlin. GoLDRING, R. (1993) Ichnofacies and facies interpretation. Palaios, 8, 403-405. GOLDRING , R. & SEtLACHER, A. ( 197 1) Limulid undertracks and their sedimentological implications. Neues 1ahrb. Geol. Palaeontol. Abh., 137, 422-442. HAND, B . M . , MIDDLETON , G .V. & SKIPPER, K. ( 1972) Antidune cross-stratification in a turbidite sequence, Cloridorme Formation, Gaspe, Quebec: Discussion. Sedimentology, 18, 135- 138. HARTLEY, A. ( 1 991) Debris flow and slump deposits from the Upper Carboniferous Bude Formation of SW England: implications for Bude Formation facies models. Proc. Ussher Soc. , 7, 424-426. HARTLEY, A.J. & WARR, L.N. (1990) Upper Carboniferous foreland basin evolution in SW Britain. Proc. Ussher Soc., 7, 2 12-216. HASLAM, H.W. & SCHRIVENER, R.C. ( 1991) A geochemical study of turbidites in the Bude and Crackington forma tions north of Exeter. Proc. Ussher Soc. , 7, 421 -423. HECHT, C.A. (1992) The Variscan evolution of the Culm Basin, south-west England. Proc. Ussher Soc. , 8, 33-38. HtGGS, R . (1983) The possible influence of storms in the deposition of the Bude Formation (Westphalian), north Cornwall and north Devon. Proc. Ussher Soc. , 5 , 477-478. HIGGS, R. (1984) Possible wave-influenced sedimentary structures in the Bude Formation (Lower Westphalian, south-west England), and their environmental impli cations. Proc. Ussher Soc., 6, 88-94. HtGGS, R . ( 1 986a) 'Lake Bude' (early Westphalian , SW England): storm-dominated sedimentation in an equa torial lake. Proc. Usshser Soc. , 6, 416-417. HtGGS, R . (1986b) A facies analysis of the Bude Formation (Lower Westphalian), SW England. PhD thesis, Uni versity of Oxford. HtGGS, R. ( 1987) The fan that never was? - Discussion of 'Upper Carboniferous fine-grained turbiditic sandstones from southwest England: a model for growth in an ancient, delta-fed subsea fan'. 1. sediment. Petrol., 57, 378-379. HIGGS, R. (1988) Fish trails in the Upper Carboniferous of south-west England. Palaeontology, 31, 255-272. HIGGS, R . ( 1991) The Bude Formation (Lower West phalian), SW England: siliciclastic shelf sedimentation in a large equatorial lake. Sedimentology, 38, 445-469. Htscorr, R.N. ( 1981) Deep-sea fan deposits of the Macigno Formation (middle- upper Oligocene) of the Gordana Valley, northern Appenines, Italy: Discussion. 1. sediment. Petrol. , 5 1 , 1015- 1033. HiscoTT, R.N. ( 1994) Traction-carpet stratification in
133
turbidites - fact or fiction? 1. sediment. Res. Sect. A: sediment. Petrol. Process. , A64, 204-208. HOLDER, M .T. & LEVERIDGE, B . E . (1986) Correlation of the Rhenohercynian Variscides. 1. geol. Soc. London, 143, 1 4 1 - 147. HUTCHINSON, D . R . , GOLMSHTOK, A . J . , ZONENSHAIN, L.P. , MooRE, T.C. , ScHoLz, C.A. & KuTGORD, K . D . ( 1 992) Depositional and tectonic framework of the rift basins of Lake Baikal from multichannel seismic data. Geology, 20, 589-592. IsAAc, K . P . , TuRNER, P.J. & STEWART, l.J. ( 1982) The evolution of the Hercynides of central SW England. 1. geol. Soc. London, 139, 521 - 53 1 . KlNG, A.F. ( 1 965) Xiphosurid trails from the Upper Carboniferous Bude Sandstones, north Cornwall. Proc. Geol. Assoc., 1626, 162- 165 . KlNG, A.F. ( 1 966) Structure and stratigraphy o f the Upper Carboniferous Bude Sandstones, north Cornwall. Proc. Ussher Soc., 1 , 229-232. KlNG , A.F. (1967) Structure and stratigraphy of the Upper Carboniferous Bude Formation, north Cornwall. PhD thesis, University of Reading. KlN G , A.F. ( 1971 ) Correlation of the Upper Carboniferous Bude Formation, north Cornwall. Proc. Ussher Soc., 2, 285-288. KuENEN, P.H. ( 1 965) Value of experiments in geology. Geol. Mijnbouw, 44, 22-36. LARUE, D.K. & PROVINE, K.G. ( 1 988) Vacillatory tur bidites, Barbados. Sedimentary Geol., 57, 21 1 -219. LowE, D . R. ( 1982) Sediment gravity flows II: depositional models with special reference to the deposits of high density turbidity currents. 1. sediment. Petrol. , 52, 279-298. MAPEO, R . B . M . & ANDREWS, J.R. ( 1991) Pre-folding tectonic contraction and extension of the Bude Forma tion, North Cornwall. Proc. Ussher Soc. , 7, 350-355. MELVIN, J . ( 1 976) Sedimentological studies in the Upper Palaeozoic Sandstones near Bude Cornwall and Walls, Shetland. PhD. thesis, University of Edinburgh. MELVIN, J. ( 1 986) Upper Carboniferous fine-grained tur biditic sandstones from southwest England: a model for growth in an ancient, delta-fed subsea fan. 1. sediment. Petrol. , 56, 19-34. MELVIN, J . ( 1987) Upper Carboniferous fine-grained turbiditic sandstones from southwest England: a model for growth in an ancient, delta-fed subsea fan: reply. 1. sediment. Petrol. , 57, 380-382. MIDDLETON , G.V. (1965) Antidune cross-bedding in a large flume. 1. sediment. Petrol., 35 , 922-927. MIDDLETON , G.V. & HAMPTON , M . A . (1976) Subaqueous sediment transport and deposition by sediment gravity flows. In: Marine Sediment Transport and Environmental Management (Eds Stanley, D . J . & Swift, D . J . P . ) , p p . 197-218. Wiley, New York. MONACO, P. ( 1992) Hummocky cross-stratified deposits and turbidites in some sequences of the Umbria-Marche area (central Italy) during the Toarcian. Sedimentary Geol. , 77, 132- 142. MoNEY, N.J. ( 1 966) Carboniferous rocks of north Devon: the Appledore Formation. Proc. Ussher Soc. , 1 , 235-238. MuTT!, E. (1977) Distinctive thin-bedded turbidite facies and related depositional environments in the Eocene Hecho Group (South-central Pyrenees, Spain).
134
R. V. Burne
Sedimentology, 24 , 107 - 13 1 . Murr1, E . & RICCI-LUCCHI, F. (1972) L e torbediti dell'Appenino settentrionale: introduzione all'analisi di facies. Mem. Soc. Geol. !tal., 1 1 , 161 - 199. NELSON , C.H. , TwiCHELL, D . C . , SCHWAB, W . C . , LEE, H . J . & KENYON, N. (1992) Upper Pleistocene turbidite sand beds and chaotic silt beds in the channelized, distal, outer-fan lobes of the Mississippi fan . Geology, 20, 693-696. NICHOLS, R .J . , SPARKS, R.S.J. & WILSON , C.J.N. ( 1994) Experimental studies of the fluidization of layered sedi ments and the formation of fluid escape structures. Sedimentology , 41, 233-253. NoRMARK, W.R. & PIPER, D .J.W. ( 1991) Initiation processes and flow evolution of turbidity currents: implications for the depositional record. In: From Shore line to Abyss: Contributions in Marine Geology in Honour of Francis Parker Shepard (Ed. Osborne, R . H . ) , Spec. Pub!. Soc. econ. Paleontol. Mineral. , Tulsa, 46, 209-229. O'CONNELL, S . , NORMARK, W.R. & KENYON, N . H . ( 1991) An entrenched thalweg channel on the Rhone Fan: interpretation from a SEABEAM and SEAMARC 1 survey. In: From Shoreline to Abyss: Contributions in Marine Geology in Honour of Francis Parker Shepard (Ed. Osborne, R . H . ) , Spec. Pub!. Soc. econ. Paleontol. Mineral . , Tulsa, 46, 259-270. OwEN, D . E . (1934) The Carboniferous rocks of the north Cornish coast and their structures. Proc. Geol. Assoc. , 45, 451 -471 . OwEN , D . E . (1950) Carboniferous deposits i n Cornubia. Trans. R. geol. Soc. Cornwall, 18 , 65 - 104. PARKER, G. (1982) Conditions for the ignition of cata strophically erosive turbidity currents. Mar. Geol. , 46, 307-327. PIPER, D.J.W. & STOw, D.A.V. (1991) Fine-grained tur bidites. In: Cycles and Events in Stratigraphy (Eds Einsele, G . , Rieken, W. & Seilacher, A . ) , pp. 360-376. Springer-Verlag, Berlin. PoSTMA, G. (1986) Classification for sediment gravity-flow deposits based on flow conditions during sedimentation. Geology, 14, 291-294. POSTMA, G. , NEMEC, W. & KLEINSPEHN , K.L. (1988) Large floating clasts in turbidites: a mechanism for their emplacement. Sediment. Geol. , 58, 47-61. PRAVE, A.R. & Du KE , W.L. (1990) Small-scale hummocky cross-stratification in turbidites: a form of antidune stratification . Sedimentology, 37, 531-539. PRENTICE, J.E. (1960a) Dinantian, Namurian and Westphalian rocks of the district south-west of Barnstaple, north Devon. Q. J. geol. Soc. London, 1 15 , 261-289. PRENTICE, J .E. (1960b) The stratigraphy of the Upper Carboniferous rocks of the Bideford region, north Devon. Q. J. geol. Soc. London, 116, 397-408. PRENTICE, J . E . (1962) The sedimentary history of the Carboniferous in Devon. In: Some Aspects of the Variscan Fold Belt (Ed. Coe, K.), pp. 93- 108. Manchester University Press, Manchester. READING , H . G . (1963) A sedimentological comparison of the Bude Sandstones with the Northam and Abbotsham Beds of Westward Ho! Proc. Ussher Soc. , 1, 68-70. READING , H . G . (1965) Recent finds in the Upper Carbon iferous of south-west England and their significance.
Nature, 208, 745 -747 . H . G . (1971) Sedimentation sequences in the Upper Carboniferous of Northwest Europe. Comptes rendues 6e Congress international de Stratigraphie et Geologie Carbonifere, Sheffield, September 11 -14, 1967, Vol. 4 , pp. 140 1 - 1412. READING , H.G. (1987) Fashions and models in sedi mentology: a personal perspective. Sedimentology, 34, 3-9. REINECK, H .-E. & WUNDERLICH, F. ( 1968) Zur unter scheidung von asymmetrischen Oszillationsrippeln und Striimungsrippeln. Senckenbergina Lethaea, 49, 321 -· 354. Ricci-LUCCHJ, F. (1975) Depositional cycles in two turbidite formations of northern Appenines (Italy) . J. sediment. Petrol. , 45, 3-43 . RoBERTS , H . H . & THAYER, P . A . (1985) Petrology of Mississippi Fan depositional environments. In: Sub marine Fans and Related Turbidite Systems (Eds Bouma, A . H . , Normark, W.R. & Barnes, N . E . ) , pp. 331 -339 . Springer-Verlag, Berlin. SAGEMAN , B . B . , WIGNALL, P . B . & KAUFFMAN , E . G . (1991) Biofacies models for oxygen-deficient facies in epi continental seas: a tool for paleoenvironmental analysis . In: Cycles and Events in Stratigraphy (Eds Einsele, G . , Rieken, W. & Seilacher, A . ) , pp. 542-564. Springer Verlag, Berlin. SANDERS, J . E . (1965) Primary sedimentary structures formed by turbidity currents and related resedimentation mechanisms. In: Primary Sedimentary Structures and their Hydrodynamic Interpretation (Ed. Middleton, G.V.) Spec. Pub!. Soc. econ. Paleontol. Mineral. , Tulsa, 12, 192-219. SAVRDA, C.E. & BoiTJER, D . J . (1991) Oxygen-related bio facies in marine strata: an overview and update. In : Modern and Ancient Continental Shelf Anoxia (Eds Tyson, R.V. & Pearson, T . H . ) , Spec. Pub!. geol. Soc. London, 58, 201 -219. Geological Society of London, Bath. ScHOLZ, C.A. & RoSENDAHL, B .R . (1990) Coarse-clastic facies and stratigraphic sequence models from Lakes Malawi and Tanganyika, East Africa. In: Lacustrine Basin Exploration: Case Studies and Modern Analogs (Ed. Katz, B . J . ) , Mem. Am. Assoc. petrol. Geol. , Tulsa, 50, 151- 168. Scorr, D . L . , NG'ANG'A, P . , JoHNSON, T.C. & RosENDAHL, B . R . (1991) High-resolution acoustic character of Lake Malawi (Nyasa) , East Africa and its relationship to sedimentary processes. In: Lacustrine Facies Analysis (Eds Anad6n, P . , Cabrera, Ll. & Kelts, K.), Spec. Publ. int. Assoc. Sediment. , No. 13, 129-145. Blackwell Scientific Publications, Oxford. SEDGEWICK, A. & MURCHISON , R.I. ( 1840) On the physical structure of Devonshire, and on the subdivisions and geological relations of its older stratified deposits. Trans. geol. Soc. London, 5, 633-705. SEILACHER, A. ( 1982) Distinctive features of sandy tempestites. In: Cyclic and Event Stratification (Eds Einsele, G. & Seilacher, A . ) , pp. 333-349. Springer Verlag, Berlin. SHANMUGAM, G. & MoiOLA, R.J. ( 1988) Submarine fans: characteristics, models, classification , and reservoir potential. Earth Sci. Rev. , 24, 383-428. SKIPPER, K. (1971) Antidune cross-stratification in a turREADING ,
·
Bude Formation subsea fans bidite sequence, Cloridorme Formation, Gaspe, Quebec. Sedimentology, 17, 5 1 -68. SKIPPER, K. & MIDDLETON , G.V. ( 1975) The sedimentary structures and depositional mechanics of certain Ordovician turbidites, Cloridorme Formation, Gaspe Peninsula, Quebec. Can. J. Earth Sci. , 12, 1934- 1952. STAUFFER, P.H. ( 1967) Grain-flow deposits and their implications, Santa Ynez Mountains, California. J. sediment. Petrol. , 37, 487-508. STOw, D . & SHANMUGAM, G. ( 1 980) Sequence of structures in fine-grained turbidites: comparison of recent deep sea and ancient flysch sediments. Sediment. Geol. , 25, 23-42. S uNDBORG, A. ( 1956) The river Klaralven, a study of fluvial processes. Geogr. Annaler, 38, 127-316. TANNER, P.W.G. ( 1 992) The duplex model: implications from a study of flexural-slip duplexes. In: Thrust Tectonics (Ed. McClay, K . R . ) , pp. 201 - 208. Chapman & Hall, London. THOMAS, J . M . (1988) Basin history of the Culm Trough of southwest England. In: Sedimentation in a synorogenic basin complex; the Upper Carboniferous of Northwest Europe (Eds Besly, B .M. & Kelling, G . ) , pp. 24-37. Blackie, Glasgow. UNRUG, R. ( 1963) Istebna Beds - a fluxoturbidity forma tion in the Carpathian Flysch. Ann. Soc. geol. Pol. , 33, 49-92. UNRUG, R. (1965) Le facies de flysch a fluxoturbidites dans les Carpathes Polonaises. Carpatho-Balkan Geological Association, VII Congress, Sofia, September, Reports, Part II, pp. 301 -303. UsSHER, W.A.E. ( 1 892) The British Culm Measures. Proc.
135
Somerset archaeol. nat. Hist. Soc. , 38, 1 1 1 -2 19. R.G. ( 1963) Distinctive types of ripple-drift cross lamination. Sedimentology, 2, 173- 188. WALKER, R . G . ( 1 964a) Some aspects of the sedimentology of the Shale Grit and Grindslow Shales (Namurian R 1c , Derbyshire) and the Westward Ho ! and Northam Formations (Westphalian, north Devon) . PhD thesis, University of Oxford . WALKER, R.G. ( 1964b) Some aspects of the sedimentology of the Westward Ho! and Northam Formations. Proc. Ussher Soc. , 1 , 85-87. WALKER, R.G. ( 1967) Upper flow regime bed forms in tur bidites of the Hatch Formation, Devonian of New York State. J. sediment. Petrol. , 4, 1052- 1058. WETZEL, A. ( 1991) Stratification in black shales: depo sitional models and timing - an overview. In: Cycles and Events in Stratigraphy (Eds Einsele, G . , Rieken, W. & Seilacher, A . ) , pp. 508-523. Springer-Verlag, Berlin. WHALLEY, J . S . & LLOYD, G . E . ( 1986) Tectonics of the Bude Formation , north Cornwall - the recognition of northerly directed decollement. J. Geol. Soc. London, 143, 83-88. WooD, A. & SMITH , A.J. ( 1959) The sedimentation and sedimentary history of the Aberystwyth Grits (Upper Llandoverian) . Q. J. Geol. Soc. London, 114, 163- 195 . YAGISHITA, K. (1994) Antidunes and traction-carpet deposits in deep-water channel sandstones, Cretaceous, British Columbia, Canada. J. sediment. Res. Sect. A: sediment. Petrol. Process. , A64, 34-4 1 . ZIEGLER, P.A. ( 1990) Geological Atlas of Western and Central Europe. Shell Internationale Petroleum Maatschappij B . V . and the Geological Society, London. WALKER,
Spec.
Pubis
int.
Ass.
Sediment. (1995)
22,
137-156
Depositional controls on iron formation associations in Canada P HI LI P F R A L I C K* and T I M O T HY J. B A R RE T Tt *Department of Geology, Lakehead University, Thunder Bay, Ontario P7B 5El, Canada; and tMineral Deposit Research Unit, Department of Geological Sciences, University of British Columbia, Vancouver, British Columbia V6T 1Z4, Canada
ABSTRACT
Understanding Precambrian iron formation depositional processes has been hindered by the lack of precise modern analogues. However, by combining a regional basin analysis of sedimentary and volcanic rocks surrounding an iron formation with detailed examination of sedimentary structures and lithic associations within an iron formation, the depositional setting and physical processes of sedimen tation can be inferred. Six iron formations present in the Canadian Shield were examined using this approach. The Palaeoproterozoic Gunflint Formation consists of strand-proximal stromatolites and oolitic shoals, transitional distally to grainstones and parallel-laminated chemical muds. This succession was laid down on a wave- and tide-dominated inner shelf. In the northern Labrador Trough, the Middle Member of the Palaeoproterozoic Baby Formation represents an outer shelf to slope environment. Here massive, graded and ripple-laminated siltstone-shale couplets dominate the succession, with iron-rich chemical sediment forming fine-grained tops to some couplets; thick assemblages of parallel laminated chemical sediments also occur. In the Beardmore-Geraldton area, a submarine environment is represented by oxide facies Archaean iron formations. The chemical units accumulated where clastic mud would normally be found in coarsening and thickening upward levee and ramp assemblages. In the Terrace Bay area, Archaean abyssal plain deposits consist of black, graphitic slate with pyrite layers and zones of chert; these sediments lie on volcanics and are overlain by a submarine-ramp clastic assemblage. Archaean volcanic-associated iron formation is represented by thin interflow sediment packages in a submarine lava-plain south of Beardmore, and as a sulphide lens in a volcanic edifice south of Schreiber. In the first five examples iron formation occupies the niche usually dominated by clastic mud. The iron formation is able to form because of a reduced clastic supply reflecting some combination of the following factors: a peneplained source region; a relative rise in sea-level creating sediment storage capacity in subaerial and shallow-water regions; development of sediment bypass systems; and cessation of active volcanism near the depositional area. The facies of iron formation in a given area is dependent on factors such as water depth (both Eh and energy), degree of basin isolation from clastic sedimentation, and amount of hydrothermal input or upwelling. Proximal to submarine hydrothermal vents, iron formation accumulation can dominate without a reduction in clastic input; the mineralogy and layering are controlled directly by the temperature, Eh and drift direction of hydrothermal discharge.
INTRODUCTION
Sediments accumulating in modern environments commonly provide a basis for comparison with ancient successions, in order to reconstruct the palaeogeography and depositional processes that operated in a region. This approach is probably the most useful technique for reconstructing ancient
depositional systems, but in the Precambrian the present is sometimes not the key to the past. In the Archaean and Palaeoproterozoic, grossly different atmospheric and water chemistry, and possibly tem perature, led to the deposition of chemical sediments that have inadequate modern analogues . Iron
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0 137
138
P. Fralick and T.J. Barrett
formations, which are chemical sedimentary rocks containing greater than 15% Fe (James, 1966), are volumetrically the most important member of this group. Modern subaqueous hot-spring activity provides clues to understanding the precipitation of Fe-rich sediments, but the mineralogy and areal extent of modern deposits is not comparable to many iron formations deposited during the Earth's early history. Theoretical chemical modelling has, to date, been unable to explain the interlayering of iron-rich sediment and chert. A broader approach is necessary when conaucting field studies of iron formation. For example, sedi mentary and volcanic rocks surrounding iron for mations provide data on depositional settings; sedimentary structures and lithic associations within iron formations give information on physical process of deposition; and regional basin analysis provides a framework within which to view iron formation genesis. Combining these techniques makes palaeo geographical reconstruction and depositional model ling feasible. Iron formations are often classified into two main groups, those of Superior type, which were deposited as thick, laterally extensive units in tectonically stable areas, with sedimentary structures and textures indicative of shallow-water environments, and those of Algoma type, which accumulated in tectonically active regions as well-layered but discontinuous units, mainly associated with volcanic rocks or deep water sediments (Gross, 1965, 1983). This rather arbitrary subdivision is similar to the geosynclinal theory, initially useful, but not suited to our present data base. It limits iron formation to two types, when in reality there are a multitude of types. Iron formation characteristics are controlled by depo sitional settings and processes that are complex multivariate systems and have little likelihood of being repeated exactly in time or space. Thus, iron formations may be similar, but will not be identical, and each is capable of providing further information on how this unusual rock type forms. Classification systems are of course useful in our nomenclature, but they must not become an end in themselves. A fuller understanding of process and depositional setting should be the primary goal of research on iron formations. This requires more detailed work on individual units. Factors involved in the deposition of iron for mation include the source of iron and silica, transport mechanisms, and controls on precipitation. Possible sources for the iron include weathering of iron-
bearing minerals, in either subaerial (Garrels et at., 1973; Drever, 1974) or subaqueous (van Hise & Leith, 19 1 1; Huber, 1959) environments, and hydro thermal ion exchange ( Goodwin, 1956; Gross, 1965; Fralick, 1987). Simonson ( 1985) has shown that lithological data from the Animikie Group and Labrador Trough do not support a cratonic source for the iron. Chemical data on trace elements, rare earths and stable isotopes do not differentiate between low-temperature weathering of mafic vol canic rocks and high-temperature hydrothermal alteration as possible sources for the iron. Iron released by both processes will be in solution in any ocean, and although hydrothermal input is obviously dominant in some situations, low-temperature leaching may have, at times, played an important role. In order for small quantities of Fe2+ to go into, and stay in solution, p02 must have been very low (Holland, 1973). A hydrothermal source has been proposed for the silica (Gross, 1965; Gross & Zajac, 1983; Fralick, 1987), although Cloud ( 1973) believed the silica in Palaeoproterozoic iron formations may have orig inated from subaerial weathering. In either case, the probable lack of silica-secreting organisms at this time would have resulted in an ocean saturated with H4Si04 (Siever, 1957; Cloud, 1973; Holland, 1973). The movement of large amounts of water into an environment with a different Eh, pH, or temperature is needed to precipitate iron and silica in the quan tities necessary to form an iron formation. This can be achieved through the hydrothermal venting of hot, reduced, acidic solutions (Fralick, 1987; Barrett et a!. , 1988), or upwelling of deep ocean waters with a low p02 on to a more oxygenated shelf (Cloud, 1973, 1983; Holland, 1973; Drever, 1974). However, these scenarios do not explain deep marine Archaean iron formation deposited distally to hydrothermal activity. In this situation, and with the oceanic bottom waters saturated with iron and silica, hydro thermal venting may produce a zone of oversatu ration and precipitation which extends away from the vent area for tens to possibly hundreds of kilo metres (Fralick, 1987). In the Palaeoprotero zoic, and possibly the Archaean, microorganisms influenced precipitation of iron and silica as they added 02 to the environment (Cloud, 1973), and possibly directly precipitated iron and silica as coatings and tests (LaBerge et al., 1987), although the latter process has been questioned by Oehler ( 1976). Alternatively, near-surface photo-oxidation of Fe2+ could have formed the iron-rich layers
Depositional controls on iron formation
(Cairns-Smith, 1978; Braterman & Cairns-Smith, 1987), and pressure reduction during upwelling could have caused silica to precipitate ( Holland, 1973). Studies of specific iron formation basins should attempt to integrate a number of the processes described above (cf. Morris, 1993), In this paper we provide six examples from the Canadian Shield which highlight the control that depositional setting had on iron formation character istics. The depositional environments described con stitute a succession from shallow shelf through slope, rise and abyssal plain to areas of active volcanism. It is apparent from these examples that iron formation type is, like clastic deposits, controlled by depo sitional setting. Of course we are not the first to emphasize this fact. Gross ( 1980), and many subsequent researchers (e.g. Ojakangas, 1983; Simonson, 1985), have stated or implied similar ideas. What we strive to emphasize here is the benefit of conducting iron formation research in a manner similar to studies of clastic depositional systems.
CASE STUDIES
Objectives
The rationale for providing this series of case studies is to emphasize the control exerted by depositional processes on iron formation attributes. The physical features that iron formations exhibit are largely the result of their depositional environment. The headings in this section are not meant to be a classification system, but are purely descriptive. We believe it is more productive to look upon each iron formation as a distinct entity reflecting the processes operative in the setting in which it formed. Shallow shelf: Gunflint Formation
Iron formation, associated with other types of chemical and clastic sediment, deposited in shallow marine settings is most common in the Palaeo proterozoic, although well-documented examples from the Archaean also exist. The Canadian Shield contains two extensive regions dominated by iron formation associated with shallow marine sediments: the Animikie Group and portions of the Labrador Trough. Iron formation within the Canadian portion of the Animikie Group is contained in the 2000 Ma Gunflint
139
Formation (Fig. 1). This sedimentary package was deposited on a peneplained Archaean surface forming the southwestern margin of Superior Province (the edge of the North American craton at that time). Minor mafic flows and extensive tuf faceous horizons within the formation indicate that the region was volcanically active at the time of basin formation and subsidence. The Gunflint For mation is underlain locally by a basal conglomerate; the Gunflint has a gradational upper contact with black shales and siltstones of the Rove Formation. The Gunflint Formation is divisible into two mem bers and averages 120 m in thickness. The lower member of the Gunflint Formation commonly has basal algal and oolitic cherts (Fig. 2). The microfossil-bearing algal structures often occur on boulders of the basal conglomerate, which pro vided a stable substrate. The remainder of the lower member commonly consists of grainstone beds sep arated by fine-grained sediment layers (Fig. 3a). Locally, this facies is termed cherty iron formation. Individual fine-grained layers are homogeneous and iron-rich, but mineralogy varies from layer to layer. Iron-oxide-chert and iron-silicate-chert mixtures are the most abundant. Sand-sized, chemical mud intraclasts forming the grainstones are commonly similar in mineralogy to interbedded fine-grained layers. Larger rip-ups of material similar to the underlying substrate are present in many beds (Fig. 3c). Internal laminations are often difficult to discern in the grainstone beds, due to the effects of dia genesis, but trough cross-stratification (Fig. 3d) is not uncommon and some sections exhibit hummocky cross-stratification (HCS). Small outcrop size and normal faulting in the area, make analysis of up-section trends within the Gunflint Formation difficult, and limited amounts of drill-core must be used. The section depicted in Fig. 2 exhibits an upward decrease, then increase in the thickness of grainstone beds through the lower member of the formation. Interstratified with the grainstones in some sections are units metres to tens of metres thick consisting of laminated, fine-grained chemical sediment (Fig. 3b), locally called slaty iron for mation, and similar in all aspects to the fine-grained material in the grainstone assemblages. The lower member is capped sporadically by breccia. In Fig. 2 this breccia is underlain by a massive layer of silt sized material. The upper member of the Gunflint Formation is similar in stratigraphy to the lower member. Basal algal cherts (Fig. 3e), often developed directly on
140
P. Fralick and T.J. Barrett c Ungava Bay
Lake Superior
�
Proterozoic
0
100km
� Archean intrusive
L...±_j metamorphic terrains
rv:v.l Archean volcanic � sedimentary terrains
top of fragments of the breccia, are succeeded by a sequence of thinning upward grainstone beds separ ated by chemical sediment. Regionally extensive tuffaceous shales, with accretionary lapilli present at some localities, are interbedded with the grainstone succession. A limestone unit caps the upper member and separates the Gunflint Formation from the shales of the basal Rove Formation. Lougheed ( 1983) attributed the depositional environments represented by the Animikie Group iron formations to supratidal, intertidal and subtidal zones of a broad shelf. The importance of subtidal and possibly lower intertidal sand shoals as depo sitional sites for the grainstone was emphasized by Simonson ( 1985). Bidirectional palaeocurrent indi cators at some sites indicate that tidal currents were responsible for the accumulation of some grain stone units (Ojakangas, 1983). Hummocky cross stratification present in other sequences indicates the importance of storm activity in creating and moving sand-sized intraclasts seaward (Fralick,
Fig.
I. (A) Location of the six iron formations discussed. (B) Regional geology of the area north of Lake Superior. (C) Regional geology of the northern Labrador Trough.
1988). Stromatolites and accretionary grains com-· monty appear to be limited to near-strand positions (Morey et at., 199 1). This spatial restriction is not due to browsing metazoans, as in the case of Phanerozoic stromatolites, but probably due to frequent current activity causing large areas of the offshore granular substrate to be unstable. Ojakangas ( 1983) has shown that water depth was the major control on the development of 'cherty' (grainstone dominated) versus 'slaty' (chemical mud dominated) iron formation. The grainstones were deposited at shallower depths by tide-and-wave/ storm driven currents (Fig. 4). The chemical muds accumulated during fair-weather times in shallow-· water regions, and at all times in the deeper offshore. Alternating successions of grainstone-dominated and mud-dominated iron formation packages suggest fluctuations in relative sea-level (White, 1954; Morey, 1983) and can be used to create sea-level curves. Upward-thinning and upward-thickening successions of grainstone beds also suggest, respect-
Depositional controls on iron formation
141
I
LEGEND
? Inter layered Grainstone
.... Q) .0
and
E
Q)
�
.... Q) c.. c.. :::J
J .... Q) .0
E
Q)
�
.... Q)
3:
0 _J
Fig. 2.
Stratigraphical section of the Gunflint Formation, Animikie Group. The upper 20% of the formation was not present in this drill hole. For detailed location see Kronberg & Fralick (1992).
.�
<1>3: -' .c: 1--
50 40 l30
·r
Mudstone
4
ively, relative rise and fall of sea-level (Fig. 2). The bed thickness trends shown in Fig. 2 suggest that sea-level fluctuations controlled the cyclic nature of the lower and upper members. The positioning of the breccia and possible vadose silt unit at the lowest sea-level stand, inferred from bed thickness, suggests that these units formed through subaerial dissolution of carbonates. Increasing volcanic activity in the upper Gunflint Formation may have resulted in a
Stromatolites Quartz Barite Vein Archean Granitic Basement Contorted
1\
20
10 0
Chert
\
Bedding
�
v I �
0
"'3:
( Volcaniclastir.)
lJ
"E
�g
Mudstone
'
+ +
+
+ +
+ .S? "0 � +
+
o�� (j) .c Ul (j)
.c 0
+ +
� c 0 <1>0-
+
Clasts
Cross-
IJ
+
Rip- up
Stratification
j +
Breccia
�
"'
c.
�
Hummocky CrossStratification
"' "' 0
>.-., c == c -UJ o u (j)
shift from chemical sedimentation to a volcaniclastic dominated system. The source of the iron- and silica-rich precipitates that form the majority of the Gunflint Formation is still controversial. Simonson ( 1985) favoured a direct hydrothermal source in the offshore with coprecipi tation of silica and iron-rich phases in the nearshore area due to ambient high levels of silica and iron in the ocean. Dissolved hydrothermal iron and
142
P. Fralick and T.J. Barrett
Photographs of inner-shelf-associated iron formation in the Animikie Group. (a) Grainstone (intraclast)-dominated succession. The grainstone beds are arranged in lenses which may cross-cut one another. Fine-grained chemical muds (m) are interlayered with sand lenses. (b) Slaty iron formation. This chemical mud-dominated succession also contains grainstone lenses but they are less abundant and smaller than in the cherty iron formation. (c) Thick-bedded cherty iron formation with abundant rip-up clasts (r). (d) Inclined rip-up clasts in a grainstone lens lying on the backset slope of a dune. (e) Rare calcium carbonate stromatolites; most other mounds are silicified. Lens cap is 5.5 em in diameter.
Fig. 3.
Depositional controls on iron formation
143
Grainstone (reworked chemical sediment)
Fine-grained chemical sediment
Block diagram of the depositional environments in which the Gunflint chemical sediments accumulated. Deeper areas not affected by currents induced by storms and tides are dominated by laminae of silica, iron silicate, iron carbonate and iron oxide. Shallower areas affected by sporadic current activity consist of interlayered grainstone and beds of fine grained chemical sediment similar to that accumulating in deeper areas. Grainstone production occurred in the shallows due to erosion and abrasion of chemical muds, with the material transported offshore during times of increased current activity. Ooid shoals developed in strand-proximal zones of maximum turbulence, and stromatolites formed along the strand where the substrate had been diagenetically hardened. Karstification occurred in subaerially exposed areas.
Fig_ 4_
silica may have been carried up to shelf depths by upwelling along cratonic margins.
Outer shelf and slope: Middle Member, Baby Formation
Examples of iron formations deposited in outer shelf and slope settings are rare. Similarly, descriptions of shallow-water and deep-sea clastic units are common but outer shelf and slope secessions are not well represented. This may be due to their more restricted distribution and lower preservation potential. The Palaeoproterozoic Baby Formation of the northern Labrador Trough (Fig. 1) provides an example of iron formation deposition in an outer shelf to slope setting. The trough separates the Superior Province from the Rae Province and under went a major orogenic episode during closure between these two land masses. Sediments of the Baby Formation were deposited prior to the major phase of orogenic activity, probably during an earlier extensional phase (Wares & Goutier, 1990; Shulski et al., 1993). They are underlain by dolomite and overlain by tholeiitic basalt. The kilometre-thick Baby Formation is dominated by a monotonous, thin-bedded succession of fine grained clastic sediment. The Middle Member
departs from this trend with the upward appearance and then domination of the unit by iron formation (Fig. 5). This interval may be correlated with iron formation present in the Knob Group to the south west of the Baby Formation. Clastic layers within the Baby Formation are commonly 0.5- 10 mm thick and non-graded with sharp lower and upper contacts (Fig. 6a). They are composed of siltstone, silty shale and occasionally very fine-grained sandstone. Silty shale layers often contain coarser silt streaks only a few grains thick. Graded units similar in other respects to the non graded layers also are common. The grading may consist of either an upward grain-size decrease with no internal laminations, or alternating siltstone and silty shale laminae that thin and fine upwards. Current-ripple-laminated layers of fine sandstone are interbedded with the fine-grained clastic suc cession. Beds are 1-4 em thick, wavy bedded and non-graded. Current ripples in the thicker units are larger and are commonly stacked. The undulating upper surfaces of the rippled units are sharply over lain by silty shale, which fills troughs and covers crests. Thicker, compound fine-grained sandstone beds are also present in the succession. These are up to 20 cm thick and are composed of vertically stacked, 1-5-cm-thick, massive to parallel lami nated, non-graded sandstone. The layers are some-
P. Fralick and T.J. Barrett
144 M
60
Fe-C, Fe-0, Fe-S
50
ss
40 SS, Fe-C
30
f,
Fe-S
Fe-C
20
_
Overburden
ss
10
Fe-0, Fe-S, Si SS, Fe-C SS, Fe-C, Si SS, Fe-P
0
SS, Fe-0, Fe-S, Si
Stratigraphical section of a portion of the Middle Member, Baby Formation. All units are internally layered on the millimetre- to centimetre-scale. Chemical sediment occurs in the shale-rich tops of some clastic beds. SS, siltstone and shale; Si , chert; Fe-S, iron silicate; Fe-C, iron carbonate; Fe-0, iron oxide; Fe-P, iron sulphide.
Fig. 5.
times separated by millimetre-scale shale drapes. Rare, thicker beds in the Middle Member consist of two types. (i) Massive, metre-scale medium-grained sandstone beds with coarse sand grains scattered throughout. Reverse grading may be developed near their tops. (ii) Organized beds, with a thin, basal, current-rippled, fine-grained sandstone overlain by millimetre-thick, alternating coarse siltstone and silty shale laminae that thin and fine upwards. The upper halves of the units are composed of normally graded silty shale. Iron formation occurs interbedded with clastics and as dominantly chemical successions within the
Middle Member of the Baby Formation (Figs 5 & 6). Chemical sediment may form discrete millimetre scale layers either between graded silt-shale couplets, or interlaminated with the shaley top of the couplet. Iron-rich minerals may be iron silicates, oxides, carbonates or sulphides. One succession of rhythmites containing thicker (average thickness= 4 em) carbonate grainstone layers is also present. This succession is separated from a chemical sedi ment assemblage by a massive dolomite, 1.5 m thick. Carbonate layers in the grainstone succession contain small- to medium-scale trough cross-stratification. Chemical-dominated successions are composed of thinly to thickly laminated iron-rich, fine-grained, sediment locally interstratified with chert or jasper layers (or more rarely lenses). Carbonate, silicate, oxide and sulphide facies iron formation are all present within the Baby Formation. Interlayering of the facies is common, although all facies rarely occur in the same interval. The iron-rich layers are centimetre- to decimetre-scale in thickness, with internal millimetre- and submillimetre-scale lami nations caused by differences in crystal size (Fig. 6c). Microscopic examination reveals that the laminae reflect changes in the amount of clastic material and chert present. The cherts interbedded with the iron-rich units are commonly 1-3 em thick and exhibit no internal laminations except for rare submillimetre-scale iron silicate layers. The Baby Formation forms part of an assemblage that has been interpreted as a passive margin suc cession, thickening towards the east (Dimroth, 1981; LeGallais & Lavoie, 1982; Wardle & Bailey, 1981). The abundant fine-grained sediments, the lack of wave or tidal deposits and the rare presence of carbonate grainstones transported by bottom cur rents suggest that deposition took place on the outer shelf or upper slope (Fig. 7). The rhythmites were deposited from low-density turbidity currents similar to those attributed to comparable units in the off shore of northeastern North America (Chough & Hesse, 1980; Hesse & Chough, 1980; Stow & Shanmugam, 1980). These sediment clouds may have been raised by sporadic slumping on the slope, or as overflows from channelized bypass systems. The latter mechanism is preferred because slump scars are rare. Thin current-rippled units represent intermittent bottom currents. Rippled units of this type are rarely described from slope deposits. The thick sandstone units provide better evidence for a slope. Their massive nature combined with reverse grading indicates that downslope grainflow processes
Depositional controls on iron formation
145
Photographs of the iron formation associated with outer-shelf-slope facies in the Baby Formation. (a) Close-up of thin-bedded sediments showing lenses and continuous layers of silt (light-coloured layers), interbedded with shales. (b) Magnetite laminae assemblages (m), interlayered with lighter siltstone and shale laminae assemblages (s). (c) Laminated magnetite. Grain-size differences produce the layering. (d) Interbedded siltstone (light, s) and magnetite (dark, m); grey units (g) represent mixtures of magnetite and clastics. (e) Reflected-light photomicrograph of pyrite bands (p) interlayered with clastic material (s). Scale bar is 0.5 mm.
Fig. 6.
146
P. Fralick and T.J. Barrett Grain-flow
Channelized turbidity flows
Block diagram of the depositional environments in which the Baby Formation chemical sediments accumulated. Thinly laminated iron-rich precipitates and chert form discrete units or are interlayered with Bouma d-e turbidites in an outer shelf and slope setting. Clastic material was delivered to this environment through channel overflow, slump generated low-density turbidity currents, and grainflows originating at times when the shelf was dominated by quartz sand. Rare iron carbonate sands were brought into this environment when the shelf was starved of clastic material. Fig. 7.
Rainout of chemical sediment
were at least in part responsible for their emplace ment. The trough cross-stratified carbonate grain stone was obviously moved into the area by fairly strong bottom currents originating on the inner shelf, where grain production would have taken place. This places the Middle Member of the Baby For mation on the outer shelf to slope break, past the mud line (Stanley & Wear, 1978) but close enough to areas affected by tide-, or storm-produced currents to receive shelf sediments during unusually high velocity flow events. The presence of iron formation in the Middle Member requires the clastic input to be greatly diminished or chemical precipitation rates to be greatly accelerated. Not enough data are present to choose between these two alternatives. Clastic supply may be controlled by: (i) the rate at which clastic materials are supplied to the shelf; and (ii) the storage capacity of the shelf combined with the efficiency of sediment bypass systems in the outer shelf to slope area. The rate of chemical precipitation was probably controlled by upwelling rates from the deep ocean. As these variables fluctuated, the system oscillated between clastic and chemical dominance. The Baby Formation has a great variety of iron rich mineral phases, which probably reflects the physiographic setting on the outer shelf. There, upwelling currents would first encounter shallower waters. This may have led to large changes in Eh, pH and dissolved concentrations through time,
resulting in the varied mineralogy of the iron for mation. The dominance of rainout processes and the variability of sediment influx and water chemistry produced an iron formation that is both thinly laminated and laterally persistent. Submarine rise: Beardmore-Geraldton Clastic Associated Iron Formation
Archaean iron-formation-bearing successions that were probably deposited on submarine rises are present in the Rainy Lake and Spirit Lake areas (Wood, 1980), Manitou Straits district (Teal & Walker, 1977), Lake St Joseph region (Meyn & Palonen, 1980), the Abitibi greenstone belt (Hyde, 1980), and the Beardmore-Geraldton terrain (Barrett & Fralick, 1985, 1989). The latter will be used as the case example (Fig. 1B). The Beardmore-Geraldton terrane consists of three metasedimentary belts, each resting on a thick volcanic assemblage. A metavolcanic terrain lies to the north and the gneisses and granites of the Quetico Subprovince lie to the south. The northern meta sedimentary belt is composed of fluvial conglom erates and sandstones (Devaney, 1987), the central belt consists of a prograding shoreline to offshore turbidites (Devaney & Fralick, 1985), and the southern metasedimentary belt contains turbidites and iron formation (Barrett & Fralick, 1985, 1989).
Depositional controls on iron formation
The belts are kilometres wide and tens of kilometres in length, with subvertical bedding younging to the north. The assemblage represents a fore-arc basin that has been stacked tectonically (Barrett & Fralick, 1989; Devaney & Williams, 1989). Oxide facies iron formation occurs interbedded with clastics in the turbiditic portion of the assemblage (Fig. 8). The turbidites are grouped into thick-, medium and thin-bedded turbidite-dominated associations, and a thin-bedded iron-formation-clastic-sediment association (Barrett & Fralick, 1989). The bulk of the succession is composed of the first three associ ations, which forms a submarine ramp, with attri butes suggesting fan development at some locations. The thin bedded, iron-formation-clastic-sediment association forms packages metres to tens of metres thick, interstratified with the ramp/fan turbidites, and is itself divisible into four iron formation lithofacies associations (IFLA) (Figs 8 & 9). Their characteristics are as follows. IFLA a: domi nantly magnetite-rich sediment with millimetre- to centimetre-scale, graded or ungraded silt interbeds (Fig. 9a, b, d, f & g). IFLA b: centimetre-scale, graded to sharply bounded silt beds, either con-
tiguous or separated by millimetre-thick laminations of magnetite-rich sediment (Fig. 9a, b, c & e). IFLA c: sand-rich composite units up to about 1 m thick, generally consisting of thin, stacked, ungraded, laminated sand beds. These units consist of medium to coarse-grained sand. The composite units are separated by intervals of magnetite or magnetite and siltstone up to 15 em thick (Fig. 9a & b). IFLA d: framework-supported polymictic conglomerate beds up to a few metres thick, interbedded with sandstone and minor iron formation, or fairly thick iron for mation and thin-bedded sands (Barrett & Fralick, 1985, 1989). The first three lithofacies associations are commonly, although not exclusively, organized into coarsening upward successions 1-50 m thich (Fig. 8) (Fralick, 1987). Sandstone assemblages are sharply overlain by IFLA a, which is gradational upwards through IFLA b to IFLA c. This package is then overlain by ramp/fan turbidites (Barrett & Fralick, 1989). The ramp-fan system represented by the Beardmore-Geraldton turbidites provided few locations where clastic influx was low enough to allow iron formation accumulation (Fig. 10). To the
B
A M
M
50
6
40
147
•
IFLAa interlaminated magnetite, siltstone, and slate
II
IFLAa interlaminated magnetite, jasper, siltstone, and slate
4
D lliJ 0
30
20
Stratigraphical sections of iron-formation-bearing units in the Beardmore-Geraldton area. (A) Outcrop section measured at the Leitch Mine (detailed location in Barrett and Fralick, 1985). (B) Outcrop section measured at Solomon's Pillars (detailed location in Fralick, 1987).
Fig. 8.
10
0
2
0
� -
Sandstone IF LAc interbedded sandstone, siltstone, slate and magnetite IFLAb interbedded siltstone, slate and magnetite
148
P. Fralick and T.J. Barrett
Photographs of iron formation associated with an inner submarine ramp-fan complex in the Beardmore Geraldton region. (a) Magnetite layers (dark) interstratified with Bouma d-e turbidites. Clastic and chemical dominated intervals are visible (IFLA a=a; IFLA b=b; IFLA c=c). (b) Interlayered magnetite and d-e turbidites forming two thin coarsening and thickening upward successions (arrows). (c and d) Alternating magnetite-rich (dark) and clastic-rich (light) laminations (prints of thin-sections). Scale bar is 5 mm. (e) Photomicrograph showing gradation from clastic-dominated bottom to magnetite-dominated top of two thin d-e turbidites. Scale bar is 1 mm. (f) Photomicrograph of interlaminated magnetite (dark) and silt (light). Scale bar is 0.5 mm. (g) Photomicrograph of a magnetite-dominated sequence. The lighter layers are mixtures of magnetite and chert. Scale bar is 0.1 mm.
Fig. 9.
south in Quetico Subprovince, where prograding channel-fed lobes merge, iron formation is rare. In the Beardmore-Geraldton area, interchannel ramp development limits iron formation sedimentation. Chemical sediment could form only in interchannel areas where ramp progradation was limited by an insufficient sediment supply. As areas of volcanic
activity sporadically shifted in the arc to the north, sediment influx rates also shifted laterally along the multiple sediment entry points to the basin. This, and probably autocydic processes of channel switching, led to the transformation of clastic dominated interchannel areas into iron-oxide dominated environments. Either levee outbuilding,
Depositional controls on iron formation
Debris-flow
149 Fine-grained chemical sediment
Block diagram of the depositional environment in which the Beardmore-Geraldton chemical sediments accumulated. Iron oxide successions interbedded with d-e turbidites developed in clastic-sediment-starved interchannel areas not receiving ramp turbidites. These successions commonly coarsen and thicken upwards, probably due either to levee or to ramp progradation into the interchannel sites.
Fig. 10.
caused by channel re-establishment in the area, or ramp progradation, due to renewed high rates of sediment supply, overwhelmed these chemical systems, causing a gradual return to a clastic dominated bottom. Conglomerates interbedded with iron formation attest to an upper ramp/fan position for the chemical sediments. Conglomerates are not present in the more distal sediments to the south (Fralick et al., 1992). Abyssal plain: Terrace Bay Clastic Associated Iron Formation
Areas described as graphitic shear zones are common throughout the Canadian Shield. They are usually black, graphitic, pyritiferous slates, metres to tens of metres in thickness, which have been sheared, due to their incompetence. Depositional environments for these units vary, but the majority probably represent abyssal oceanic muds which accumulated in settings distal to active sediment sources. The area to the east of Terrace Bay (Fig. 1B) contains a number of graphitic slate assemblages. They appear to represent fault and fold repetition of the transition from sea-floor volcanics to overlying clastics (Fralick & Barrett, 1991; Eriksson etal., 1994). The Terrace Bay sediments occupy a basin bordered to the northeast and southwest by arc related volcanics. The sediments also overlie a volcanic pile kilometres in thickness. Thin, pyri tiferous black slates occur between some of the flows near the top of the pile (Fig. 12a) and a fine-grained clastic-chemical sedimentary unit, up to 40 m in thickness, directly overlies the volcanics. This unit consists of black, graphitic slate with intricately
3
�
Felsic porphyry intrusive
�-
lnterlaminated pyritic-carbonaceous slate and pyrite
• �
Massive pyrite Chert
Representative stratigraphical section of a portion of the Kingdom iron formation in the Terrace Bay area. The 40-m-thick iron formation is underlain by altered, intermediate, pillowed volcanics and overlain by turbidites. Finely interlaminated carbonaceous slate, pyrite and chert form most of the iron formation, with occasional pyrite layers up to 20 cm in thickness (detailed location in Schnieders, 1987). Fig. 11.
150
P. Fralick and T.J. Barrett
Fig. 12.
Photographs of iron formation associated with an abyssal plain assemblage in the Terrace Bay area. (a) Pyrite and chert form a zone (p) between two pillows (v). The pillow selvages have been highly silicified (si). These chemical sediments and volcanics lie directly below the iron formation (Kingdom Occurrence in Schnieders, 1987). (b) Interlayered slate (s), pyrite (p) and chert (c). Deformation has caused some disruption but the fine layering is still visible. (c) Interlayered pyrite (p) and black slate (s). (d) Turbidite succession above graphitic slate-chemical sediment zone.
laminated pyrite layers and zones of chert (Fig. 1 1; Schnieders, 1987; Barrett et a!., 1988). The pyrite occurs as massive layers (Fig. 12c) up to 25 em thick; as thin, millimetre-scale laminations in the mudrock and chert (Fig. 12b & c); as centimetre-scale spheres floating in the mudrock and chert (Fig. 12a & b); and as disseminated pyrite cubes. The cherts form zones, up to tens of centimetres thick, of either pure silica or mixtures of silica, clay and/or pyrite (Fig. 12b). The black slates are composed of quartz, sericite, carbonate, chlorite, pyrite and graphite, with scattered, floating, angular quartz and feldspar grains of possible tuffaceous origin, and rare volcanic shards. Bouma d-e turbidites are occasionally inter bedded with the metalliferous succession. A metasandstone-slate assemblage (kilometres in thickness) overlies the succession of metalliferous sediments. Beds are commonly graded from medium-
or coarse-grained sandstone to siltstone or slate and exhibit features typical of turbidites (Fig. 12d). The beds vary in thickness from several centimetres to several metres. Successions of turbidite beds range from a-dominated to d-e-dominated; and thinning fining and thickening-coarsening upward success ions are present, although uncommon. Deposition in the area began with subaqueous volcanism building a thick extrusive succession. Waning volcanic activity allowed sediment to accumulate between successive flows. With total cessation of volcanic activity, a blanket of fine grained sediment was deposited (Fig. 13). Venting hydrothermal fluids produced alteration zones in the underlying volcanics and added Si, Fe and trace amounts of other metals to the fine-grained mud raining down on the bottom. Outbuilding of a sand rich submarine fan-ramp complex ended deposition
Depositional controls on iron formation
15 1
Sand lobes
Block diagram of the depositional setting in which the Terrace Bay chemical sediments are interpreted to have accumulated. Intermediate to mafic volcanic sea floor was covered by carbonaceous muds interlayered with iron sulphides and chert. Prograding turbiditic fan/ramps built out over the chemical-clastic mixture. Fig. 13.
sediment and clay
Intermediate and mafic volcanics
of the metalliferous muds. The metal-rich deposits are interbedded with the outer-fan/ramp turbidites of this complex (Fralick & Schnieders, 1986; Schnieders, 1987). Submarine lava plain: Beardmore Volcanic Associated Iron Formation
Thick lava-plain successions are common in Superior
Province. Examples from the Abitibi Subprovince comprise major portions of cycles I and II of Dimroth et at. ( 1982), the basal komatiitic and tholeiitic portions of supergroups described by Jensen ( 1985) and, in particular, the Kinojevis Group (Jensen, 1978a, b, 198 1). Sedimentary rocks associated with the lava-plain successions are not abundant and comprise thin, interflow units, Fralick ( 1987) described a series of interflow sedimentary units
A
B
Layered siltstone and shale Massive siltstone
E2J � 1-
Layered pyrite Pyrite and layered chert Pyrrhotite and disrupted chert
f+:+l �
Felsic dike Intermediate volcanics
Fig. 14.
Fralick
\A/\ �
II
Mafic intrusive Pillowed and massive flows Volcanic ash Chert lnterlaminated magnetite and chert
Stratigraphical sections representative of volcanic-associated iron formation in (A) the Schreiber area (location in 1989), and (B) the Beardmore-Geraldton area (location in Fralick, 1987).
et al.,
152
P. Fralick and T.J. Barrett
present in a volcanic pile underlying a turbiditic succession in the Beardmore area (Fig. 1B). Oxide and carbonate iron formation constitute the domi nant lithologies present in the interflow sediment (Fig. 14B). Layers of magnetite (Fig. 15c & d) or · side1:ite (Fig. 15a & b) range from submillimetre, to centimetres in thickness, with interbedded chert (Fig. 15a, b, c & d) sometimes attaining thicknesses of metres. Laminations within beds are not common. Chlorite-rich units (with plagioclase and quartz) are often associated with the iron-rich laminae, forming bundles of alternating iron-rich and chlorite-rich layers. Clastic supply to the area was limited to volcani clastic ash, which forms the chloritic layers and isolated grains in the iron formation. The chemical sediment layers tend to be purer and thicker than the iron formations previously described. This is probably the result of limited clastic supply together
with proximity of the hydrothermal vent sources (Fralick, 1987; Fig. 16). The thickness of the inter flow sediment packages is controlled by four factors. Time duration between flows is obviously important, but the rate of hydrothermal emission, proximity to the active vents and bottom-current patterns are also major controls on unit thickness (Fralick, 1987). Submarine extrusive edifice: Morley occurrence
Submarine volcanoes produce relief of the sea floor. The Canadian Shield contains many such suc cessions, usually dominated by intermediate and felsic volcanics. Associated iron formations may be similar to those described in the section on lava plains, although they have a tendency to be inter· stratified with thicker successions of pyroclastic rocks due to the more explosive nature of the volcanism. Massive sulphide build-ups are also common in this
Photographs of volcanic-associated iron formation in the Beardmore area. (a) lnterlayered chert, c, and iron carbonate, a, at the Empire Mine, south of Beardmore. (b) Close-up of (a). The dark layers are siderite. (c) Interflow sediment consisting of chert, c, and magnetite, m. The thickest magnetite layer is loaded into the underlying chert. (d) Photomicrograph of interlayered magnetite-rich, m, and chert-rich, c, laminae. Scale bar is 0.5 mm.
Fig. 15.
Depositional controls on iron formation
153
Off-axis volcanism
proximal chemical sediment
sediment and clay Fig. 16.
Block diagram representing the depositional setting of iron formation types similar to those near Schreiber and Beardmore-Geraldton. Thick successions of pyrite and chert formed in locations close to vents, probably associated with axial valleys, off-axis volcanism or rifted arcs. Iron oxides, silicates, carbonates and chert were associated with either less i ntense and/or lower temperature portions of discharging hydrothermal zones, or were precipitated distal to major hydrothermal point sources.
setting. Iron-rich massive sulphide deposits are found in clastic- and volcanic-dominated abyssal plain assemblages, but tend to be more prolific in felsic dominated piles. The Morley occurrence will be used as an example of this type of iron formation. It is located south of Schreiber in a volcanic pile immediately west of the area discussed for the abyssal plain association (Fig. 1B). The volcanic succession containing the iron for mation was deposited about 2. 7 Ga ago, and rep resents a large, arc-type edifice (Schnieders, 1987; Fralick & Barrett, 199 1). The Morley deposit is a lenticular chemical sedimentary unit up to 7 m thick which is underlain by intermediate flows and pyro clastic rocks, and overlain by thin turbidites and structurally emplaced mafic flows (Fralick et at., 1989). The lower portion of the unit (Fig. 14A, 6 -; 10 m) contains pyrite with interbanded light and dark chert: the upper half consists of bedded to laminated pyrite (Fig. 14A, 10- 1 1.5 m). Within the pyrite-rich upper part of the succession, a variety of bedding structures are developed. The pyrite layers contain delicate internal laminations of pyrite and carbonaceous chert from 0.02 to 1 mm thick. Near the tops of individual pyrite laminae the proportion of chert and disseminated clastic debris is greater. Colloform pyrite domes up to 3 em across and 2 em thick are also present. Millimetre-scale mudstone laminae thin over the small pyrite domes and thicken in flanking depressions. Discordant pyrite growth structures on domes, together with inclined pyrite crusts and microslumps, indicate that pyrite accumu lation produced an irregular microrelief on the sea-
floor (Fralick et at., 1989; Fralick & Barrett, 199 1). Clastic supply to this area was limited (Fig. 16), although some fine-grained material, of probable volcaniclastic origin, was being delivered. Each lamina in the pyrite beds probably reflects a short term hydrothermal injection into a stagnant bottom layer of water. Upward-thinning bundles of laminae may correspond to medium-duration hydrothermal events. The domal structures, and high carbon con tent of the sediment provide evidence for relatively deep-water organic mats during chemical precipi tation (Fralick et at., 1989).
DISCUSSION
The case studies illustrate that iron formations could form in a wide range of marine environments, given the right conditions. Shallow shelves appear to be 'all or nothing ' environments. Here sediments are distributed on the shelf commonly by non channelized flows. Clastics delivered to the near shore will be spread over large areas in the offshore, limiting iron formation development. To form iron formations, clastic delivery must be minimal, such that chemical sedimentation has the opportunity to dominate. This leads to the development of Superior Type iron formation with great lateral extent, and sedimentary structures indicative of shallow water. Limited clastic delivery can be achieved if sediment storage sites are available near the strand, thereby preserving a chemical-dominated offshore. An example of this occurs in Gunflint correlatives in
154
P. Fralick and T.J. Barrett
Michigan, where clastic tidal flats with magnetite layers served as depositional sites for sands, silts and clays, leaving subtidal areas dominated by iron for mation and chert (LaBerge et al . , 1992). Other shelves dominated by chemical sediment display features quite different from the Gunflint succession. Carbonates and cherts of the Archaean Steep Rock Group exhibit prolific stromatolite development (Wilks & Nisbet, 1988) and lack grain stones. This may be due to quiet-water conditions or other, as yet unexplained, factors. Waves, storm surges and tidal currents operating on shelves, com bined with relative sea-level changes, variation in clastic influx , climate and mixing rates with offshore water, all govern the physical attributes of the chemical sediments deposited in this setting. Physical controls on chemical sedimentation on outer shelves and slopes are less well understood. Clastic supply is obviously a major control. Sea-level rise may be a key to limiting clastic supply as it provides more storage capacity on the shelf. It there fore becomes important to study the linkage in depositional response between coeval inner shelf, outer shelf and slope systems during periods of sea level rise. The development of sediment bypass zones can also produce interchannel areas dominated by chemical sediment. Upper slope interchannel areas are poorly described in the literature. This hinders evaluation of sediment bypass as a control on iron formation development. In submarine fan and ramp environments, devel opment of sediment bypass systems has been documented as a major control on formation of areas in which chemical sediment can accumulate. The development of a channel does not always prevent interchannel ramp progradation from flooding off-channel areas with clastic material (Barrett & Fralick, 1989). Changes in location of volcanic activity alter sediment supply rates along the ramp, switching depositional systems from chemical to clastic and visa versa. Bottom conditions become more unsuitable for iron formation accumu lation further downslope on the ramp or fan. In this area, flows become non-channelized, spreading out over the bottom. Sediment-starved areas are uncommon and likewise so is iron formation. To deposit iron formation here, major portions of the clastic supply system must be shut off. This may be accomplished through sea-level rise creating more sediment storage capacity in upslope environments or climatic change in the hinterland causing sediment transport systems to dry up.
Portions of oceanic abyssal plains removed from major clastic supply provided the most st:�ble sites for iron formation development. Here Precambrian sedimentary successions are commonly dominated by iron formation and fine-grained clastics. Factors controlling chemical sediment deposition in other areas were of little consequence on the abyssal plains. A mixture of iron formation and fine-grained clastic sediment could continue to be deposited until vol canism or outbuilding of a clastic pile buried this lithofacies. Depositional environments in volcanic terrains vary as much as in their clastic counterparts. Although only two examples of volcanic-associated iron formation are discussed here, they do, however, illustrate some general characteristics of chemical sediment deposited in this setting. Layering of chemical sediments in volcanic terrains reflects the dominant effect of discrete hydrothermal venting events. Layer thickness and type are inferred to be controlled by variations of temperature and com position of venting fluids (Fralick, 1987 ; Fralick et al., 1989). Interbedded ash layers may reflect magma recharge events with related increases in hydrothermal activity. These types of iron for mation should provide the best source of data on the relationship between iron formation bedding attri butes and hydrothermal activity. Layered poly metallic massive sulphide deposits are likewise an excellent source of this type of information. The above discussion highlights our relative lack of knowledge of physical controls of iron formation accumulation. Further work emphasizing the linkage between various scales of depositional process and iron formation attributes is needed. More work is also needed on chemical controls. Studies of major and trace element chemistry have been of limited value, especially where whole-rock samples rather than individual layers have been analysed. Stable isotope and REE studies have proven more interesting, although more data are needed on mono mineralic samples. Experimental modelling of the chemical systems responsible for Fe and Si precipi tation appears to be the missing link at present. In particular, we need to study the way in which progressive saturation of stable bottom-water layers by hydrothermal injections controls the nature and sequence of precipitation of iron-rich minerals. Only by integrating sedimentological, geochemical and experimental studies of iron formation can we further unravel the processes that formed these intriguing sediments.
Depositional controls on iron formation ACKNOW LEDGEMENTS
DEVANEY, J .R. & WILLIAMS,
We are particularly grateful to the Thunder Bay staff of the Ministry of Northern Development and Mines (Mines and Mineral Division) for helpful discussions and information on occurrences of iron formation in this area. Field-work in northern Quebec was assisted by Bob Wares. Useful com ments and suggestions on an earlier version of the manuscript were provided by Richard Hyde, Guy Plint and an anonymous reviewer. Figures were drafted by Sam Spivak and the word processing was conducted by Wendy Bourke. This research was supported by the Natural Sciences and Engineering Research Council of Canada.
REFERENCES
T.J. & FRALICK, P.W. (1985) Sediment redepo sition in Archean iron formation: examples from the Beardmore-Geraldton greenstone belt, Ontario. J. Sediment. Petrol. , 55, 205-212. BARRETT, T.J. & FRALICK, P.W. (1989) Turbidites and iron formation, Beardmore-Geraldton, Ontario: application of a combined ramp/fan model to Archaean clastic and chemical sedimentation. Sedimentology , 36, 221-234. BARRET!', T.J . , FRALICK, P.W. & JARVIS, I. (1988) Rare earth-element geochemistry of some Archean iron for mations north of Lake Superior, Ontario. Can. J. Earth Sci. , 25, 570-580. BRATERMAN , P.S. & CAIRNS-SMITH , A.G. (1987) Iron photo precipitation and the genesis of the banded iron formation s. In : Precambrian Iron-formations (Eds Appel, P.W.U. & LaBerge, G.L.), pp. 215-245. Theophrastus Publications, S.A., Athens. CHOUGH, S.K. & HESSE, R. (1980) The northwest Atlantic Mid-Ocean Channel of the Labrador Sea: III. Head spill deposits from turbidity currents on natural levees. J. sediment Petrol. , 50, 227-234. CLOU D , P. ( 1973) Paleoecological significance of the banded iron-formation . Econ. Ceo!. , 68, 1135-1143. CLOUD, P. (1983) Banded iron-formation - a gradualist's dilemma. In: Iron-formation Facts and Problems (Eds Trendell, A.F. & Morris R.C.), pp. 401-416. Elsevier, New York. CAIRNS-SMITH , A. G. (1978) Precambrian solution geo chemistry, inverse segregation, and banded iron for mations. Nature, 148, 27-35. DEVANEY, J . R. (1987) Sedimentology and stratigraphy of BARRETT,
the northern and central metasedimentary belts in the Beardmore- Geraldton
area
of
northern
Ontario.
Unpublished MSc thesis, Lakehead University, Thunder Bay, Ontario, 227 pp. DEVANEY, J.R. & FRALICK, P.W. (1985) Regional sedimen tology of the Namewamenikan Group, northern Ontario: Archean fluvial fans, braided rivers, deltas and an aquabasin. In : Current Research, Part B, Paper 85- J B , pp. 125-132. Geological Survey of Canada, Ottawa.
155
H.R. (1989) Evolution of an Archean subprovince boundary: a sedimentological and structural study of part of the Wabigoon -Quetico boundary in n orthern Ontario. Can. J. Earth Sci. , 26, 1013-1026. DIMROTH , E. (1981) Labrador Geosyncline: type example of early Proterozoic cratonic reactivation. In : Precambrian Plate Tectonics (Ed. Kroner, A.), Developments in Precambrian Geology 4, pp. 331-352. Elsevier, Amsterdam. DIMROTH , E., IMREH, L., ROCHELEAU, M. & GOULET, N . (1982) Evolution of the south-central part of the Archean Abitibi Belt, Quebec. Part I: stratigraphy and paleo geographic model. Can. J. Earth Sci. , 19, 1729- 1758. DREYER, J . l . , (1974) Geochemical model for the origins of Precambrian banded iron formations. Ceo/. Soc. Am. Bull. , 85 , 1099-1106. ERIKSSON , K. A., KRAPEZ, B. & FRALICK, P.W. (1994) Sedimentology of Archaean greenstone belts: signatures of tectonic evolution. Earth Sci. Rev. , 37, 1-88. FRALICK, P.W. (1987) Depositional environment of Archean iron-formation: inferences from layering in sediment and volcanic hosted end members. In : Pre cambrian Iron- Formation (Eds Appel, P. & LaBerge, G.), pp. 251-266. Theophrastus Publications, S.A. Athens. FRALICK, P.W. (1988) Microbial bioherms, Lower Proter ozoic Gunflint Formation, Thunder Bay, Ontario. In: Reefs. Canada and A djacent A reas (Eds. Geldsetzer, H.H.J., James, N . P. & Tebbutt, G.E.), Mem. Can. Soc. petrol. Geol., Calgary, 13, 24-29. FRALICK, P.W. & BARRETT, T.J. (1991) Precambrian depositional systems along the southwestern edge of the Superior Craton. Geological Association of Canada, Annual Meeting Fieldtrip A3 Guidebook, 54 pp. FRALICK, P.W. & SCHNIEDERS, B.R. (1986) Depositional environments and tectonic implications of an Archean volcanic-metalliferous slate-turbiditic succession, north ern Ontario, Canada, 12th International Sedimentological Congress, Programme with Abstracts, Canberra, p. 110. FRALICK, P.W., BARRETT, T.J., JARVIS, K.E., JARVIS, I., ScHNIEDERS, B.R. & VANDE KEMP, R. (1989) Sulfide facies iron-formation at the Archean Morley occurrence, northwestern Ontario: contrasts with oceanic hydro thermal deposits. Can. Mineral. , 27, 601-616. FRALICK , P.W., Wu, J. & WILLIAMS, H.R. (1992) Trench and slope basin deposits in an Archean metasedimentary belt, Superior Province, Canadian Shield. Can. J. Earth Sci. , 29, 2551-2557. GARRELS, R.M., PERRY, E.A. J R . & MACKENZIE, F.T. (1973) Genesis of Precambrian iron formations and the development of atmospheric oxygen. Econ. Ceo/. , 68, 1173-1179. GooDWIN , A.M. (1956) Facies relationships in the Gunflint iron-formation. Econ. Ceo!. , 5 1 , 565-595. GRoss, G.A. (1965) Geology of iron deposits in Canada, Vol. 1, general geology and evaluation of iron deposits. Ceo/. Surv. Can. econ. Ceo/. Rep . , 22, 181 pp. GRoss, G.A. (1980) A classification of iron formations based on depositional environments. Can. Mineral. , 18, 215-222. GRoss, G.A. (1983) Tectonic systems and the deposition of iron-formation. Precam. Res. , 20, 17 1-187.
P. Fralick and T.J. Barrett
156
GRoss, G.A. & ZAJAC , I.S. (1983) Iron formations in fold belts marginal to the Ungava Craton. In: Iron formation: Facts and Problems (Eds Trendell, A.F. & Morris, R.C.), pp. 253-294. Elsevier, New Y ork. HESSE, R. & CHOUGH, S.K. (1980) The northwest Atlantic Mid-Ocean Channel of the Labrador Sea: Il. D eposition of parallel laminated levee-muds from the viscous sub layer of low density turbidity currents. Sedimentology , 27, 697-711. HoLLAND, H.D. (1973) The oceans: a possible source of iron in iron-formations. Econ. Geol. , 68, 1169-1172. HuBER, N . K. (1959) Some aspects of the origin of the Ironwood iron formations of Michigan and Wisconsin. Econ. Geol. , 54, 82-118. HYDE, R.S. (1980) Sedimentary facies in the Archean Timiskaming Group and their tectonic implications, Abitibi Greenstone Belt, N ortheastern Ontario, Canada. Precam. Res. , 12, 161-195. JAMES, H.L. (1966) Chemistry of the iron-rich sedimentary rocks. U.S. geol. Surv. Prof. Pap . , 440 J, p. 61. JENSEN , L.S. (1 978a) Geology of Thackeray, Elliott, Tannahill and D okis Townships, District of Cochrane, Ontario. Ont. geol. Surv. Rep . , 165, 71 pp. JENSEN, L.S. (1978b) Geology of Stoughton and Marriott Townships District of Cochrane. Ont. geol. Surv. Rep . , 173, 73 pp. JENSEN , L.S. (1981) A petrogenic model for the A rchean Abitibi
Belt
in
the
Kirkland
Lake
area,
Ontario .
PhD thesis, University of Saskatchewan, Saskatoon, Saskatchewan, 520 pp. JENSEN, L.S. (1985) Stratigraphy and petrogenesis of Archean metavolcanic sequences, southwestern Abitibi subprovince, Ontario. In: Evolution of A rchean Supra crustal Sequences (Eds Ayres, L. D., Thurston, P.C., Card, K. D. & Weber, W.), Geol. Assoc. Can. Spec. Pap. , 28, 65-87. KRONBERG, B.l. & FRALICK, P.W. (1992) Geochemical alteration of felsic Archean rocks by Gunflint Formation derived fluids, Quetico-Superior region, northwest Ontario. Can. I. Earth Sci. , 29, 2610-2616. LABERGE, G.L., ROBBINS, E.l. & HAN , T.-M. (1987) A model for the biological precipitation of Precambrian iron-fonnations - a: geological evidence. In: Precambrian Iron-formations (Eds Appel, P.W . U . & LaBerge, G.L.), pp. 69-96. Theophrastus Publications, S.A., Athens. LABERGE, G.L., 0JAKANGAS , R.W. & KICHT, K. (1992) Early Proterozoic and Archean rocks of the Gogebic Range. Institute of Lake Superior Geology, 38th Annual Meeting, Part 2, Fieldtrip Guidebook. Hurley, Wisconsin. LAGALLAJS, C.J. & LAVOIE , S. (1982) Basin evolution of the Lower Proterozoic Kaniapiskau Supergroup, central Labrador miogeocline (trough), Quebec. Bull. Can. petrol. Geol. , 30, 150-166. LouGHEED, M.S. (1983) Origin of Precambrian iron formations in the Lake Superior region. Geol. Soc. Am. Bull. , 94, 325-340. MEYN, H.D. & PALONEN , P.A. (1980) Stratigraphy of an Archean submarine fan. Precam. Res. , 12, 257-285. MOREY, G.B. (1983) Animikie Basin, Lake Superior region, U.S.A. In : Iron-formation: Facts and Problems (Eds Trendell, A.F. & Morris, R.C.) pp. 13-68. Elsevier, New Y ork.
MoREY, G.B., SouTHWICK, D . L. & ScHOlT LE R , S.P. (1991) Manganiferous zones in Early Proterozoic iron formation in the Emily district, Cuyuna Range, east central Minnesota. Minn. geol. Surv. Rep. Invest. , 391, 42 pp. MoRRIS, R.C. (1993) Genetic modelling for banded iron formation of the Hamersley Group, Pilbara Craton, Western Australia. Precam. Res. , 60, 243-286. OEHLER, J .H. (1976) Hydrothermal crystallization of silica gel. Geol. Soc. Am. Bull. , 87, 1143-1152. 0JAKANGAS, R.W. (1983) Tidal deposits in the early Proterozic basins of the Lake Superior region. The Palus and the Pokegama Formations; evidence for subtidal shelf deposition of Superior-type banded iron-formation . In : Early Proterozoic Geology of the Lake Superior Region (Ed. Medaris, L. G.) , Geol. Soc. Am. Mem., Boulder, 160, 49-66. ScHNIEDERS, B. R. (1987) The geology of sulfide-facies iron formations and associated rocks in the Lower Steel River-
Unpublished MSc thesis, Lakehead University, Thunder Bay, Ontario, 196 pp. SHULSKI, T. , WARES, R.P. & SMITH, A. D. (1993) Early Proterozoic (1.88-1.87 Ga) tholeiitic magmatism in the New Quebec orogen. Can. I. Earth Sci. , 30, 1505-1520. SIEVER, R. (1957) Silica budget in the sedimentary cycle. Am. Mineral. , 42, 821 - 841. SIMONSON , B.M. (1985) Sedimentological constraints on the origins of Precambrian iron formations. Geol. Soc. Am. Bull. , 96, 244-252. STAN LEY, D.J. and WEAR, C.M. (1978) The 'mud-line': an erosion-deposition boundary on the upper continental slope. Mar. Ceo!. , 28, 19-29. STOw, D.A.V. & SHANMUGAM, G. (1980) Sequences and structures in fine-grained turbidites; comparison of recent deep-sea and ancient flysch sediments. Sediment Geol. , 25, 23-42. TEAL, P.R. & WALKER, R.G. (1977) Stratigraphy and sedimentology of the Archean Manitou Group, north western Ontario. Geol. Surv. Can. Pap . , 77-1A , 181-184. VAN H 1 s E , C.R. & LEITH, C.K. (1911) The geology of the Lake Superior region. U.S. geol. Surv. Monogr. , 52, 641 pp. WHITE, D.A. (1954) Stratigraphy and structure of the Mesabi range, Minnesota. Minn. geol. Surv. Bull. , 38, 92. WARDLE, R . J . & BAILEY, D . G. (1981) Early Proterozoic sequences in Labrador. In: Proterozoic Basins in Canada (Ed. Campbell, F.H.A.), Geol. Surv. Can. Pap., Ottawa, 81-10, 33 1-358. WARES, R.P. & GouTIER, J . (1990) Deformational style in the foreland of the northern New Quebec Orogen. Geosci. Can. , 17, 244-249. WILKS, M.E. & NISBET, E.G. (1988) Stratigraphy of the Steep Rock Group, northwest Ontario: a maj or Archean unconformity and Archean stromatolites. Can. J. Earth Sci. , 25, 370-391. WooD, J. (1980) Epiclastic sedimentation and stratigraphy in the N orth Spirit Lake and Rainy Lake areas: a comparison. Precam. Res. , 12, 227-255. Little Steel Lake area, Terrace Bay, Ontario.
Spec. Pubis int. Ass. Sediment. ( 1995) 22, 157-193
Facies models in volcanic terrains: time's arrow versus time's cycle G EO F F R E Y J . O RTON Department of Geology, McMaster University, Hamilton, Ontario L8S 4Ml, Canada
ABSTRACT
Vertical and lateral facies variations in volcanic terrains are abrupt owing to the sudden input of additional sediment from one or more volcanic vents. Further complications arise from the influence of volcanic activity on the subsidence history of the basin. A detailed study of one small depositional system from the Ordovician Llewelyn Volcanic Group near Tryfan Fach, North Wales is used to illustrate the problems this causes for developing facies models. The Tryfan Fach Member is made up of four facies associations characterized by their geometries, composition, and sedimentary structures. Each is assigned to a particular depositional setting: marine shelf, braided stream, floodbasin and alluvial fan. The basal portion comprises a comparatively thick , mudstone-dominated succession deposited in a quiet-water marine setting largely below fair-weather wave base. In contrast, coarse-grained, rhyolite-bearing sandstones were deposited in shallow, flood prone, southward flowing bedload-dominated braided streams and as unconfined sheet floods. These sandstones amalgamate to form a laterally extensive sheet c. 15m thick and at least 2 km wide, which lies with sharp contact on subjacent marine mudstones. Sandstones pass gradationally upwards into interbedded coarse and fine sandstones, laminated vitric siltstones with accretionary lapilli, and mudstones, deposited on a low-gradient coastal plain. Conglomeratic alluvial fan deposits, derived from older deposits of granite, tuffaceous sediment and rhyolite, occur along the northeast margin of the basin. Although several features can be explained by envisaging the whole succession as the product of one linked depositional system , the differences in sediment composition and palaeocurrent trends raise problems. These cast doubt on the strict application of Walther's Law to the total succession, and demand at least three genetically unrelated depositional systems. The rhyolitic braidplain to floodbasin succession is attributed to subaerial aggradation of primary and reworked pyroclastics, and had a different source from subjacent marine mudstones and the adjacent alluvial fan. The sharp basal contact of the braidplain sandstones is interpreted as an erosional unconformity, and is inferred to have resulted partly from volcano-tectonic uplift in advance of the volcanic eruption. The contemporaneous progradation of an epiclastic alluvial fan from the opposite side of the basin is related to accelerated but more differential basin subsidence during volcanism. Within the syneruptive deposits it could not be established whether facies change reflected sorting processes on the alluvial plain or progressive eruption of finer grained ash. As a consequence, vertical sequences are difficult to interpret, palaeo geographical maps are speculative, and the timing of relative sea-level changes could not be assessed fully.
INTRODUCTION
Vertical-sequence analysis, essentially an extension of Walther's Law, is probably the single most important advance in sedimentology in recent decades (Dott, 1983) . It provides a powerful tool with which to interpret changes in sedimentary
environments through time from two-dimensional sequences (de Raaf et al. , 1965 ; Visher, 1965). As Walther stressed, the law can be applied only to successions with gradational , non-erosive contacts between facies and/or environments (see Middleton,
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
157
158
G.J. Orton
1973). Also implicit in its application is the assump tion that the control on changes in grain size and/or facies largely lies within the area of deposition. External parameters (e.g. sediment supply, relative sea-level) change gradually rather than abruptly; continuity rather than discontinuity is assumed. Although modern facies models do allow for 'cata strophic' events (e.g. storms) , these catastrophes are assumed to recur with the same frequency in the past, present and future . And the thickness of sedimentary deposits resulting from catastrophes is normally not sufficient to alter radically the pre-existing geography and patterns of sediment dispersal. Things are not so simple in volcanic terrains. B y their very nature, explosive volcanic eruptions are catastrophic and rapidly change the supply of sedi ment available within a catchment area. The grain size and type (crystals, lithics, glass shards) of sedi ment supplied can change with equal rapidity during the course of a volcanic eruption owing to changes in the rate of magma extrusion, volatile content of magma erupted or the amount of interaction with external water in the vent (see Heiken & Wohletz, 199 1 ; White, 199 1 ) . During eruptions, patterns of sediment dispersal are not always controlled by pre existing topography; ash-fall deposits in particular are controlled by wind patterns and are distributed over broad areas independent of topography and , to some extent, gravity. Depositional basins around volcanoes are often infilled from multiple sources. The growth of intrabasinal volcanoes can form physiographic barriers that isolate depositional systems and limit the fetch (and hence size) of waves reaching coastlines. Relative sea-level can change quickly owing to uplift or subsidence resulting from a volcanic eruption. As a consequence, ancient volcaniclastic suc cessions are notoriously difficult to interpret. Two concepts jockey for recognition in every ancient succession: (i) each bed in the succession represents a unique historical event (i.e. a new eruption, a new type of eruption) with no genetic relation to the bed before; (ii) the succession represents a 'group of facies genetically related to one another and which have some environmental significance' (Collinson, 1969). The first perspective relates facies change to changing source parameters. Sedimentology essentially loses its predictive capability and ability to define environments. The second is an ahistorical perspective in which laws, processes and the pace of change remains predictable . I think these two
concepts, labelled by Gould ( 1987) as time's arrow (a linear succession of unique events) and time's cycle (recurrent patterns in a world that remains essentially unchanged), embody the major issues within analysis of every volcaniclastic succession . A small-scale but well-exposed sequence from the Ordovician Llewelyn Volcanic Group near Tryfan Fach, North Wales is used to illustrate the differences and interaction between these two concepts. At this locality, a 15-m-thick sheet of coarse-grained rhyo litic volcaniclastics is bounded by marine mudstones or fine-grained sandstones, and lies a few kilometres away from conglomeratic alluvial fan deposits. The main problem addressed in this paper is whether the rhyolitic sandstones are reworked pyroclastics and indicate penecontemporaneous volcanism, and whether the other stratigraphical units had the same source. Through analysis of sediment composition , it can be shown that there is no genetic relationship between vertically and laterally adjacent strati graphical units. Each represents a separate depo sitional system. The implications for facies modelling and reconstructing environments in volcanic settings are discussed.
TERMINOLOGY
Sands rich in volcanic debris are of two dominant types according to the method of clast formation . Pyroclastic material (shards, pumice, crystals) is generated by explosive volcanism , and is often contributed directly to the sedimentary record as primary pyroclastic deposits. In contrast, epiclastic volcaniclastic materials result from erosion and weathering of older volcanic rocks, including lithified tuffs, and do not usually reflect contemporaneous volcanicity. A 'grey area' within this classification concerns reworking of unconsolidated pyroclastic debris. Many authors (particularly Fisher, 196 1 ; Fisher & Schmincke, 1984) regard recycled pyro clastic material, redeposited by wind or water as secondary pyroclastics; others (e.g. Cas & Wright, 1987) suggest that pyroclastic particles reworked by water or other agents should be called epiclastic. In older and more deformed successions, dis tinction becomes even fuzzier, and one is often lucky if one can demonstrate that some particles were produced by explosive volcanism. A further complication is that many volcaniclastic deposits consist of mixtures of both syneruptive pyroclasts and epiclastic material (sensu Fisher & Schmincke ,
Facies models in volcanic terrains
1984) . The recognition of pyroclastic debris where such material is diluted by and mingled with sands of other origins is not easy. Resolving this problem , however, is far from an exercise in semantics but is crucial in determining the nature of the volcanic contribution to sedimentation. Calling all pyroclastic material that has suffered additional transport 'epiclastic' hides the fact that volcanicity may have been active at that time, and breaks the genetic lineage between deposits and their source. In this study the term pyroclastic is used where the composition, texture, variability, and geometry of the deposits all suggest that clasts were produced by a contemporaneous volcanic eruption. Pyroclastic material that was retransported by 'normal' hydro logical processes shortly after its initial deposition is referred to as reworked pyroclastics in order to emphasize that sedimentation was still occurring in response to volcanism.
GEOLOGICAL FRAMEWORK
During the latter part of the Ordovician, North Wales was part of an extensional or transtensional marginal basin (Campbell et al. , 1988; Kokelaar, 1988) sited on continental crust comprising accreted volcanic arcs (Thorpe , 1979). The Precambrian crust formed part of a small microcontinent, Eastern Avalonia (Soper & Hutton, 1984) , derived from continental Gondwanaland, and separated from the North American continent (Laurentia) and Baltica by the Iapetus ocean and the Tornquist's sea. Palaeo magnetic and faunal reconstructions record the northward movement of Avalonia throughout Early Ordovician to Early Devonian times (Cocks & Fortey, 1982; van der Voo, 1983, 1988) , with temperate southerly latitudes (c. 30-35°S) indicated for mid to late Ordovician (Caradoc) times (Torsvik & Trench, 1991). Alluvial fan facies suggest a humid climate (Orton, 199 1 ) . Ordovician sedimentation i n North Wales was dominated by muddy facies. Volcanism and depo sition of associated coarse-grained clastic material was confined largely to the Caradoc series, and confined to a NE-SW oriented rift termed the Snowdon graben. An orthogonal array of deep seated fractures cutting the ensialic basement (Campbell et al . , 1988; Kokelaar, 1988) controlled the position of eruptive centres within the graben. Reactivation of these and shallower fractures as faults, often in conjunction with volcanism , deter-
159
mined configuration of sub-basins, rates of fault block subsidence, the distribution of volcanic and sedimentary units, and depositional settings. The Snowdon graben can be divided into at least two structural basins, 10- 15 km wide, about Llanberis Pass (Fig. 1) based on the distinctive petrography of their contained sediment (Orton, 1990) and the distribution of basalt intrusions (Campbell et a!. , 1988) . The basin northeast of Llanberis Pass has been referred to as the Tryfan depocentre (Orton, 1990) . Volcanism, and the deposition of coarse clastic material, varied in time and space. It can be divided into two eruptive cycles that are reflected in two volcanic groups (Howells et at. , 199 1 ) : a lower Llewelyn Volcanic Group in northeast Snowdonia (Fig. 1) and an upper Snowdown Volcanic Group in southwest Snowdonia. Although this activity was confined to just two chronostratigraphical stages (Soudleyan and Longvillian) , about 5% of total (72 Ma) Ordovician time , its products comprise about half the total thickness of the Ordovician sequence. The average rate of sediment accumu lation during the Soudleyan and Longvillian approached 1 m lOOO yr-1 (Orton, 199 1 ) . The earliest volcanic activity in the Llewelyn Volcanic Group developed from at least four, partly contemporaneous , centres. Rhyolite lavas and silicic ash flow tuffs (Conwy Rhyolite, Braich tu Du Formation) , trachyandesite lavas and tuffs (Foel Fras Formation) and basaltic-andesite lavas (Foel Grach Basalts) were all erupted (Fig. 1). Extrusive rocks are associated mainly with marine mudstones and show little evidence of reworking; they were probably ponded in subsiding, fault-bounded areas of the sea-floor (Howells et al. , 199 1 ) . However, the latest activity in the Llewelyn Volcanic Group, the Capel Curig Volcanic Formation, was dominated by larger scale subaerial eruptions of silicic magma, giving rise to widespread ash flow tuff deposits. Volcaniclastic successions underlying the Capel Curig Volcanic Formation, herein referred to as the Tryfan Formation (Figs 1 & 2), reflect a compara tively reduced amount of volcanic activity. The Tryfan Formation can be divided into at least eight members (Fig. 2), based on differences in sedi mentary facies and/or environments, and the amount, petrography, chemical composition and location of any volcanicity associated with sedimen tation. Each member represents a distinct depo sitional episode with no genetic relation to the episode before. The Gwern GofTuff, and underlying fluvial- deltaic deposits were derived from a source
160
G.J. Orton
N
t
� CapeiCt.rig �"J Volcanic Formation (;:::;::::::1 Tryfan Formation
r:Mi:!;(!Jil
-
Foel Fras Volcaric CC>I"\l)lex
Foel Grach Basalt Fm. � &aichTuDu � Volcanic Formation � Coowy Rhyoite Formation
�
o-
-skm
Stbvolcaric ntrusions Important faults
Fig. 1. Llewelyn Volcanic Group: distribution of formations and related subvolcanic intrusions. Box locates Fig. 2.
Intrusions in the south identified as follows: MP, Mynydd Perfedd; BC, Bwlch y Cywion; T, Talgau; CL, Carnedd Llewelyn (modified from Howells et al. , 1991). On inset map, shaded area denotes approximate position of Snowdon graben, MS refers to the Menai Straits.
to the east-southeast (Orton, 1988) . Progradation and/or transgression of coarse-grained fluvio-deltaic systems was the dominant mode of basin infill above the Gwern Gof Tuff. Each of these prograded toward the south and southeast, indicating a 'flip' in the polarity of infill of the Tryfan depocentre. The depo sitional package of concern here lies directly above the Gwern Gof Tuff and is referred to as the Tryfan
Fach Member (Figs 2 & 3 ) . It varies in thickness from 30 to 120 m , and has been recognized only within the Tryfan anticline (Figs 1 & 2) , over an extremely small area ( < 20 km2) . Exact information on the time-scale spanned by the Tryfan Fach Member cannot be provided. However, it was of the order of a few hundred thousand years and certainly not millions of years.
161
Facies models in volcanic terrains 68
I
I
*
fossillocalities
500
m
66
Fig. 2. Simplified geology of the Tryfan anticline (modified after British Geological Survey, 1985, unpublished 1:10000
sheets; Orton, 1990). Numbers on border refer to UK Ordinance Survey National Grid (grid square SH) and are 1 km apart. Circled letters on map refer to location of the main sedimentary sections, whereas numbers on stratigraphical log (and associated textures) denote the three basinal facies associations: 1, marine shelf mudstones; 2, Rhyolitic braided stream sandstones; 3, floodbasin sandstones .
FACIES ASSOCIATIONS
The stratigraphical record of the Tryfan Fach Member can be divided into four facies associations, each assigned to a particular depositional setting. As cleavage obscures sedimentary facies in fine grained lithologies (unless rhyolitic) , environmental interpretations are based largely on data collected from sedimentary rocks of medium silt grade and coarser. Marine shelf mudstones
The basal portion of the Tryfan Fach Member in the south consists of poorly exposed mudstone with sharp-based massive or massive to horizontally laminated beds (to 1 0 cm) of coarse siltstone and fine-grained sandstone. Upper surfaces of sandstone beds are either sharp or grade into overlying mud-
stone. One bed of coarse-grained lithic sandstone containing rounded rhyolite pebbles (Fig. 4, 49 m) also occurs. Rare brachiopods (Rostricellula) indicate a marine setting (Tunnicliff & Rushton, 1980). Thicker horizons (to 2 m thick) of rhyolitic fallout tuff (Fig. 3) consist of thin, 8 -1 0 cm thick, massive to graded beds of vitric siltstone. The succession generally becomes sandier upwards with bioturbated sandy siltstone giving way to inter stratified very fine sandstone (beds 10- 100 em) and mudstone (beds to 20 cm) . Most sandstone beds are sharp based, massive, sometimes normally graded, and rarely infill isolated-scours 30 em deep. Beds near the top of the succession , however, contain ripple cross-stratification, small-scale (to 15 em) trough cross-stratification , planar horizontal lami nation , and more rarely undulating lamination. Vv'here trough cross-stratification occurs, sets are often separated by discontinuous siltstone lenses.
162
G.J. Orton
Facies associations
0">
en u::
� ;
100
'
a: (!) :2 I u
z
u..
>a: f-
�� ....
Thinly bedded floodbasin sandstones
"
Rhyolitic braided stream sandstones
60
en u::
40
Marine shelf m udstones
20
GWERN
GO F
TUFF
Fig. 3. Composite log through the Tryfan Fach Member (from the vicinity of log C, Fig. 2) showing stratigraphical relationships between facies associations. Ruled vertical bars denote position of more detailed sedimentary logs. The Gwern Gof Tuff (base of section) is about 40 m thick and only its top portion is depicted. The top of the braided stream association forms the popular rock-climbing slab at Tryfan Fach.
163
Facies models in volcanic terrains Facies interpretation 65
5cm
Small scale trough and ripple Cross stratification
? Shoreface
Wave swash on sand bar?
Low-angle planar lamination
------ -- Fairweather wave base-??
Bioturbated siltstone
55
Ripple cross stratification
Alternation of rapid deposition by storms and waning alluvial floods and slow suspension sedimentation
Pebbly sandstone
Graded sandstone beds
45
1m
I
Fig. 4. Section from mid- to upper portion of the marine shelf mudstone association. Scale is in metres above Gwern Gof Tuff. From section B, east of Tryfan Fach, SH 67205995. Log is continued in Fig. 5. See Fig. 3 for stratigraphical position of log.
164
G.J. Orton
The top few metres of the association are again fine-grained (Fig. 5 ) . Mudstones are slightly tuf faceous and a laterally continuous comparatively pure unit of vitric siltstone (50- 100 em thick) lies directly beneath the coarse-grained rhyolitic volcaniclastic material of the braided stream associ ation (Figs 5 & 6) . Mudstones contain thin ( < 30 cm) tabular sandstone beds, most of which display bio turbation. Although the sandstone beds do not pinch and swell they could not be traced between sections only a few hundred metres apart. Indistinct parallel lamination and more rarely trough cross stratification were the only sedimentary structures observed. Overall, facies and fossils indicate deposition in a quiet-water muddy environment. Environmental interpretation of Caradoc fauna is facilitated by comparison within coeval strata in the nearby (< 100 km) Berwyn Hills (Pickerill & Brenchley, 1979). Here Rostricellula is associated with the Macrocoelia subcommunity of a Dinorthis com munity and is considered to have colonized high energy non-turbid, well-oxygenated environments on silty substrates under moderately high rates of sedimentation. A maximum water depth of 25 m was assigned to the Macrocoelia subcommunity based on sedimentary structures, and the presence of the fossil-boring Vermiforichnus (Pickerill , 1976) . The bottom part of the Tryfan Fach Member was dominated by deposition of mud from suspension. Coarser grained sediment was introduced period ically in decelerating flows. These could represent suspension clouds deposited during the waning stages of a storm (Nelson, 1982) or underflows (hyper pycnal effluent) from river-derived flood events. Although underflows are documented most com monly where river waters enter freshwater lakes, they also occur when bedload-dominated streams enter marine basins. Wright et al. (1986) confirmed that hyperpycnal underflows carrying silt occur off the mouth of the Huanghe (Yellow) River. Under flows carrying sand develop when floodflows from the San Lorenzo River (California) scour the river mouth and pass offshore as a plane jet (Hicks & Inman, 1987) and are suspected also to be a common phenomena on fan-deltas along fjords (Prior et al. , 1987 ) . I f storm o r river-introduced silts and sands d o not exhibit evidence of reworking by waves, it is usually inferred that beds reflect deposition below effective wave base. A comparable facies described from other Ordovician successions in North Wales was
assigned to an 'outer shelf setting' (Fritz & Howells , 199 1 ) . Care must be taken, however. As noted by Pickerill & Hurst ( 1983), if suspension clouds produced by storms or river floods contain , in addition to sand and silt-grade material, sufficient mud to provide cohesiveness, modification by later currents could be precluded even though deposition occurred above normal, even 'fair-weather' , wave base. The upper sandier part of the succession contains unequivocal evidence for deposition in shallower water. Planar lamination (Fig. 4, c. 61 m) may reflect wave swash processes. However, a similar facies , termed quasi-planar lamination, described from Lower Cretaceous lower shoreface to shelf deposits of Montana has been attributed to deposition under single-event, high-energy combined-flow conditions (Arnott, 1993 ) . Outcrop quality is not adequate to discern whether lamination remains perfectly planar or gently undulates, although the position of the lamination at the base of an upward-shoaling unit lends some credence to a combined-flow origin . Undulatory lamination, considered to form under intense oscillatory flow (Allen, 198 1 ) or combined flow (Myrow & Southard, 1991) occurs nearby at about the same stratigraphical level and provides additional evidence for wave processes. The twice observed, sandstone-filled scours are similar in size and orientation (but not abundance!) to gutter casts described by Myrow ( 1992) from shallow subtidal deposits in Newfoundland. In his model the subtidal zone , dominated by fine-grained sediment, is largely a zone of sediment bypass in which high-velocity sediment-laden flows erode shore-normal scours preserved as gutter casts. The trough cross-stratification of the sandiest deposits (Fig. 4, 63-66 m) is thought to be produced by waves and/or river mouth floods. The absence of facies such as lenticular and flaser bedding, reactivation surfaces, rhythmic bedding and mud drapes suggest that tidal processes were absent or unimportant. The small size of the cross-stratification, inter-bedding of mudstone, and absence of swaley or hummocky cross-stratification indicates low near shore wave power. This requires further comment. Wave power at a coastline depends on the maximum deep-water wave energy and its shallow-water fric tional attenuation, which is a function of the sub aqueous slope. Younger shorefaces of the Tryfan Formation commonly contain abundant hummocky and swaley cross-stratification, undulating lami nation, and large wave ripples (e.g. Orton, 1988;
. c
E
� 1-"',
ri®� :·�
500
m
8-------
.. ' n·v
-� --8- --- --------- sil one - -ts ---t... ...
120
ri o o-- ---D
,.._
�
400
m
. �-;·
� _.,._. �
�·
�
marker
0
0
�
0
"'i>= o --..
u:;:;:;;;;:::s.;;;-
l2
0
,
o-
o
0
400
m
FLOODBASIN
•
•
118
"'
------1
o--
-
�)�
,
B
� ("') �·
.
--...i'.J-· -..!!!!..� .. .... ..
;:s
"'J"=
Channel fill
Channel fill
��
0 0 0
n=ll
BRAIDED STREAMS
KEY
I I
Indistinct stratification
MUDSTONES
I�
I
1:.:;5'
2 ;:, Bioturbation
MARINE
� "' 0
128
.----
0
Graded sandstone Ripple cross-strat
c:;· <;;
�
5' "'
Horizontal lamination
. . Trough cross-strat . .., v\ Ashfall tuff , --- -- ....... Bedded sandstone
�>
�. " v v
3-10 0 11-20 0 21-30 0 >30 0
Mudstone
SET SIZE INCM
Fig. 5. Detail of facies variation in upper part of Tryfan Fach Member on west limb of Tryfan anticline. See Fig. 2 for location of sections. Section B is a continuation
(without a stratigraphical gap) of section B depicted in Fig.
4. ...... 0\ Vl
166
G.J. Orton
Fig. 6. Contact between (A) fine
grained bioturbated sandstones of the marine shelf association (beneath hammer) and (B) trough cross-stratified coarse-grained volcaniclastic sandstones of the braided stream association (above hammer), which often contain a thin (30-50cm) bed of siliceous mudstone. General fluvial palaeocurrent is to the left and towards the reader. Hammer is 30cm long.
Fritz, 199 1 ) . As it seems unlikely that deep-water wave energy changed dramatically over a few hundred thousand years, the low wave power documented throughout the Tryfan Fach Member is related to a low subaqueous slope. At the time of emplacement of the Gwern Gof Tuf (Fig. 3) , the offshore slope must have dipped to the west northwest. In contrast, the rhyolitic braidplain deposits of the Tryfan Fach Member indicate a slope to the east-southeast. Sometime during the inter vening time period, that is when the marine shelf association was being deposited, the slope must have rotated, either tectonically or by deposition, through the horizontal. Low-gradient muddy shelves are highly efficient at attenuating wave energy (cf. Wells & Coleman, 198 1 ). On the Surinam coast line the total wave-energy loss through slow shoaling across the inner shelf ranges from 9 3 to 9 6% , and most waves do not reach the shoreline nor break. Rhyolitic braided-stream sandstones
A multistorey, 12- 15 m thick , laterally extensive (about 2 km) horizon of poorly sorted granule conglomerate to fine sandstone directly overlies the muddy coastal sediments (Figs 3, 5, 6 & 7). Tuffaceous siltstone rip-up clasts (to 5 em) and sub angular rhyolitic pebbles (to 2 cm) are common . The sandstones were deposited as broad tabular sheets 20-250 cm thick separated by thin ( < 20 cm) beds of vitric siltstone or interbedded siltstone and fine
sandstone. Fine-grained intervals are laterally continuous for distances of about 200 m, with the exception of a tuffaceous, green-coloured siltstone that extends for at least 900 m (Fig. 5). Sandstone sheets are characterized by extremely abrupt vertical changes in grain size and texture . Vertical-sided , laterally stepped erosional scours up to 30 em deep sometimes occur along their base. In thicker sheets, shallow trough cross-stratification is the dominant sedimentary structure , with rare climbing ripple cross-lamination and horizontal lamination. Palaeo current distribution is unimodal towards the southeast. Many thick sheets fine upwards , from conglomeratic coarse sandstone to siltstone (Fig. 8). When this occurs the set size of cross-stratification decreases upwards ( c. 20 cm to Scm) in conjunction with grain-size changes. Thinner sandstone beds display a wider range of sedimentary structures. Coarse-grained beds are commonly normally graded , fining up from conglomeratic sandstone to siltstone over short (< 50 em) vertical distances. Finer grained beds are massive, have massive bases with cross stratified tops, horizontal lamination, or undulating lamination. The geometry of the above sandstone packages , their internal structures, coarse texture and uni modal palaeocurrents indicate deposition by uni directional tractional currents under a broad range of flow strengths. A subaerial setting is assumed because: (i) there is no evidence (e.g. wave ripples) within fine-grained beds for marine processes ,
Facies models in volcanic terrains
167 Facies interpretation
Scm 5cm
Stacked 'channel' fills
Small scale trough Cross-stratification
�
"
Low-angle indistinct stratification
....
�:..:...:_ ,_ �,�·
A
/
/ � ··<-'"·�.., io -.--- ,
10cm Sea-green siltstone horizontal lamination and rippling
z 0
t=
lnfill of deeper fluvial channel
0 (/) (/)
Aggradation in shallow unconfined braided stream
20cm ;...;,.:...��·
Shallow trough cross-stratification
1 5 cm
lnfill of minor 'channels'
30cm
<(
10cm
a: a:l
Graded sandstone beds
'Overbank' floods
Undulating lamination
Horizontal lamination with small scours
Upper flow regime sheetfloods
Fig. 7. Complete section through
braided stream facies association. From Section D, east slope of Tryfan, SH 6683 5945. See Figs 3 & 5 for stratigraphical position of log. Numbers to right of log denote thickness of cross-strata. In thin beds that fine upwards the set size indicated refers only to sets at the base of the beds.
Rhyolitic siltstone
Ashfall
tuff
Bedded sandstone
M
(ii) there is no bioturbation as occurs within other facies associations, (iii) sandstones are extremely poorly sorted, (iv) tidal indicators (e.g. mudstone drapes) are absent.
Si
Sa
G
The thicker sandstones are best interpreted as infill of fluvial , probably braided, channels (Miall, 1977). In relation to modern braided rivers, these sandstones resemble to some extent the South
168
G.J. Orton
Fig. 8. Channel-fill sequence from a 2-m-deep pebbly
braided stream (braided stream association). Base of sequence is just beneath the field of view. Height of cross stratification decreases upwards, from 20cm to 4cm. Flow of fluvial current was towards the reader.
Saskatchewan River (Cant & Walker, 1978) in being dominated by trough cross-stratification. The South Saskatchewan maintains average channel depths of 3m (maximum 5 m) , in which trough cross-stratified sands accumulate by migration of sinuous crested dunes. Fining, upward sequences are generated as the channel shallows due to lateral migration. As this occurs, the trough sets are overlain and/or interbedded with stacked sets of planar cross stratified sand formed by cross-channel migration of shallow bar complexes. Unlike the South Saskatchewan, the Tryfan Fach Member contains no planar cross-stratification. Although puzzling this is not unique and numerous other ancient braided stream successions, such as the Jurassic Westwater Canyon Formation
(Campbell, 1976; Godin, 199 1 ) , or Pennsylvanian South Bar (Rust & Gibling, 1990a) and Boss Point (Browne & Plint, 1994) Formations are similarly dominated by trough cross-stratification. The absence of planar cross-stratification has been explained in at least two ways. Some suggest that lateral migration of deeper channels across a braided sand-flat would lead to the preferrential preservation of three-dimensional dunes at the expense of cross channel bars characterizing shallower channel por tions (Campbell, 1976). Alternatively, channel avulsion, causing rapid abandonment of the pre viously active tract, combined with flood-stage erosion of top-stratum would prevent development of fining upward waning channel fill sequences and result in a sheet-like sandbody dominated by trough cross-strata (Rust & Gibling, 1990a; Browne & Plint, 1994) . Several important observations suggest a differ ent interpretation for the trough cross-stratified 'channel' fill units of the Tryfan Fach Member. 1 There is little evidence, such as channel forms, deep scours or channel base lags, that bedforms were confined within deep channels. Instead, fluvial cycles rest on virtually planar erosion surfaces or wedge out at their base against broad, shallow channels. 2 Although sediments are poorly sorted, cosets often fine upwards with a corresponding decrease in set size (Figs 7 & 8) . In contrast , no clear grain size trends in the facies cycles are present within the Boss Point Formation and co-sets are inferred to represent partial-channel fills (Browne & Plint, 1994). Although trough cross-sets in the Boss Point range from 50 em to 120 em in thickness, vertical variation in set-size is not reported. 3 The set size is small enough that sandstone sheets could represent complete 'channel' fills. A broad relationship between bedform height (H) and flow depth (h) in rivers has been established (e.g. Allen, 1984, fig. 8-20). Preserved set thickness will be considerably less than the height of dunes originally present at flood stage because of relaxation during falling stage (Allen, 1 974) and truncation by sub sequently migrating scours. Consequently, Rust & Gibling (1990a) suggest adding 50% as a conservative estimate of the amount lost. This approach gives a flood-stage dune height of about 30 em for the largest dunes at the base of cosets and a few centimetres for bedforms at their tops. The mean relationship (H!h 0. 15) of Allen ( 1984) yields a flood-stage flow depth of about 2 m for the large basal dunes, a =
Facies models in volcanic terrains
depth comparable to the thickness of the cosets. The broad, extremely low-angle, and shallow nature of the trough cross-stratification (Fig. 8) indicates either extremely rapid flows and/or shallow flow depths (Jopling, 1965 ) . There are similarities with proglacial sandur deposits (e.g. Ruegg, 197 1 } . For all o f the above reasons, there seems iittle doubt that most of the fining upward trough cross stratified sandstone sheets record virtually uncon fined flood events. The thickness and type of sequence resulting from such flood events would depend on the depth and duration of the flood, and the amount and grain size of sediment available. Three possibilities are considered . The thickest sheets require that a range of grain sizes are available and that flow-depths change in a comparatively gradual fashion so that bedforms have time to respond to the decreasing flow strengths. These cosets represent the most prolonged (and most widespread} discharges, and may have involved some channel incision or migration. Thinner sand stone sheets where trough cross-stratification grades rapidly into ripple cross-stratification would develop where floods waned extremely rapidly, whereas thin incomplete (top missing) successions could result if flood waters simply ran out of sediment to deposit. Although longitudinal sections are rarely pre sented, some of the cross-strata (termed low-angle indistinct stratification on logs) appears similar to scour-fill bedding described from deposition of pebbly sand sheets under high-discharge , shallow flow conditions (cf. Smith , 1986, 1987, 1988) . Sand stone sheets with this type of cross-strata are extremely poorly sorted, rarely fine upwards, and set size is typically less than 10 em (cf. Fig. 7, top of log). These features suggest that an abundant sedi ment supply 'forced' rapid aggradation of low relief sinuous-crested dunes on the river bed. The thin sandstone beds without cross-stratifi cation contain sedimentary structures indicating high-velocity shallow-water sheetflood conditions. The massive to stratified facies, for instance, resembles hyperconcentrated flood flow deposits described from Holocene and ancient alluvial sequences of both volcanic (Smith, 1986; Maizels, 1989) and non-volcanic (Simpson & Eriksson, 1989} settings. Beds containing horizontal lamination would represent upper flow regime plane-bed con ditions. Following Rust & Gibling (199Gb) the beds with undulating lamination can be interpreted as vertically aggrading antidunes that may have formed in 10 -15 em of water flowing at 1 m s- 1 . These 4
169
deposits could occur on the floodplain or within channel belts at times of falling stage when water depths were locally very shallow. Facies of modern , shallow, sand-bed braided rivers differ considerably according to discharge and dis charge variability, depth of flood waters, and grain size of sediment transported . Large perennial , comparatively sluggish rivers with a finer sediment load are usually characterized by migration of large lobate linguoid bars , which deposit superimposed suites of planar-tabular cross-bedding. Examples include the Platte River of Nebraska (Crowley, 198 3) or outwash fans in Iceland (Boothroyd & Nummedal, 1978) . In contrast, with smaller, ephemeral, flood-prone streams with considerable short-term variability in discharge there is either insufficient time or insufficient depths of water for large linguoid bars to develop. This can be again illustrated by outwash fans in Iceland. Catastrophic floods caused by subglacial volcanic eruptions (jokulhlaups) have discharges many orders of magnitude larger than the normal summer discharges (Maizels, 1989) . Sandy deposits that result are structureless, horizontally laminated, or trough cross-stratified and lack stacked sets of planar cross stratified sand formed during normal discharges. Sinuous-crested megaripples (St) were also the most common bedform in channelized stream flood in central Australia (Williams, 197 1 ) . I n summary, I conclude that Tryfan Fach Member 'rivers' varied from a few centimetres to perhaps 2 m in depth, and were virtually unconfined. Sedi mentation proceeded through vertical aggradation beneath shallow ephemeral sheet-floods rather than by incision, lateral migration, and avulsion of more perennial channels. The term 'sheet braide d' (Cotter, 1978) is an apt description of the type of 'river' envisaged for the Tryfan Fach Member. This fluvial style is not common at the present day, except in arid areas where ephemeral runoff forms a network of shallow, interlacing poorly defined channels. Ancient examples that are dominated by trough cross-stratification include the Devonian Peel Sound Formation (Miall & Gibling, 1978) and perhaps distal successions of the Miocene Ellensburg Formation (Smith, 1988). Thinly bedded floodbasin sandstones
A finer grained interval of interbedded sandstones and mudstones gradationally overlies the cross stratified rhyolitic sandstones sheet (Fig. 9). Beds
170
G.J. Orton
Facies interpretation ��Rhizocorallium
. �:
�I " v
-..---1
""I "
...,...
I
::·:'/}';;-.':'·:."?:·'�?1i 1r
!;:.�:.:.:.t.r:-:.�"..':.Y-�·;);;.�
-------
V-
-··- ··
V
:z-x7)
�
��'�··
Massive-sandstone bedded siwilnlttteshrtloaccranemineandattioenarydsandstlapionel i
nated tsiLamiulfftsatceous one Bioturvbeatsandst ed ones massi 15cm Small scale tstrough ratificcratoiossn lHoramiiznontatioanl Horlamiiznontatioanl and bioturbation 10cm
Skolithos
I --'lJ0==-1 lr ....:. �'
1mIM Si Sa
one wisiBedded tltshtonesandst lenticles
___ _
1
G
are sheet-like on outcrop scale with even thin ( 10 em) beds traceable for 100 m. Sandstone units vary sub stantially in composition and thickness. Very coarse sandstones (beds 20-50 cm thick) are either massive (Fig. lOA) or contain small-scale (lO cm) trough cross-stratification, which yields palaeocurrents toward the southeast. Coarse-grained lithologies sometimes load into underlying mudstones (Fig. lOA). Climbing ripple cross-lamination and horizon tal lamination are the domjnant lithofacies within the finer grained and thinner ( 1 0- 15 em) sandstone
Debris flow Reworashfalkedl Quisuspensi et-watoenr sedimentation Minorchannel fluviasl Sheetshalfllooodsw watinteor Fluvial channel Amalsheetgamatfloodsed
Fig. 9. Representative section from
floodbasin facies association. From section C, just south of wire fence on east slopes of Tryfan, SH 6696 5960. See Fig. 3 for stratigraphical position of log and Fig. 2 for location of section.
beds. The intervening siliceous siltstones display horizontal or undulating lamination and small wave ripples (Fig. lOA). Lamination in fine-grained beds often contains water-escape structures or is con voluted. Subhorizontal burrows in the top of the succession (Fig. lOB ) (Cruziana ichnofacies: Rhizocorallium) indicate a low-energy depositional environment, but cannot be used to infer water depths or salinity because Rhizocorallium has been described from coastal lagoon and freshwater set tings (Fiirsich & Mayr, 1 98 1 ) .
Facies models in volcanic terrains
171
Fig. 10. Facies o f floodbasin facies
association. Photograph is located 11m above base of Fig. 9. (A) Climbing? wave ripples, undulose and horizontal lainination within siliceous siltstones. Deposits lie about 85 em stratigraphically above position where photograph B was taken. Massive beds (M) may be bioturbated. Coarse-grained crystal-lithic sandstones at the top of the photograph represent the base of a thin (50em) debris flow deposit. Lens cap is 6 em in diameter. (B) Horizontally laminated tine-grained vitroclastic (i.e. predominantly devitritied glass shards) tine-grained sandstones and siltstones. Notice lack of grading or organization to sedimentary units. Letters denote: R, subhorizontal Rhizocorallium burrows; W, vertically-aggrading wave ripples. Coin is 3.5 em in diameter.
Facies indicate a quiet subaqueous environment. They are termed floodbasin deposits because of their fine grain-size , because there is little evidence (e.g. desiccation cracks, palaeosols) for prolonged subaerial exposure, and because they appear to have accumulated simultaneously across a basin several kilometres wide. Rare occurrences of accretionary lapilli within siltstones provide evidence for some penecontemporaneous rhyolitic volcanicity. The evenly laminated siltstones indicate episodic sedi mentation out of suspension. Bioturbation and wave rippling (Fig. lOB) within siltstone units indicate that they represent an amalgamation of several smaller events. These quiet-water conditions were period-
ically interrupted by input of coarser sediment. Fine-grained horizontally laminated or rippled sand stones, and coarse-grained massive sandstones probably represent subaqueous high concentration sheet flows (cf. Dam & Andreasan, 1990). The thin coarse-grained sandstone beds with trough cross stratification are not unlike beds from the previously described braided-stream association, which were attributed to waning alluvial floods and are inferred to represent a similar phenomena. Whether the floodwaters developed on an alluvial plain cr whether beds represent the subaqueous extension of the flood events is a moot point. Plausible settings for the floodbasin association
172
G.J. Orton
include: (i) an interdistributary bay, (ii) lagoons along the delta front behind barrier islands, (iii) lakes on a floodplain adjacent to an active alluvial tract and (iv) a transgressed (or flooded) low-gradient alluvial plain or braidplain. The first environmental setting is dscounted because there is no evidence (deeply incised channels, levees) for the distributary channels. If the fine-grained succession represented lagoon deposits onlapping a braidplain during a marine transgression then one would expect: (i) some evidence for landward-direct currents (e.g. from storm wash overs), and (ii) the lagoon deposits to be overlain by sandstones representing the protective barrier islands/ bars. Yet all palaeocurrents are seaward directed (Fig. 5) and the floodbasin association is overlain by a black laterally extensive mudstone interpreted as offshore marine (Orton, 1990) . A lacustrine origin, the third alternative, is more difficult to fully disprove. Lakes can develop on poorly drained areas of an alluvial floodplain and receive sediment from nearby river tracts. Quiet water suspension deposition of mud and silt occurs during normal river discharges , with sheets of coarser sediment introduced during large river floods, when crevasse splays form. The grain size of sediment within crevasse splays reflects that carried by the parental river tract, and can include coarse sands (e.g. Pollard et al. , 1982). Palaeocurrent directions from crevasse-splay deposits are usually at a high angle to the associated river deposits, and crevasse splay sandstones often display coarsening or thickening upward cycles according to position with respect to the active river tract (e.g. Galloway, 198 1 ) . If the floodbasin association is considered in relation to the underlying rhyolitic fluvial succession , the above lacustrine alternative seems improbable. The most important observations are: (i) the contact between the fluvial and floodbasin succession is gradational (over about 2 - 3 m), (ii) the lateral extent (perpendicular to palaeoslope) for the braided stream and floodbasin associations is the same, (iii) palaeocurrent directions from the floodbasin association are consistent with those obtained from the underlying fluvial association, and (iv) little sys tematic vertical change in facies or grain size could be discerned within the floodbasin sections. I therefore conclude that the floodbasin associ ation represents a submerged low-gradient alluvial plain. During most of the time, quiet-water sus pension sedimentation of silt and mud occurred. However, when flash floods and sediment gravity
flows figured prominently on a nearby alluvial plain, the alluvial input was denser than basinal waters (hyperpycnal flow) allowing flood waters to pass unchecked into the ?marine realm. Mixing between floodwaters and ambient waters appears to have been extremely limited (cf. Dam & Andreasan, 1990) and the subaqueous flood deposits are internally similar to subaerial sand and gravel sheets deposited earlier (braided stream association) on an alluvial plain. The association has many features in common with distal floodbasin/lacustrine deposits described from a rapidly subsiding Devonian strike-slip basin (Hornelen Basin, Norway) (Steel & Aasheim, 1978) . Similarities include the thickness of depositional units, occurrence of horizontally laminated, rippled and trough cross-stratified sandstones inter stratified with mudstones, convoluted lamination, and absence of obvious intermediate (2-20 m) or larger scale cyclicity. • Conglomeratic alluvial fan sandstones
Conglomeratic alluvial fan deposits characterize the northeast margin of the Tryfan depocentre and are well exposed between Craig yr Ysfa and Carnedd Llewellyn (Fig. 1 ) . The alluvial succession that is correlated with the Tryfan Fach Member (Fig. 1 1 ) overlies bioturbated siltstones t o very fine-grained rippled and horizontally laminated sandstones thought to reflect deposition in a shallow quiet water environment (Howells & Leveridge , 1980). Alluvial-fan facies rest erosively on these bioturbated siltstones, and a scour at (, :ast 50 m wide and more than 1 .5 m deep has been identified (Fig. 1 1 , 2 1 .5 m). The lowermost alluvial sediments (Fig. 1 1 , 2 1 . 5 3 6 m ) consist o f either (i) poorly sorted medium grained to conglomeratic sandstone with trough cross-stratification or (ii) scours infilled by normally graded , conglomeratic sandstone to fine sandstone. Palaeocurrent directions are all to the west in keeping with those obtained from younger fan successions at Craig yr Ysfa (Orton, 199 1 , fig. 4) . The trough cross-stratified sandstones suggest deposition by braided streams. Compared with the cross-stratified sandstones of the braided stream facies association these sandstones (i) have larger foresets (sets to 70 cm) ; (ii) less commonly display upwards changes in facies, grain size, or set size; and (iii) contain planar cross-stratification (Sp ). These observations imply that rivers or floodwaters were comparatively deep, that channel avulsion and/or flood-stage erosion
173
Facies models in volcanic terrains
60 ·+---.....,
Wave reworking Wave swash Fan abandonment Small braided streams Thin sheetfloods ·J --50 <>.o•,•- •J'ng",\ G ms Disedilutme ent gravity flows ) Sb � Sm �--�- - :.� ¢� -� --·-�:· __G_m_G______ ' :_:_:_:_:,_f_:_::_:_�_t_:_:_____ Proxi malal al l u vi es fi loewsd fan vi s cous debri ��Stream flow Poorlsy flchannel debri ows ized Mid-fan Ext r e me var i a ti o n i n St 10cm flow regime llonw-fil sofinuosimobityle _:�1) �:�4 brai--d--ed channel s----- � ) ------� � � � � 30 Channel izedgradiviltuyteflows _) Sg sedi m ent and shal l o w ( 1 m deep! �_:, st r e amfl o w --:� alfanDilsutalvial Stem30 -.... llonwfil siofnuosi3-4 tmy channel deep �::;:= �::::r:-;'St ��70cm Flinutovilaolwsheetwavefloenerods gy Shoreface §�, 20 lake Sh ' Of shore' Bioturbation M Sl Sand G )-
----=r _J ------
. •.
.·
G
--7.-:�:: :·- : :-: -:-. - :;_���" :� :_
·.
.
_
•
_.A...-. ----;:..-' <:_... J �
::o·,.::-.<;:-:<:1
-�� ----------------j
�--�
�
d:
O o • C UI
___,)
� .. _ ./ ...A: _ ·� - < __ /�/
� · --
__/'-...� .
BROAD
CHANNEL
8HYOLITE
--------·
' /
/
-----/
,1P
------
VOLCANICITY
rs
Fig. 11. Alluvial fan facies association . From section A, between Carnedd Llewellyn and Craig yr Ysfa, SH 6885 6450.
Distances are metres above the base of the Tryfan Formation . Note occurrence of rhyolitic ash-fall tuffs in fine-grained deposits beneath the coarse-grained alluvium .
174
G.J. Orton
of topstratum prevented the formation of fining upward cycles , and that in-channel bars (hence Sp) occasionally developed (see earlier discussions). The normally graded conglomeratic sandstones resemble hyperconcentrated flow deposits (e.g. Smith, 1986; Maizels, 1989). Deposits in the Tryfan Fach Member are either internally structureless or contain poorly developed trough cross-stratification within the finer grained upper portions. The latter relationship indicates an initial rapid fallout of suspended load and deposition by direct suspension sedimentation, followed transitionally by traction sedimentation within the more dilute low-density upper portions (see also Lowe, 1982). The inter calation of such debris flows within finer grained stream flow deposits indicates rapid variation in fluvial discharge conditions on the alluvial fan. The overlying alluvial fan sediments (Fig. 1 1 , 3651 m) are dominated by unstratified conglomeratic sandstone occurring in tabular beds 20 -50 cm thick. They commonly contain subvertical rhyolite pebbles (4 em size) or rip-up clasts of siliceous siltstone (30 em long). The sharp non-erosive bases, occurrence of well-defined units, lack of sorting or grading and nature of fabric are all in accord with deposition by cohesive matrix-supported debris flows. The overall succession can be interpreted in terms of increasing proximity to sediment supply. Although deposits could represent a valley-fill, with axial braided streams and laterally derived debris flows, there is no evidence for this physiography in the sedimentary record. Instead, a simpler model is favoured involving progradation of a small alluvial fan westward into the Tryfan depocentre . The main tenance of deep channels in a distal fan setting is not common because channel bifurcation usually results in flow expansion, progressive shallowing, and eventually unchannelized flow. It implies both an abundant supply of water and confinement of flow by channel banks. The uppermost alluvial sediments of the Tryfan Fach Member (Fig. 1 1 , 5 1 -60 m) are better sorted and finer grained. Trough cross-stratified sandstones yield palaeocurrents in a variety of directions. Thin beds of fine-grained sandstone display planar horizontal to low-angle cross-cutting lamination and contain concentrations of magnetitic heavy minerals. Together these facies suggest reworking of the fan surface , perhaps by waves, following its abandonment.
FACIES ORGANIZATION AND DISTRIBUTION
Facies relations in the Tryfan anticline
The first three facies associations are best displayed on the northwest limb of the Tryfan anticline, where four sections, spaced about 400 m apart have been measured (Figs 2, 5 & 12) . These can be correlated with a section (F) on the opposite limb of the anti cline, about 3 km to the southeast (after unfolding) . Along the northwest limb of the anticline , the marine shelf association thickens northeastward from 30 to 70 m. Although vitric siltstone beds are thicker and have less admixed epiclastic material in the southwest , sandstones are coarsest and most prevalent in central sections. On the southeast limb (section F) , the total succession is muddier. How ever, a 1 m thick, medium-coarse-grained sandstone containing moderate-scale trough cross-stratification (25-cm sets) lies 1 8 m beneath the top of the marine mudstone. Although reliable palaeocurrents could not be obtained, a fluvial origin seems unlikely given its stratigraphical position. As there is no evidence for tidal processes during deposition of the Tryfan Formation , this lends further credence to the idea that the marine shelf association represent a low gradient platform on which wave energy decreased shoreward. The relative thicknesses of braided stream and floodbasin components (Fig. 9) indicate that the most persistent axis of coarse fluvial deposition was along the southwest side of the braidplain. Fluvial discharge and attendant sediment/water ratios also appear to have been more variable along this side. For instance, the thickness of cross-stratified sand stone sheets implies that most 'channels' were quite shallow; thin debris flows and sheetfloods were an important part of aggradation on the alluvial plain. In contrast, to the northeast, successions indicate infill of several deeper (3 -4 m) 'channels' and sheet flood deposits are absent. Despite good exposure (> 70% outcrop) more detailed stratigraphical cor relation was not possible. To the southeast (section F) , the cross-stratified braidplain sandstone sheet can be correlated with an erosively based , 2.5 m thick, fining upwards suc cession of trough cross-stratified medium- to fine grained rhyolitic sandstone (Figs 12 & 13). Fine siltstone occurs as thin discontinuous lenses between individual sets, indicating numerous slack-water periods during aggradation. The cross-stratified
. . ... - · ·····
;:;;;-
� "' �50
�
0 >:>.. "'
m
<:;;:; · � 0
"' ;:: ;:; · ;;:-
.,
'0
�
FACIES
ASSOCIATION
�
'V vv ·v FLOODBASIN v v ·v vv ·vv v
BRAIDED STREAM
Ill Eg�·jj
MARINE SHELF
ALLUVIAL FAN
,; :
�'Vv V
vv vv 'VV 'vv
�
;:; · "'
l KM
c
�
60
.···
. o
,, ::) :'7 E (,../ ,
._.•
F
,_.,,,./ ; ; : _.....�. .
.
.,
... . · ' r:::;J CAPEL CURIG . c::J VOLCANIC FM.
6
,..,.., GWERN GOF TUFF 70
Fig. 12. Distribution of facies associations in Tryfan anticline. Inset shows location of sections described in this paper. Horizontally ruled vertical bars denote location of detailed sections shown in Figs 4, 7 , 9 & 13.
...... -..) Ul
G.J. Orton
176
20
Interbeddedandbiositurltsbtatoneed sandstone
Transgrn teossiofvfsehore floodbasi
Trmedioughumcrtooss-finse-trgartaifiineedd sandstone mudstone iwinterthbeds Thinly bedded sandstone
Fluvial flood currents
15
?
Shoreface
Fossiliferous sandstone Mudstone
Offshore Fig. 13. Section through upper
M Si Sand sandstones overlie a 2-m-thick sharp-based suc cession of massive (?bioturbated) and horizontally bedded sandstones, and are overlain by interbedded fine -coarse-grained sandstones, siliceous siltstones, and interlaminated sandstones and siltstones. Overall the succession mimics the facies suc cessions seen on the northwest side of the anticline and interpretation of section F (Fig. 13) is made
portion of Tryfan Fach Member, southeast limb of Tryfan anticline. See Fig. 12 for location of section .
against the background of these sections . Two interpretations warrant further comment. The cross-stratified rhyolitic sandstones are similarly interpreted as representing distal fluvial currents . If the cross-sets were wave produced , their size suggests a high nearshore wave energy, yet wave power was low enough that mudstone could be deposited during interludes between input of coarser
Facies models in volcanic terrains
sediment. This does not negate the possibility, however, that some or all of the cross-stratified succession may have been deposited below sea-level. The second important difference is that the cross stratified sandstones lie above fine-grained sand stones, and rest erosively on them. On both sides of the anticline, the floodbasin association gradationally overlies the fluvial deposits, indicating a reduction in the amount of coarse sedi ment deposited on the braidplain. Successive sand stone beds, however, do not become finer grained ur-section . They merely become thinner or less common (i.e. more widely separated in vertical section). Bounding surface between the braided stream and marine shelf associations
One fundamental question concerns the position of the marine shelf association with respect to a shore line. At least two interpretations are possible, largely because of uncertainty concerning the amount and source of the sand along the coastline. Neither model is entirely satisfactory. The first alternative treats the sandy portion of the marine shelf association as part of a shoreface. The mudstone-dominated interval that caps the as sociation may represent the shoreline itself. Walker and Harms ( 1975) and Hamblin ( 1992) have de scribed a tideless muddy coastal deposit with some similarities. However, in the marine shelf association no indicators (e.g. change in colour or reddening of mudstones, desiccation cracks, palaeosols, beach/ foreshore deposits) for subaerial or near subaerial conditions were observed, and even the uppermost deposits appear to be fully marine. The second alternative considers the sandy succession to be a small and offshore component of an even muddier coastline. On such coastlines bed thicknesses and grain sizes first increase than decrease away from the shoreline. This pattern arises because of in situ resuspension of ambient muddy sediment with fines transported landward combined with offshore transport of coarser sediment during storms (cf. Hill & Nadeau, 1989; Myrow, 1992) . In this scenario , the uppermost mudstone-dominated interval of the marine shelf association would correspond to both muddy shoreline and nearshore subtidal deposits. Yet it contains no evidence (flaser and wavy bedding, current ripples, gutter casts) for a nearshore setting. In both alternatives the transition between the two associations appears too abrupt, with expected
177
facies successions either missing or thinner than expected. Even if the braided streams prograded into a low-wave-energy basin one would expect more widespread shoreface or delta-front deposits between the alluvial fan/braidplain and interfan/ offshore mudstones (e.g. Vos & Eriksson, 1977; Fernandez et al. , 1988; Bardaji et al. , 1990). Consequently, the abrupt juxtaposition of 'fluvial' and 'marine' along the northwest limb of the Tryfan anticline, as well as the scoured base to the more distal ?fluvial sandstones suggests a relative sea-level fall during sedimentation, with possible removal of upper shoreface/nearshore deposits by erosion. 'Uplift' appears to be greatest in the northwest, where braidplain deposits rest with sharp contact on marine mudstones. The succession to the southeast (Fig. 13) suggests more continuous aggradation, although the condensed nature and sharp base to probable shoreface sandstones indicates that relative sea-level was still falling ( cf. Plint, 1988). A similar abrupt alternation between braid plain and shallow water environments has been described from other strike-slip basins, including the Pennslyvanian Boss Point Formation (Browne & Plint, 1994) and English Westphalian B strata (Haszeldine & Anderton, 1980) . In the Boss Point, braidplain packages rest erosively on, and locally cut deep ( 15 m or more) scours into underlying lacustrine mudstones. A reason for the sharp contact was not given , although uplift due to episodic tectonic source area rejuvenation seems likely. Extent of the deposit
It is pertinent here to dispel the notion that the deposits described represent a small or distant component of a much larger volcaniclastic apron. In the first place, deposition of the Tryfan Fach Member occurred only to the northeast of Llanberis Pass (Fig. 1 ) . The Snowdon region was uplifted and subaerial throughout deposition of the Tryfan Formation (Orton , 1991 ) and contemporaneous marine mudstones further to the southwest belong to a separate depocentre. Although contempor aneous mudstone-dominated successions occur 10 km to the southeast, distal 'feather-edge' sand stones from the Tryfan Fach depositional system have not been recognized. In light of the rate at which the braided stream sandstones pinch out across the Tryfan anticline, correlation is not suspected. Less than 5 km to the northwest , adjacent to the Nant Peris fault (Fig. 1 ) , all Ordovician successions
178
G.J. Orton
are stratigraphically condensed (Trythall et a!. , 1987 ; Orton, 1990). The entire Tryfan Formation is about 50 m thick, compared with more than 700 m in the Tryfan anticline. It consists of mudstone or fine grained sandstone with no sandstones being similar in composition to the rhyolitic sandstones of the Tryfan Fach Member. These must correlate with a zone of non-deposition and/or an unconformity near Nant Peris. Ordovician successions also occur an additional 30 km to the northwest, on Anglesey (Fig. 1 , inset) . Although substantial Precambrian strike-slip move ment occurred along the narrow stretch of water (Menai Straits) that separates Anglesey from the Welsh mainland (Gibbons, 1983) , lithostratigra phical correlation of Cambrian rocks across the Menai Straits indicates little or no movement of Anglesey relative to the Welsh mainland since the Early Cambrian (Reedman et a!. , 1984). Ordovician successions on Anglesey consist of deep-marine muds and/or mass flows deposited in smaLl fault bounded basins (Bates, 1972). These include volcanogenic massive sulphide deposits hosted in rhyolitic volcanics and pillow basalts of probable Caradocian age (Pointon & lxer, 1980; Westhead, 1991) . Thus, although there might be some Cardocian volcanism on Anglesey, there was certainly no upstanding subaerial volcanic arc that could have sourced the Tryfan Fach Member. For the above reasons, it is concluded that the depositional system responsible for the Tryfan Fach Member affected areas within only a few kilometres of Tryfan Fach itself, and that the deposits of the Tryfan anticline represents a substantial portion of the area traversed by the system.
EV IDENCE FOR VOLCANISM
Characteristic petrographic features indicative of explosive volcanism include glass shards, pumice and scoria, bipyramidal �-quartz, and euhedral zoned feldspars. To those not familiar with volcani clastic deposits, it may seem an easy task to establish when volcanism occurred. However, volcanic clasts differ widely in size, shape, density and durability, with mean settling velocities varying by several orders of magnitude. During reworking, hydraulic processes can quickly remove most of the less dense or smaller ash and pumice fraction. Volcaniclastic deposits are also more susceptible to diagenesis than any other type of sand because of the chemi-
cal instability and reactiVIty of their framework grains. Glass suffers particularly, with chalcedony, opal, zeolite and clay minerals being alteration by-products (e.g. Davies et al. , 1979; Surdam &: Boles, 1979) . The end result of sorting and alteration processes in distal exposures is usually a crystal- lithic sand- stone, with framework grains set in a fine-grained, variably altered and/or recrystallized, non-descript matrix (e.g. Runkel, 1990; Cather & Folk , 1991) . Thus many features, including the geometry of the deposit, its sedimentary facies, the variability of its composition, shape and sorting parameters , and the extent that sediment composition differs from the 'norm' for sandstones within the basin, are needed to establish penecontemporaneous pyro·· clastic volcanism. Petrography
Alluvial fan deposits
Alluvial fan deposits contain significant mono crystalline quartz (24% ) , little feldspar (8%) and lithic fragments dominated by aphanitic or sparsely porphyritic quartz-bearing rhyolite (20%) , siltstone (26% ) , or granitic (3% ) lithic fragments (Table 1) . Composition indicates derivation from a mixed granite , tuffaceous sediment, and rhyolite volcanic source. Although volcanics still comprise about half of the lithic population, lithics are thought to be derived from older rhyolites of the Llewellyn Volcanic Group. Braided stream and floodbasin deposits
Sandstones contain large (to 4 mm) euhedral sani·· dine (25%) and smaller plagioclase crystals (17% ) . The percentage of monocrystalline quartz (5% ) is low compared with the 'norm' for the Tryfan Formation (Table 1) . Lithic fragments (to 1 em) are extremely angular and dominated by rhyolitic volcanics. These are mostly porphyritic two-feldspar felsic volcanics (32%) with phenocrysts of albite oligoclase and simply twinned alkali feldspar set in a recrystallized quartzo-feldpathic matrix that some·· times displays a trachytic texture . Quartz pheno·· crysts are rare and deeply embayed when they do occur. Other lithic clasts include aplite, ?rhyolitic tuff, or perlitically fractured glass (8% ) . Clasts are set in a quartzo-feldspathic matrix ( c. 20% of totall rock) . The texture of the matrix and lithic fragments
179
Facies models in volcanic terrains Table 1.
Composition of sandstones Tryfan Fach Member Tryfan Fach braidplain (n = 7) Mean
Monocrystalline quartz Polycrystalline quartz Chert K-feldspar Plagioclase Rhyolite Rhyolitic tuff and perlite Aplite Mafic volcanics Siltstone Granite Opaques Other heavy minerals
4.8 2.6 1.6 25.3 17.5 31.9 8.5 1.1 2.4 1.9 0.2 1.3 0.4 0.43 0. 94 0. 14
PF ratio LvL ratio QF ratio
SD
Mean
SD
2.9 3.9 1.7 15.2 9.8 14.5 8.2 1.3 3.7 2.5 0.4 2.0 1.0
23.4 10.6 11.2 5.3 2.3 10.7 Absent Absent 3.8 26.0 2.7 0.5 0.2
9.8 5.2 6.4 3.2 1.2 6.0
0. 19 0. 09 0. 1 1
9 : 43 : 47 5 : 43 : 52 5 : 89 : 8 6 : 83 : 1 1
QFR QmFLt QpLvLs LvbLvrLsp
Craig yr Ysfa alluvial fan ( n = 5)
0.33 0.51 3.3 46 : 8 : 46 24: 8 : 68 17 : 44: 39 6 : 36 : 58
2.5 12.5 3.6 0.6 0.3 0.12 0. 13 0.7
Tryfan Formation (n = 101) Mean 32.6 6. 1 7.8 14.0 6.5 7.2 1.8 1.2 1.7 14.8 0. 9 2.9 0.4 0.36 0.58 2.4
SD 15.6 5.4 7.2 1 1.4 6.5 12.3 5.8 1.5 3.9 14.8 1.6 5.8 0. 9 0.25 0.30 1.9
41 : 21 : 38 34: 21 : 45 16 : 51 : 33 4: 46 : 50
PF, plagioclase/total feldspar ratio ; LvL, proportion of volcanic lithics within lithic population; QF ratio, quartz/total feldspar ratio. Q, total quartzose grains; F, total feldspar grains; R, unstable lithic fragments; Qm, monocrystalline quartz; Lt, total lithics R + polycrystalline quartz; Qp, polycrystalline quartz; Lv, volcanic lithic fragments; Ls, sedimentary lithic fragments; Lvb, mafic volcanics; Lvr, rhyolitic volcanics; Lsp, sedimentary and plutonic lithics. =
is similar and it is often difficult to tell the two apart. Although geochemical data are not available, the crystal population (both free and within lithics) suggests derivation from a rhyodacite to rhyolite magma. The abrupt appearance of a monomict sheet consisting of feldspar crystals and angular volcanic lithics within an environment previously receiving mudstone suggests a penecontemporaneous volcanic eruption. Unequivocal evidence for explosive pyro clastic volcanicity is provided by the vitric siltstones containing accretionary lapilli. Although the nature of the eruption is hard to pin down, the high pro portion of non-vesicular lithics and crystals, low abundance of vitric-material, absence of pumice, and presence of accretionary lapilli suggest small scale phreatomagmatic eruption(s) (cf. Sheridan & Updike, 1975; Heiken & Wohletz, 199 1 ) . The rhyolitic deposits have a combined volume of at least 0.2 km3. As crystal content in volcanic rocks
rarely exceeds 50% (Ewart, 1979), the crystal content of sandstones must have been enriched, either by fractionation processes accompanying the eruption and deposition of the tephra and/or hydraulic sorting during fluvial reworking. The erupted volume of magma would therefore be much larger than the volume of deposits in the Tryfan Fach region, perhaps double. Although still small, the volume is still well in excess of the amount of pyroclastics (c. 0. 15 km 3) extruded in the 1980 eruption of Mount St Helens (Rowley et al. , 198 1 ) . Pre-eruptive marine mudstone association
The fine-grained sandstones intercalated with the marine mudstones are moderately sorted quartzo feldspathic and lithic arenites. No samples were point-counted owing to the fine grain size and dif ficulty of identifying components, but composition appears similar to the 'norm' for the Tryfan For-
180
G.J. Orton
mation (Table 1 ) . The anhedral nature of framework grains, the high percentages of chloritic matrix and the dominance of sedimentary lithics suggests derivation from a source of tuffaceous sedimentary rocks and mudstone. Although rhyolitic ash-fall deposits occur near the base of the succession (Fig. 3) there is no evidence elsewhere for pene contemporaneous volcanicity, with the exception of the vitric siltstones at the very top of the succession. Possible position of the volcanic source
In many cases volcaniclastic successions can be interpreted in terms of proximity to a central volcano (e.g. Smith, 1988) by comparison with well-studied modern examples. In the Tryfan Fach Member, the absence of primary volcanic deposits (larvas, pyro clastic flow deposits) , coarse breccias and debris flows is consistent with deposition in a distal setting. Similar monolithological syneruptive sands occur on the most distal ring plains that flank major central volcanoes at distances of 30- 100 km from volcanic vents. I begin by discounting these distances. Within the Llewelyn and Snowdon Volcanic Group, the absence of systematic facies variation across the entire out crop belt, as well as the common identification of intrabasinal vents, argues against derivation of volcaniclastic material from a line of volcanoes beyond the reach of the present outcrop. Instead the stratigraphy reflects a rather haphazard assemblage of small volcanoes, none of which had appreciable subaerial relief, erupting into and within subsiding sedimentary lowlands (Howells et al. , 1 99 1 ) . Although the Welsh basin i s typically portrayed a s a back-arc basin, this is not essential, and such low lying volcanic fields can also occur due to intra-arc extension (e.g. White & Robinson, 1992). In such settings, 'distal' or more accurately 'low-gradient' deposits (e.g. Tryfan Fach Member) are often locally interstratified with more 'proximal' facies (e.g. lavas, pyroclastics). In North Wales, most large Ordovician volcanic eruptions had vents within a few tens of kilometres, or less, of the sites of deposition of their primary products (excluding fallout tuffs) (Kokelaar, 1988, 1992; Howells et al. , 199 1 ) . It is not necessary to suppose that smaller eruptions had more distant volcanic sources. The absence of a correlative coarse rhyolitic unit amongst more proximal successions about 5 km away (see above 'Extent of the deposit' section) also lends support for a local source for the Tryfan Fach Member.
This remains to be located . There are several subvolcanic intrusions in the Llewelyn Volcanic Group for which comagmatic extrusive correlatives have not been recognized (Fig. 1) and which invite comparison. These lie about 5 km or more (after unfolding strata) to the west of Tryfan Fach along the northeast trending Nant Peris fault. The largest intrusion (Bwlch y Cwyion) consists of microgranite (rhyodacite/dacite) with later marginal intrusions of rhyolite (Howells et a!. , 1991) whereas smaller intrusions are rhyodacites , granodiorites and feld spar porphyries.
REASSEMBLING THE STRATIGRAPH ICAL P IECES
In the Tryfan Fach Member, each facies association is in some way different, in composition and/or facies, to adjoining facies associations. It will be shown that the meaning one attaches to facies changes between and within associations depends on whether the rhyodacitic sandstones are attributed to penecontemporaneous explosive volcanism. One epiclastic system: time's cycle
The simplest solution, and the one that would be adopted if the petrography of sandstones and composition of rhyolite clasts was unknown, is to integrate the four facies associations into one linked depositional system or systems tract (Fig. 14) . As noted by Hutton ( 1788) it is with 'pleasure that [we] observe order and regularity in the works of Nature, instead of being disgusted with disorder and confusion'. In this interpretation , a tributary alluvial fan at Craig Yr Ysfa would feed sediment on to an axial braidplain confined to the southwest by additional uplift along Llanberis Pass. Younger alluvial systems in the Tryfan Formation exhibit this drainage pattern (Orton, 199 1 ) . The sandier upwards marine sheU association would represent offshore to coastal plain environments reflecting deposition in front of, adjacent to , or between alluvial fan salients that were deprived of coarser sediment. Tectonic uplift along the basin margin resulted in progradation of the alluvial fan and a derivative braidplain into the marine basin. Lesser amounts of uplift within the basin resulted in a relative sea-level fall and braid plain sandstones rest with sharp contact on a condensed nearshore succession . Following tectonic
SEDIMENT COMPOSITION
FACIES
PALAEOGEOGRAPHY
ENVIRONMENT
RELATIVE SEA-LEVEL
LACUSTRINE
BRAIDPLAIN
� "' �·
;:!
Q
�
1::;;: · -.: Q
;::;-
., ;:s
r; · �
� �-
Gwern Gof Tuff
OUTER 51-ELF
l
vvvv vvvv vvvv
pyroclastic
COM'OSITION
[I] '•'• •'• ••••• • • • ..-·.·.·
reworked pyroclastic
epiclastic
+ Low + _ High V+ + r· I
Fig. 14. Epiclastic interpretation of vertical successions. Block diagram (looking northeast) shows palaeogeography if morphological elements are genetically
related. An unchanging marine transgression is assumed throughout. Place names are: TF, Tryfan Fach; CY, Craig yr Ysfa. Facies associations as in Fig.
12.
...... 00 ......
182
G.J. Orton
uplift, a more gradual shift in the position of the aLluvial fan lobe aLlowed floodbasin waters to gradually transgress on to the fan/braidplain surface. Floodbasin siltstones therefore increase in pro portion upwards at the expense of coarse sandstone beds. There are many pieces that do not fit into this puzzle: (i) the composition of braided stream deposits differs from sandstones of the alluvial fan at Craig Yr Ysfa, implying derivation from a differe;1t aLluvial fan or from a more distant axial source; (ii) coarse rhyolitic sand is never present in the marine shelf association - one would expect
occasional high-velocity floods to introduce coarse material on to the shelf; (iii) the thickness and 'purity' trends in reworked ash-fall tuffs at the top of the marine shelf association , as well as grain-size trends across the braidplain (Fig. 5) suggest a source to the southwest rather than to the northeast. Together these features cast doubt on the strict application of Walther's Law to the total sequence and seems to demand three weakly or unlinked depositional systems that developed in advance of, and in conjunction with penecontemporaneous volcanism (Fig. 15).
· :-. . · · · · :·.. ·
SEDIMENT COMPOSITION
FACIES
ENVIRONMENT •
·· · �
. :;, , J:�· ·[?i;��t��)· �t:�l;-
ALLUVIAL PLAIN
BRAIDED SAND·Sf£ET
NNER SHELF
OUTER SH::LF
Fig. 15. Pyroclastic (or syneruptive) interpretation of the vertical succession. Block diagrams (also looking northeast) depict palaeogeography if Walther's Law cannot be applied to the full succession and morphological elements need to be considered separately. Two types of 'reworked pyroclastic materials' are portrayed on 'sediment composition' column in order to illustrate uncertainty about the number of volcanic events. Darker tone indicates pyroclastic materials that have been reworked more extensively (i.e. they were deposited a greater length of time after the volcanic eruption). Several (A, B and C) possible interpretations of relative sea-level change are presented for the syneruptive period. See text for discussion. See Fig. 14 for key to sediment composition.
Facies models in volcanic terrains Three unrelated depositional systems: time's arrow
Progradation of a muddy shelf and postulated volcano-tectonic uplift
An initial supply of fine-grained silts and mudstones (largely non-volcanic) was delivered to a low-wave energy basin, probably by small channels or sheet floods (cf. Wells, 1983). Sand distribution suggests that the source lay to the northwest to west (Fig. 1 5 , bottom block diagram). Reworked rhyolitic ash-fall tuffs were derived from a similar direction and may represent small eruptions precursor to the main volcanic event. The lateral variation in the amount of sand suggests that sediment input occurred from localized point sources with minimum lateral reworking at the shoreline. Widespread tectonic uplift preceding and associated with volcanicity resulted in a relative sea-level fall. Hence coarse grained sandstones derived from the volcanism rest with a sharp, probably erosional contact on a condensed nearshore succession . Five to 15 m of uplift would be needed to cause erosion of upper shoreface/nearshore deposits. The timing of uplift with respect to volcanism in the Tryfan Fach Member is not thought to be fortuitous and some of the uplift is related to buoyant forces associated with the intrusion and/or migration of magma. All large rhyolitic eruptions in North Wales show evidence for tens of metres of tectonic uplift in the vicinity of volcanic centres in advance of eruptions (Howells et al. , 199 1 ; Orton, 1991) and there is no reason why smaller amounts of uplift might not precede smaller rhyolitic eruptions. A volcano-tectonic control on uplift and/or development of regressive cycles has also been suspected within other ancient sedimen tary, particularly extensional, basins (cf. Leeder, 1983 ; Henry & Price, 1984; Watkins, 1986; Leckie & Cheel, 1989; Sloan & Williams, 199 1 ) . Our understanding of the influence o f volcanism on basin subsidence is limited by the short time span < 100 yr) over which detailed observations from modern volcanoes have been recorded and the difficulty in ancient successions to separate volcano tectonic from tectonic effects. Most modern data is from volcanic fields where large silicic caldera forming eruptions previously occurred (Newhall & Dzurin, 1988). Vertical ground movements are best recorded at calderas located by the sea, where elevation changes, forming land , are very con spicuous. Ground deformation in calderas occurs readily because of the existence of large shallow
183
(3 - 15 km) magma chambers, with rates of uplift ranging from 0.002 to almost 1 m per year (cf. McKee et al. , 1984; Hill et al. , 1985; Luongo & Scandone, 199 1 ) . These rates obviously are extremes or we would not notice them. The unique longer term (2000 yr) historic record from Campi Flegrei caldera (Italy) shows that episodes of unrest are often separated by decades of quiescence (Lyell, 1853, Caputo, 1979; Dvorak & Gasparini, 199 1 ) . Here, 10 centuries o f gradual subsidence ( 1 1 m) was followed by more gradual uplift beginning in the eleventh century. In 1538, a mere 0.03 km 3 of ash and scoria erupted from Monte Nuovo. This was preceded, however, by 40 years of episodic uplift (and land formation) culminating in 7 m of uplift in the two days preceding the eruption at a site (Pozzuoli) several kilometres away. As the Tryfan Fach Member was not deposited within or nearby such a caldera, it is pertinent to query whether the above data is relevant. Extrusive rocks of the lowermost Llewelyn Volcanic Group were ponded in subsiding, probably fault bound, sectors of the sea-floor (Howells et al. , 199 1 ) , the thickest accumulations (both about 1000 m) being rhyobtic lavas and ash-flow deposits of the Conwy Rhyolite and trachyandesite lavas and tuffs of the Foe! Fras Volcanic complex (Fig. 1 ) . As has been noted recently (Lipman, 1984; G.P.L. Walker, 1984) , it is a semantic issue whether the rectangular volcano tectonic depressions that contain these deposits should be termed caldera. The final volcanism of the Llewelyn Volcanic Group, the Capel Curig Volcanic Formation involved eruption of two rhyolitic ash flow tuffs (each > 20 km3) from a postulated vent in the north near Conwy, as well as smaller eruptions (c. 1 km3) from volcanic centres just north of Carnedd Llewelyn and just south of Tryfan (Fig. 1 ) . The continuity o f compositional variation within the Llewelyn Volcanic Group suggests that various intrusive and extrusive products were all originally comagmatic and related by closed-system fractional crystallization (Howells et al. , 199 1 ) . This necessi tates a parental magmatic system that stretched 25 km from the Llanberis Pass fault toward Conwy, and extended about 10 km in a transverse direction. This magma 'chamber' may have been an elongate mid- to subcrustal beam or dyke controlled by the dominant structural grain , a shape documented by geophysics from modern continental rifts (e.g. Cordell et al. , 1985 ) . Petrochemical and exper imental data (Ball & Merriman , 1989) show that rhyolites derived from this magmatic system evolved
184
G.J. Orton
at relatively shallow crustal levels (PH,o < 500 bars) . Thus many of the conditions needed for volcano tectonic 'unrest', such as earlier caldera-forming eruptions, existence of large , shallow magma chambers , volcanism in an extensional graben-type structure, existed within the Snowdon graben at Tryfan Fach time. Aggradation of synerup tive pyroclastics
Sediment composition within braidplain and flood basin settings all indicate deposition in conjunction with a period of explosive rhyodacitic volcanicity (Fig. 15, top block diagram). The sudden increase in sediment grade and abundance resulted in a brief period of aggradation within shallow ephemeral channels. This braidplain prograded rapidly as a sheet across the pre-existing shoreline and/or exhumed marine mudflat into a low-wave-energy basin, with facies suggesting a more direct input of volcanically produced sediment loads on to the southwest portion of the braidplain. It is not certain whether sandstones represent pyroclastic material reworked from a series of small eruptions rather than a single eruption (left- and right-hand sediment composition columns of Fig. 15). Consequently, the thin siliceous mudstone beds within braidplain deposits could reflect: (i) periods of quiet-water sedimentation following lateral migration of the fluvial channels (the most 'traditional' interpretation) , or (ii) fallout of fine grained ash following episodes of rapid syneruptive aggradation of coarse clastic material. Fine ash may have been erupted at the same time as the underlying coarse, clastic material but initially dispersed within an eruption cloud. There are also several interpretations that can be presented for the braidplain to floodbasin transition. 1 The upwards decrease in the proportion of coarse grained sandstones may reflect gradual waning of the coarse sediment supply, perhaps due to erosive lowering of the volcanic edifice and/or near-vent subsidence following the volcanic eruption. In this model, the maximum basinward advance of the syneruptive braidplain is 'picked' as the top of the coarse-grained rhyolitic sandstones. Fine-grained floodbasin siltstones would largely represent matrix of the pyroclastic flows, liberated by fluvial rework ing, and transported to the site of deposition by low gradient streams. In this model, the floodbasin successions are distal and should interfinger else-
where (more proximally, or laterally) with the coarser grained braided stream deposits. If one was trying to determine relative sea-level changes, the top of the braided sandstone sheet might be picked as a marine flooding event and used to infer a rise of relative sea-level (sea-level curve B , Fig. 15). 2 A contrasting scenario results if one allows volcanism to continue and the type of volcanism to change with time. Time lines would essentially parallel bedding contacts rather than cut across them as in the first scenario . In this case, it is assumed that all vitric siltstones are primary, or only slightly reworked, pyroclastic materials. Rather than pro ducing coarser grained pyroclastic material, the volcano produced fine-grained ash. The occurrence of accretionary lapilli suggests that this change may have been triggered by access of water to the vent. Fluvial channels did not become more distant (i.e. distal) from the source up-section, they just became overwhelmed by aggradation of fine-grained primary and reworked fallout ash. Hence, coarse sandstone progressively formed a smaller proportion of the sediment deposited on the alluvial plain. Laminated ashfall siltstones may have been deposited in ponded depressions on the alluvial plain; a marine trans gression may have occurred only at the very top of the succession, after all volcanism had ceased (curve C, Fig. 15). Two observations favour the second interpret ation. Firstly, sandstones at the top of the floodbasin association (Fig. lOA) are as coarse as those lower down. Secondly, lamination in many of the siltstones/ fine-grained sandstone beds is often . diffuse (Fig. lOB) indicating that fine sediment from one event was still being deposited when coarsest grains from the next event arrived . Near-source ash-fall deposits commonly display such weak bedding (e.g. Branney, 1991) owing to fluctuations in the energy of the eruption, direction and strength of winds or over lapping of several eruptive pulses. In contrast, sedimentary laminae in lakes are commonly sharp based (e.g. Hamblin, 1992) because they reflect infrequent episodes (e.g. seasonal varves) of clastic input. A symbiotic alluvial fan ?
The relationship of the alluvial fan deposits to the other two genetic units is most perplexing. One might at first expect that the two stratigraphical units were not contemporaneous. However, as four
185
Facies models in volcanic terrains
coarsening upward alluvial fan sequences can be recognized in the Tryfan Formation at Craig yr Ysaf (Orton, 1988, Fig. 4) , and four braidplain -deltaic successions have been recognized in the Tryfan anti cline it seems reasonable to make a one-to-one correlation. None of the other alluvial fan suc cessions contain feldspar-phyric rhyolitic volcanics that could suggest otherwise . Rhyolitic ash-fall deposits within fine-grained 'shoreface' deposits that underlie the coarse-grained alluvial fan deposits lend further support to this correlation. The evidence suggests that the alluvial fan pro graded at about the same time as volcanism was active, or shortly thereafter, but from the opposite side of the basin. Tectonic changes associated with the volcanism must have produced enough relief between Craig yr Ysfa and the Tryfan depocentre to allow fan progradation. Migration of magma may have been the mechanism driving this subsidence pattern. Magma migration should have two effects: (i) it will produce relative uplift in the immediate vicinity of the volcanic vent; (ii) fault blocks that lose some of their supporting 'cushion' of magma could subside relative to surrounding fault blocks, even those more distant from the volcanic centre. In some ways this process is analogous to the movement of salt, laterally and upwards, toward the position of least overburden pressure. Magma migration is documented most commonly in extensional volcanic settings. It has been invoked in the Rio Grande rift to explain progressive increase in the gradient ( 1 m km- 1 to 7 . 1 m km- 1) of rift drainage systems from late Pliocene to Holocene times (Bachman & Mehnert, 1978) and present-day subsidence that is similar in size but opposite in sign to uplift in the Socorro area of New Mexico further along the rift axis (Reilinger & York, 1979). A more convincing illustration is provided by the 1912 eruption of Novarupta (Alaska) where a depression 600 m deep (volume 6 km3) formed at Mount Katmai 10 km away from the eruption site, implying a complex magmatic plumbing system (Hildreth, 1987). In the basaltic Bedded Pyroclastic Formation (North Wales), episodes of subsidence preceded periods of copious magmatism and rapid sediment accumulation (Kokelaar, 1992); this subsidence pattern could be explained by lateral migration of magma away from deposition sites and toward eruptive vents. Within a Variscan pull-apart basin in Spain, Besley & Collinson, ( 1991) also documented accelerated but more differential subsidence during
active volcanism and concluded that 'volcanic activity was related in a close but subtle way to the subsidence history of the basin, and thus to the sedimentology of its fill'.
THE STRA T IGRA P H ICAL RECORD OF VOLCANISM
Sedimentary sequences associated with active volcanoes are often considered to consist of two elements (G. Smith , 199 1 ; White, 199 1 ) : (i) syn eruptive sequences, which result from primary volcanic deposition and immediate post-eruptive reworking, (ii) inter-eruptive sequences, which record deposition without significant influence of volcanic activity where normal sediment delivery processes are dominant (Fig. 16) . The syneruptive condition is a perturbation brought about by the volcanic eruption and includes not only the time scale of the eruption but also the period following cessation of volcanism while hydrological and hill slope conditions remain disturbed. In humid cli mates, and without vegetation, the syneruptive condition may continue for a long time, at least one or two decades, until most loose debris from the eruption is removed completely. In the Tryfan Fach Member it includes the braided stream and flood basin facies associations. Syneruption sedimentation is in large part aggradational and the resultant stra tigraphy is an event stratigraphy with time planes essentially parallel to bedding surfaces. Around stratovolcanoes
Most conceptual models for subaerial volcaniclastic alluvium are based largely on sedimentation within volcaniclastic aprons deposited adjacent to relatively high-standing volcanic arcs (e.g. van Houten 1976; Vessell & Davies, 198 1 ; Smith, 1987, 1988). During volcanic eruptions the instantaneous production of large volumes of sediment, combined with enhanced and more variable runoff, result in sedimentation principally by aggradation of high-sediment-load flood and debris-flow processes. Inter-eruption periods are often characterized by erosion and channel incision; incision (Vessell & Davies, 1981) following syneruptive episodes was an integral aspect of the predictive facies model presented by G . A . Smith ( 1991) for continental volcaniclastic sediments. Inter-eruption streams draining the high-standing
186
G.J. Orton
volcanic highlands are commonly of gravel-bedload character. The overall geometry of the succession that results is controlled by the relative importance of syneruptive versus inter-eruptive conditions, which in turn are controlled by eruptive frequency and rates of tectonic subsidence (G . A . Smith, 199 1 ; Fig. 16a). Within a humid low-relief continental rift
The Tryfan Fach Member succession differs from the above scenario in three ways: (i) debris flow deposits are largely absent from the syneruptive sedimentary record; (although debris flows occur in a penecontemporaneous basin-margin alluvial fan, this occurrence was unrelated to high sediment loads due to volcanism); (ii) incision, if anywhere, occurred at the base of syneruption deposits; and (iii) fine grained muds, rather than gravels, were deposited after or between volcanic events. These differences are thought to arise largely because the Tryfan Fach sedimentary system developed within a humid region, actively subsiding, continental rift. Syneruptive facies and debris flows
Three reasons are given for the paucity of debris flows.
A: STRATOVOLCANOES
1 Slope was insufficient. Rhyolitic volcanic fields typically have a subdued topography. In North Wales, the stratigraphical architecture and the dis tribution of products from volcanic cones suggest that maximum relief was a few hundred metres (Howells et al. , 199 1 ; Orton, 199 1 ) . Walton ( 1986) also argued that the paucity of debris-flow deposits in the Tertiary rhyolitic volcaniclastic sediments of West Texas resulted from the low-gradient of source areas. At the Taupo volcanic centre , although debris flows occur in post-eruptive reworked pyroclastic deposits (R. Smith, 199 1 ) , their development depended on the moderate slopes ( 10°-20°) of the inward dipping pre-existing topography surrounding Lake Taupo. This topography is a tectonically uplifted ridge of Upper Palaeozoic- Mesozoic grey wackes that has nothing to do with the volcano itself (Cole , 1984). 2 Depositional site was too distal. Cross-stratified sands, not unlike those of the braided stream associ ation, occur within most distal volcaniclastic suc cessions. As volcanoes range from a few tens of metres (e.g. tuff rings) to several kilometres (e.g. some stratovolcanoes) in height, the term 'distal' has no singular meaning in terms of distance from source , with 'distal' facies developing as close as a few kilometres to more than 100 km from the volcano.
8: LOW RELIEF HUMID RIFT · ·:�_ :·:::.:s:�-:;-�-: :·: :: :};����.-: _-.=�-:: �--�_-��·:;-::-�� ��� -� .� �::.:.C.. ::.�.l.�:�;�<��-:., ;,:;.).�:� i..�:�.-:��:<; - -��- -0�_-.· .·..:; � · ·.".: ;.:� · - ·-�·:.-:·:· ::: . .·o . •:a. ,..,uo "-o.-��-o.an.: .•:. -·
f:t�: .��:�i�?:�\:�-?�_�.� �??��
,. -.
.. .
..�-
.-c
••
primary pyroclastics
SYN ERUPTIVE
debris flow hyperconcentrated flows braided streams
Fig. 16. The stratigraphical record of volcanicity: (A) around stratovolcanoes. Pre-eruptive and post-eruptive successions are not separated (from G.A. Smith, 1991). (B) Fluvial-deltaic successions of the Tryfan Formation, North Wales. Lower example would represent Tryfan Fach Member.
187
Facies models in volcanic terrains 3 Pyroclastic materials were reworked from a series of small eruptions rather than a single eruption. Although most small monogenetic volcanoes are short-lived , multiple soil horizons described from the 'flanks' of some volcanoes (Wells et al. , 1990) , direct observation of volcanic activity (Foshag & Gonzalez, 1956) and ancient examples (e.g. Kokelaar, 1992) indicate repeated eruptions over a long time period (decades to hundreds of thousands of years) with reuse of the same magma plumbing system. Due to the small size of individual eruptions, and time between eruptions, primary products (pyroclastic flows, debris flows) could be completely reworked into cross-stratified sandstones.
Bounding surface between syn- and post-eruptive deposits
All sedimentary basins are the result of isostatic adjustment of the lithosphere to a tectonic driving force , enhanced by subsequent sediment loading (Watts, 1989), and in this case modulated by the presence of an underpinning, buoyant magma body. The amount and rate of surface deformation is controlled ultimately by the strength of the litho sphere (Karner & Dewey, 1986) and gravity controls on the upwards migration and accumulation of magma (G.P.L. Walker, 1989). If the lithosphere is highly fractured, or magma is of low density and/or viscosity, deformation will be compensated locally. Unconformities can develop within any sedimen tary basin, their occurrence depending on rates of subsidence, rates of sediment production, and eustatic sea-level change. The only base-level scenario that can prevent syneruption deposits of the Tryfan Fach Member from being incised after aggradation is to invoke a rapid rise in relative sea level in conj unction with the explosive volcanism. Given the short time available, a eustatic control seems unlikely, and the sea-level rise is best related to accelerated subsidence of the underlying fault block. Although this subsidence could be purely tectonic and have no relation to the volcanism, the coincidence of the two events seems hard to explain. Several other authors have also noted that vol canism was associated with accelerated and/or dif ferential basin subsidence (Kokelaar, 1988; Besley & Collinson, 199 1 ; see above). This genetic linkage between rates of basin subsidence (e.g. relative sea level rise) and sediment load (e.g. an eruption) is uncommon in normal sedimentary successions. To some extent, however, particularly in strike-slip
basins, the details of the relative timing of sub sidence and volcanism is a 'chicken-and-egg' type of question. Did increased transtension and associated subsidence cause the volcanic eruption or did the volcanic eruption occur and allow or drive renewed subsidence? Inter-eruptive sedimentation
Inter-eruption or 'background' facies and facies geometries depend on the topography, relief, type of bedrock , climate, subsidence history, etc. of the drainage basin. Almost any facies could conceivably occur. In low-relief extensional basins fine-grained mud commonly is deposited after volcanic events on alluvial plains or within lacustrine to playa environments (e.g. Besley & Collinson, 199 1 ; Buesch, 199 1 ; Turbeville, 1991 ; White & Robinson, 1992; this study) . These environments arise because of the low topography of the basin, because syn eruptive deposition flattens and infills topography, because low-gradient drainage systems are easily blocked or re-routed by lava flows or debris flows, or because volcanism drives basin subsidence.
FACIES MODE L L ING : T IME ' S ARROWS V ERSUS T IME ' S CYCLE
The stratigraphical record allows us to define 'bite sized' pieces or facies (R. G. Walker, 1990). In facies modelling we assume 'there is a system and order in Nature' (R.G. Walker, 1992) and attempt to re assemble these 'bites' into a group of facies that are genetically related to each other and have some environmental or palaeogeographical significance. The two end-member conditions, time's arrow (a unilinear succession of unique events) and time's cycle (recurrent patterns in a world that remains essentially unchanged) are categories of our invention, erected here to illustrate the problems of interpreting volcaniclastic successions. Although both approaches, by themselves, are incorrect (as noted by Gould, 1987), one may become dominant. Two variables unite to determine the degree to which Walther's Law is applicable; these are the completeness of the sedimentary record (the num ber of time gaps), and the extent to which the sedi ment source changed, in terms of the amount and grain size of supply, throughout deposition (Fig. 17) . The two end-member conditions are: continu-
188
G.J. Orton
SEDIMENT COMPOSITION
STRATIGRAPHIC "BITES"
COMPLETENESS OF RECORD
WALTHER'S MARITAL STATE
t
(/) (/) w z � u None
A lot
AMOUNT SOURCE CHANGED BETWEEN SUCCESSIVE BEDS
�
PURE BLISS
DIVORCED
Fig. 17. Application of Walther's Law to syneruptive and inter-eruptive volcaniclastic successions (using ideas from Dott,
1983; Gould, 1987; Walker, 1990, 1 992). Walther's marital state is discussed further in the text.
ous sedimentation and/or an unchanging source (Walther's Law fully applicable) ; episodic sedi mentation and/or a constantly changing source (Walther's Law inapplicable) . As all environments, including non-volcanic ones, are to an extent influenced by episodic events (Dott, 1983 ) , the metaphor that becomes dominant depends on the time frame over which observations are made. The time's arrow metaphor embodies the entire Tryfan Fach Member. The coarse-grained cross stratified and thinly bedded rhyodacitic sandstones are unrelated to and lie with probable erosional unconformity on subjacent marine mudstones. The up-sequence progression of facies on this time-scale is essentially unpredictable, as pre-existing sediment dispersal systems were overwhelmed and probably
transformed by introduction of coarse-grained rhyodacitic material associated with the explosive volcanic eruption. The time's cycle metaphor is most applicable within the pre-eruptive marine mudstones and the epiclastic alluvial fan succession. Because source was comparatively constant and unchanging (or changing gradually) a facies model can be provided for the whole succession, vertical sequences can be interpreted, and palaeogeographical maps can be presented . Although unique events did occur (e.g. the ash-fall beds, flood events from the alluvial plain) , their magnitude was insufficient to radically alter pre-existing sediment dispersal systems and/or environments. Episodic hydrological and oceano graphic events (storms, river mouth floods) become
Facies models in volcanic terrains
'normal' over the time period (tens of thousands of years) involved and Walther's Law is fully applicable. In non-volcanic terrains , and in the epiclastic successions noted above, it is implicit (de Raaf et a/. , 1965) or explicit (Busch , 1971) that any progressive facies association with gradational facies contacts is the result of one related set of depositional conditions. However, within syneruptive reworked pyroclastic deposits, gradational grain-size and facies changes are virtually meaningless (or at least very difficult to interpret! ) in two-dimensional sequences unless the eruptive history of the volcano is known. This is rare, even in historic eruptions. Many volcanic eruptions consist of a series of closely spaced but smaller events, with the amount, grain size, density, shape and/or 'distribution of eruptive products gradually changing owing to progressive changes in the amount of magma vesiculation, the eruptive flux rate, vent parameters , and the supply rate of water. Although each bed is unique, the succession of syn eruptive beds may not be random but linked by gradual (?and predictable) changes in source region parameters. The problems this causes in sequence analysis are well illustrated in the Tryfan Fach Member by the braidplain to floodbasin transition, where it could not be determined whether facies change reflected sorting processes on the alluvial plain or progressive eruption of finer grained volcanic ash. Environments were difficult to reconstruct and palaeogeographical maps could not be produced. As a consequence the timing of other controls on the stratigraphy (e.g. sea-level change) could not be fully assessed . Thus although one cannot totally divorce Walther's Law from these volcaniclastic sequences it certainly must be applied with caution, with the relationship best described as an unhappy marriage (Fig. 17) .
1 89
a nearby, but unrelated, source. However, within syneruptive pyroclastics it could not be determined whether coarse sandstones were reworked from a single or several eruptions, and whether rhyolitic siltstones on the alluvial plain represent fine-grained material extracted from older pyroclastic flows during reworking or influx of ash from new volcanic eruptions. The key to unlocking the story hidden within such volcaniclastic sequences is to recognize when volcanism was active. The basic question that is addressed is whether facies changes in a vertical sequence reflect changing energy levels within the depositional setting (Walther's Law is applicable) or changing source parameters because of or during an eruption (Walther's Law does not work ) . In this regard, the study of volcaniclastic materials is still in its infancy, and is the area where much new research is needed. The above understanding is possible only if volcanism-induced eruptive and emplacement processes can be recognized and separated from 'background' conditions of sediment supply (see also McPhie & Allen, 1992). Even when the eruption is observed (e.g. Mount St Helens) , it may be dif ficult to distinguish between primary and reworked pyroclastic material, and to recognize individual eruptive episodes in the depositional record. These distinctions becoming increasingly difficult to make when studying the ancient record , where syneruptive sheets may be difficult to separate (G. Smith, 1991) and material from smaller eruptions becomes admixed with 'background' inter-eruption detritus (Hackett & Houghton, 1989). In these successions, it may not be possible to determine whether facies change reflects a change in type and/or energy of the environment, or a change in the type of sediment produced by the volcano, or both. We should be prepared to admit this and realize the implications for the accuracy of environmental and/or sequence interpretations .
CONCLUSIONS
Coarse-grained sandstones of the Tryfan Fach Member represent aggradation and progradation of reworked pyroclastic materials deposited in a braid plain setting in conjunction with penecontempor aneous ?phreatomagmatic rhyodacitic volcanicity. It can be shown with reasonable certainty that these deposits have no genetic relation to either muddy shelf deposits , upon which they lie with probable erosional unconformity, and/or adjacent alluvial fan deposits, which prograded at the same time from
ACKNOWLEDGEMENTS
Research formed part of a doctoral thesis ( 1990) completed under the helpful guidance of Harold Reading (Oxford University) and Roger Suthren (Oxford Brookes University) . Later visits to North Wales were financed by the Monash University Development Fund and the Australian Research Council. Over the years stimulating field discussions in North Wales with Dairmid Campbell, Ray Cas,
190
G.J. Orton
Bill Fritz, Malcolm Howells, Adrian McArthur and Tony Reedman have contributed to my understand ing. Ideas gathered in North Wales have benefited by additional research experiences in Australia (Exhibition of 1851 Fellowship) and Japan (STA & AIST Fellowships) where similar environmental problems were encountered. Ideas presented in this paper were greatly sharpened by thorough and constructive reviews by Gary Smith, James White, and the editor, Guy Plint.
Bull. A m . Ass. petrol. Geol. , 55, 1 137- 1 154. CAMPBELL,
C .V. ( 1976) Reservoir geometry of a fluvial sheet sandstone . Bull. Am. Assoc. petrol. Geol. , 70, 1009 - 1020.
CAMPBELL,
S.D.G. , HowELLS, M.F. , SMITH , M . & REEDMAN, A.J. ( 1 988) A Caradoc failed rift within the Ordovician marginal basin of Wales. Geol. Mag. , 125,
257-266. CANT, D.J.
REFERENCES ALLEN,
J.R.L. ( 1974) Reaction, relaxation and lag in natural sedimentary systems . Earth Sci. Rev. , 10, 263-
342. ALLEN,
Petrol. , B64, 40-59.
BuESCH, D . C. ( 1991) Changes in depositional environ ments resulting from emplacement of a large-volume ignimbrite. In: Sedimentation in Volcanic Settings (Eds Fisher, R.V . & Smith, G.A.), Spec . Publ. Soc. econ. Paleontol. Mineral. , Tulsa, 45, 139- 154. BuscH, D . A. ( 1971 ) Genetic units in delta prospecting.
& WALKER, R.G. ( 1 978) Fluvial processes and facies sequences in the sandy braided South Saskatchewan River, Canada. Sedimentology , 25, 625-· 648.
CAPUTO,
J. R.L.
( 1 984)
Sedimentary
Structures:
their
Elsevier, Amsterdam, pp . ALLEN, P.A. ( 1981) Wave-generated structures in the Devonian lacustrine sediments of south-east Shetland and ancient wave conditions . Sedimentology, 28, Character and Physical Basis. 663
369-379.
ARNOTT, R.W .C . ( 1 993) Quasi-planar-laminated sandstone beds of the Lower Cretaceous Bootlegger Member, North-Cental Montana: evidence of combined-flow sedimentation. J. sediment. Petrol. , 63, 488-494. BACHMAN, G.O . & MEHNERT, H.H. ( 1 978) New K - Ar dates and the late Pliocene to Holocene history of the central Rio Grande region, New Mexico. Geol. Soc. Am. Bull. , 89, 283-292.
BALL, T.K. & MERRIMA N, R.J. ( 1989) The petrology and geochemistry of the Ordovician Llewelyn Volcanic Group, Snowdonia, North Wales. British Geological Survey Research Report, No. 89/ 1 . HMSO, London . BARDAJI , T. , DABRJO, C.J., GoY, J.L., SoMOZA, L. & ZAZO, C . ( 1 990) Pleistocene fan deltas in southeastern Iberian peninsula: sedimentary controls and sea-level changes. In: Coarse-grained Deltas (Eds Colella, A. & Prior, D.B.), Spec. Pubis int . Ass . Sediment., No. 1 0 , pp. 129- 1 5 1 . Blackwell Scientific Publications, Oxford. BATES, D . E .B . ( 1 972) The stratigraphy of the Ordovician rocks of Anglesey . Geol. J. , 8, 29-60. BESLEY, B .M . & COLLIN SON , J.D. ( 199 1 ) Volcanic and tectonic controls of lacustrine and alluvial sedimentation in the Stephanian coal-bearing sequence of the Malpas Sort Basin, Catalonian Pyrenees . Sedimentology , 38, 3 - 26.
BooTHROYD, J.C. & Nu MME DA L , D. ( 1978) Proglacial braided outwash: a model for humid alluvial fan deposits. In: Fluvial Sedimentology (Ed. Miall, A.D.), Mem. Can. Soc . petrol. Geol., Calgary, 5, 641 - 668. BRANNEY, M.J . ( 1 99 1 ) Eruption and depositional facies of the Whorneyside Tuff Formation, English Lake District: an exceptionally large magnitude phreatoplinian erup tion. Geol. Soc. Am. Bull. , 103, 886-897. BROWNE, q.H. & PuNT, A . G . ( 1 994) Alternating braid plain and lacustrine deposition in a strike-slip set ting: the Pennsylvanian Boss Point Formation of the Cumberland Basin, Maritime Canada. J. sediment.
M. ( 1 979) Two thousand years of geodetic and geophysical observation in the Phlegran Fields near Naples. Geophys. J. R. astron. Soc. , 56, 3 19-328. CAS, R.A . F. & WRIGHT, J .V. ( 1987) Volcanic Successions: Modern and Ancient. Allen & Unwin, London, 528 pp . CATHER, S.M. & FoLK, R.L . ( 199 1) Pre-diagenetic sedimentary fractionation of andesitic detritus in a semi arid climate: an example from the Eocene Datil Group, New Mexico. In: Sedimentation in Volcanic Settings (Eds Fisher, R .V. & Smith, G.A.), Spec. Publ. Soc. econ. Paleont. Mineral. , Tulsa, 45, 2 1 1 -226. CoCKS, L. R .M . & FORTEY, R.A. ( 1 982) Faunal evidence for oceanic separations in the Palaeozoic of Britain. J. geol. Soc. London, 139, 465 -478.
CoLE, J .W . ( 1 984) Taupo-Totorua depression: an ensialic marginal basin of North Island, New Zealand. In: Marginal Basin Geology: Volcanic and Associated Sedimentary and Tectonic Processes in Modern and
(Eds Kokelaar, B.P. & Howells, M .F.), Spec . Publ . geol. Soc. London, No. 16, pp. 109120. Geological Society of London, Bath. CoLLINSON, J . D. ( 1 969) The sedimentology of the Grindslow Shales and the K.inderscout Grit: a deltaic complex in the Namurian of northern England. J. A ncient Marginal Basins
Sediment. Petrol. , 39, 194- 221 .
CORDELL, L. , LoNG, C.L. & JONES, D.W. ( 1985) Geo·· physical expression of the batholith beneath Questa caldera, New Mexico. J. geophys. Res . , 90, 263-269. COTTER, E. ( 1 978) The evolution of fluvial style, with special reference to the central Appalachian Palaeozoic. In: Fluvial Sedimentology (Ed. MiaU, A.D.), Mem . Can. Soc. petrol . Geol. , Calgary, 5, 361 -384. CROWLEY, K.D. ( 1 983) Large scale bed configurations (macroforms), Platte River Basin, Colorado and Nebraska: primary structures and formative processes. Geol. Soc. Am. Bull. , 94, 1 17 - 133.
DAM, G. & AN DREASAN , F. ( 1990) High energy ephemeral stream deltas: an example from the Upper Silurian Holmestrand Formation of the Oslo region, Norway. Sediment. Geol. , 66, 197-226.
DAVIES, O . K. , ALMON, W. R. , BoNIS, S. B . & HuNTER,. B . E . ( 1 979) Deposition and diagenesis of Tertiary Holocene volcaniclastics, Guatemala. In: Aspects of Diagenesis (Eds Scholle, P.A. & Schluger, P.R.), Spec. Publ. Soc. econ. Paleont. Mineral. , Tulsa, 26, 281 -306.
Facies models in volcanic terrains DE RAAF, J.F.M. , READING, H. G. & WALKER, R.G. (1965) Cyclic sedimentation in the Lower Westphalian of North Devon, England. Sedimentology , 4, 1 -52. Dorr, R.H. (1983) 1982 SEPM presidential address: Episodic sedimentation - how normal is average? how rare is rare? does it matter? I. sediment. Petrol. , 53, 5-23. DvoRAK, J.J. & GASPARINI, P. (1991) History of earth quakes and vertical ground movements in Campi Flegrei caldera, southern Italy: comparison of precursory events to the AD 1538 eruption of Monte Nuovo and of activity since 1968. I. volcano!. geotherm. Res. , 48, 199-222. EwART, A. (1979) A review of the mineralogy and chemistry of Tertiary- Recent dacitic , latitic, rhyolitic and related sialic volcanic rocks. In: Trondhjemites, Dacites and Related Rocks (Ed. Baker, F.), pp. 113 - 121. Elsevier, Amsterdam. FERNANDEZ, L . P . , AGUEDA, J.A . , COLMENERO, J . R . , SALVADOR, C.!. & BARBA, P . (1988) A coal-bearing fan delta complex in the Westphalian D of the Central Coal Basin, Cantabrian Mountains, northwestern Spain: implications for the recognition of humid-type fan deltas. In: Fan Deltas: Sedimentology and Tectonic Setting (Eds Nemec, W. & Steel, R.J.), pp. 286-302. Blackie, Glasgow. FISHER, R.V. (1961) Proposed classification of volca niclastic sediments and rocks. Geol. Soc. Am. Bull. , 72, 1395- 1408. FISHER, R. V. & SCHMINCKE, H.U. (1984) Pyroclastic Rocks. Springer-Verlag, Berlin, 472 pp. FosHAG , W.F. & GoNZALEZ-R, J. ( 1956) Birth and devel opment of Paricutin volcano, Mexico. U.S. geol. Surv. Bull. , 965-D, 355-489. FRITZ, W.J. (1991) Theoretical modelling of large wave ripples in volcaniclastic sediments, Ordovician Llewelyn Volcanic Group , North Wales. Sediment. Geol. , 74, 241 -250. FRITZ, W.J. & HOWELLS, M.F. (1991) A shallow marine volcaniclastic facies model: an example from sedimentary rocks bounding the subaqueously welded Ordovician Garth Tuff, North Wales, U . K. Sediment. Geol. , 74, 217-240. FORSICH, F.T. & MAYR , H. (1981) Non-marine Rhizo corallium (trace fossil) from the Upper Freshwater Molasse (Upper Miocene) of southern Germany. Neues Jahrb. Geol. Palaeontol. Monatsh. , 198 1 , 321 -333. GALLOWAY, W.E. (1981) Depositional architecture of Cenozoic Gulf Coast plain fluvial systems. In: Recent and Ancient Non-marine Environments: Models for Exploration (Eds Ethridge, F. G. & Flores, R.M.), Spec.
Pub!. Soc. econ. Paleont. Mineral., Tulsa, 3 1 , 127- 155. W. (1983) Stratigraphy, subduction and strike slip faulting in the Mona Complex of North Wales - a review. Proc. Geol. Assoc. , 94, 147 - 164. GoDIN, P.D . (1991) Fining-upward cycles in the sandy braided-river deposits of the Westwater Canyon Member (Upper Jurassic), Morrison Formation, New Mexico. Sediment. Geol. , 70, 6 1 - 82. GouLD, S.J. (1987) Time's Arrow, Time's Cycle: Myth and Metaphor in the Discovery of Geological Time. Penguin Books, New York, 222 pp. HACKEIT, W.R. & HouGHTON , B.F. (1989) A facies model for a Quaternary andesitic composite volcano: Ruapehu, New Zealand. Bull. Volcano!. , 5 1 , 5 1 - 68. GIBBONS,
191
A. P. ( 1992) Half-graben lacustrine sedimentary rocks of the lower Carboniferous Strathlorne Formation, Horton Group , Cape Breton Island , Nova Scotia, Canada. Sedimentology , 39, 263 -284. HASZELDI N E , R.S. & AN DERTO N , R. (1980) A braidplain facies model for the Westphalian B Coal Measures of north-east England. Nature, 284, 5 1 - 53 . HEIKEN , G. & WoHLETZ, K. (1991) Fragmentation pro cesses in explosive volcanic eruptions. In: Sedimentation in Volcanic Settings (Eds Fisher, R.V. & Smith, G.A.), Spec. Pub!. Soc. ccon. Paleont. Mineral. , Tulsa, 45, 19-26. HENRY, C.D. & PRICE, J.G. (1984) Variations in caldera development in the Tertiary volcanic field of Trans Pecos Texas. I. geophys. Res. , 89, 8765-8786. HICKS, D.M. & INMAN , D.L. (1987) Sand dispersion from an ephemeral river delta on the central California coast. Mar. Geol. , 77, 305-3 18. HILDRETH , W. (1987) New perspectives on the eruption of 1912 in the Valley of Ten Thousand Smokes, Katamai National Park, Alaska. Bull. Volcano!. , 49, 680-693. HILL, D.P. , BAILEY, R.A. & RYALL, A.S. (1985) Active tectonic and magmatic processes beneath Long Valley Caldera, Eastern California: an overview. J. geophys. Res. , 90, 1 1 1 1 1 - 1 1 120. HILL, P.R. & NADEA U , O . C. (1989) Storm dominated sedimentation on the inner shelf of the Canadian Beaufort Sea. I. sediment. Petrol. , 59, 455-468. HowELLS, M.F. & LEVERIDGE, B.E. (1980) The Capel Curig Volcanic Formation. Institute of Geological Sciences Report, No. 80/6, pp. 1 -23. HMSO, London. HoWELLS, M . F. , REEDMAN, A.J. & CAMPBELL, S.D.G. (1991) Ordovician (Caradoc) Marginal Basin Volcanism in Snowdonia (North- west Wales) . British Geological Survey, HMSO, London, 191 pp. HunoN, J. (1788) Theory of the Earth. Trans. R. Soc. Edinburgh , l , 209-305. JOPLING, A.V. (1965) Hydraulic factors controlling the shape of laminae in laboratory deltas. I. sediment. Petrol. , 35, 777-791. KARNER, G.D. & DEWEY, J .F. (1986) Rifting: lithospheric versus crustal extension as applied to the Ridge Basin of southern California. In: Future Petroleum Provinces of the World (Ed. Halbouty, M.T.), Mem. Am. Assoc. petrol. Geol., Tulsa, 40, 317-337. KOKELAAR, B.P. (1988) Tectonic controls of Ordovician arc and marginal basin volcanism in Wales. I. geol. Soc. London, 145, 759-775. KoKELAAR, B.P. (1992) Ordovician marine volcanic and sedimentary record of rifting and volcanotectonism: Snowdon, Wales, UK. Geol. Soc. Am. Bull. , 104, 1443 - 1455. LECKIE , D.A. & CHEEL, R.J. ( 1989) The Cyprus Hills Formation (upper Eocene to Miocene): a semi-arid braidplain resulting from intrusive uplift. Can. I. Earth Sci. , 26, 1918-1931. LEEDER, M. R. ( 1983) Lithospheric stretching and North Sea clastic sourcelands. Nature, 305, 510-513. LIPMAN , P.W. (1984) Roots of ash-flow calderas in western North America: windows into the tops of granitic batho liths. I. geophys. Res. , 89, 8801 -8841. LowE, D.R. ( 1982) Sediment gravity flows: II. Depo sitional models with special reference to the deposits of high-density turbidity currents. J. sedime'ru. Petrol. , 52, HAMBLIN ,
192
G.J. Orton 'offshore bars' in the Cardium Formation of Alberta :
279-297. G.
LuONGO,
&
SCAN DONE,
R.
( 1991)
Campi Flegrei.
J. Volcano!. Geotherm. Res. , 48, 227 pp. LYELL, C . ( 1853) Principles of Geology 9th edn. John Murray, London, 835 pp. MA.tzELS, J. ( 1989) Sedimentology, palaeoflow dynamics, and
flood
history
of
jokulhlaup
deposits:
palaeo
hydrology of Holocene sediment sequences in southern Iceland
sandur
deposits.
J.
sediment.
59,
Petrol. ,
their relationship to relative changes in sea level. I n : Sea-level Changes: a n Integrated Approach (Eds Wilgus , C . K . , Hastings, H . W . , Kendall, C . , Posamentier, H . W . , Ross, C . A . & Van Wagoner, J . C . ) , Spec. Pub!. Soc. econ. Paleont. Mineral . , Tulsa, 42, 357-370. POINTON , C . R . & IXER, R . A . ( 1 980) Parys Mountain mineral deposit, Anglesey, Wales: POLLARD, J . E . , STEEL, R . J .
204-223. McKEE, C . P . , LOWENSTEI N , P . L . , DESAINTOURS, P . , TALA!,
geology and ore
mineralogy. Trans. Inst. Min. Metal/. , 89, 143 - 155 .
&
UNDERSUD, E. ( 1 982) Facies
sequence and trace fossils in lacustrine fan delta deposits ,
some regional implications. Sediment. Geol. , 2 1 , 85 -
Hornelen basin, Middle Devonian, Western Norway. Sediment. Geol. , 32, 63-87. PRIOR, D . B . , BORNHOLD, B . D . , WiSEMA N , W . J . & LOWE, D . R . ( 1987) Turbidity current activity in a British Columbian fjord. Science, 237, 1330- 1333. REEDMAN, A .J . , LEVERIDGE, B . E . & EVANS, R . B . ( 1 984) The Arfon Group ('Arvonian') of North Wales. Proc. Geol. Assoc. , 95, 3 1 3 - 32 1 . REILINGER, R . E . & YORK, J . E . ( 1 979) Relative crustal subsidence from levelling data in a seismically active part of the Rio Grande rift, New Mexico . Geology , 7 , 139 - 143.
127 .
ROWLEY , P . D . , KUNTZ, M . A .
B . , htKARAI, I.
&
Motu, J . L . (1984) Seismic and ground
deformation crises at the Rabaul caldera: prelude to an eruption? Bull. volcano!. , 47, 397 -41 1 . McPHIE, J .
&
ALLEN, R . L . ( 1992) Facies architecture of
mineralized submarine volcanic sequences: Cambrian Mount Read Volcanics, Western Tasmania. Econ. Geol. , 87, 587-596. MlALL, A . D . (1977) A review of the braided river depo sitional environment . Earth Sci. Rev. , 13, 1 - 16. MlALL, A . D .
& GlBLI N G , M . R .
( 1978) The Siluro-Devonian
clastic wedge of Somerset Island, Arctic Canada, and
MIDDLETON , G . V . (1973) Johannes Walther's Law of the correlation of facies. Geol. Soc. Am. Bull. , 84, 979-988. MYROW, P . M . ( 1992) Bypass-zone tempestite facies models and proximality trends for an ancient muddy shoreline
&
&
In:
D . R. ) . U . S . geol. Surv. Prof. Pap . , 1250, 489 - 5 12 . RUEGG,
G.H.
( 1971)
Features of
SouTHARD, J.B. ( 1991) Combined-flow
model for vertical stratification sequences in shallow
5 - 24.
marine storm-deposited beds. J. sediment. Petrol. , 6 1 ,
volcanogenic sedimentation; sedimentary
NELSON, C.H. ( 1 982) Modern shallow-water graded sand Bouma sequences and turbidite systems .
a mimic of
.1. sediment.
Petrol. , 52, 537-545. NEWHALL, C.G. & DzuRJSIN, D. ( 1988) Historical unrest at large calderas of the world . U . S . geol . Surv. Bull . ,
braidplain deltaic sequences, North Wales: a conse quence of varying fluvial clastic input. I n : Fan Deltas:
Sedimentology and Tectonic Setting (Eds Nemec, W. Steel, R . J . ) , pp. 23-49. Blackie, Glasgow. G.J.
( 1 990)
&
Volcaniclastic sedimentation in a
Caradocian marginal basin, North Wales . Unpublished PhD thesis, Oxford U niversity. ORTO N , G .J. ( 1991) Emergence of subaqueous depositional environments in advance of a major ignimbrite eruption, Capel Curig Volcanic Formation, North Wales:
an
example of regional volcanotectonic uplift? Sediment.
Ordovician of central Wales. Geol. Mag. , 1 13, 159- 164.
& BRENCHLEY,
P.J. ( 1979) Caradoc marine
benthic communities of the south Berwyn Hills, North PtCKERJLL, R . K .
& H uRST, J . M .
depositional
environments,
the
lower
analysis of two Eocene
Bend
Region ,
Texas.
J.
sediment. Petrol. , 60, 747-760. RusT, B . R . & GtBLI N G , M . R . ( 1990a) Braidplain evolution in the Pennsylvanian South Bar Formation near Sydney, Nova Scotia. J. sediment. Petrol. , 60, 59-72. RusT, B.R.
GtBLIN G , M.R. (1990b) Three-dimensional
&
antidunes as HCS mimics in a fluvial sandstone: the
Llandovery
Scotia. J. sediment. Petrol. , 60, 540-548. SHERIDAN ,
M.F.
&
UPDIKE,
( 1983) Sedimentary facies, and
faunal
(Silurian)
R.G.
( 1975)
Sugarloaf
Mountain tephra - a Pleistocene rhyolitic deposit of base-surge origin in Northern Arizona. Geol. Soc. Am. Bull. , 86, 5 7 1 - 5 8 1 . SIMPSON , E . L . & ERIKSSON , K . A . ( 1 989) Sedimentology of the U nicoi Formation in southern and central Virginia: evidence for late Proterozoic to Early Cambrian rift-to·· passive margin transition. Geol. Soc. Am. Bull. , 101 , 42-54. SLOAN, R.J.
&
WILLIAMS, B .P . J . ( 199 1) Volcano-tectonic
control of offshore to tidal flat regressive cycles from the of southwest Ireland.
In:
Sedimentation, Tectonics and Eustacy: Sea-level Changes at Active Plate Margins (Ed . Macdonald , D . I . M. ) , Spec. Pub!. int. Ass. Sediment . , No. 12, pp. 105 - 1 19 . Blackwell Scientific Publications, Oxford . SMITH, G . A . ( 1986) Coarse-grained nonmarine volcani··
Wales. Palaeontology , 22, 229-264.
of
Big
Dunquin Group (Silurian)
Geol. , 74, 25 1 - 288. PtCKERJLL, R . K . ( 1976) Vermiforichnus borings from the PICKERJLL, R . K .
aprons,
Pennsylvanian South Bar Formation near Sydney, Nova
1855, v . 1 , 598 pp. ORTO N , G . J . ( 1988) A spectrum of mid-Ordovician fan and
ORTON ,
Pleistocene
RuNKEL, A . C . ( 1990) Lateral and temporal changes in
202-210. layers from storm surges, Bering shelf:
middle
sandur deposits in the Netherlands. Geol. Mijnbouw, 56 ,
and shelf. J. sediment. Petrol. , 62, 99 - 1 15. MYROW, P.M.
MACLEO D , N . S . (1 981) The 1 980 Eruptions of Mount St. Helens (Eds Lipman, P . W . & Mullineaux,
Pyroclastic-flow deposits.
associations
Beechill
Cove
clastic sediment: terminology and depositional process.
Geol. Soc. Am. Bull. , 97, 1 - 10. SMITH, G.A. ( 1987) The influence of explosive volcanism
Formation, Arisaig, Nova Scotia. Can. J. Earth Sci. , 20,
on
1761- 1779.
(Neogene) in Central Oregon. J. sediment. Petrol. , 57 ,
PuNT, A . G . ( 1988) Sharp-based shoreface sequences and
fluvial
6 1 3 - 629.
sedimentation:
the
Deschutes
Formation
193
Facies models in volcanic terrains SMITH , G.A. (1988) Sedimentology of proximal to distal volcaniclastics dispersed across an active foldbelt: Ellens
South Africa. Sediment. Geol. , 18, 161 - 174. WALKER, G . P . L . (1984) Down-sag calderas, ring faults,
berg Formation (late Miocene) , central Washingto n .
caldera sizes, and incremental caldera growth. J. geophys.
Sedimentology, 3 5 , 953-977. SMITH, G.A. ( 1991) Facies sequences and geometries i n
Res. , 89, 8407-8416. WALKER, G . P . L . ( 1989) Gravitational (density) controls on volcanism , magma chambers and intrusions. Aust. J.
continental volcaniclastic sequences. I n : Sedimentation
in Volcanic Settings (Eds Fisher, R.V. & Smith, G . A . ) , Spec. Pub!. Soc. eco n . Paleontol. Mineral . , Tulsa, 45, 109 - 122.
Earth Sci. , 36, 149-165. WALKER, R.G. ( 1 990) Facies modelling and sequence stra
SMITH , R . C . M . ( 1991) Landscape response to a major
WALKER, R . G . ( 1 992) Facies, facies models and modern
ignimbrite
eruption ,
Taupo
volcanic
center,
New
Zealand . In: Sedimentation in Volcanic Settings (Eds Fisher, R . V .
Smith, G . A . ) , Spec. Pub!. Soc. econ .
&
HurroN , D .H . W . ( 1 984) Late Caledonian
&
sinistral displacements in Britain: implications for a three plate collision model. Tectonics, 3, 791 - 794. STEEL, R.J.
&
AASHEIM, S . M . ( 1978) Alluvial sand depo
sition in a rapidly subsiding basin, Devonia n , Norway. In: Fluvial Sedimentology (Ed. Miall, A . D . ) , Mem. Can. Soc. petrol. Geol . , Calgary, 5, 386-412. SuRDAM, R.C.
& BoLES, J . R .
stratigraphic concepts. I n : Facies Models: Response to
Sea Level Change (Eds Walker, R . G .
( 1979) Diagenesis of volcanic
&
James, N . P . ) ,
p p . 1 - 14 . Geological Association o f Canada, St John's. WALKER, R . G .
Paleontol. Mineral . , Tulsa, 45, 123 - 138. SoPER, N.J.
tigraphy. J. sediment. Petrol. , 60, 777-786.
tidal
HARMS, J . C . ( 1975) Shorelines of weak
&
activity:
Upper
Devonian
Catskill Formation,
Central Pennsylvania. In: Tidal Deposits: a Casebook of
Recent Examples and Fossil Counterparts (Ed. Ginsburg, R . N . ) , pp. 103 - 108. Springer-Verlag, New York. WALTON, A . W . ( 1 986) Effect of Oligocene volcanism on sedimentation i n the Trans-Pecos volcanic field of Texas. Geol. Soc. Am. Bull. , 97, 1 192- 1207. WATKINS,
R.T.
( 1986)
Volcano-tectonic
control
on
sandstones. I n : Aspects of Diagenesis (Eds Scholle, P . A .
sedimentation i n the Koobi Fora sedimentary basin. In:
Schluger, P . R . ) , Spec. Pub!. Soc. econ. Paleontol.
Sedimentation in the African Rifts (Eds Frostick, L . E . ,
&
Renault, R . W . , Reid, I .
Mineral . , Tulsa, 26, 227-242. THORPE, R . S . ( 1979) Late Precambrian igneous activity in
Southern Britain . I n : The Caledonides of the British
Isles - Reviewed (Eds Harris, A . L . , Holland , C . H . & Leake, B . E . ) , Spec. Pub!. geol. Soc. London, No. 8 ,
& TRENCH, A . ( 1991) The Ordovician history
of the Iapetus Ocean in Britai n : new palaeomagnetic R.J.B . ,
EccLES ,
C.,
MOLYNEUX,
S.G.
&
TAYLOR, W . E . G . ( 1987) Age and controls of ironstone deposition (Ordovician) North Wales. Geol. J. , 22, 3 1 - 43 .
sediment loads:
& RuSHTO N , A . W . A .
( 1980) Faunas from
sediments associated with the Capel Curig Tuffs: One inch sheet
106 (Bangor). British Geological Survey
COLEMA N , J . M . ( 1981) Physical dynamics
&
South America. J. sediment. Petrol. , 51, 1053 - 1063. WELLS ,
N.A.
( 1983)
B.N.
(1991) The influence of ephemeral
processes on pyroclastic sedimentation in a rift-basin, volcaniclastic- alluvial sequence, Espanola basin, New Mexico. Sediment. Geol. , 74, 139 - 155 . North America, Gonwana, and intervening displaced terranes: comparisons of paleomagnetism with paleo climatology and biogeographic patterns. Geol. Soc. Am.
Bull. , 100, 3 1 1 -324. VAN HouTE N , F . B . (1976) Late Cenozoic volcaniclastic deposits, Andean foredeep, Columbia. Geol. Soc. Am.
Bull. , 87, 481-495.
&
DAVIES,
sand-poor
&
Lewin , J . ) , Spec. Pubis int. Assoc.
Sediment . , No . 6, pp. 393-403. Blackwell Scientific Publications, Oxford . WELLS, S . G . , McFADDEN , L . D . , RENAULT, C . E . volcanism
&
CROWE,
in
the
Yucca
Mountain
area,
southern
Nevada: implications for the proposed high-level radio active waste disposal . Geology, 18, 549-553. WESTHEAD,
VAN DER Voo, R. (1988) Paleozoic paleogeography of
R.K.
Transient streams i n
redbeds. In: Modern and Ancient Fluvial Systems (Eds
B . M . (1990) Geomorphic assessment of late Q uaternary
Open File Report, PD 80/12. TuRBEVILLE,
implications for the origin of offlap/
133 - 144. WELLS, J.T.
Colliso n , J . D .
TUNNICLIFF, S . P .
VESSELL,
of London, Bath. WAITS , A . B . ( 1989) Lithospheric flexure due to prograding
and fine-grained sediment dynamics, coast of Surinam,
constraints. J. geol. Soc. London , 148, 423 -425. TRYTHALL,
Tiercelin, J . - J . ) , Spec. Pub!.
on lap patterns in sedimentary basins. Basin Res. , 2 ,
p p . 579-584. Geological Society of London, Bath. ToRSVJK, T.H.
&
geol. Soc. London, No. 25 , 85-98. Geological Society
S .J .
(1991) Prospects at Parys Mountains.
Geol. Today , 7 , 130 - 135. WHITE, J . D . L. (1991) The depositional record of small, monogenetic volcanoes within terrestrial basins. I n : Sedimentation in Volcanic Settings ( E d s Fisher, R . V . & Smith, G . A . ) , Spec. Pub!. Soc. econ. Paleontol. Mineral . , Tulsa, 45, 155- 174. WHITE,
J.D.L.
&
ROBINSON,
P.T.
( 1992)
Intra-arc
sedimentation in a low-lying marginal arc, Eocene Clarno D.K.
( 1981)
Non-marine
Formation,
sedimentation i n an active fore-arc basin. In: Recent
89 - 1 14 .
and Ancient Non-marine Environments: Models for Exploration (Eds Ethridge, F . G . & Flores, R . M . ) , Spec. Pub!. Soc. eco n . Paleontol. Mineral . , Tulsa, 3 1 , 3 1 - 45 . V!SHER, G . S . (1965) Use o f the vertical profile i n environ
WiLLIAMS,
central
G.E.
Oregon.
( 1971)
Flood
Sediment. deposits
of
Geol. ,
80,
sand
bed
ephemeral streams of central Australia. Sedimentology , 17, 1 -40. WRIGHT, L .D . , YANG, Z . S . , BORNHOLD, B . D . , KELLER,
mental reconstruction. Bull. Am. Assoc. petrol. Geol. ,
G . H . , PRIOR, D . B .
49, 4 1 - 6 1 .
pycnal plumes and plume fronts over the Huanghe
Vos, R . G .
&
ERIKSSON , K . A .
( 1 977) An embayment
model for tidal and swash deposits occurring within a fluvially dominated middle Proterozoic sequence in
&
WiSEMAN , W . J . ( 1986) Hyper
(Yellow River) delta front . Geomarine Letters, 6, 97106.
Tectonics and Sedimentation
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
Spec. Pubis int. Ass. Sediment.
(1995) 22, 197-217
Coarse-grained lacustrine fan-delta deposits (Pororari Group) of the northwestern South Island, New Zealand: evidence for Mid-Cretaceous rifting M A L C O L M G. LAIRD Institute of Geological and Nuclear Sciences, c/o Department of Geology, University of Canterbury, Christchurch, New Zealand
A BS T R A C T
Mainly coarse-grained non-marine sediments of the Pororari Group, dominated b y breccia and conglomerate, and of mid- to late Albian age, are widespread west of the Alpine Fault in the South Island of New Zealand. They represent an unconformity-bounded sequence, resting with marked angular discordance on Early Palaeozoic sediments of the Greenland Group or non-conformably ori mid-Palaeozoic and Early Cretaceous granites, and overlain with regional discordance by coal measures of Maastrichtian age. In the southern Paparoa Range, where the Pororari Group is in excess of 2000 m thick, it can be divided into three major facies assemblages: (1) matrix- and clast-supported breccia; (2) pebbly sandstone and massive or graded sandstone; and (3) laminated dark mudstone with thin graded sandstones. Facies assemblage 1 is inferred to represent the deposits of debris flows and sheet flows on alluvial fans; facies assemblage 2, the deposits of debris flows and turbidites in a dominantly lacustrine environment; and facies assemblage 3 was deposited in a distal lacustrine environment. The three facies assemblages intertongue, and are inferred to represent elements of a fan-delta system debouching into a lake. The palaeocurrent pattern deduced from clast imbrication in breccia units, and from sedimentary structures in the lacustrine deposits, indicates a dominantly NNE-SSW direction of sediment transport. The consistent and rapid thinning of lithostratigraphical units from north to south is compatible with the palaeocurrent results. Palaeocurrent plots and clast lithologies suggests that several point sources supplied coarse sediment during· the history of deposition of the Pororari Group, each giving rise to distinct but overlapping fans. The great thickness of coarse deposits (a minimum of 1200 m for the breccia assemblage alone) and the angularity of clasts suggests that these sediments were generated by contemporaneous tectonic activity. Fluctuations in lake level also may have been controlled by active faulting at the basin margin, although climatic changes may have played a part. The fan-delta system is likely to have occupied an active subsiding half-graben. The consistently south-southwest directed sediment gravity flows suggest that the active fault was likely to have been oriented in a WNW-ESE direction, downthrown to the south, and with an implied north-northeast direction of tectonic extension. A similar extension direction for the Paparoa Range area during early. Albian times has been deduced from independent structural research on a closely associated metamorphic core complex, and is also compatible with the WNW E SE orientation of normal faults bounding adjacent offshore undrilled half-grabens infilled with deposits of prdbably mid-Cre��ceous age. The coarse deposits of the Pororari Group are likely to have been associated with an early period of rifting in the Albian, pre-dating by perhaps 25 Ma the separation of New Zealand from Gondwana in the late Campanian or early Maastrichtian.
INTRODUCTION
Coarse-grained non-marine sediments, o f mainly mid- to late Albian age, dominated by breccia and
conglomerate, are widespread throughout the southern and western South Island of New Zealand.
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0 197
198
M.G. Laird
Most are poorly exposed, and detailed descriptions are sparse. One exception is the Kyeburn Formation, which lies in the southeast of the South Island (Bishop & Laird, 1976; Fig. 1), and which represents talus to braided river and lacustrine deposits of mid Albian to at least late Cenomanian age, emplaced in a locally developed, triangular, fault-bounded depression. Another well-described exception is the Puysegur Formation of the southwestern South Island (Lindqvist, 1990; Fig. 1), which is inferred to have been deposited in terrestrial, lake margin, and lacustrine environments in a fault-controlled basin. All of these mid-Cretaceous sediments rest uncon formably on older rocks, and clearly post-date the Early Cretaceous Rangitata Orogeny, which marked a significant change of tectonic pattern in the New Zealand region. From the Permian to the Early Cretaceous, most rocks formed under the influence of convergent-margin tectonics and comprise incom plete remnants of magmatic arcs, fore-arc basins, trench-slope basins, .;:tnd accretionary complexes (Bradshaw, 1989). The later stages of this regime, in the Early Cretaceous, were accompanied by wide-
spread compressional deformation, metamorphism, uplift, erosion and calc-alkaline plutonism. Over much of New Zealand the rocks of the Rangitata Orogen are separated by a major uncon formity from younger, less-deformed strata. This unconformity marks the change in tectonic regime from convergent margin to extension, attributed by Bradshaw (1989) to collision with the New Zealand margin of the spreading ridge between the Phoenix and Pacific Plates. Mid Cretaceous extensional tec tonism is best exemplified by the post-unconformity development, in the mid Cretaceous, of half grabens throughout much of the New Zealand region, both onshore and offshore. Subsidence continued along the bounding faults in the western and southern South Island, at least in some instances, from mid Albian until Santonian or early Campanian times (Laird, 1993, 1994). Both the Kyeburn and the Puysegur Formations are inferred to occupy basins formed east of the Alpine Fault in mid-Albian times under the extensional tectonic regime. On the west coast of the South Island, west of the Alpine Fault, outcrops of breccia (collectively known
N
s
Kyeburn •
Fig. 1. Locality map of the South
Puysegur Point
100 km
Island of New Zealand showing ( in black) the main outcrop areas of mid-Cretaceous non-marine coarse clastic deposits. B, Beebys Conglomerate.
199
Cretaceous rift deposits of New Zealand as the Hawks Crag Breccia, forming part of the Pororari Group) are scattered but widespread over a NE-SW extent of approximately 320 km (Fig. 1). The breccia is matrix- or clast-supported with a sparse sandy matrix, and is often reddish-brown in colour. Structure is generally simple, most suc cessions dipping uniformly at angles up to 50°, or folded into simple, open synclines. Thickness of the Pororari Group varies considerably, but a maximum is c. 4 500m. The age of the group has been determined solely from pollen. Palynological studies (Raine, 1984) indicate that the bulk of the Pororari Group lies within the Lycopodiacides bullerensis assemblage, with the youngest rocks reaching into the lower part of the overlying Trichotomosulcites subgranulatus assemblage. Absence of angiosperms suggests that the Pororari Group lies within the New Zealand Motuan Stage (mid- to late Albian). The original upward time range of the Group in the region is unknown, being constrained only by the local occur rence of unconformably overlying coal measures of Maastrichtian age. However, a possible correlative of the Pororari Group, lying c. lOOkm northeast of the Paparoa Range (Beebys Conglomerate; Fig. 1) has yielded a poorly constrained microfloral age in the range late Cenomanian to early Santonian (Johnston, 1990), indicating that Pororari Group deposition may have extended locally well into the Late Cretaceous. The origin and depositional setting of the Hawks Crag Breccia has been the object of considerable speculation in the past: explanations have included glacial or glaciofluvial action (McKay, 1883), talus or alluvial fan formation (Morgan & Bartrum, 1915), or the deposits of air-lubricated landslides (Korsch & Wellman, 1988). In most occurrences, these breccia units make up the entire exposure, and the mechanism of emplacement and geometry of the bodies has been difficult to interpret because of their apparently uniform nature over large areas. How ever, in the Paparoa Range in the northwest of the South Island, where the most extensive areas of outcrop occur, other lithofacies are also present, and these throw light on sedimentary processes and environments of deposition of both the Hawks Crag Breccia and of the Pororari Group as a whole. This study examines the sedimentology of the Pororari Group in the southern Paparoa Range, where the breccia lithofacies is associated with fine-grained deposits (Fig. 2).
G E OL O G IC A L S E T TI N G
The Pororari Group comprises an unconformity bounded sequence resting on Early Palaeozoic turbidites of the Greenland Group, or on mid Palaeozoic and Early Cretaceous granitoids. It is overlain by the Paparoa Coal Measures of Maastrichtian age. In the southern Paparoa Range, the Pororari Group crops out in two extensive areas on the eastern and western flanks of the range, and in small exposures in Fox River to the north (Fig. 2). In the western outcrops, four lithostratigraphic units can be recognized (Laird, 1988; Figs 2 & 3): (1) a locally occurring basal unit, dominated by inter bedded matrix- and clast-supported breccia, massive or normally graded sandstone, and laminated mud stone (Watson Formation), followed gradationally by (2), thick-bedded matrix- and clast-supported breccia (Hawks Crag Breccia), followed by (3), inter bedded thin-bedded sandstone and matrix-supported breccia and conglomerate beds (Bovis Formation), which are in turn overlain by (4), laminated carbon aceous mudstone (Bullock Formation). The eastern outcrops lack the basal unit, and Hawks Crag Breccia rests directly on basement rocks. The landscape is extremely rugged, and covered with dense temperate rainforest (Fig. 4). This limits the accessible and usable exposure to the gorges and tributaries of the main streams, which have provided the measured stratigraphical sections (Fig. 3).
D E S C RIPT I O N OF
A N D I N T E R PR E T A TI O N
FACIES
A S S E MBLA G E S
Three facies assemblages are recognized within the southern Paparoa Range outcrops: (1) matrix- and clast-supported breccia; (2) pebbly sandstone, and massive or graded sandstone; and (3) laminated dark mudstone with thin graded sandstones. Assemblage 1: matrix- and clast-supported breccia
Facies descriptions This facies assemblage makes up the entire Hawks Crag Breccia, and also occurs within the Watson Formation, both as the basal 165m exposed in Tin dale Creek and as a 100-m-thick unit near the base of the formation in Pororari River (Fig. 3). Breccia units increase in frequency near the top of the Watson Formation, which passes gradationally
200
M.G. Laird
LJ Younger strata
;
N
t
Unconformity
Pororari Group Bullock Formation Bovis Formation Hawks Crag Breccia Watson Formation
Unconformity
0 Basement _.- Fault
0
2
3
4
5 km
Fig. 2. Map of part of the southern Paparoa Range showing Pororari Group outcrop and palaeocurrent plots. See Fig. 12
for key to the palaeocurrent plots.
upwards into the Hawks Crag Breccia. Maximum thickness of the facies assemblage is hard to estimate, but in the Bullock Creek area in the north, where it is thickest, it is likely to reach 1500 m.
The assemblage consists dominantly of poorly sorted matrix-supported breccia consisting of sub angular to subrounded clasts (often outsize) of base ment material in a coarse sandy matrix (Fig. 5).
Cretaceous rift deposits of New Zealand N
PAPAROA COAL MEASURES
M
201
s
Bullock Creek
Slaty M ·Creek
BULLOCK FORMATION
1
''LACUSTRINE (distal)
BOVIS FORMATION
120D-fi:I:i'Iii
M _
Pororari River
Facies assemblage
Tindale Creek
''LACUSTRIN E ' ' 2000 (proximal)
3 1000
M
I-�<( <(cr: �
1800
-' -
�0 l9cr:
ALLUVIAL
a._
(/)(.9:'! <(t:lw <(I�cr: Ugj
LL
--- -w
FAN
I
<(cr:cr: 0cr: 0
Facies assemblage
/
1
/
Obscurred by Quaternary ' deposits
a._
800
r----
LACUSTRINE
I
I I
600 I
ALLUVIAL
5;i= I-<( <(�
KEY Clast-supported breccia Pebby sandstone
I
z zo
Massive or graded sandstone
FAN
I I I
s:� LL
Mudstone
I I
LACUSTRI NE GREENLAND GROUP/ GRAr-.'ITE
I I
° Fault +------
3·2 km-------1-8 km--+
Fig. 3. Stratigraphical columns showing relationships between the major formations and facies assemblages of the Pororari
Group.
Fig. 4. Hogback and cliffed
topography developed on Hawks Crag Breccia (Pororari Group), with steep east-facing dip slopes. The prominently bedded unit in the upper middle of the photograph is 20-30m thick. The valley of the upper Pororari River is in the foreground: crest of the Paparoa Range is in the background.
202
M.G. Laird
Fig. 5. Thick matrix-supported (upper) and clast-supported (lower) breccia beds, Fox River mouth. Beds young towards the left. Hammer for scale . Photograph courtesy of L. Homer.
Fig. 6. Succession of thin clast supported breccia beds, Fox River mouth. Note normal grading in the bed below the geological hammer. Some of these beds may represent the deposits of highly sediment charged sheet flows. Beds young upwards. Hammer for scale.
Clast-supported breccia makes up a minority of the beds (Fig. 5). Breccia beds most commonly lie in the range 1 -6 m thick, but some breccia units (probably composite) may reach 20-30 m in thickness and extend with little thickness change for at least several hundred metres (Fig. 4). Most lack internal stratifi cation and bedding is in many cases indistinct. Mud stone is absent, and thin sandstone beds up to 20 em thick occur only rarely. In rare instances clast supported breccia forms well-defined beds, usually
less than 1m thick but reaching a maximum of 3 m in thickness, and separated by thin (up to 10 em) beds of poorly sorted coarse to pebbly sandstone (Fig. 6). Poorly developed symmetrical grading, reverse grading, or normal grading occurs locally (Figs 6 & 7), although this is uncommon. Individual clasts reach a maximum size of 3m, and in some instances show imbrication and preferred orientation of long axes. Carbonaceous material is common, particularly at
Cretaceous rift deposits of New Zealand
203
Fig. 7. Thick, matrix-supported
breccia bed, with thin sandstone bed at base, Pororari River. Note lack of internal structure, but the presence of symmetrical grading. Beds young to left. Hammer for scale.
the interfaces between beds, and small coalified logs occur rarely. At the mouth of Fox River several thick beds of poorly sorted breccia show crude fining-upward grading into thin graded sandstone (in some instances showing synsedimentary defor mation structures) and then into coaly material. No root horizons have been recognized, and all plant material appears to have been transported from its source. Shallow scours or channels, up to 1m deep, occur locally. Other sedimentary structures, apart from rare synsedimentary deformation features, are notably absent. Interpretation The preponderance of poorly sorted, poorly strati fied matrix-supported breccia with common outsize clasts, a subordinate proportion of clast-supported breccia, the rarity of sedimentary structures within breccia beds, and the absence of cross-stratification and only minor presence of thin well-bedded sand stones, suggests that mass wasting events played a predominant role in the formation of assemblage 1. The major facies of matrix-supported breccia closely resembles the deposits of debris flows, which domi nate sedimentary processes on many modern alluvial fans, whereas the subordinate clast-supported breccia is characteristic of associated debris flow levees (cf. Pierson, 1980; Blair & McPherson, 1992). Thin, moderately well-sorted sandstone beds and thin-bedded breccia and sandstone couplets, forming a minor proportion of assemblage 1, closely resemble facies prominently represented in some modern and
Cenozoic alluvial fans (cf. Ballance, 1984; Blair, 1987), and attributed to the deposits of sheet flows emplaced during catastrophic floods. The breccia beds at Fox River mouth, which pass upwards into graded sandstone and then into coaly material, may represent the results of flash floods and the delayed settling out of fines, with transported vegetation being the last to settle. Assemblage 1 breccias are thus inferred to have been deposited on a series of alluvial fans, dominated by debris-flow depositional processes, closely similar to the humid-temperate climate, late Quaternary and modern alluvial fans that are characteristic of the mountainous regions of New Zealand (cf. Pierson, 1980, 1981; McArthur, 1987) and elsewhere (e.g. Kochel & Johnson, 1984). The abundance of debris flows, and the poor sorting and angularity of many of the clasts suggests that source slopes were steep and lay in close proximity. The common occurrence of carbonaceous material, including rare coalified logs, indicates that the source slopes were vegetated. No fine-grained subfacies potentially representa tive of interdistributary deposits (cf. Lewis & Ekdale, 1991) has been recognized within the assem blage, and fans were therefore likely to have been coalescing. A ssemblage 2: pebbly sandstone and massive or graded sandstone
Facies descriptions This facies assemblage makes up most of the Watson
204
M.G. Laird
and the Bovis Formations. It consists dominantly of matrix-supported conglomerate and pebbly sand-. stone, graded or massive very fine-grained to coarse grained sandstone, and alternating sandstone and mudstone successions. Thickness is variable, but an aggregate thickness of 600 m is reached in the Watson Formation exposed in Pororari River, and a maximum thickness for the Bovis Formation is approximately 400 m (in Slaty Creek}. Clasts in the conglomerate and pebbly sandstone of assemblage 2 are notably better rounded than in the breccia of assemblage 1, and all beds are matrix-supported. Conglomerate and pebbly sandstone beds, which are up to 2 m thick, lack internal sedimentary struc-
tures, and most are ungraded. Clasts commonly occupy specific horizons within each bed, but random distribution also occurs. Clasts are of basement gran ite or Palaeozoic sandstone, and range up to 0.5 m in size. Imbrication is relatively common (Fig. 8). Inter leaved very fine- to coarse-grained, moderately well sorted sandstones vary in thickness up to 1 m, and include both graded and ungraded beds (Fig. 9). Some have scoured bases, and show sole marks consisting mainly of grooves and less common prod marks or flutes. Primary current lineations and cur rent ripples are also present on bedding planes and parallel-lamination is common in the finer sandstone beds. Isolated horizons of current-rippled sandstone
Fig. 8. Imbricated pebbles of
granite and of Greenland Group sediments supported by a matrix of very coarse to gritty sandstone, Bovis Formation, Slaty Creek (at approximately 500m, Slaty Creek column, Fig. 3). Palaeocurrent direction from right to left (N-S). Beds young towards top right. Note the thin interbeds of carbonaceous mudstone. Hammer for scale.
Fig. 9. Amalgamated beds of fine
sandstone showing no internal stratification or grading, in the lower portion of the Watson Formation, Pororari River (at 120m in column, Fig. 12). Beds young towards top. Pocket knife is 9cm long.
Cretaceous rift deposits of New Zealand are common in the intervening mudstones (Fig. 10). Load casts and flame structures are common at the interface between sandstone and mudstone beds (Fig. 11), suggesting rapid deposition, and some thin sandstone layers are intersected by small low-angle synsedimentary thrust faults. The thrusting is in the same sense as the overturning of flame structures and of other palaeocurrent indicators. Widely sep arated burrowed horizons also occur. Some variation within the facies assemblage is evident between the Watson and Bovis Formations. The Watson Formation is characterized by inter leaved components of facies assemblages 1 and 2, although assemblage 2 is dominant (Fig. 12). The
Fig. 10. Mudstones of the Watson Formation containing thin massive or laminated beds of fine-grained sandstone, and horizons of isolated ripples (same locality as Fig. 9). Beds young upwards. Pocket knife is 9cm long.
Fig. 11. Load casts and flame structures in very fine-grained sandstone beds, lower portion of the Watson Formation, Pororari River (at 110m in column, Fig. 12). Palaeocurrents flowed from left to right (N-S). Beds young towards top. Pocket knife is 9cm long.
205
formation is particularly well-exposed in Pororari River and south-flowing tributaries, although its basal contact is faulted. In the adjacent Tindale Creek, the basal unit consists of clast-supported breccia consisting entirely of Palaeozoic sandstone clasts resting uncomformably on Palaeozoic strata (Fig. 3). This basal unit, and a petrographically similar 100-m-thick unit of clast-supported-breccia occurring about 300 m above the basal fault in Pororari River, are considered to represent an inter calation of Assemblage 1. The remainder of the formation is dominated by alternating sandstone and mudstone, including common slumped and hydroplastically deformed beds (Fig. 13). These beds
206
M.G. Laird
KEY v Sole mark
Slump u Channel
fj, 17 V
Fining-upward units Coarsening-upward units
Sequence
Palaeocurrent Plots:
• Sole mark W Parting lineation • Slumping direction [] Cross lamination --+
Imbrication direction
Lithology and Facies Assemblage
Palaeocurrents
Palaeoenvironment
M 0 <{<{ a:
u U (flu
Clast-supported breccia, dominated by granitic clasts (A 1)
�w
�� <{
ALLUVIAL FAN
J:
�----- N
--------1-----------------------+------------� � Alternating sandstone and mudstone, with minor clast-and-matrix BRAID PLAIN supported breccia and conglomerate. AND Rare cross-bedding. Channelling LACUSTRINE · · and small fining-upward sequences ': FAN common in the upper part (A2) n = 11
N
Alternating thin-bedded sandstone and mudstone, organized into coarsening upward sequences (A2)
LACUSTRINE FAN
Clast-supported breccia, clasts consisting of Palaeozoic sandstone(A1)
ALLUVIAL FAN
z 0
�
n
:::i:
=
21
a: 0 lL z 0 (/)
u
� �
\
Alternating thin-bedded sandstone and mudstone (A2)
LACUSTRINE FAN
v
Fig. 12. Detailed stratigraphical column of the Watson Formation, Pororari River, showing main sedimentary characteristics, palaeocurrent plots, and inferred palaeoenvironments. Al, A2, facies assemblages 1 and 2.
Cretaceous rift deposits of New Zealand
207
Fig. 13. Slumped strata near the
base of the Watson Formation, Pororari River (at approximately 20m in column, Fig. 12). Beds young towards top of photograph. Pocket knife is 9em long.
commonly alternate with packets of thick-bedded, graded or massive sandstone. Some 460 m above the faulted base of the section, 6 m of pebbly mudstone occurs, containing indurated sandstone pebbles up to 20 em, and large rafts of thinly bedded sandstone up to 140 cm long, some showing evidence of syn sedimentary deformation. This unit is interpreted to be a mass flow deposit. Above this prominent slump horizon, a series of thickening and coarsening upward units occurs, varying in thickness from 15 to 110m. The base of each unit consists of laminated mudstone, passing upwards into alternating mud stone and thin sandstone beds, and finally up into thick, often amalgamated sandstone beds up to 1 m
Fig. 14. Thick, massive fine-grained
sandstone beds (upper one with a scoured base) in a fining upwards succession, middle portion of the Watson Formation, Pororari River (at 650m in column, Fig. 12). Note pseudonodule horizon below the lower sandstone bed. Beds young towards top of photograph). Pocket knife is 9em long.
thick (Fig. 14). Small scours or channels up to 0. 5 m deep, infilled with fine-grained sandstone, in some instances slumped, occur at rare intervals in the succession (Fig. 15). A change in lithofacies occurs at about 800 m above the base of the Pororari River section. At this horizon, 5 m of cross-bedded sandstone occurs, the cross-bedded sets averaging 10-20 em in thickness. Some sets have suffered hydroplastic deformation. At about the same level, thin beds of matrix- or clast-supported graded breccia, the clasts consisting of granite fragments, start to appear, commonly with erosive bases (Fig. 16). Units of alternating sandstone and mudstone, although still common,
208
M.G. Laird
Fig. 15. Channel cut in mudstone and infilled with slumped sandstone, Watson Formation, Pororari River (at 440m in column, Fig. 12). Beds young towards top of photograph. Hammer for scale.
Fig. 16. Thin graded breccia beds in the upper part of the Watson Formation (at approximately 820m in column, Fig. 12). Note slumped laminated mudstones between the breccia beds. Beds young to the left of the photograph.
become less frequent upwards. The top 10-20 m of the Watson Formation is dominated by matrix supported conglomerate, carbonaceous sandstone, and thin coaly layers in a series of fining upward units 1. 5-3 m thick. Each unit comprises a basal conglomerate consisting of moderately well-rounded pebbles of basement rocks, predominantly granite, up to 15 em in diameter, commonly resting on a scoured surface, passing upwards into carbonaceous parallel- or cross-bedded very coarse sandstone, and then into thin beds of carbonaceous siltstone or coal seams up to 20 em thick (Fig. 17). The fining upwards units pass with rapid gradation upwards into thick
massive matrix-supported breccia beds compnsmg the basal beds of the Hawks Crag Breccia. In the Bovis Formation, assemblage 2 forms part of one systematically fining upwards succession, with only minor intercalations of other facies associations compared with the Watson Formation. There is an upward gradation from matrix-supported breccia of assemblage 1 into metre-bedded pebbly coarse sand stone of assemblage 2, containing scattered sub angular and subrounded clasts of basement rocks up to 1m (Figs 8 & 9), with minor interbeds of decimet:re to metre-bedded medium- to coarse-grained massive sandstone. The proportion of conglomerate to sand-
Cretaceous rift deposits of New Zealand
209
Poorly-stratified matrix-supported conglomerate, scoured at base Cross-bedded carbonaceous sandstone
Poorly-stratified matrix-supported conglomerate, scoured at base
Carbonaceous siltstone(lenses)
Poorly-stratified matrix-supported conglomerate, scoured at base, and fining upwards
Carbonaceous sandstone Coal Carbonaceous sandstone
Poorly-stratified matrix-supported conglomerate, scoured at base
Fig. 17. Sketch of conglomeratic
fining upwards units at the top of the Watson Formation, Pororari River (940m in column, Fig. 12).
stone beds decreases upwards, and in its upper portions the succession consists almost entirely of massive or normally graded sandstone beds separ ated by thin layers of carbonaceous mudstone (Fig. 18). Some of the massive sandstone beds contain pebble trains or scattered clasts, the clast size decreasing upwards in the succession. The top few tens of metres of the Bovis Formation consists of centimetre to decimetre graded very fine to fine grained sandstone alternating with mudstone. Initially thin mudstone interbeds between sand stones thicken upwards until near the top of the unit, the predominant lithofacies is alternating mud stone and fine-grained graded or massive sandstone. Flame-structures are commonly developed at the base of sandstone beds, and show a uniform direction of overturning. Sole structures, mainly grooves, but a few flutes, also occur. Burrows occur locally.
Poorly-stratified sandstone
Interpretation The high proportion of mudstone throughout assem blage 2 in the Watson Formation and in the upper portions of the Bovis Formation, preponderance of sediment gravity flow deposits and scarcity of traction current deposits, and the occurrence of rare bur rowed horizons suggests sedimentation in a standing water body. The highly carbonaceous nature of the deposits and the absence of shelly fossils suggests that this is likely to have been a lake. The very poor sorting of the matrix-supported breccia beds, the presence of megaclasts in some beds, and absence of sedimentary structures suggests deposition of these beds by subaqueous debris flows. Common loads and flame structures at the interface between sand stone and mudstone beds suggest rapid deposition of the coarse sediments, and the increasingly common
210
M.G. Laird
Fig. 18. Thin, massive to poorly
graded fine- to medium-grained sandstone beds with thin partings of carbonaceous mudstone, near the top of the Bovis Formation, Slaty Creek. Beds young towards top of photograph. Hammer for scale.
presence of normally graded bedded sandstones upwards in the Bovis Formation suggests deposition by turbidity currents. The common occurrence of slump deposits in the Watson Formation suggests the presence of a significant slope throughout much of its depositional history. The thin fining upwards successions near the top of the formation in Watson Creek are inferred to have resulted from a high-.energy braided stream system developed at the toe of a Hawks Crag alluvial fan, which prograded over a vegetated coastal flat between the advancing alluvial fan and the standing water body. Miall's (1978) 'Scott-type' braided stream model, considered typical of proximal braided stream deposits on alluvial fans, is prob ably a close analogue. The Scott River was also represented as typically containing minor interbeds of cross-bedded sand and small fining upwards cycles. A similar environment has been inferred by Hamblin (1992) for parts of the Lower Carboniferous Horton Group of Nova Scotia.
Slaty Creek. Near the base, assemblage 3 contains mudstone interbedded with decimetre-bedded fine to very fine-grained, parallel laminated sandstone. Sandstone beds are commonly normally graded, and have sharp bases and tops. This facies passes rapidly up into finely laminated, dark, carbonaceous mudstone (Fig. 19). Sole marks, consisting mainly of grooves, but also with rare prod and flute casts, occur on the base of some sandstone beds. Isolated vertical burrows occur at some horizons. Rare slumped units up to 1. 5 m thick also are present near the base of the assemblage. In Slaty Creek, a horizon of coal 0. 5 m thick, bounded above and below by dark mudstone with prominent vertical burrows, occurs a few metres above the gradational contact between assemblages 2 and 3. The coal could not be traced more than 5 m laterally and is almost certainly lenticular. No root horizons are present, and the coal is inferred to represent allochthonous material. Interpretation
Assemblage 3: laminated mudstone
Facies descriptions This facies assemblage, which makes up the bulk of the Bullock Formation, consists dominantly of dark brown to grey, highly carbonaceous laminated mud stone. It passes upwards with rapid gradation from alternating thin sandstones and mudstones of assem blage 2, attaining a thickness in excess of 500 m in
The ubiquitous presence of highly carbonaceous mudstone, absence of traction current deposits, and absence of marine shelly fossils but presence of burrowed horizons, suggests deposition in a standing body of water, inferred to be a lake. The rare occurrence of sediment gravity flow deposits and slumps suggests that deposition occurred distal to the sediment distributary system, and probably well out on the floor of the lake basin.
Cretaceous rift deposits of New Zealand
211
Fig. 19. Laminated carbonaceous
mudstone, Bullock Formation, Slaty Creek. Beds young towards top of photograph. Hammer for scale.
D EPO SIT IO N A L
SYS T E M
O F THE PO R O R A R I G R O U P
The three facies assemblages intertongue, and are obviously closely related portions of a common depositional system. The Watson, Hawks Crag Breccia, and Bovis Formations of the western Paparoa Range, and their associated facies assem blages also show rapid systematic thinning from north to south (Fig. 3). This, together with the nature of the interrelationships between the facies assemblages supports the interpretation of the Pororari succession as a lacustrine fan-delta system. The alluvial fan deposits of assemblage 1 are inferred to have fed either directly , or via a narrow braid plain, into a lake, where deposition continued on a fan consisting of assemblage 2 proximal deposits of the Watson and Bovis Formations. Lacustrine, mainly quiet-water conditions, are represented by the assemblage 3 mudstones of the Bullock Forma tion. A simplified reconstruction showing the inferred relationship of assemblages to each other, and the evolution of the fan delta system, is given in Fig. 20. The most informative profile is that given by the Pororari River and its tributaries (Fig. 12). Here there is clear evidence of a fluctuating lake shoreline, with irregular progradation of the subaerial portion of an alluvial fan into the lake, and subsequent periods of transgression. The presence of 165m of basal breccia in Tindale Creek suggests that depo sition started with an alluvial fan. Subsequent
inundation by a relative rise in lake level and depo sition of proximal lacustrine sediments is inferred to account for association 2, with a brief re-establish ment of an alluvial fan represented by association 1 deposits from 32 5 to 425 m in the Pororari River section (Fig. 12). The lacustrine deposits of the Watson Formation were increasingly encroached on and finally overwhelmed by the alluvial fans making up the Hawks Crag Breccia, which represented the major alluvial fan development of the Pororari Group. The end of Hawks Crag Breccia times shows increasing incursion and swamping of the fan by first proximal and then distal lacustrine associations in all areas of outcrop. The thickening and coarsening upward successions in the middle part of the Watson Formation may have been autocyclic, i.e. represented periodic diversion of the main feeder channels of a lacustrine delta and the build-up of a new pro-delta lobe, with accommodation provided by continued subsidence of the basin. However, the repetitive nature of the successions, and the evidence for major changes in base level outlined above, suggest that a more likely cause was fluctuations in lake level. The mudstone at the base of each cycle is inferred to represent a flooding event caused by a rise in lake level, the subsequent development of the thickening upward succession caused by lobes of sediment prograding into the lake during relative high-stand. The fluctu ations in lake level could have been brought about by a variety of mechanisms, including tectonic activity on a fault or faults controlling the depo-
212
M.G. Laird
N
s Lake surface a
Cl Cl
Cl
Cl
l:.
Cl
Cl
Cl
Cl
Cl
Cl
Cl
Alluvial fan
l:. l:.
Cl Cl
l:.
l:.
Cl
l:.
l:. l:.
Cl l:.
Cl
l:.
l:. l:.
l:.
l:.
BRECCIA l:.
l:. Cl
Cl l:.
CRAG
l:. l:.
l:. Cl
Cl Cl
l:.
Cl
Assemblage 1 Cl
l:. l:.
l:.
l:. l:.
l:. l:.
Cl
Fig. 20. Palaeoenvironmental/tectonic sketch (not to scale), showing the inferred relationships of the facies assemblages
making up the Pororari Group. a-a' shows the approximate location of the Pororari River measured section (Fig. 3).
sitional basin, or by climatic cycles. Cyclic suc cessions present in the dominantly lacustrine early Mesozoic Newark Supergroup of eastern North America have been attributed to climate-controlled high-frequency fluctuations in the depth of rift lakes (e.g. Olsen, 1986, 1991).
P A L A E O C U R R E N T PAT T E R N
Palaeocurrents (summarized in Fig. 2 ) were deter mined from a wide variety of sedimentary features, including grooves, prods, flutes, primary current lineation, channel margins, and cross-lamination, and palaeoslopes from slump fold axes in the lacus trine facies associations 2 and 3. Palaeocurrent indicators are relatively common in the Watson Formation, and in the Bovis and Bullock Formations of the western Paparoa Range. Although no suitable bedding surfaces were exposed for measurement of palaeocurrent directions in the Bovis Formation of the eastern Paparoa Range, numerous overturned flame structures, and local occurrence of imbrication in pebbly beds indicate a consistent general direction
of transport. Only a small number of directional features were recorded in the Bullock Formation in the eastern Paparoa Range. The palaeocurrent and palaeoslope trend in the Watson Formation is generally from north to south. However, in the thick, well-exposed succession in the Pororari River and its tributaries, plotting palaeocurrents in segments according to facies groups clearly shows that significant variations occur within the succession (Fig. 12). The lower, poorly exposed 300 m, consisting of alternating sandstones and mudstones of association 2, has palaeocurrents directed towards the south or south-southwest. By contrast, the similar facies higher in the column between 430 and 800 m in the section has palaeocur rents directed towards the south-southeast. The sole imbrication measurement recorded in the inter vening 100-m unit of assemblage 1 breccia (between 330 and 430 m, section Fig. 12), although statistically insignificant, is also directed to the south-southeast, suggesting a source in common with the overlying sediments, but differing from the source providing the earlier sediments. The 100-m thick breccia unit is thus inferred to represent an abrupt incursion of
213
Cretaceous rift deposits of New Zealand an alluvial fan into the lake. The monolithological nature of clasts, which consist of indurated basement sandstone, and their high degree of angularity, suggests a very proximal source. The same fai'!_. prob ably continued to feed clastic detritus in 'a 0uth southeasterly direction into the lake up until:a�:'reast the 800 m mark (Fig. 12), an inference supportetl by the fact that the clasts in the 6-m bed of slumped pebbly mudstone 20m above the breccia (at 460 m, Fig. 12) also consist largely of indurated basement sandstone. The upper part of the Watson Formation, between 800 and 950 m in Fig. 12, although dominated by assemblage 2 facies, includes thin beds of breccia and conglomerate containing granitic clasts, indi cating a different provenance. This mixture of sources is reflected by the wide variation in current directions recorded, ranging through southeast to southwest. Southwest and west-southwest directions are indicated by clast imbrication on two surfaces in the overlying Hawks Crag Breccia (Fig. 12), and a south-southwest direction of transport is recorded overall for the Hawks Crag Breccia in the study area (Fig. 2). Palaeocurrent data were not available from the Watson Formation in the adjacent Tindale Creek, the only other area of major outcrop, thus these changes of palaeocurrent trend could not be confirmed as a regional phenomenon. However, the inference is that at least three distinct sources and fans influenced sedimentation: the earliest from the north or north-northeast; the next from a Palaeozoic area to the north-northwest, and the upper part of the Watson Formation is increasingly influenced by debris derived from a granitic source to the north east. Derivation of indurated sandstone clasts in the second fan from the north-northwest is consistent with the presence of Early Palaeozoic turbidites of the Greenland Group cropping out to the west of the Pororari Group. However, basement rocks cur rently lying to the northeast of the Hawks Crag breccia and extending for many kilometres consist entirely of foliated gneiss of the Charleston Meta morphic Group (Nathan, 1978; Laird, 1988), a lith ology almost absent from the Hawks Crag Breccia, which is dominated by granitic clasts. This problem of derivation is discussed below under Tectonic Implications. Clast imbrication is the only usable palaeocurrent indicator in the alluvial fan assemblage represented by the Hawks Crag Breccia, which is dominated by granitic clasts. Suitable surfaces for recording imbri-
s
cation were relatively uncommon however, and only a few measurements were obtained. Although there is some variation, a dominant NNE-SSW, or NE-SW trend is apparent. In the Bovis Formation, palaeocurrents are widely variable (Fig. 2), but still show a generally southerly trend. The information from sole marks and other features is strongly reinforced by the direction of overturning of flame structures, which occur at the base of many sandstone beds. Sole marks from the Bullock Formation show a closely constrained and consistent trend from north-northeast to south southwest (Fig. 2). The consistent thinning of the Watson Formation, Hawks Crag Breccia, and Bovis Formations from north to south in the western Paparoa Range also supports the inference of a transport direction in a generally N-S direction.
T E C T O NIC
1M PLIC A T I O N S
I t has been widely recognized that fan-deltas are preferentially located in areas of active tectonism, and are commonly controlled by an active fault or faults (e.g. Ethridge & Wescott, 1984). The great thickness of the clastic wedge (a minimum of 1200 m for the Hawks Crag Breccia alone), and the relative angularity of clasts, suggests that they were gener ated by an active fault system in close proximity. The evidence suggests that the fan-delta system occu pied a half-graben, downthrown to the south (Fig. 20). The facies relationships, with thick coarse sub aerial deposits concentrated in the north, and rapid thinning and fining to the south, i.e. downcurrent, also closely fits the model proposed by Surlyk (1978) for fan-delta sedimentation in a half-graben. Move ment on the inferred bounding fault to the north continued during much of the deposition of the Pororari Group, and was responsible not only for generating the coarse facies of the Hawks Crag Breccia and the associated lacustrine sediment grav ity flows, but also for the necessary subsidence to provide accommodation for the great thickness of the deposits. Contemporaneous fault movement may also have been a factor in causing the fluctuations in lake level noted earlier, although climate also may have played a role. Judging from the consistently s-ssw directed sediment gravity flows, both sub aerial and subaqueous, the active fault scarp was likely to have been oriented in a westerly or west-
214
M.G. Laird South Island New Zealand
N
t
� "i
10
0
-
km
KEY -
c;:::?
9
;>
Late Cenozoic fault Mid Cretaceous fault(?) Palaeocurrent azimuth
Pororari group Greymouth
[I] m j...: - : l :
Upper lake beds Hawks Crag Breccia Lower lake beds
northwest direction with an implied N-NNE direc tion of tectonic extension (Fig. 21). As noted previously, the similarity of Lithofacies, stratigraphical position, and palaeocurrent pattern of the Hawks Crag Breccia and, in particular, the overlying lake beds in the now widely separated outcrops on the eastern and western flanks of the Paparoa Range suggests that they were formed in the same depositional basin. However, the thick nesses of the Hawks Crag Breccia and the overlying
Fig. 21. Map showing generalized
palaeocurrent trends and inferred alignment of the mid-Cretaceous fault bounding the half-graben containing the Pororari Group.
lacustrine strata in the outcrop area on the eastern flank of the Paparoa Range, which lies south southeast of that on the western flank, i.e. downcur rent, do not conform with the southward thinning pattern of units clearly shown by the western out crops. The thickness of the Hawks Crag Breccia of the eastern outcrop is similar to the maximum thick ness in the north of the western unit, but the over lying strata of association 2 are more than twice the thickness of the western maximum. This is hard to
Cretaceous rift deposits of New Zealand explain by invoking a change in sedimentation pat tern, as the palaeocurrent pattern is essentially indis tinguishable between the two outcrop areas. A more likely explanation is that transcurrent movement along a north-northeast trending fault (Hawera Fault?) subsequent to deposition has offset the eastern part of the half graben to the south by at least 10-15 km. It was noted previously that the clasts making up the Hawks Crag Breccia consist almost entirely of granite , inferred to have been derived from a proxi mal source to the northeast. The basement rocks to the northeast and north, however, consist for many kilometres of foliated gneiss of the Charleston Meta morphic Group, a lithology that occurs only rarely in the Hawks Crag Breccia. An added apparent complexity is the evidence provided by Tulloch & Palmer (1990) that granite samples typical of the bulk of the Hawks Crag Breccia at the Fox River mouth and from several large exposures in Bullock Creek are similar to the Buckland Pluton, the closest outcrop of which lies 20 km to the northeast. The clasts are distinct from the Meybille Granite , which crops out to the immediate south of Fox River mouth. The inference drawn by Tulloch & Palmer (1990) is that, at the time of their deposition, either the Fox River and Bullock Creek deposits were situated 20-2 5 km to the northeast, or the eroded upper parts of the Buckland Pluton had a consider ably greater horizontal extent. Tulloch & Kimbrough (1989) inferred that the Greenland and Pororari Groups, together with the granite from which the latter had b'een derived, were part of an upper plate that detached from a metamorphic core complex (Charleston Metamorphic Group) which comprised a lower plate. On this hypothesis, dislocation of the Hawks Crag Breccia from its original source is possible. A possible analogue of the southern Paparoa Range half-graben has been recognized in seismic profiles offshore from Greymouth, 30 km to the southwest (Bishop, 1992; Fig. 21). This half-graben is also aligned WNW-ESE, and is inferred to be infilled largely with Pororari Group sediments. It is clearly bounded to the north by a southward-dipping normal fault, and to the south by a pinchout. It is divided by major north-northeast striking transfer faults into several segments. The orientation of the half-graben and bounding normal fault clearly sup ports the direction of tectonic extension suggested for the southern Paparoa Range half-graben. The inferred E-W to SE-NW oriented half-graben
215
infilled by the Beebys Conglomerate (Johnston , 1990; Fig. 1) is also likely to have been formed during the same period of north-northeast extension. These data fit well with the north-northeast exten sion direction for the Paparoa Range area during late Aptian to early Albian times deduced from structural research on the basement noted above. Tulloch & Kimbrough (1989) provided evidence that stretching lineations associated with low-angle normal detachment faults separating the upper from the lower plate defined a regionally consistent north northeast extensional trend, which coincides with that inferred from the half-graben. It has been suggested (Laird, 1993, 1994) that half-graben formation and synchronous infilling were associated with a New-Zealand-wide early period of extension beginning in the mid-Albian, pre-dating by approximately 2 5 Ma the separation of New Zealand from Gondwana and sea-floor spreading in the Tasman Sea.
CONCLUSIONS 1 The Pororari Group of the southern Paparoa Range represents a fault-controlled lacustrine fan delta complex. The Hawks Crag Breccia and equiv alent matrix- and clast-supported breccias represent deposits of humid-temperate climate alluvial fans, while intertonguing and overlying finer grained sedi ments were deposited predominantly in a lacustrine environment. 2 Mass-flow processes dominated both alluvial fan and lacustrine sedimentation, suggesting that tec tonic activity continued throughout most of the period of deposition of the Pororari Group, dying away only towards the end. The abundance of debris flow deposits in the alluvial fan association, and the poor-sorting and angularity of many of the clasts suggests that source slopes were steep and lay in close proximity. The ubiquitous occurrence of debris-flow deposits, turbidites and slumps in the proximal lacustrine succession indicates the presence of a significant slope on the subaqueous fan, which may account for the limited presence of deposits transitional between the alluvial and sublacustrine fans. 3 Fluctuations in lake level are indicated by episodic transgressions and regressions of the lake with res pect to the adjacent alluvial-fan system, and by the presence of repetitive thickening and coarsening upwards successions in the proximal lacustrine
216
M .G. Laird
deposits of the Watson Formation. These fluctuations may have been caused by active faulting at the basin margin, by climatic changes, or by a combination of both. 4 The fan-delta system is inferred to have occupied an actively subsiding half-graben. The consistently south-southwest directed sediment gravity flows and rapid southward thinning of the sedimentary wedge suggests that the active basin margin fault was likely to have been oriented in a WNW-ESE direction. The implied NNE-SSW direction of tectonic exten sion compares well with a similar extension direction determined for this area from structural research on basement rocks, and with that of a probable mid Cretaceous half-graben recognized on offshore seis mic profiles. The orientation of the half-grabens is essentially parallel to the future spreading axis associ ated with the opening of the adjacent Tasman Sea. This opening occurred at about 80 Ma, suggesting that initial rifting and half-graben formation occurred 20-25 Ma before sea-floor spreading began.
ACKNOWLEDGEMENTS
I am grateful to D . Lewis, J . Lindqvist and A. Tulloch for stimulating discussions, both in the field and in the office, which helped to clarify my ideas on processes of emplacement and geological setting of the Pororari Group. Thorough reviews by T. Astin, S. Flint, D. Lewis, J. Lindqvist and G. Plint greatly improved the manuscript and are appreciated.
REFERENCES
P.F. (1984) Sheet-flow-dominated gravel fans of the non-marine middle Cenozoic Simmler Formation, central California. Sediment. Geol. , 38, 337-359. BISHOP, D.J. (1992) Extensional tectonism and magmatism during the middle Cretaceous to Paleocene, North Westland, New Zealand. N.Z. J. Geo/. Geophys. 35, 81-91. BISHOP, D.G. & LAIRD, M.G. (1976) Stratigraphy and depositional environment of the Kyeburn Formation (Cretaceous), a wedge of coarse terrestrial sediments in Central Otago. J. R. Soc. N.Z. , 6, 55-7 1. BLAIR, T.C. (1987) Sedimentary processes, vertical stratifi cation sequences, and geomorphology of the Roaring River alluvial fan, Rocky Mountain National Park, Colorado. J. sediment. Petrol. 57, 1 - 18. BLAIR, T.C. & McPHERSON, J.G. (1992) The Trollheim alluvial fan and facies model revisited. Geol. Soc. Am. Bull. , 104, 762-769. BRADSHAW, J.D. (1989) Cretaceous geotectonic patterns in the New Zealand region. Tectonics, 8, 803-820.
BALLANCE,
F.G. & WESCOTT, .W.A. (1984) Tectonic setting, recognition and hydrocarbon reservoir potential of fan delta deposits. In: Sedimentology of Gravels and Con·· glomerates (Eds Koster, E.H. & Steel, R. J . ), Mem. Can. Soc. petrol. Geol., Calgary, 10, 217-235. HAMBLIN, A.P. (1992) Half-graben lacustrine sedimentary rocks of the Lower Carboniferous Strathlorne Formation, Horton Group, Cape Breton Island, Nova Scotia, Canada. Sedimentology, 39, 263-284. JOHNSTON , M.R. (1990) Geology of the St Arnaud District, Southeast Nelson (Sheet N29). New Zealand Geological Survey Bulletin 99. New Zealand Geological Survey, Lower Hutt, New Zealand, 119 pp. KocHEL, R.C. & JOHNSON, R.A. (1984) Geomorphology and sedimentology of humid-temperate alluvial fans, central Virginia. In: Sedimentology of Gravels and Con glomerates (Eds Koster, E . H. & Steel, R.J.). Mem. Can. Soc. petrol. Geol., Calgary, 10, 109-122. KoRSCH , R.J. & WELLMAN, H.W. (1988) The geological! evolution of New Zealand and the New Zealand region. In: The Ocean Basins and Margins, Vol. 7B (Eds Nairn, A.E.M., Stehli, F.G. & Uyeda, S. ), pp. 411-482. Plenum, New York. LAIRD, M.G. (1988) Sheet S37 Punakaiki. Geological map of New Zealand 1 : 63 360. New Zealand Geologi cal Survey, Department of Scientific and Industrial Research, Wellington. LAIRD, M.G. (1993 ) Cretaceous continental rifts. New Zealand Region. In: Sedimentary Basins of the World. South Pacific Sedimentary Basins (Ed. Ballance, P.F. ), pp. 37-49. Elsevier, Amsterdam. LAIRD, M.G. (1994) Geological aspects of the opening of the Tasman Sea. In: The Evolution of the Tasman Sea Basin (Eds van der Lingen, G.J., Swanson, K. & Muir, R.J.), pp. 1-17. Balkema, Rotterdam. LEWIS, D.W. & EKDALE, A.A. (1991 ) Lithofacies relation ships in a late Quaternary gravel and loess fan delta complex, New Zealand. Palaeogeogr. , Palaeoclimatol. , Palaeoeco/. , 81, 229-251. LIN DQVIST, J.K. (1990) Puysegur Group: a mid Cretaceous lacustrine fan-delta complex, Balleny Basin, southwest Fiordland. Geol. Soc. N. Z. Misc. Pub/. , SOA, 83. McARTHUR, J.L. (1987) The characteristics classification, and origin of Late Pleistocene fan deposits in the Cass Basin, Canterbury, New Zealand. Sedimentology, 34, 459-471. McKAY, A. (1883) On the geology of the Reefton District, lnangahua County. Geol. Survey Reports during 1882 , Vol. 15, pp. 142-144. Wellington. MLALL, A.D. (1978) Lithofacies types and vertical profile models in braided river deposits: a summary. In: Fluvial Sedimentology (Ed. Miall, A.D.), Mem. Can. Soc. petrol. Geol., Calgary, 5, 597-604. MORGAN , P.G. & BARTRUM, J.A. (1915) The geological and mineral resources of Buller-Mokihinui Subdivision, Westport Division. Geological Survey Bulletin No. 1 7 (new series) . New Zealand Department of Mines, Geo logical Survey Branch, Wellington. NATHAN , S. (1978) Sheet S31 & Part S32 Buller- Lyell. Geological Map of New Zealand 1 : 63 360. New Zealand Geological Survey, Department of Scientific and Indus trial Research, Wellington. OLSEN, P.E. (1986) A 40-million-year lake record of early Mesozoic orbital climatic forcing. Science, 234, 842-848. ETHRIDGE,
Cretaceous rift deposits of New Zealand P.E. ( 1991) Tectonic, climatic, and biotic modu lation of lacustrine ecosystems - examples from Newark Supergroup of eastern North America. In: Lacustrine Basin Exploration. Case Studies and Modern Analogs (Ed. Katz, B.J.) , Mem. Am. Assoc. petrol. Geol. , Tulsa, 50, 209-224. PIERSO N , T.C. ( 1 980) Erosion and deposition by debris flows at Mt Thomas, north Canterbury, New Zealand. Earth Surf Process. , 5, 227-247. PIERSON, T.C. ( 1981) Debris flows. An important process in high country gully erosion. 1. Tussock Grasslands Mountain Lands Inst. , Rev. , 39, 3- 14. RAINE, J.l. ( 1984) Outline of a palynological zonation of Cretaceous to Paleogene terrestrial sediments in West Coast region, South Island, New Zealand. New Zealand OLSEN,
217
Geological Survey Report 109 Department of Scientific and Industrial Research, New Zealand. SuRLYK, F. ( 1978) Submarine fan sedimentation along fault scarps on tilted faultblocks (Jurassic-Cretaceous bound ary, East Greenland). Gr¢n. geol. Unders. Bull. , 128, 108 pp. TULLOCH , A.J. & KIMBROUGH, D.L. ( 1989) The Paparoa Metamorphic Core Complex, New Zealand: Cretaceous extension associated with fragmentation of the Pacific margin of Gondwana. Tectonics, 8, 1217- 1234. TuLLOCH, A.J. & PALMER, K. (1990) Tectonic implications of granite cobbles from the mid-Cretaceous Pororari Group, southwest Nelson, New Zealand. N.Z. 1. Geol. Geophys. , 33, 205 -217.
Spec. Pubis int. Ass. Sediment. (1995)
22, 219-236
Sedimentation and tectonics of a synrift succession: Upper Jurassic alluvial fans and palaeokarst at the late Cimmerian unconformity, western Cameros Basin, northern Spain N IG E L H . P L A T T Geco-Prakla Schlumberger, Schlumberger House, Buckingham Gate, Gatwick Airport, West Sussex RH6 ONZ, UK
ABSTRACT
The Upper Jurassic (late Kimmeridgian to Berriasian ?) Senora de Brezales Formation of the western Cameros Basin, northern Spain, comprises a laterally variable succession of continental conglomerates, sandstones and pedogenetic carbonates deposited in wadi-type channels, alluvial fans, sandflats and palaeosols in a semi-arid environment. This succession rests on a complex unconformity surface that developed in response to relative falls in sea-level and strong Late Jurassic extensional faulting. Footwall uplift resulted in truncation of underlying strata at the crestal axes of fault blocks. The clastic rocks of the Senora de Brezales Formation represent the erosion products of the Jurassic anq older strata. Lateral facies variations reflect the changing lithology beneath the unconformity surface, whereas strong lateral thickness variations record fault control on sedimentation. In areas of limited erosion, karst surfaces developed on subaerially exposed lower Kimmeridgian limestones. Where these were remo ved, erosion of upper Oxfordian marginal marine sandstones led to the deposition of red continental sandstones. In other areas, erosion led to the incision of channels into a pediment of Middle Jurassic carbonates. The channels were filled with conglomerates largely deri ved from erosion of the Jurassic marine limestones. The location of channels was strongly influenced by NE-SW faults. Laminar and nodular calcretes formed directly on the unconformity surface, in interchannel areas, and within the o verlying clastic succession. The Senora de Brezales Formation is an outcrop analogue for Upper Jurassic clastic successions present in the subsurface of basins to the west of Britain on the northern Biscay margin . Field studies highlight the complex lateral facies variations within this synrift succession and underline the control of seismic-scale and subseismic-scale faults on erosion and sedimentation patterns at the late Cimmerian unconformity.
INTRODUCTION AND GEOLOGICAL SETTING
Many of the Mesozoic basins bordering the North Atlantic display complex unconformities and thick continental successions recording major rifting events during the Late Jurassic to Early Cretaceous (Tankard & Balkwill, 1989; Hiscott et a!., 1990). Improved understanding of the stratigraphy and tectonic evolution of rifts bordering the Bay of Biscay has highlighted the similarities between the offshore and onshore basins of Spain (Garcfa-Mondejar
1985; Platt, 1989a, 1990; Platt & Pujalte, 1994), Ireland (Petrie et a!., 1989; Shannon, 1991) and southern England (Kamerling, 1979; Chadwick, 1985; Evans, 1990; Ruffell & Coward, 1992). Upper Jurassic clastic rocks are reported from the offshore basins to the southwest of the UK and Ireland on the northern Biscay margin (Millson, 1987; Evans, 1990; Shannon, 1991; Moore, 1992), but there has been limited sedimentological study of
et a!.,
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
219
220
N.H. Platt
these subsurface successions to date. In northern Spain, on the southern Biscay margin, Upper Jurassic-Lower Cretaceous continental clastic de posits are well-exposed, permitting detailed sedi mentological study and mapping of lateral facies and thickness variations within the Upper Jurassic synrift succession (Platt, 1986; Clemente & Perez-Arlucea, 1993; Gomez Fernandez & Melendez, 1994). This paper presents the results of a sedimen tological study of the Upper Jurassic Senora de Brezales Formation of the western Cameros Basin (Fig. 1), outlining the tectonic controls on alluvial fan deposition and palaeokarst development at the late Cimmerian unconformity.
the modern Florida Everglades or to the marshlands of southern Iraq (Platt, 1989b, c; Platt & Wright, 1991, 1992). The Rupelo Formation is overlain by oncoidal limestones, sandstones and mudstones of the ?Valanginian-Barremian Hortigtiela For·· mation (Fig. 2), and a thick ?Barremian-Aptian series of alluvial clastic rocks comprising the sandy-· muddy Piedrahita de Mufi6 Formation and the conglomeratic-sandy Salas Group. This succession is capped by alluvial sandstones and conglomerates of the upper Albian to lower Cenomanian Utrillas Formation.
LITHOFACIES STRATIGRAPHY
The stratigraphy of the western Cameros Basin, described by Platt (1986, 1989a, b) and Platt & Pujalte (1994), is outlined in Fig. 2. A thick suc cession of Early-Middle Jurassic marine carbonates is truncated by a complex series of unconformities (Mensink & Schudack, 1982; Schudack, 1987; Platt et al., 1991). In the south, Bathonian Callovian limestones are discordantly overlain by (?Callovian-) Oxfordian sandstones and cal carenites of the San Leonardo Formation (Wilde, 1988) and Oxfordian to basal Kimmeridgian car bonates of the Talveila Formation (Dfaz et al., 1983; Platt, 1986; Thalmann, 1989). Sedimentary cycles within this complex and heterogeneous succession of marginal marine carbonates and littoral clastics (Platt, 1986) may be compared with those present in the Corallian of southern Britain (Talbot, 1973; Sun, 1989). The middle Oxfordian to Kimmeridgian shallow to marginal-marine rocks are absent in many areas, reflecting erosion prior to deposition of the Senora de Brezales Formation, which is a laterally variable succession of continental deposits present through out the western part of the basin (Platt, 1989a). The Senora de Brezales Formation comprises poly genetic conglomerates, red sandstones and calcareous palaeosols, and is probably of Kimmeridgian Berriasian age (Platt, 1986; Wright et al., 1988). The Senora de Brezales Formation reaches a maximum thickness of 75 m to the north of Espej6n (Figs 1 & 3), and is overlain by lacustrine-palustrine lime stones of the Rupelo Formation (Berriasian). The Rupelo Formation is interpreted as a succession of freshwater carbonates deposited in a low-gradient, low-energy system of lakes and swamps similar to
The Senora de Brezales Formation shows strong lateral variability (see discussion below), but in all cases comprises three main facies: sandstones, con glomerates and carbonate rocks. Figure 3 presents a representative logged sedimentological section from the Senora de Brezales Formation. Red sandstones
Red sandstones reach a maximum of 40 m in thick ness and are present: 1 forming the lower part of the Senora de Brezales Formation to the southwest of the San Leonardo Fault (Fig. 1); 2 above basal Senora de Brezales Formation lime stone conglomerates in the northeast of the study area. The sandstones are typically dark brick-red in colour and are of medium grain size. They have subrounded grains, are generally quartzose, and contain scattered quartz pebbles 1-4 mm in dia meter. The sandstones are generally structureless (Fig. 4A), although trough cross-bedding, thin pebbly lags and plane lamination are locally discern ible (Fig. 4B). White-red mottling and/or 1-2 cm diameter burrows are developed at a few localities. Vertical cylindrical structures 5-10 em in diameter are also present locally, as are 2-cm-diameter tubu lar structures parallel to bedding. In the west, at Hortezuelos and Mamolar (Fig. 1), the sandstones are cut by vertical and horizontal cracks. Brecciated, nodular carbonate horizons 0.5-1 m thick are also present. These are commonly mottled and display floating quartz grains, angular, branching spar-filled cracks and grey carbonate stringers 1 mm in width. These carbonate facies are described in detail below.
221
Alluvial fans, palaeokarst and tectonics
A
1. (A) Palinspastic reconstruction of the North A tlantic and Bay of Biscay region for Late Jurassic times showing the location of major Mesozoic basins. The Cameros Basin is in central Iberia, to the south of the Biscay rift. (B) Map of the western Cameros Basin showing major structural elements and localities mentioned in the tex t: C, Castrovido; H, Hortigiiela/Valparaiso; Ho, Hortezuelos; HR, Huerta del Rey; J, Jaramillo Quemado; LG , La Gallega; M, Mamolar; Mo, Moncalvillo; ML, Mambrillas de Lara; PB, Pinilla de los Barruecos; PM, Pinilla de los Moros; Q, Quintanilla de las Vinas; R, Rupelo; SI, Salas de los Infantes; SL, San Leonardo de Yagiie; T, Talveila; QCF, Quintanilla Castrovido Fault; JCF, Jaramillo Covarrubias Fault; SLF, San Leonardo Fault. (C) Inset showing area covered by map B . Fig.
�
QCF
�
R .'......._ '-... ... '-... "'-.. :' He :' J """ • ... /JCF PM •
... "'
• a · ML
•
.·
Q__J9km
Mo • �
-- basin margin
--'--- normal fault ___.___ · · · · · · · ·
thrust
transfer fault
Interpretation
The homogeneous character of the sandstones is consistent with strong bioturbation. Primary sedi mentary structures are preserved in only a few
places, but the rare presence of lags and mesoscale cross-bedding suggests that the sandstones were waterlain. Mud is absent; flows were prol::ably short lived with no deposition from suspension. Thin sec tion observation shows that the red colour is the
W CAMEROS
BASIN- STRATIGRAPHY
(/) ::::l a::O
w w a..U ::::>1w a: u a..<{
Albian
Aptian
Barremian (/) ::::l a::O
ww :s:;u o<�: ....Jtu a: u
!'-'""" Hauterivian
Valanginian
Berriasian
Tithonian
lfiii±�;ijj£/�,��;P""�
�--------��:0�::;i;�:�i:�:,�r:: �:��i.;�:��:f}:+:?:>:xj;fj}/j;;����:;;.; Kimmeridgian
,;;.;..;.
S:
�
·
·
Talveila��
Oxfordian
-,
Callovian
KEY
Evaporite
Marl
Oncoidal limestone
Sandstone
Organic-rich limestone
Conglomerate
2. Upper Jurassic to Lower Cretaceous stratigraphy of the western Cameros Basin. Fig.
Limestone
223
Alluvial fans, palaeokarst and tectonics
z 0
�
� a: Ladera
f2 0
Member
...J LU a..
�
Las Vinas Member
KEY
(ill!illJ�illn
(/) LU ...J
limestone and marl
COI
Polymict conglomerate
w 0:::o
Brecciated limestone
< w 0�
Reticulate calcrete
Laminar calcrete Burrows and roots Cross-bedded red sandstone
ou.. •Z LU (/)
3. Logged sedimentological section from the Senora de Brezales Formation, section near chapel of Nuestra Senora de Brezales, 2 km north of Espej6n.
Quartz conglomerate Charophyte-gastropod
0
Massive red sandstone
j���� Marine limestone
Fig.
Marine Jurassic
product of iron oxide grain coatings, typical of con tinental 'red bed' deposits. This interpretation is consistent with the sedimentological evidence of subaerial exposure. This includes cracks, interpreted as sheet cracks developed during desiccation, mottles and tubular structures of varying morphology and size, interpreted as 1-2-cm-diameter burrows and 2-10-cm-scale roots. The brecciated, nodular and mottled fabrics of the carbonates are typical of pedo genetic horizons, favouring an interpretation as cal cretes. The presence of floating quartz grains indicates the in situ growth of replacive calcite. The red sandstones are thought to represent ephemeral stream deposits. The climate was prob ably semi-arid, with sporadic rainfall permitting only short-lived stream flow; in the intervening periods the sandstones were homogenized by burrowing and root action. The occasional rainfall permitted the growth of hardy scrub vegetation and periodically allowed the establishment of high or perched water tables, favouring the development of palaeosol pro files and pedogenetic carbonates (see also below). Desiccation and strongly oxidizing conditions were responsible for the red colour of the sediments and
prevented the preservation of organic material so that the roots are preserved only as moulds. Conglomerates
Conglomerates are common in the Senora de Brezales Formation. These are poorly sorted, with subangular to subrounded clasts, mostly of carbonate and generally varying in size from 2 to lOcm. Locally, the clasts reach 30-100 em, as at Jaramillo Quemado (Fig. 4C). The matrix consists of medium grained red sandstone with 1-2-cm rounded quartz pebbles, which are more abundant in the southwest. Conglomerates commonly occur in channels. The channels are 10-50 m across, have erosive bases and are 2-3 m deep (Fig. 5). Coarse conglomerates at the channel bases pass up into red, trough cross bedded sandstones. Many channels are cut directly into marine Jurassic carbonates, as in the northeast of the study area. At Castrovido (Fig. 1), channelized conglomerates rest on lower Callovian limestones, passing up into interbedded conglomerates and red sandstones. Channelized conglomerates also rest unconformably on lower Callovian marine
224
N.H. Platt
�
f2 "" �
'i:i'
·"' � ;:;
� �
� C) ,.,. �
�
� ;:; .,__
�
0 ;:;
�·
4. Lithofacies and sedimentary structures within the Senora de Brezales Formation. (A) sandstones- massive. Huerta del Rey. Person for scale. (B) sandstones- cross-bedded. Huerta del Rey. Hammer for scale. (C) conglomerates- very coarse conglomerates containing locally derived clasts of lower Callovian limestone. Jaramillo Quemado. Lens cap for scale. (D) conglomerates- stacked, fining up sheet-flood conglomerates. Arrows mark bases of sheet-flood units. La Gallega (Navas road) . Hammer for scale. (E) carbonates- cross-cutting laminar calcrete present within conglomerate, Hortezuelos. Lens cap for scale. (F) carbonates- stacked, horizontally bedded laminar calcretes in sandstone matrix, Mamolar. Lens cap for scale. (G) karst surface- karstic cavity developed in Lower Kimmeridgian limestones, filled by limestone breccia and red internal sediment, Talveila. Hammer for scale. (H) karst surface- red sandstone, marked by arrow, piped down into lower Callovian limestone surface, Pinilla de los Moros. Hammer for scale. Fig.
1-'3
V1
�
0\
:<: �
� :::::
5. Channels of polymict conglomerate cut into red sandstones and marls, Espej6n. The arrows mark the bases of three distinct channels. These channels are not fault-bounded. Bedding is vertical, younging away from camera: The photograph was taken standing on Jurassic marine limestones and looking southwards. The prominent ridge running left to right across the photograph above the Senora de Brezales Formation channels is composed of non-marine limestones of the o verlying Rupelo Formation (R). The village and cornfields beyond lie on Valanginian-Cenomanian clastic rocks (KL), whereas the mountain on the horizon consists of Upper Cretaceous marine limestones (KU). Fig.
Alluvial fans, palaeokarst and tectonics
carbonates at Quintanilla de las Vinas and Hortigiiela in the northwest, an area where red sandstones are absent. In the southwest, conglomerates occur inter bedded with sandstones or with palaeosols some metres above the base of the Senora de Brezales Formation. Many channels incised into the basal unconformity are bounded on one or both margins by minor NE SW trending faults. Lateral variations in conglomer ate thickness also commonly occur across these faults. In contrast, channelized conglomerates pre sent at higher levels within the Senora de Brezales Formation are not fault-bounded. At Espej6n {Fig. 1), a number of individual channels are spectacularly exposed. The channels are asymmetric, with strongly erosive bases. The coarsest conglomerates fill the base of the channel, passing upwards and towards the channel margin into red, trough cross-bedded sandstones. Sheet-like conglomerates are also common; these show planar erosive bases and typically occur in 2-5-m beds (Fig. 4D). Several pulses of conglomer ate occur at Senora de Brezales; these are traceable laterally for 20-50 m. At Hortezuelos (Fig. 1), conglomerates near the base of the succession are monomict, dominated by angular 2-4-cm clasts of grey micrite. Elsewhere, the conglomerates are polymict ('polygenetique' of the French usage; Salomon, 1982). Clast lithologies include: 1 fossiliferous yellow-orange packstones containing 0.5-mm quartz grains and brachiopod fragmentsthese are derived from the marine lower Callovian; 2 pink to grey micrites, commonly partially recrystallized- these may represent subaerially weathered clasts of marine carbonate from the Talveila Formation, Middle Jurassic, or Liassic; 3 pink oolitic grainstones- these are especially common in the southwest of the study area, and were probably derived from Middle Jurassic shallow water carbonates; 4 rarer, small ( < 1 em) clasts of red sandstonethese may be derived from red sandstones in the Senora de Brezales Formation, or from Triassic red beds. The conglomerates show few sedimentary struc tures, and are mostly clast-supported. Matrix supported conglomerates occur at a few localities, as at Senora de Brezales and La Gallega (Navas road), but even in these cases there is little mud and the matrix is comprised of sandstone.
227
Interpretation
The channels provide evidence of powerful, strongly erosive flows which were probably intermittent events associated with sporadic heavy rainstorms. The wadi-type channels were incised into the uncon formity surface on uplifted blocks, and locally, as at Espej6n, into finer grained deposits of the Senora de Brezales Formation. Away from the fault scarps, flows spread out into sheet-floods, depositing tabular units of conglomerate and sandstone. Carbonate rocks
Pedogenetic carbonate horizons are interbedded with the red sandstone facies of the Senora de Brezales Formation (see above). Carbonate horizons also rest directly upon marine Jurassic limestones in interchannel areas. Laminated limestones are present in the north of the study area, lying either directly on the uncon formity surface (as at Castrovido and Mambrillas de Lara; Fig. 1), interbedded with or cross-cutting the sandstones and conglomerates (as at Hortezuelos and Mamolar; Fig. 4E & F) or developed as exten sive sheets resting on conglomerate bodies (as at Quintanilla de las Vinas). The laminae are irregular and undulating, consist ing of 1-mm alternations of light and darker grey (rarely black) micrite. Many of the laminated hor izons display a fabric of horizontal 1-mm pale beige, concentrically laminated carbonate tubules 0.5 mm in diameter filled with sparry calcite. At Senora de Brezales and at Espej6n (Fig. 1), a succession of marly, brecciated and heavily mottled carbonate horizons is interbedded with sheet-flood and channelized conglomerates. These sediments display a complex reticulate fabric produced by horizontal 2-20-mm layers of laminar calcrete and vertical tubular structures 1-10 em in diameter and up to 1 m in length. Some of the vertical structures are composed of homogeneous micrite, but others show an internal fabric similar to that of the laminar horizons. Between these structures, the mottled marly, or sandy, matrix commonly contains abun dant fine carbonate strings and spar-filled angular, branching cracks. At Senora de Brezales, stacked laminar horizons form mound-like structures 5 m across with a relief of 0.5 m (fig. 9a in Wright et a/., 1988).
228
N.H. Platt
Interpretation
The carbonate horizons within the Senora de Brezales Formation are interpreted as calcareous palaeosols. The laminated limestones are similar to laminar calcretes described from modern exposure surfaces in Florida by Multer & Hoffmeister (1968). Although similar in some respects to laminar stro matolites, an absence of diagnostic algal structures, such as calcified filaments (Read, 1976), and their association with pedogenetic features, suggests inter pretation as biogenic laminar calcretes or 'rhizolites' (Wright et a!., 1988). The fine spar-filled tubules are interpreted as root structures (rhizoliths). The beige micrite coating may either be a replacement of the root tissue itself (internal septae commonly form the characteristic 'alveolar texture' of Esteban, 1974), or may be made up of layers of concretionary car bonate precipitated around the root. A more detailed description of the laminated horizons is to be found in Wright et a!. (1988). The thickness and abundance of laminar calcrete horizons towards the top of the formation point to long periods of root colonization, both on the surface and beneath it. These times were punctuated by occasional input of fine-grained distal alluvial clastic material, producing an alternation of laminar cal cretes with mottled sandstone or marl. Desiccation and the action of larger roots were probably respon sible for most of the observed brecciation of the carbonate horizons. Brecciated, mottled carbonate horizons are also common within the basal part of the overlying Rupelo Formation, and in places there may be an upward transition between the Rupelo and Senora de Brezales formations. Although Platt ( 1986) con sidered these to be two essentially separate units, they may be partial lateral equivalents.
GEOMETRY OF LATE JURASSIC UNCONFORMITIES
The study of Upper Jurassic facies distributions is complicated by the presence of at least two import ant unconformities, one beneath the San Leonardo Formation (probably late Callovian to middle Oxfordian in age) and a later one beneath the Senora de Brezales Formation (probably intra Kimmeridgian). Further discontinuities may exist within the San Leonardo Formation. The superpos ition of these unconformities causes problems in
unravelling the depositional history of the study area. Effects of local faulting were certainly signifi cant in the localized preservation of strata beneath the unconformities. The Senora de Brezales Formation rests on the most important intra-Mesozoic unconformity in the study area. This unconformity records a period of uplift, faulting and erosion that began some time after the early Kimmeridgian. It marks a major change in basin configuration, and a change from relatively uniform, gentle subsidence and generall.y shallow marine conditions from the Rhaetian to early Kimmeridgian, to the more rapid, strongly localized subsidence and continental sedimentation that characterized the latest Jurassic and Early Cretaceous. The combined effects of tectonic activity and ero sion during this period resulted in the formation of a faulted surface with significant relief. Thus the Senora de Brezales Formation lies unconformably on a variety of different strata, from Bathonian- Callovian shallow-marine limestones, to upper Oxfordian to lower Kimmeridgian marginal marine sandstones and limestones of the San Leonardo and Talveila Formations. Where the Senora de Brezales Formation rests on Kimmeridgian and Callovian limestones, it reslts variably upon karst surfaces on palaeohighs and on an erosive channelled base in the lows. Where sand stones of the San Leonardo Formation are truncated by the unconformity, the basal contact is planar without the development of palaeokarst or significant relief (Platt, 1986). Karst surface
The Senora de Brezales Formation lies on karstified lower Kimmeridgian limestones at Talveila (Mauthe, 1975; Fig. 40), Pinilla de los Barruecos and Horte zuelos (Fig. 1). At all these localities, red internal sediment fills cavities and joints within the underlying carbonate rocks. At Hortezuelos, the karstified sur face is cut by a small fault with a throw of 1 m. Here, deeper karstification and/or partial erosion of the fault scarp led to the local deposition of a 1-2-m lens of coarse chaotic conglomerates. At Castrovido in the north of the study area (Fig. 1), a karst surface is developed on marine Callovian limestones in interchannel areas. The marine Jurassic carbonate surface is locally dis coloured and brecciated and laminar calcretes rest on the unconformity. A similar situation occurs
Alluvial fans, palaeokarst and tectonics
at Mambrillas de Lara (Fig. 1), where the lower Callovian limestone surface is draped by laminar calcretes penetrating downwards for 30 em along joints and cracks (see Wright et al., 1988). At Pinilla de los Moros (Fig. 4H), a karstified surface of lower Callovian limestone is penetrated by vertical cracks 1 em across filled with red sandstone. At these localities, subaerial exposure and erosion prior to deposition of the Senora de Brezales For mation resulted in karstification of underlying lime stones. Red internal sediment was piped into joints and karstic cavities; laminar calcretes formed at the surface and along joints and cracks beneath. Planar contact
To the north of Hortezuelos, and at Mamolar and La Gallega (Fig. 1), red sandstones of the Senora de Brezales Formation rest on yellow sandstones of the San Leonardo Formation and paleokarst is not devel oped. A S-cm-thick, dark red, ferruginous-cemented sandstone layer is locally present at the contact. The poorly defined contact between the red sand stones and the underlying yellow sandstones suggests that the red sandstones may have been at least partially derived by reworking of the San Leonardo Formation. Channels
Channels filled with polymict conglomerates occur at the base of the Senora de Brezales Formation in the north of the study area, where the San Leonardo and Talveila Formations are absent. The channels record powerful stream erosion of the limestones beneath. The siting of channels on the downthrown sides of NW-SE trending faults (Fig. 6) suggests that they formed in structural lows produced by the Late Jurassic faulting.
TECTONIC CONTROLS ON FACIES DEVELOPMENT
The style of sedimentation within the Senora de Brezales Formation strongly reflected the pattern of fault blocks and the lithologies exposed. Where the sandstones of the San Leonardo Formation were eroded, planation of the contact ensued and rework ing led to the deposition of red sandstones. Karst surfaces developed on fault-bounded structural highs, where Middle Jurassic and lower Kimmerid-
229
gian marine carbonates were exposed. Erosion of these highs acted to equalize the pediment relief. A schematic NW-SE stratigraphical section for the Senora de Brezales Formation in the north of the study area (Fig. 6) demonstrates the tectonic control on facies that is evident after restoration to the base of the overlying Rupelo Formation. The presence at Jaramillo Quemado of coarse basal con glomerates with clasts of lower Callovian limestone up to 1 m in size (Fig. 4C) probably records erosion of a faulted high to the west of the Jaramillo Covarrubias Fault, a major NE-SW transfer struc ture (Platt, 1990). At this locality there may be a gradation to karst conglomerates, where brecciation was followed by only limited transport, so that the locally derived clasts were deposited more-or-less in situ, infilling an irregular karstic relief. Red sand stones are restricted to the area to the east of the fault, whereas immediately to the west, between Valparaiso and Mambrillas de Lara, karst surfaces are developed on the marine Jurassic, and the Senora de Brezales Formation is thin or absent. Moving further to the west, conglomerates are present firstly on isolated fault blocks and eventually in fault controlled channels. Many of the conglomerate-filled channels are bounded on one or both sides by NE-SW faults (Fig. 6); good examples occur at Castrovido, Quintanilla de las Vinas, and between San Leonardo de Yagiie and Hontoria del Pinar (Fig. 1). These minor faults are parallel to the Jaramillo-Covar rubias Fault. At La Gallega, a NE-SW fault with a throw of 25 m on the western side of the Navas road bounds a depression filled with Senora de Brezales Formation conglomerates and sandstones. Many more minor faults occur, typically at a 100-m spacing. These faults have throws of only 2-10 m, but still controlled lateral variations in the subcrop to the unconformity and exerted strong local influence on facies development and thicknesses within the Senora de Brezales Formation. Changes in facies and thickness also occur across NW-SE faults. Although the exposure is not con tinuous in the dip direction and the basin structure only allows comparison between outcrops on parallel anticlines several kilometres apart, general contrasts are evident. Thick successions with red sandstones up to 40 m thick occur near the southwestern basin margin, whereas thinner successions are present to the north. In the north of the study area, the Senora de Brezales Formation is thicker to the south of the Quintanilla-Castrovido Fault than immediately to
230
N.H. Platt
NW
SE
m ottled I brecciated lim estone
l
MARINE JURASSIC PEDIMENT JCF Quintanilla de las Vinas
KEY
Mambrillas de Lara
?Berriasian
Valparaiso Hortiguela
Jaramillo Quemado
Rupelo Formation (Las Vinas Member)
Pinilla de los Moros
Castrovido
non-marine limestone & marl
J.+.�+----············-·······-·-············-··············
...............................................................................................................
red sandstone
?Kimmeridgian - Berriasian
Senora de Brezales Formation
..................................................•....•............•....•••.................................•
Lower Callovian
top of Marine Jurassic
conglomerate
laminar calcrete
ee;;r;<;i;l·
.................•....................................
marine limestone
Fig. 6. Cartoon strike section between Quintanilla de las Vinas and Castrovido, illustrating tectonic control on lateral changes in facies development within the Senora de Brezales Formation. Section restored to the base of the Rupelo Formation to eliminate the effects of post-depositional faulting of Lower Cretaceous and Tertiary age (see Platt, 1990). Horizontal distance between Quintanilla de las Vinas and Castrovido is 30 km; maximum vertical thickness of the Senora de Brezales Formation is 50 m.
the north. Red sandstones are well-developed to the east of the Jaramillo-Covarrubias Fault between Jaramillo Quemado and Castrovido but are very thin or absent immediately to the northwest of the fault. These lateral variations suggest that both NW-SE and NE-SW fault sets were probably active during deposition of the Senora de Brezales Formation, defining a complex mosaic of fault blocks with differing subsidence histories and different detailed stratigraphies.
EROSION, CLAST DERIVATION AND DEPOSITIONAL HISTORY
The composition of the conglomerates within the Senora de Brezales Formation strongly reflects the subcrop to the unconformity (Fig. 7). The irregular
clast shapes and abundance of coarse carbonate clasts suggests limited clast transport; clast derivation was chiefly from erosion of the marine Jurassic limestones of underlying or adjacent fault blocks. The vertical succession may therefore record the progressive erosional history ('unroofing') of the source areas. 1 Locally, as at Hortezuelos, a basal conglomerate is present, containing clasts only of grey micritic limestone. These clasts appear to have been derived from laminar calcretes deposited in interchannel areas and from Talveila Formation lagoonal carbonates. 2 Red sandstones above consist mainly of quartz, probably including a component reworked from the sandstones of the San Leonardo Formation. 3 Polymict conglomerates above these contain abundant reddened oolitic limestone clasts, indi-
231
Alluvial fans, palaeokarst and tectonics
CLAST DERIVATION
DEPOSITIONAL HISTORY Valanginian Cenomanian
4
quartz sandstones and conglomeratesprogressive unroofing and erosion of
• :1111������ �;;;���;,��=�:.
···············
·
:���
Formation
.....................
Senora de Brezales Formation
Talveila Formation:
����!��� ���!�: �; ���!?.���
-----
-
--
-
_
_
-
-
-
-----
San Leonardo Formation:
.....'!��-��!��!-�-��!�-�-�-���-�!��!:�---Middle Jurassic: shallow-marine limestones
limited clastic supply
--------------------- --·-············-·----------------·
Lower Jurassic: marine limestones and marls
polymict conglomerates reworking of marine Jurassic, especially Middle Jurassic red sandstones - reworking of San Leonardo Fm
monomict basal conglomerate ��:-y-���!�-�-?!.!.����i!� .F.�.: .................
--------------------------------
Triassic: continental, mostly sandstones
.
7. U pper Jurassic-lower cretaceous clast derivation . Clastic rocks- vertical succession illustrating likely derivation of clasts from the successive 'unroofing' and erosion of underlying stratigraphical units. See text for explanation. The schematic section shown is a composite based on outcrops on the northeastern flank of the Hortezuelos anticline.
Fig.
eating erosion and reworking of Bajocian-Callovian shallow marine carbonates. Their deposition may therefore date the 'unroofing' of the underlying Middle Jurassic. Similar polymict conglomerates containing abun dant coarse limestone clasts occur in a number of areas, in each case recording local clast derivation from the marine Jurassic. Examples occur in the north of the study area, as described above (Fig. 6). The San Leonardo Formation is absent in the north, but the local presence of red sandstones, as at Castrovido, may record transport of eroded material elsewhere from the San Leonardo Formation or from basement lithologies. In the southwest of the study area, the presence of quartz pebbles may record erosion of quartzitic material from the San Leonardo Formation or from basement rocks on the Duero block to the south. However, widespread erosion of Triassic sandstones and Palaeozoic quartzites probably did not com mence until much later (Platt, 1989a). The low clastic content in the freshwater carbonates of the overlying Rupelo Formation suggests that this phase of peneplanation was complete by the Berriasian, although the catchment area geology was still appar ently dominated by carbonate lithologies. The first wide-spread erosion of the Triassic and Palaeozoic is probably dated by the appearance of sandstones within the Barremian-Aptian Hortigiiela and Pied rahita de Mufi6 Formations (Figs 2 & 7). These rocks do not contain Jurassic limestone clasts-
this may reflect lengthier transport in more perma nent river systems developed under more humid conditions.
ENVIRONMENTAL SYNTHESIS
Figure 8 presents a schematic depositional model for the Senora de Brezales Formation, showing the main environments represented and the control of the unconformity subcrop and of tectonics on facies development. Sedimentation took place under a semi-arid climate. Small-scale wadi-type channels were incised directly into the underlying marine limestone pediment, but much of the sandstone and conglomerate was deposited from laterally extensive sheet-floods spreading out from these channels. Pedogenetic carbonates developed on the pediment in interchannel areas, and also formed as surficial layers and nodular brecciated horizons on or within the exposed alluvial clastic deposits. Bioturbation and root penetration acted to destroy primary sedimentary structures, particularly within the sandstones. The association of channels and sheet-floods is consistent with deposition in small-scale alluvial fans. Basin-bounding NW-SE faults and NE-SW trans fer faults both influenced sedimentation, which acted to infill fault-bounded depressions; clastic supply came dominantly from the erosion of neigbouring uplifted blocks.
N.H. Platt
232
pedogenesis - laminated crusts
& cross-graben wrench or "transfer" faults
laminated crusts on interchannel areas of pediment
\
paleosols on alluvial substrates
small fans sourced at fault intersections
channels cut into pediment located along transfer faults
basin-bounding extensional normal faults
_.U I
2
Oxfordian arenites
erosion of Mid. Jurassic sequence - polymict conglomerates localised erosion of Lr Kimmeridgian limestones - monomict conglomerates
8. Schematic depositional model for the Senora de Brezales Formation, illustrating sedimentary environments represented and the influence of tectonics and the underlying geology on facies development and clast derivation.
Fig.
REGIONAL PERSPECTIVE
Upper Jurassic clastic rocks
The Kimmeridgian-Tithonian saw a major re gression in Iberia and across much of northwest Europe. In the Lusitanian Basin of Portugal (Wison et al., 1989), tectonic movements led to an acceler ation of subsidence; the lacustrine-lagoonal car bonates of the Oxfordian Caba'
Brezales Formation occur at the base of the laterally equivalent Aguilar Formation (see Platt & Pujalte; 1994). Here, too, the subcrop to the Late Jurassic unconformity was dominated by marine Jurassic limestones developed in facies similar to those of the Cameros. Faulting and continental erosion then permitted the deposition of locally derived, tec tonically controlled conglomerates of similar composition. Upper Jurassic continental clastic sediments are also reported from several basins to the west of Britain on the northern Biscay margin, where they form potential hydrocarbon reservoir targets. Exam ples occur in the Celtic Sea Basins (Millson, 1987; Petrie et at., 1989; Shannon, 1991) and the Brittany Basin of the Southwestern Approaches (Evans, 1990). in these areas, a major period of Late Jurassic extension led to fragmentation of the basins into fault blocks, which were tilted and eroded. These Late Jurassic fault block structures form a major exploration target in the Porcupine Basin of offshore
233
Alluvial fans, palaeokarst and tectonics
Ireland (MacDonald et al., 1987). In the Porcupine, the Kimmeridgian saw a change in tectonic regime, as widespread broad subsidence gave way to fault controlled subsidence and sedimentation. This evol ution is parallel to that recorded by deposition of the Senora de Brezales Formation in the western Cameros Basin. Moore (1992) recognized an Upper Jurassic synrift to post-rift transition succession up to 500 m thick within faulted sub-basins of the Porcupine Basin. These rocks are characterized on seismic data as mounded, fan-style deposits resting on a faulted unconformity, passing laterally and upwards into parallel, onlapping reflectors of a ?marginal marine succession. Moore suggests that these fans may have significant reservoir potential, particularly in the sand-prone proximal-mid- and mid-fan regions. Proximal areas of the fan contain poor reservoir quality unsorted conglomerates (Shell well 34/19), whereas the distal part is expected to contain fine siltstone and shale. The clastic fans were bounded by faults on one side, passing basinwards into finer grained distal fan deposits and onlapped by Lower Cretaceous shales. Although this succession was deposited in a marginal marine setting, the facies geometries developed are essentially identical to those inferred for the alluvial clastics of the Senora de Brezales Formation (Fig. 9) and the overlying low-gradient lacustrine deposits of the Rupelo Formation. In the North Celtic Sea (Millson, 1987), inter bedded marginal marine carbonates and littoral and continental clastics of the mid- to upper Oxfordian
Clastic fans
Timoleague Formation show similarities with the heterogeneous marginal marine succession devel oped in the San Leonardo and Talveila Formations. Like the western Cameros, the North Celtic Sea saw a change to continental conditions during the Kimmeridgian. Here, the coarse- to fine-grained, locally conglomeratic sandstones of the Clonakilty Formation may correspond with the clastics of the Senora de Brezales Formation in the Cameros. This interpretation is further supported by Millson's (1987) descripton of thin, concretionary, silty micritic carbonate horizons, possibly representing calcretes, within the Clonakilty Formation. Late Jurassic unconformities
Many authors have recognized a Late Jurassic to Early Cretaceous unconformity in northwest Europe (see e.g. Ziegler, 1990). This 'late Cimmerian' unconformity, which is variably dated as earliest Cretaceous or Volgian-Tithonian, is known from the North Sea (Rawson & Riley, 1982), Porcupine Basin (Croker & Shannon, 1987; Tate, 1993), North Celtic Sea (Millson, 1987; Shannon, 1991; McCann & Shannon, 1993), and Cantabria (Platt & Pujalte, 1994). Accurate dating of this surface is commonly hindered by problems of stratigraphical resolution in marginal/restricted Upper Jurassic facies; for exam ple, in the study area, the basal Senora de Brezales Formation unconformity can be dated only as post lower Kimmeridgian and pre- lower Berriasian. Superposition of several unconformities means that the time gap represented by the unconformity is
Onlapping infill
Clastic fans
Basal unconformity
9. Sketch showing inferred stratigraphical relationships in the Upper Jurassic Senora de B rezales Formation of the western Cameros Basin. Figure is modified after a cartoon by Moore ( 1992) depicting seismic relationships in the Upper Jurassic synrift to post-rift transition of sub-basins within the Porcupine Basin (Moore, 1992). The two successions share a common tectonic evolution and are thought to show e ssentially identical geometry. Fig.
N.H. Platt
234
of variable extent in different areas. In basinal areas of the North Sea and the Wessex Basin, the 'base Cretaceous unconformity' is represented by a con densed section of Volgian-Ryazanian (Tithonian Berriasian) age (Rawson & Riley, 1982; Ruffell, 1992). However, on structural highs and in areas adjacent to the Biscay rift, Lower Cretaceous rocks may rest directly upon Liassic or Triassic rocks, as occurs in western parts of the Celtic Sea basins (Millson, 1987), the Western Approaches (Evans 1990) and northern Cantabria (Garcia-Mondejar et a!., 1985; Hines, 1988). Study of the Cameros Basin indicates that the complex association of superimposed unconformities in the areas around Biscay is the product of vari ations in relative sea-level (notably falls in the early Oxfordian and ?mid-Kimmeridgian and rises in the late Oxfordian and ?late Kimmeridgian to Berriasian), coupled with widespread Late Jurassic extensional block faulting. The effect of these, reflected in the basal contact, deposition and deri vation of the Senora de Brezales Formation, was to produce truncation during times of lower base level, particularly on the footwall crests of extensional fault blocks, and complex sedimentary onlaps during infilling of local accommodation space created by faulting or rises in relative sea-level.
were derived from localized reworking of lower Kimmeridgian limestones, red sandstones from erosion of upper Oxfordian sandstones, and polymict conglomerates were derived from erosion of the entire Jurassic succession, in particular the Middle Jurassic. The Senora de Brezales Formation offers a useful outcrop analogue for Upper Jurassic clastic suc cessions in basins bordering the northern Biscay margin to the west of Britain. A depositional model developed for the Senora de Brezales Formation (Fig. 9) serves to illustrate the importance of seismic scale and subseismic-scale fault movements in con trolling sedimentation and facies distributions and highlights the strong lateral variability so character istic of synrift deposition.
ACKNOWLEDGEMENTS
My sincere thanks go to Dr Harold Reading for support and encouragement. I am grateful to the students from Oxford and Bern Universities who helped with detailed mapping and logging. Helpful reviews and comments were provided by Guy Plint, Alastair Ruffell, Bruce Sellwood and Pat Shannon. My special thanks go to Adrian Forrer for discussion and good humour in the field.
CONCLUSIONS
The Senora de Brezales Formation is a laterally variable Upper Jurassic succession of continental red-beds deposited under a semi-arid climate. Con glomerates and sandstones were deposited in wadi type channels and by sheet-floods in small-scale alluvial fans. Bioturbation, desiccation and root action led to homogenization of the waterlain sand stones during protracted subaerial exposure. Pedo genetic carbonates represent palaeosols and laminar calcretes formed within and at the base of the clastic succession. The Senora de Brezales Formation rests on the late Cimmerian unconformity, which truncates a thick succession of marine Jurassic carbonates. Lateral variations in thickness in the Senora de Brezales Formation reflect Late Jurassic extensional block faulting; many channels are fault-bounded and the coarsest conglomerates occur adjacent to major faults. The deposition of the Senora de Brezales Formation recorded erosion of the under lying marine Jurassic; monomict conglomerates
REFERENCES
( 1 985) End Jurassic-early Cretaceous sedimentation and subsidence (late Portlandian to Barremian), and the late Cimmerian u nconformity. I n :
CHADWICK, R . A .
Atlas of Onshore Sedimentary Basins i n England and
Wales (Ed. Whittaker, A . ) , pp. 9-25. British Geological Survey, Keyworth. CLEMENTE, P. & PEREZ-ARLUCEA, M. ( 1993) Depositional architecture of the Cuerda del Pozo Formation, Lower Cretaceous of the extensional Cameros Basin, north central Spain. J. sediment. Petrol. , 63, 437 -452. CROKER, P.F. & SHANNON, P . M. ( 1987) The evolution and hydrocarbon prospectivity of the Porcupine Basin, Off shore Ireland. In: Petroleum Geology of North West Europe (Eds Brooks, J. & Glennie, K . ) , pp. 633- 642. Graham & Trotman, London. DfAz, M. , GoY, A . & YEBENES, A . ( 1983) Un complejo de isla barrera-lagoon del Jurasico superior (Talveila, Soria).
X Congreso Nacional de Sedimentologfa, Menorca,
Abstracts, 2.58-2.62 . C . M. ( 1 974) Caliche textures and 'Microcodium'. Boll. Soc. geol. !tal. , 92 (Suppl . ) , 1 05- 125. EvANS, C.D. R . ( 1 990) United Kingdom Offshore Regional
EsTEBAN,
Report: the Geology of the Western English Channel and its
Western
Approaches.
British Geological Survey,
Alluvial fans, palaeokarst and tectonics
235
HMSO, London.
Western Cameros Basin , Province of Burgos, northern
J . , HlNES, F.M . , PU J ALTE , v. & READING, H.G. ( 1985) Sedimentation and tectonics in the western Basque- Cantabrian area (northern Spain) during Cretaceous and Tertiary times. Excursion 9. In:
Spain.
GARclA-MONDEJAR,
Excursion Guidebook 6th European Regional Meeting, International Association of Sedimentologists , Lleida, Spain
(Eds Mila, M.D. & Rosell , J . ) , pp. 308-392. J.C. & MELENDEZ, N. (1994) Estra tigrafia de Ia 'Cuenca de Cameros' (Cordillera Iberica Noroccidental, N de Espana) durante el transito Jun1sico Cretacico . Revista de Ia Sociedad Geologia de Espana , 7, 12 1 - 139. HILL, G. (1989) Distal alluvial fan sediments from the Upper Jurassic of Portugal: controls on their cyclicity and chanel formation. 1. geol. Soc. London, 146, 539-555. HINES, F.M. ( 1988) The sedimentation, tectonics and stra
GOMEZ FERNANDEZ,
tigraphy of the Cretaceous/ Tertiary sequence of north western Santander, northern Spain.
Unpublished PhD thesis, University of Oxford, 250 pp. HISCOlT, R . N . , WILSON, R.C.L . , GRADSTEIN, F. M . , PUJALTE, V . , GARciA-MoNDEJAR, J . , BouDREAU , R . R . , e t al. (1990) Comparative stratigraphy and subsidence history of Mesozoic rift basins of North Atlantic. Bull. Am. Assoc. petrol. Geol. , 74, 60-76. KAMERLING, P. (1979) The geology and hydrocarbon habi tat of the Bristol Channel Basin. 1. petrol. Geol. , 2, 75-93 . MACDONALD, H., ALLAN, P.M. & LOVELL, J.P.B. ( 1 987) Geology of oil accumulation in Block 26/28, Porcupine Basin, Offshore Ireland. In: Petroleum Geology of North West Europe (Eds Brooks; J . & Glennie, K . ) , p p . 643-65 1 . Graham & Trotman, London. MAUTHE, F. ( 1975) Palaokarst im Jura der Iberischen Ketten (Prov. Soria, Nordspanien). Neues 1ahrb. Geol. Palaeontol. Abh. , 150 , 354-372. McCAN N , T. & SHANNON, P.M . ( 1993) Lower Cretaceous seismic stratigraphy and fault movement in the Celtic Sea Basin. First Break , 1 1 , 335-344. MENSINK, H., & ScHUDACK, M. ( 1982) Caliche, Boden bildungen und die Palaogeographischen Entwicklung in der westlichen Sierra de los Cameros. Neues 1ahrb. Geol. Palaeontol. Abh . , 163, 49-80. MILLSON, J . A . ( 1987) The Jurassic evolution of the Celtic Sea Basins. In: Petroleum Geology of North West Europe (Ed. Brooks, J. & Glennie, K.), pp. 599...:.610. Graham & Trotman, London. MooRE, J .G. ( 1 992) A syn-rift to post-rift transition sequence in the Main Porcupine Basin, offshore western Ireland. In: Basins on the Atlantic Seaboard: Petroleum and Basin Evolution (Ed. Parnell, J . ) , Spec. Pub!. geol . Soc. London, No. 62, pp. 333-349. Geological Society of London, Bath. M u LTER, H.G. & HoFFMEISTER, J . E . ( 1968) Subaerial laminated crusts of the Florida Keys. Geol. Soc. Am. Bull. , 79, 183- 192. PETRIE, S . H . , BROWN, J . R . , GRANGER, P.J. & LOVELL, J.P.B. ( 1989) Mesozoic history of the Celtic Sea basins. In: Extensional Tectonics and Stratigraphy of the North Atlantic Margins (Eds Tankard, A.J. & Balkwill, H . R . ) , Mem . A m . Assoc. petrol. Geol . , Tulsa, 46, 433-444. PLATT, N.H. ( 1986) Sedimentology and tectonics of the
Unpublished PhD. thesis, University of Oxford, 250 pp. PLATT, N.H. (1989a) Continental sedimentation in an evolving rift basin: the Lower Cretaceous of the Western Cameros Basin (Northern Spain). Sediment. Geol. , 64, 9 1 - 109. PLATT, N . H . ( 1989b) Lacustrine carbonates and pedo genesis: sedimentology and origin of palustrine deposits from the Early Cretaceous Rupelo Formation, Western Cameros Basin, Northern Spain. Sedimentology , 36, 665-684. PLATT, N.H. ( 1989c) Climatic and tectonjc controls on sedimentation of a Mesozoic lacustrine sequence : the Purbeck of the West Cameros Basin, Northern Spain. Palaeogeogr. Palaeoclimatol. Palaeoecol. , 70, 187- 197 . PLATT, N.H. ( 1 990) Basin evolution and fault reactivation in the western Cameros Basin, Northern Spain. 1. geol. Soc. London, 147, 165- 175. PLATr, N.H. & PuJALTE V. ( 1994) Correlation of Upper Jurassic- Lower Cretaceous of continental sequences from the southern Biscay margin northern Spain. 1. geol. Soc. London, 1 5 1 , 7 15 -726. PLATT, N . H . & WRIGHT, V.P. ( 1991 ) Lacustrine carbonates: facies models, facies distributions and hydrocarbon aspects. In: Lacustrine Facies Analysis (Eds Anad6n, P. & Cabrera, LI.), Spec. Pubis int. Assoc. Sediment . , N o . 13, pp. 57-74. Blackwell Scientific Publications, Oxford. PLATT, N.H. & W RIG HT , V.P. (1992) Palustrine carbon ates, calcretes and the Florida Everglades: towards an exposure index for the fresh-water environment? 1. sediment. Petrol. , 62, 1058- 107 1 . PLATr, N . H . , BURNS, S . J . , PLATT, J . D . & WIGNALL , P.B. ( 1991 ) Organic-rich deposits from the Callovian of the Sierra de Ia Demanda, Northern Spain. Eclogae geol. Helv. , 84(2), 483-508. RAwsoN , P . F. & RILEY, L.A. ( 1 982) Latest Jurassic-Early Cretaceous events and the 'late Cimmerian' unconform ity in the North Sea area. Bull. Am. Assoc. petrol. Geol. , 66, 2628-2648. READ, J . F. ( 1976) Calcretes and their distinction from stromatolites. In: Stromatolites (Ed. Walter, M . R . ) , pp. 255 -27 1 . Developments i n Sedimentology, 20, Elsevier, Amsterdam. RuFFELL, A.H. & CowA RD , M . P. ( 1 992) Basement tecton ics and their relationship to Mesozoic megasequences in the Celtic Seas and Bristol Channel area. In: Basins on the Atlantic Seaboard: Petroleum Geology, tology and Basin Evolution
Sedimen
(Ed. Parnell, J . ) , Spec. Pub!. Geol. Soc. London, Vol. 62, pp. 385-394. RuFFELL, A . H . ( 1992) Early to mid-Cretaceous tectonics and unconformities of the Wessex Basin (southern England). 1. geol. Soc. London, 149, 443-454. SALOMON, J. ( 1 982) Les formations continentales du Jurassique superieur-Cretace inferieur (Espagne du Nord, Chaines Cantabrique et NW Iberique). Mem. Geol. Univ. Dijon , 6, 228 pp. ScHUDACK, M. ( 1987) Charophytenftora und fazielle Entwicklung der Grenzschichten mariner Jura/Wealden in den Nordwestlichen Tberischen Ketten (mit Verglei chen zu Asturien und Kantabrien). Palaeontogr. Abt. Ser. B, 204, 1 - 180.
236
N . H. Platt
P.M. ( 1991 ) The development of Irish sedimen tary basins. J. geol. Soc. London , 148, 1 8 1 - 189. SuN S.Q. (1989) A new interpretation of the Corallian (Upper Jurassic) cycles of the Dorset coast, southern England. Geol. Mag. , 24, 139- 158. TALBOT, M.R. ( 1973) Major sedimentary cycles in the Corallian Beds (Oxfordian) of Southern England. Palaeogeogr. Palaeoclimatol. Palaeoecol. , 14, 293-317. TANKARD, A.J. & BALKWILL, H . R. (eds) ( 1 989) Extensional SHANNON,
Tectonics and Stratigraphy of the North Atlantic Margins.
Mem. Am. Assoc. petrol. Geol . , Tulsa, 46, 639 pp. TATE, M.P. ( 1993) Structural framework and tectono stratigraphic evolution of the Porcupine Seabight Basin, offshore western Ireland. Mar. petrol. Geol. , 10, 95 - 123. THALMA N N , C. ( 1989) Die Geologie entlang der San Leonardo- Oberschiebung zwischen San Leonardo de Yagiie und Cubilla, NW-Keltiberische Ketten, Soria, Spanien .
Provinz
Unpublished Diplom-Arbeit, Universitiit Bern, Teil B, pp. 126-270. VALENZUELA, M . , GARclA-RAMOS J . C . , GONZALEZ-LASTRA,
C. ( 1 986) The Jurassic sedimen tation in Asturias. Trab. Geol. Univ. Oviedo, 16, 1 2 1 - 132. WILDE, S. (1988) The Bathonian and Callovian of the northwest Iberian Range: stages of facial and palaeo geographical differentiation on an epicontinental plat form. Cuad. Geol. Iberica, 14, 1 13- 142. WILSON, R.C.L . , HISCOTT, R.N. , WILLIS, M . G . & GRADSTEIN , F.M . (1989) The Lusitanian Basin of west central Portugal: Mesozoic and Tertiary tectonic, strati graphic and subsidence history. In: Extensional Tectonics and Stratigraphy of the North Atlantic Margins (Eds Tankard, A.J. & Balkwill, H.R.), Mem. Am. Assoc. petrol. Geol . , Tulsa, 46, 341 -361. WRIGHT, V . P . , PLATI, N . H . & WIMBLEDON, W . A . ( 1 988) Biogenic laminar calcretes: evidence for calcified root mat horizons in palaesols. Sedimentology, 35, 603-620. ZIEGLER, P.A. ( 1 990) Geological Atlas of Western and Central Europe, 2nd (and completely revised) edn, 2 vols. Shell Internationale Petroleum Maatschappij BY, The Hague. 239 pp. J. & SUAREZ DE CENT!,
Spec. Pubis int. Ass. Sediment. (1995) 22, 237 - 2 64
The use of geochemical data in determining the provenance and tectonic setting of ancient sedimentary successions: the Kalvag Melange, western Norwegian Caledonides RODMAR RAVNAS
and
HAR ALD FURNES
Geological Institute, Univers ity of Bergen, Allegt. 41, N-5007 Bergen, Norway
ABSTRACT
The inferred Silurian Kalvag Melange is a thick (> 2500 m ) bedded olistostromal melange, which formed due to gravitational resedimentation driven by tectonic activity along the steep submarine slope of a basin margin and was deposited in the lower slope or base-of-slope environment. The melange consists of olistoliths, more than 2.8 km in length and up to 0.5 km in thickness, of shallow-marine deposits, deep-marine turbidites, bedded cherts, pebbly mudstones, calcareous rocks, lava-flow basalts and andesites, rhyolites and rhyolitic ignimbrites embedded in a sheared olistostromal 'groundmass'. The latter comprises chiefly debris-flow deposits. Interfingering channel-fill conglomerates and gully-fill turbidites with erosional contacts to the olistostromal 'groundmass' are present. The channel-fill conglomerates are dominated by debris of lithic-volcanic arenites and volcanic rocks. Olistoliths of volcanics and the volcanic debris in the channel-fill conglomerates were derived from alkaline, calc alkaline , !AT, MORll and boninitic sources. Geochemical and isotope data from the volcanic rock olistoliths reflect an evolution from magmatism associated with initial rifting to incipient back-arc spreading, and finally to volcanic arc magmatism. Geochemical analyses of conglomerate debris and sedimentary-rock olistoliths suggest derivation from an evolved, mature island arc, and an ophiolitic source. The volcanic arc is thought to have formed by rifting of accreted terranes along an active continental margin, and the MORJJ-type and boninitic debris most likely represent derivation from an older, accreted ophiolitic terrane. The sedimentological and geochemical data jointly suggest that the deposition of the Kalvag Melange took place in a back-arc basin , formed by progressive cannibalization of the volcanic arc apron . Based on the present knowledge of the development of marginal basins within the outboard terranes of the Caledonian orogen, the basin-evolution history recorded by the Kalvag Melange can be correlated with the history of other late Ordovician to early Silurian marginal basins.
INTRODUCTION
The primary control that tectonics exerts on the production, d ispersal, accumulation and lithification of sediments has long been recognized (Pettijohn et al., 1987). Provenance regions, and especially those forming parts of the overriding plate at destruc tive plate margins, can u ndergo significant and rapid uplift. Much information is lost due to subsequent erosion, deformation and/or tectonic displacement, and a complete in s itu record of the source rocks is rarely preserved. The sedimentary record of adjacent sedimentary basins can provide some constraints toward redressing this preservational bias (e.g.
Dorsey, 1988; Smith et al., 1988; Haughton & Halliday, 1991; Marsaglia et al., 1992). Intercalated volcaniclastic rocks and clasts in coarse conglom erates can be characterized and sometimes dated, providing supplementary geochemical and isotopic information about the hypothetical provenance area and the original plate tectonic setting of the depo sitional basin. The mineralogical and geochemical composition of terrigenous sediments and sedimentary rocks depend chiefly upon provenance, relief and climate of the source region(s), weathering, transport
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
237
238
R. Ravnas and H. Fumes
processes, basin geometry and morphology, depo sitional e nvironment and diagenesis (e. g. Bhatia, 1983; Basu, 1985; Helmold, 1985; McBride, 1985; Mutti, 1985; Normark, 1985; Ricci Lucchi, 1985; Dorsey, 1988; Johnsson & Stallard, 1989; McLennan et a/., 1990). Plate-tectonic setting has been advo cated as the primary control on sediment com position (Blatt et a/., 1980; Pettijohn et a/., 1987) and the link between plate-tectonic setting and modal framework of modern sediments (mainly sands and sandstones) appears empirically robust (e. g. see Crook, 1974; Schwab, 1975; Dickinson & Suczek, 1979; Ingersoll & Suczek, 1979; Dickinson & Valloni, 1980; Valloni & Mezzardi, 1984; Dickinson, 1985; Valloni, 1985). The nature of the primary framework grains can, however, be modified significantly, obliterated or destroyed during post depositional diagenesis, burial, deformation and/or metamorphism. The geochemical composition, element ratios and isotope composition of sediments seem less depe ndant upon these post-depositional processes (Hiscott, 1984; Robertson & Henderson, 1984; Argast & Donnelly, 1987; Larue & Sampayo, 1990), and as such have proved powerful and more reliable in provenance studies and discrimination of plate-tectonic setting of ancient sedimentary successions (Bhatia & Taylor, 1981; Bhatia, 1983; Bhatia & Crook, 1986; Roser & Korsch, 1986; McLe nnan et a/., 1990). The present paper discusses the provenance and tectonic setting of an olistostromal melange, the inferred Silurian Kalvag Melange of the western Norwegian Caledonides (Fig. 1), on the basis of data obtained from the melange's ' groundmass', olisto liths (of volcanic and sedimentary rocks) and cobbles and boulders of interbedded polymictic conglom erates. Secondly, an attempt is made to further detail the reconstruction and probable evolution of the melange's provenance region and depositional setting by integrating the present knowledge of the island arc/ophiolite terranes and marginal basins of the Scandinavian Caledonides.
GEOLOGICAL SETTING
The area between Solund and Bremanger (Fig. 1) forms part of the westernmost Norwegian Caledo nides. Three tectono-stratigraphical units are recognized: the lower, middle and upper tectonic units (Brekke & Solberg, 1987; Andersen et a/., 1990). The lower tectonic unit comprises Pre-
cambrian migmatitic orthogneisses (Jostedalen Complex), paragneisses and associated augen gneisses, amphibolites, eclogites and meta anorthosites (Fjordane Complex), and low- to medium-grade metamorphic sedimentary, volcanic and plutonic rocks (Askvoll Group) of the Western Gneiss Region. A quartz diorite from the Askvoll Group has yielded a middle Proterozoic magmati.c U-Pb age of 1640. 5 ± 2.3Ma (Skaret a/., 1994). The middle tectonic unit is separated from the lower by the extensional Kvamshesten Detachment Zone. It consists of the Dalsfjorden Suite, composed of various types of syenitic to charnockitic ortho gneisses, granites and gabbros (Kolderup, 1921), and the unconformably overlying pre-Silurian and Silurian metasediments of the H�yvik and Herland Groups, respectively. The H�yvik Group experienced polyphase deformation and associated metamorphism prior to deposition of the u nconform ably overlying Herland Group (Brekke & Solberg, 1987; Andersen et a/., 1990). The upper tectonic unit comprises the Solund Stavfjord Ophiolite Complex and its cover of meta sediments and metavolcanics (Fumes eta/., 1990). It is separated from the middle tectonic unit by a composite terrane, the Sunnfjord Melange, inter preted to have developed during obduction of the ophiolite complex upon an e arly to middle Silurian continental margin represented by the Herland Group (Berg, 1988; Andersen et a/., 1990; Alsaker & Fumes, 1994). The Solund-Stavfjord Ophiolite Complex has yielded a U - Pb age of 443 ± 3 Ma (Du nning & Pedersen, 1988). Single zircon U-Pb datings from its cover sediments have yielded both Archaean, Proterozoic and early Ordovician ages (Pedersen & Dunning, 1993). The Kalvag Melange is separated from the gneisses, metasediments and metamorphic plutonic and volcanic rocks of the lower, middle and upper tectonic units to the north by a major thrust zone (Hartz et al., 1994). The metasediments on Bremanger have bee n correlated with the H�yvik Group (Fumes et a/., 1990; Hartz et a/., 1994). The metamorphic igneous rocks include a number of small serpentinite bodies, amphibolites, green stones, gabbros and schists (Bryhni et a/., 1981; Hartz et a/., 1994), which may possibly correlate with the Sunnfjord Melange and/or the Solund· Stavfjord Ophiolite Complex and its cover. More over, Hartz et a/. (1994) recognized major E-W striking, southward dipping extensional shear zones along the northern and central parts of
TECTONOSTRATIGRAPHY Devonian
�
UPPER TECTONIC UNIT
Sunnfjord Melange (SM)
� Herland Gp.
� Hoyvik Gp.
� Dalsfjord Suite
Western Gneiss Region
Granodiorite
V7777i1 � � �
Coarse conglomerates Bedded olistostromal facies
� Qm G:::::J
OHstoHhs
. Q Q
Arenitic shallow-marine sediments (protolith blocks) Chert
r=l L=...J
Deep-marine sandy turbidites Ignimbrite/rhyolite
O sasalt ---Fault
0 '
.
1
km
0
5
10
15
20
25 km
g.
�
!}
[ill] �e�����j��::;:g ���f
The KalvAg Melange
"' 0
s
Devonian
'0' ...,
Sunnfjord Melange
Melange �gmundmass"
Cl
� I:..!..:....J
Kalvc\g Melange
� Gabbronorite/diorite
� �
LOWER TECTONIC UNIT
-diorite
Younger igneous instrusions
A L_j
MIDDLE TECTONIC UNIT
c:;· �
Q �����g��;f�� • D
LEGEND
�
LEGEND
undifferentiated
[ill
�
....L.....J..
.....1-..1-
(SSOC
&
nic � s cover/S�1?)
Cover of metasediments and lavas/intrusions
UPPER TECTONIC UNIT
� s· �·
�
Solund-Stavfjord OphiolitE Complex (SSOC)
..:
cs
"' ;:s :::> ;:s
Herland Group (HG) Hoyvik Group
R
Dalstjord Suite Western Gneiss Region undifferentiated
�
LOWER TECTONIC UNIT
Detachments between Lower. Middle and Upper Plates
Fig. 1. Simplified geological map and tectono-stratigraphy of the So lund- Bremanger area, compiled from Fumes eta/. ( 1990), Hartz eta/. (1994) and Osmundsen & Andersen ( 1994) , and simplified geological map of Frs;lya, modified from Bryhni & Lyse ( 1985).
t;5
\0
240
R. Ravnas and H. Fumes
Bremangerlandet (Fig. 1) that separate major plates, analoguous to the tectono-stratigraphy of the Sunnfjord area (Brekke & Solberg, 1987; Andersen et al. , 1990; Andersen & Jamtveit, 1990; Osmundsen & Andersen, 1994). The Kalvag Melange has been interpreted in terms of a tectonized olistostrome formed in an island arc or accretionary prism setting (Bryhni & Lyse, 1985; Fumes et al., 1990), and was tentatively included in the upper tectonic unit by Fumes et al. (1990). The melange succession is intruded by two plutons of gabbronoritic/dioritic and granodioritic composition (Fig. 1). The older gabbronorite/diorite has yielded a Sm-Nd age of 380 ± 26 Ma (Fumes et al., 1989), thus providing a minimum age for the melange formation. The uppermost stratigraphical level of the region is represented by early to middle Devonian clastic sediments, deposited in late to post-Caledonian collapse basins (Hossack, 1984; Norton, 1986; Seranne & Seguret, 1987; Andersen & Jamtveit, 1990). In the Bremanger area these sediments rest unconformably upon the pre-Devonian rocks as well as on the granodioritic pluton.
THE KALVAG MELANGE
The Kalvag Melange forms a more than 16-km-wide rock unit on the island s of Fqllya and Bremanger (Fig. 1). On Fr¢ya, the Kalvag Melange makes up a minimum 2500-m-thick succession consisting of clasts and large detached blocks (olistoliths) of native and exotic rocks (sensu Raymond, 1984) d ispersed in a pelitic 'groundmass' , and interbedded conglom erates and sandstones (Figs 1 & 2) (Ravnas, 1991). Graptolites and tabulate corals of probable Ordovician-Silurian age were reported by Reush (1903) and Kolderup (1928), but have not been confirmed by later studies. The melange succession shows no pervasive shear fracturing, but local zones of pervasively sheared matrix occur. Strong tec tonic deformation, including folding, faulting and boudinage, is limited to discrete, isolated zones (fault zones and shear fracture zones). Primary structures within olistoliths and the fragmental fabric of the melange 'groundmass' are generally well preserved within the contact aureoles of the two plutons, allowing distinction between different types of olisto liths and recognition of their primary depositional environments. The melange as a whole can be considered in
terms of its two principal components: the bedded olistostromal 'groundmass', and the olistoliths that are scattered in this groundmass. The olistoliths, which may exceed 2. 8 km in length and up to 0. 5 km in thickness, comprise shallow-marine (offshore to shoreface) sandstones and mudstones, deep-marine sandy turbidites, bedded cherts, pebbly mudstones, calcareous rocks, subaerial and subaqueous lava·· flow basalts and andesites with interstitial calcareous rocks and intercalated sediments, rhyolite and rhyolitic ignimbrite (Fig. 2A-E). The olistoliths are interpreted to have been emplaced as large slides and slumps. The melange 'groundmass' comprises mainly matrix-rich sedimentary breccias and gravelly mudstones (Fig. 2F & G) that are inferred to represent chiefly cohesive debris-flow deposits. The interbedded conglomerates and sandstones (Fig. 2H), interpreted to have been deposited from gravelly non-cohesive debris-flows and high-density turbidity currents and sandy high- to low-density turbidity currents respectively, commonly have sharp, erosive basal contacts, locally displaying large loadcasts. They form broad, lenticular bodies that locally interfinger laterally with the olistostromal facies of the melange's 'groundmass', and are inter preted to represent submarine slope channel- and gully-fill deposits. Black shales and subordinate fine·· grained turbidites, the latter probably representing overbank deposits of the submarine channels, form minor constituents of the melange succession. Locally the 'groundmass' is intruded by fine- to medium-grained sandstone dykes and sills. The Kalvag Melange is interpreted to have been formed due to gravitational resedimentation driven by tectonic activity along the steep submarine slope of a basin margin, and deposited in the lower slope to base-of-slope environment (Ravnas, 1991). Several points of evidence suggest that the entire melange succession chiefly is of sedimentary origin. On the island of Fr¢ya the melange d isplays a pervasive, although subtle, bedding defined by textural and compositional variations. The general stratigraphical way-up direction is recognizable, and there seem to be some large-scale thickening/ thinning upwards trends. These trends are correla tive throughout the island area, marked predomin antly by the stratigraphical thickness distribution of the m ass-flow facies (Ravnas, 1991). The slide and slump units (olistoliths) stand out as isolated, out sized peaks on this background, and are correlative (Fig. 3) only when associated with the maxima of the mass-flow facies thickening. The olistoliths that are
Geochemical data for determining provenance
encased in the bedded 'olistostromal' matrix show primary depositional contacts, and are in some cases draped by black shales. In some places, gradations down the inferred palaeoslope from sedimentary rock olistoliths of slide/slump origin into debris-flow deposits can be seen. The matrix of the melange is thought to have been derived from the fine-grained interlayers of the shallow marine sediments in addition to slope mud (see below). Partly similar clast lithologies are present in the interfingering channel-fill conglomerates and the olistostromal 'groundmass' or debris-flow deposits. Locally the overbank and channel-fill deposits show evidence of slumping prior to deposition from the subsequent flow, and the olistostromal 'groundmass' deposits locally are found as olistoliths and conglomerate debris. The spatial distribution of facies associations and olistoliths (Fig. 3) suggests that the melange developed during a series of successive resedimen tation events (Ravnas, 1991). Basal and internal penecontemporaneous, 'soft type' deformation (folding, faulting and boundinage) of sedimentary-rock olistoliths is attributed to their emplacement by sliding and slumping while still in a semi-consolidated state. The volcanic-rock olistoliths show little internal deformation, apparently due to their inherently rigid character. Sandstone dykes show cross-cutting relationships with the sheared olistostromal 'groundmass' and locally are found to intrude the deformed basal part of sedimentary-rock olistoliths. High-pressure-low-temperature miner alogy has not been observed, neither in olistoliths nor in the sheared matrix. The melange has been divided broadly into a lower part comprising mainly sedimentary slides, slumps and cohesive debris-flow deposits, and an upper part comprising mainly cohesive debris-flow deposits and subordinate slumps, fine-grained turbi dites and resedimented coarse conglomerates. The lower part contains few or no volcanic-rock olisto liths, whereas the upper part shows an upward increase in the content of volcanic-rock olistoliths.
GEOCHEMISTRY
Chemical analyses of samples from olistoliths, conglomerate debris and the olistostromal 'ground-· mass' were carried out by X-ray fluorescence. The glass bead technique of Norrish & Hutton (1969) was used for major element oxides, and pressed powder pellets for the trace elements, using inter-
241
national basalt standards and the recommended values of Govindaraju (1984) for calibration. Possible post-depositional alteration of basaits, vol caniclastic materials and sediments have been dis cussed in a number of studies. The elements Ti, P, V, Zr, Y, Nb and Cr, with the addition of Si, Fe, Mg, La, Th, Ce and Co in sedimentary rocks, are considered relatively stable during low-grade meta morphism (e.g. Cann, 1970; Coish, 1977; Shervais, 1982; Bhatia, 1983; Staudigel & Hart, 1983; Hiscott, 1984; Robertson & Henderson, 1984; Bhatia & Crook, 1986; Cas & Wright, 1988; Larue & Sampayo, 1990) and are used systematically through out this study. Ca, Na and K in sedimentary rocks show increased mobilization with increasing meta morphic grade (Larue & Sampayo, 1990) and thus are less useful as indicators of primary composition. The melange succession, forming mainly a biotite quartz hornfels (Bryhni & Lyse, 1985), shows prograde contact metamorphism towards the two plutons, especially the gabbronoritic/dioritic pluton, which has a border zone up to 150 m thick of migma tized and foliated host rock (Fumes et a!., 1989). To minimize the effect of remobilization of elements due to the increase in metamorphic grade, most samples were collected well away from the two intrusive bodies. To avoid possible systematic com positional variations of sediments related to vari ations in grain size (Roser & Korsch, 1986; Argast & Donnelly, 1987), chiefly medium- to fine-grained arenites (greywackes) were selected for analyses. Conglomerate debris of channel-fill conglomerates
Clasts of well-rounded igneous and well-rounded and angular sedimentary rocks are present, rep resenting gabbro, diorite, metabasalts, trachyande site/dacite, granite, rhyodacite/rhyolite, bedded chert, quartzite, subarkose, lithic arenite, lithic volcanic arenite and wacke, black shale and calcareous rocks. The chemical analyses of rep resentative samples are presented in Table 1. The metabasalts can be divided into three different types based on their geochemical characteristics (Fig. 4, Table 2). These are MORB (mid-ocean ridge basalts), tAT (island arc tholeiites) and rocks approaching boninitic composition. All types show negative Nb anomalies and exhibit either a flat pattern or show slight or pronounced depletion of P, Zr, Ti andY in the MORB normalized, multi-element variation diagrams (Fig. 4A), typical of back arc basin basalts, tAT and boninites respectively
242
R. Ravnas and H. Fumes OLISTOLITH COMPOSITION Alkaline potassic meta basalts Subalkaline MORB-Iike metabasalts Calcareous rocks Bedded chert Transitional type metabasalts Large olistoliths of native and exotic blocks
Ignimbrite and IAT metabasalts Deep-marine sandy turbidites
E
8 "'
'Background' sedimentary mass of bedded olistostrome (gravitational melange); ---contacts sedimentary, non-ophiolitic matrix and rock clasts/blocks
N
Shallow marine siliciclastic sediments
Alkaline potassic metabasalts, Ignimbrite and bedded chert
Shallow marine siliciclastic sediments
Fig. 2. (Above. ) Composite profile through the Kalvag Melange, compiled from measured sections along the eastern and
western coast of Frll\ya. (Right. ) Locations of field photographs are indicated by letters. (A) Flow-folded, welded ignimbrite. Lens cap is 6cm in diameter. (B) Massive andesitic and rubbly aa-type basaltic flows. Length of hammer is 40cm. (C) Shallow-marine (shoreface) sediments; interbedded inferred wave-formed ripple cross-laminated and parallel stratified sandstones. Scale bar is 5 em. (D) Normally graded, sandy Bouma-type Tabc and Tab turbidites. Lens cap is 6em in diameter. (E) Evenly bedded, turbiditic ribbon chert. Length of hammer is 40 cm. (F) Disintegration of olistolith of shallow-marine deposits into (G) clast-rich, debris-flow deposits of the melange 'groundmass'. Scale bars of (F) and (G) are 20cm and Scm respectively. (H) Clast-supported conglomerate representing inertial, non-cohesive, high-density gravelly debris-flow. Note polymictic clast population. Length of hammer is 40cm.
Geochemical data for determining provenance
243
t
B A
c ----
-
----
2000
? --
--
. ..:.; ��-�:-<�:
_.·����:�Jg�6�c��IC RO��§ts !D
---<< . ...:-:-:�.::.:�:�
· : -. ·�:-: !J . ·. : .- :·- : : : ·(::-: :. ·� .. . ·. .· . ·. · · : ?. · . · ·
1500
. ·
. ·
-
+---- 2.0km ___
2.5km
•
?
.
1500:
.
.
Stratigraphic height
.
.. .
.
uncertain
�)1 --��-�-� /
CHANNEL-/GULLEY FILL DEPOSITS ..,
1500
... ... ... . . . . . . .
.--:--=--.� �.-.-,.
------.
___
:;_,
_.,..-,""'":"-:- .-.-:
.
E
: :- ;,·,
_.;./ · Stratigraphic\..·...:....:...:_> height
o o"';--
2.8km -------!�
'/ /
<:... . .
oo
"�-·����OLI TH OF DEEP-MARINE SANDY TURBIDITES
� �
--<;_.:.-·� .
D
-o--
:
!:;·
·
"' ;:s "'-
500
uncertain
·
·� : �: :;i�::.:::.�::::-:-:::.-:-::.:-6
("
.� : 17 & & .::
Fig. 3. Spatial distribution of olistoliths within the Kalvag Melange. Possible correlations are marked by dashed lines . Sections A and B are generalized logs
interpreted from the map, whereas sections C, D and E are simplified versions of measured sections described by Ravnas (1991).
:;_,
"' ""' ;:s
� � � �
Table 1. Representative geochemical analyses of conglomerate debris Sample no. R8-157 R8-163 R8-164 R8-166 R9-17 R9-21
Location
Debris type
Si02
MORB
49,60
MORB
51,37
Mulen
MORB
50,52
Mulen
MORB
Mulen Mulen
Mulen Mulen
Ti02
Al203 Fe203" MnO
1,65
14,43
1,18
16,58
1,47
13,65 15,13
.50,23
1,55
MORB
49,00
1,91
MORB
49,82
MORB
14,34
1,63
14,19
MgO 6,92
CaO 9,01
12,60
0,29
9,34
0,23
5,79
10,62
12,98
0,24
6,59
9,23
11,62
0,26
6,53
9,80
13,05
0,34
6,77
9,93
12,51
0,30
7,21
9,07
Na20
1<20
P205
LOI
Total
v
�
�
�
�
�
�
&
y
li
0,79
99,29
232
203
54
45
20
41
157
35
103
175
308
39
0,60
98,8�
237
257
56
62
3
109
78
6
72
20
1�
38
0,34
99,65
221
346
50
101
3
101
7
88
0,88
99,05
269
314
54
51
23
117
195
39
114
288
114
167
53
11
295
170
43
35
36
149
48
18
22
1,00
0,13
2,84
0,32
0,10
2,76
0,66
0,12
3,59
0,44
0,17
2,48
0,25
0,10 0,21
1,66
98,66
2,86
2,65
1,12
0,13
0,30
0,70
98,66
99,34
234
48,47
2,30
14,35
13,87
0,28
6,63
7,65
2,13
R9-16
Mulen
MORB
53,74
1,40
14,88
10,22
0,23
5,09
8,51
2,73
1,12
0,32
0,13
0,95
98,19
195
R8-155
Mulen
IAT
53,22
0,79
22,86
4,83
0,08
2,32
7,13
4,62
1,07
0,11
1,87
98,89
99
3,81
1,41
1,20
99.40
6,91
4,87
3,71
2,70
'o.o5
1,49
98,47 100,36
R9-15
R8-156 R8-162
Mulen
Mulen Mulen
51,87
0,55
IAT
49,71
0,58
IAT
0,26
3,91
4,80
0,03
17,06
14,48
18,05
10,23 10,10
0,24
5,88
10,10
3,21
0,25
0,13
0,55
0,23
9,04
3,74
1,62
1,34
0,03
3,96
98,42
0,19
6,23
4,13
1,57
0,03
1,91
98,57
6,30
3,98
3,89
1,48
0,07
2,56
7,93
0,03
R8-165
Mulen
IAT
53,64,
0,98
15,28
R9-18
Mulen
boninitic
51,33
0,33
13,58
13,23
52,58
0,28
17,71
10,51
0,17
11,32
3,55
12,56
3,59
76,13
0,29
0,16
12,00
2,21
75,03
0,22
11,36
3,22
71,06
0,27
12.55
5,52
ign./rhyol.
76,09
0,22
11,37
dac./tr.and.
69,24
0,29
14,37
67,05
0,43
16,52
3,77
100,60
0,02
0,35
98,51
0,_03
98,00
0,05
1,39
R8-161 R9-30 R8-167
Mulen Mulen Mulen Mulen Mulen
R8-168
Mulen
R8-169
Mulen
granite (co) granite (co) granite (m) granite (m) 1gn./rhyol.
R9-38
Mulen
dac.ltr.and.
R8-170
Mulen
chert
R8-171
Mulen
chert chert
0,33
0,04 0,06
2,89
26
47
4
374
11
28
27
13
328
165
73
163
10
24
29
14
M8
313
10
24
0,04
0,55
99,86
40
0,07
1,16
99,65
38
5,08
0,80
0,08
0,59
99,82
50
3,98
0,40
0,03
1,44
99,50
33
0,14
0,75
5,20 5,51
18
2,37
0,10
1,44
2,34
2,66
0,04
0,39
2,88
2,56
0,06
0,94
4,55
5,27
0,40
47
1,74
4,50
4,18
0,42
0,23
1,02
98,56
0,07
99,93
69
45
0,27
0,01
0,04
0,08
0,22
0,01
O,Q2
0,23
101,83
0,53
0,02
0,02
0,30
102,05
M H
� U
0,13
101,44 99,76
�
�
0,90
101,53
30
16
0,51 0,63
0,02
0,04
0,02
0,12
1,04
0,08
0,04
0,01
0,26
0,96
1,24
0,07
0,02
0,16
1,52
2,03
0,18
R9-28
Muten
99,65
0,03
0,34
0,01
R9-36
Mulen
chert ang.
92,54
0,07
3,67
0,61
R8-160
Mulen
subarkose?
92,06
0,07
4,26
0,32
0,53
0,50 0,51
0,04
0,01 0,02
0,53
0,30
101,05
29 25
53 23
51
71
48
251
57
22
14
201
27
14
145
32
14 42
12
87
10
39
220
7
18
21
95
24
42
506
13
8
20
306
24
38
161
9
10
216
15
3
47
213
19
29
18
35
174
13
20
35
71
24
43
310
10
202
18
79
19
35
105
9
25
96
11
91
4
23
31
67
17
26
24
231
5
63
18
5
29
26
8
15
17
13
5
12 13
§
38
186
5
[
10
141
20
38
62
13
231
12
150
11
78
47
82
33
70
16
255
18
218
14
144
69
118
56
10
3
28
11
3
143
55
5
47
9
4
75
48
6
84
6
47
59
23
54
57
17
16
14
�
922
215
12
6
66
13
"' 0 (") � "'
4
1192
23
4
C)
16
12
14
13
8
15
12
15
5
21
21
23
7
25
40
46
Th
35
8
67
10
17
82
18
20
10
12
2,30
1,17
4,32
43 41
11
27
0,58
Mulen
0,59
2,99
9
154
0,04
0,03
R9-27
chert
0,97 1,09
66
78
0,69
0,38
0,22
72,48
R9-24
0,13
66
158
0,44
4,57
75,36
R9-23
77,45
349
7
0,64
granite (co)
232
15
4,65
1,51
Mulen
17
51
1,85
11,24
R8-159
24
42
0,39
2,76
43
100,02
0,01
12,41
25
100,12
0,03
0,18
312
0,70
0,32
0,25
72.69
2
1,15
5,81
54,89
granite (co)
73
0,08
2,88
diorite
Mulen
75
277
0,04
0,61
Mulen
R8-158
40
197
0,63
0,03
R8-154
651
33
0,75
81
480
98,80
0,03
836
1
0,90
1,13
28 17
0,02
3,12
18
36
38
1,98
6,77
25
32
32
98,30
3,04
'5,36
39
10
34
0,58
7,40
0,17
150
550
-47
32
170
0,16.
9,37
566
21
144
9,18
16
15
163
13
116
17,04
345
53
51
161
96,29
0,30
4
12
28
99
110
185
98,59
50,23
260
57
1,30
gabbro
19
34
�
233
128
22
Mulen
147
95
52
R8-153
260
17
130
54
0,20
21
23
61
173
9,95
6
18
24
27
99,23
16,93
43
37
24
63
27
0,26
61
160
71
83
49
120
53,61
45
705
68
52
boninitic
42
171
�
19
178
62
Mulen
2
-�
=n
�
31
175
57
R8-172
17
8
�
250
199 1050
boninitic
14,45
66
48
177
3,94
Mulen
R9-22
60
88
•
5
27
44
39 41
11
50
14 15
11
R9-29
Mulen
subarkose?
91,18
0,11
4,59
1.24
0,30
1,04
1,84
0,21
0,02
0,47
101,02
M
�
17
14
10
39
111
10
R8-151
Mulen
p. mudst.
55,95
0,77
17,42
6,98
0,09
3,08
4,40
4,54
1,54
0,15
3,54
98,47
234
196
31
89
39
130
72
261
24
156
11
36
30
517
13
17
73,74
0,47
10,46
3,83
241
15
89
4
19
26
432
17
10
2,38
99,24
37
0,06
2,49
69
1,51
0,11
62
2,73
1,00
53
2,78
2,69
14
2,05
2,90
151
0,04
1,53
171
5,52
0,04
98,03
231
149
21
69
33
95
65
133
22
103
5
28
32
363
22
9
0,07
1.62
1,94
2,52
1.22
0,09
0,75
100,18
89
54
16
23
84
51
93
24
1529
43
52
47
265
22
26
140
180
38
45
29
105
69
262
34
255
14
90
40
727
29
129
187
33
36
26
97
55
220
29
242
11
72
45
627
30
Mulen
bl. mudsh.
R9-33
Mulen
bl. mudsh.
68,44
0,50
12,01
R8-175
Mulen
x-lam arenite
76,70
1,62
8.81
4,84
R9-31
Mulen
lit.-volc. aren.
R8-152
R9-32
Mulen
tit.-volc. aren.
58,92
0,97
16,04
8,17
0,08
3,38
2,76
4,45
1,63
0,10
1,75
98,24
63,31
0,94
13,96
7,86
0,10
3,12
3.02
3,31
1,34
0,09
1.18
98,22
146
65
f} 1i)
�· s �'1::3
\3
12
39
!:)
'Cj' ....
"" "' ;:s !:> ;:s
�
10
LOI, loss on ignition; Fe.o; , total iron as Fe.03 + 1 . 1 x FeO.
�
A
8 COMENOITE
CD
100
PANTELLEAITE
\ PHONOLITE '
10
0,1
10
- - -� --- � I BASALT / . ., . / --------
0,01
.... .. . . -
;;_�// -��2�-----
' '
............
0
Zr/Ti02
®
RHYOLITE
0
0,1
100
I I
ANDESITE
····-......
ALKALI BASALT
SUB-ALKALINE BASALT
�
0,001 /
-l-------+--+---1 ®
1000
10
0,1
0,01
Nb/Y
0,1
,,,,
I I I I
0,01
®
100
ppm Cr
I I I I I I
100
\
\,.rJl
10 ID a: 0 :IE
>< 0 0 a:
D
[]
10
0,1
0
100
IAT
I
\
WPB \
I I I\ I I \ I \ \ I \ :
I\
+------+-'�-���--_L-�
100
10
®
10
0,01
\ \ \
t, t: II
ppmY
'
Trachyandesilic/dacilic debfis
10
'
\
WITHIN PLATE Ti02(%)
I
I \
\
'
1
' I I
0,1
+---..C.:..---1----_____,
0,1
®
100
Granitic and rhyolite debris 10
10
100 ppmZr •
a •
0,1
o
Sr
K
Rb
Ba
Nb
Ce
p
Zr
Ti
y
Cr
•
MORB melabasatts Boninilic melabasalts and gabtxoic rocks tAT metabasatts granitic and rhyolitic debris trachyandesitesl dacites
1000
247
Geochemical data for determining provenance
Table 2. Classification of metabasaltic and gabbroic conglomerate debris according to the discriminant plots of Fig. 4B
Discriminant plot Nb/Y-Zr/Ti02 Y-Cr Zr-Ti02 (%) Total
A
B
c
D
Metabasaltic debris
Metabasaltic debris
Metabasaltic debris
Gabbro/diorite debris
Subalkaline andesite/ basalt MORB/WPB MORB (LATfWPB)
Subalkaline andesite/ basalt LAT (MORB) LAT
Subalkaline andesite/ basalt LAT Boninite
Subalkaline andesite/ basalt !AT Boninite
MORB
LAT
Bonini tic
Boninitic
MoRB, mid ocean ridge basalts; WPB, within plate basalts; LAT, island arc tholeiites.
(Saunders & Tarney, 1979; Basaltic Volcanism Study Project, 1981; Hickey & Frey, 1982; Cameron et al., 1983; Thompson et al., 1984; Crawford, 1989; Wilson, 1989). This tripartite subdivision is also suggested by the discriminant plots of Fig. 4B. The debris of phaneritic, mafic volcanic rocks (gabbro and diorite) show a similar chemical composition to the boninitic samples, as suggested by their nearly identical MORB normalized, multi-element variation diagram patterns (Fig. 4A) and the discriminant plot of Fig. 4B. The gabbro and diorite are interpreted accordingly to represent the plutonic equivalents of the boninitic metavolcanics (see further discussion below). Although the value of geochemical analyses of acidic volcanic rocks has been questioned (e.g. see discussion in Cas & Wright, 1988, and references therein), the chemical composition of the trachy andesite/dacite cobbles testifies to their alkaline affinity (Fig. 4) and demonstrates their incompati bility with the granitic and rhyolitic rock debris. The rhyolitic and fine- to coarse-grained granitic clasts display a similar mineralogy (particularly diagnostic is the presence of blue quartz in both types) and nearly identical chemical compositions, and are interpreted to represent extrusives and their intrusive equivalents. Their geochemical signature (e.g. pro nounced negative Nb, P and Ti anomalies, see Fig. 4A) indicates derivation from a depleted tAT parent (Pearce et al., 1984; Holm, 1985), and suggests erosion of low-K granitic rocks, typical of island arcs (Ewart, 1982). Moreover, the rhyolitic debris exhibits a mineralogical and chemical composition
close to that of the rhyolite and rhyolitic ignimbrite olistoliths, thereby suggesting a common protolith (see interpretation below). Bedded chert occurs as large angular blocks up to 1. 7m across. Angular cobbles of 'quartzite' are, based on mineralog.ical and geochemical criteria, interpreted to represent disrupted chert beds derived from the ribbon chert. The subarkosic debris are geochemically similar to chert, but differ with respect to their coarser grain size and higher proportion of plagioclase. However, detrital sodic plagioclase locally is present in some chert beds of the larger bedded chert olistoliths, suggesting that a common parent cannot be ruled out. Well-rounded quartzite clasts are recrystallized completely, and their protolith (whether representing vein quartz, quartzitic sandstone or chert, see below) cannot be determined. Well-rounded cobbles of lithic arenites differ mineralogically and geochemically from the angular debris of lithic-volcanic arenites. The lithic arenites consist, in decreasing order, of quartz, lithic sedi mentary grains (quartzite/quartz schist), sodic plagioclase, alkali feldspars, ilmenite, zircon and minor apatite and opaques. Modal ap.alyses indicate either a continental craton or recycled orogen/ collision orogen as provenance (Dickinson, 1985; Valloni, 1985), as is also suggested by their chemical composition. Lithic-volcanic arenites, on the other hand, are texturally, mineralogically and geochemi cally similar to the olistoliths of shallow-marine sandstones and deep-marine turbidites (see description and interpretation below), and are
Fig. 4. (Opposite.) (A) MORB normalized, multi-element variation diagrams of conglomerate debris of volcanic rocks.
(1) MORB metabasalts, (2) LAT metabasalts, (3) boninitic metabasalts and debris of gabbroic rocks, (4) trachyandesites/ dacites, and (5) granites and rhyolites. The order of the elements and the values of the normalizing constants are from Pearce (1983), except for Cr, which is from Pearce (1980). (B) Discriminant plots for conglomerate debris of ( 1) volcanic rocks, and (2) and (3) mafic volcanic rocks. Discriminant plot (B . 1 ) is from Winchester & Floyd (1977) , whereas discriminant plots (B.2) and (B.3) are from Pearce ( 1980).
248
R. Ravnas and H. Fumes
accordingly interpreted to have been derived from such deposits. Angular clasts of pebbly mudstone and black shales are geochemically similar to the olistoliths of pebbly mudstone and interbedded black shales of the melange succession, respectively. Furthermore, black shales are present only in conglomerate deposits resting with erosive basal contacts on black shale, thus suggesting the intrabasinal provenance of this clast type. Well-rounded pebbles of calcareous rocks have not been analysed geochemically. Although an extrabasinal origin of this clast type cannot be excluded, they may have been produced by erosion of lava-flow basalts with calcite-filled interstices now present as large olistoliths in the melange. Olistoliths of volcanic rocks
Geochemical analyses of samples from the different olistoliths of lava-flow basalts and andesites (Table 3) show considerable variations, allowing subdivision into alkaline, mildly alkaline to subalkaline MORB like and island arc basalts (Fig. 5 & Table 4). In addition there are some transitional types. Isotope analyses were performed on samples from olistoliths of ignimbrite and alkaline and subalkaline MORB-like basalts/andesites, and the results are presented in Table 5. The Es, and ENct values are calculated for several ages (470, 450 and 430Ma) based on the interpretation that the Kalvag Melange forms part of the Caledonian island arc/marginal basin terrains (see discussion below). The high E5, values of the lava-flow basalts/andesites indicate contamination, possibly from (a) syn- or post-lithification hydro thermal solution(s) from which the calcitic infills of vesicles and interstices precipitated (Ravnas, 1991). As the volcanic-rock olistoliths occur scattered in the olistostromal groundmass and do not form parts of a continuous succession, the geochemistry of the different volcanic-rock olistolith types are described separately, followed by a brief discussion of their possible primary tectonic setting. Alkaline basalts comprise subaerial compound flows of pillowed and scoriaceous, porphyritic and vesicular, greenschist facies metabasalts. Phenocrysts are augite/diopside, saussuritisized plagioclase, primary sanidine and microcline pseudomorphs, probably after leucite. Calcite-filled amygdules are common. The chemical composition, characterized by high content of alkali/alkaline earth metals, Ti02, Zr and Nb, denotes a strong alkaline affinity of these
metabasalts. The high K content, together with the inferred primary texture and mineralogy jointly suggest that these rocks belong to the potassic ig neous rock suite (Foley et at., 1987; Wilson, 1989), probably representing shoshonites transitional into leucite-tephrites (Bates & Jackson, 1987). The original tectonic setting of these metabasalts is difficult to assess. The high contents of Ti02 and Nb, and to a lesser extent Zr, typical of ocean island basalts (O!B), argue against a subduction-related setting (e.g. Perfit et al., 1980; Morris & Hart, 1983). However, high Ti02 and Nb values are reported from alkaline basalts of back-arc, intraplate settings undergoing extension (e.g. Thorpe et a/., 1984; Briggs et a/., 1990) and subduction-related ultrapotassic rocks (Thompson, 1977; Edgar, 1980; Thompson eta/., 1984). The MORB normalized, multi element variation diagrams resemble closely the pattern shown by continental flood-basalts (CFB) (e.g. Thompson et a/., 1984; Wilson, 1989), and differ from those of OtB by their lack of pronounced positive Nb anomalies. Provided these metabasalts represent potassic rocks, the chemical composition indicates a transitional character between lamproites and potassic/ultrapotassic rocks of active orogenic zones (i. e. group I and III potassic rocks of Foley et at. (1987)), whereas the content of Nb relative to Zr strongly suggests a subduction-related origin (see discussion in Thompson & Fowler, 1986; and their Fig. 9). The calculated ENct value (Table 5) coincides with values reported from both OtB, potassic rocks, island arc and active continental margin volcanics., as well as some rift-related continental flood-basalts (e.g. Faure, 1986; Wilson, 1989; and references therein). Accordingly, rift-related, continental within-plate or back-arc, or alternatively subduction·· related potassic volcanics provide the best analogues for these alkaline metabasalts. Subalkaline MORB-like basalts are present as com·· pound accumulations of greenschist facies aphyric to slightly phyric, vesicular, subaerial aa and pahoehoe, and subaquatic, pillow lava-flows. The flow-units range in composition from basaltic to andesitic, with the pahoehoe- and pillowed-type flows representing the former. Saussuritisized plagioclase and abundant calcite-filled amygdules are present in all flow types. Despite some variations in geochemical signatures between different olistoliths (see Fig. 5 and dis·· cussion below), all metabasalts classify as sub·· alkaline MORB-like (Fig. SB, Table 3). The MORB normalized, multi-element variation diagram pat terns (Fig. SA), however, indicate a slight alkaline
Table 3. Representative geochemical analyses of volcanic rock olistoliths Si02_
Ti02
Al203 Fe203"
Alk.basalt
47,25
2,88
21,65
9,31
0,18
3,88
7,91
Alk.basalt
51,83
2,47
19,68
5,16
0,18
2,35
13.75
Sleneset
Alk.basalt
52,36
2,49
19,67
5,49
0,23
2,19
13,69
0,96
R9-13
Sleneset
Alk.basalt
52,60
2,59
19,97
4,96
0,20
2,42
9,92
R9-14
Sleneset
Alk.basalt
53,62
2,33
18,05
5,23
0,18
2,52
10,33
R9-52
Kalvegj.n.
MORB-Iike
53,74
2,30
23,19
6,27
0,06
3,63
4,49
2,94
R9-53
Kalvegj.n.
MORB-Iike
54,56
2,16
22,84
6,27
0,06
3,42
4,40
3,49
R9-54
Kalvegj.n.
MORB-Iike
58,44
1,96
19,64
8,74
0,13
3,99
0,00
0,88
R9-55
Kalvegj.n.
MORB-Iike
53,44
2,17
22,65
7,71
0,12
3,69
2,74
3,19
3,86
R9-56
Kalvegj.n.
MORB-Iike
57,57
2,00
20,74
7,00
0,21
3,47
2,33
2,97
3,23
RB-48
Botnan.
MORB-Iike
50,13
1,60
18,15
12,54
0,16
7,02
6,17
1,84
RB-50
Botnan.
MORB-Iike
52,64
1,58
19,60
10,12
0,08
5,90
3,93
4,99
RB-51
Botnan.
MORB-Iike
51,15
1,59
18,55
12,04
0,11
8,40
3,31
4,04
0,64
RB-54
Botnan.
MORB-Iike
52,62
1,64
20,32
10,58
0,20
4,30
3,04
3,51
3,50
R9-58
Botnan.
MORB-Iike
50,33
1,72
19,39
11,05
0,10
8,12
4,06
3,77
R9-59
Botnan.
MORB-Iike
51,90
1,53
18,49
9,50
0,10
6,51
6,29
4,08
R9-60
Botnan.
MORB-Iike
54,05
1,63
19,80
8,94
0,06
5,69
4.97
3,56
Sample no.
Location
Rock type
RB-8
Sleneset
RB-9
Sleneset
R9-12
MnO
MgO
CaO
Na20
v
�
�
N
�
239
�
�
&
v
�
m
45
70
35
21
100
42
87
32
125
57
106
26
89
13
101
25
72
12
93
158
101
209
47
376
49
119
41
371
37
104
83
146
127
265
36
113
17
121
98
591
38
129
18
120
76
514
42
105
�
�
154
143
680
56
567
30
175
37
34
519
29
174
257
47 35
74
456
31
202
37
101
409
34
180
36
�
�
70
34
841
40
54
637
20
34
36
550
33
56
41
870
36
38
41
905
14
17
26
34
17
29
41
167
14
17
23
29
546
10
20
26
45
510
17
17
46
307
26
13
309
32
24
178
K20
P205
LOt
Total
1,80
4,61
0,54
1,00
100,01
272
0,95
3,36
0,49
1,13
100,24
193
188
2,53
0,51
1,93
100,11
199
178
1,71
5,02
0,53
1,94
99,92
230
198
0,89
6,41
0,63
1,52
100,18
199
169
3,08
0,30
7,31
100,01
2,46
0,31
6,69
99,98
286
439
30
128
5,86
0,38
9,40
100,01
255
455
41
160
0,45
7,80
100,02
268
409
34
178
0,50
12,54
100,03
238
430
28
224
2,27
0,14
3,26
100,03
283
28
52
21
20
123
40
316
36
95
0,96
0,21
3,70
100,00
210
647
52
277
84
107
20
407
29
87
0,20
4,70
100,03
229
633
60
266
74
114
12
338
26
90
29
12
113
0,31
7.87
100,01
249
437
49
239
86
122
78
404
38
102
11
34
38
431
1,24
0,21
3,42
99,99
243
742
59
306
52
124
25
380
29
98
10
36
20
220
9
1,39
0,21
1,33
100,01
198
560
47
212
56
99
29
388
26
87
35
20
254
10
0,85
0,41
6,50
99,96
216
574
41
247
67
108
16
446
30
84
9,
10
19
27
160
247
292
440
34
38
35
111
241
15
0
!}
Mulen
Trans. bas.
51,24
1,12
18,87
7,55
0,35
3,78
15,61
0,75
0,12
0,28
3,84
99,66
156
27
41
38
88
3
454
26
48
4
20
28
61
20
Minnet
Trans. bas.
50,60
1,91
18,59
12,07
0,24
9,17
2,78
3,22
1,11
0,31
2,33
100,00
229
302
59
93
28
104
40
446
29
157
10
70
11
787
11
R9-89
Mulen
Trans. bas.
55,44
0,99
15,78
10,71
0,29
5,10
6,53
4,60
0,25
0,27
1,99
99,94
189
257
47
85
48
110
5
26
81
6
34
27
264
12
R9-90
Mulen
Trans. bas.
53.15
1,21
18,05
10,09
0,26
5,64
6,06
4,29
0,87
0,29
1,07
99,91
203
290
43
56
88
89
24
221· 361
27
103
35
27
1251
0,53
0,25
0,72
100,01
168
235
47
94
16
104
15
367
25
64
30
15
117
17
384
51,00
1,03
tAT
49,09
1,09
tAT
50,45
1,01
Sjeneset
tAT
50,63
1,14
Sjeneset
tAT
49,63
1,04
tAT
RS-19
Sj0neset
RB-20
Sjeneset
RB-21 R9-1
!.:)
� ..., � �
� ;:;: �-
·
5,45
9,69
2,95
0,24
6,35
4,27
1,89
4,00
0,19
6,71
100,03
262
77
53
32
63
141
111
447
20
55
24
0,26
7,13
5,82
1,76
2,78
0,28
4,53
100,06
242
54
58
26
39
131
82
329
19
44
36
21
264
8
11.42
0,22
6,56
7,22
3,13
1,26
0,27
5,37
100,09
239
50
49
25
50
130
32
475
22
51
23
16
195
13
11,39
0,22
6,74
7,05
2,49
1,84
0,21
4,19
100,02
234
64
50
31
44
140
78
248
17
48
27
30
448
1
1,22
2,15
0,01
2,70
98,62
28
2
18
201
63
140
94
400
59
47
1742
22
11
682
10,67
0,30
21,03
11.90
17,66
12,93
18,25 19,41
18,14
� s. �
[
RS-95
Sjeneset
CJ
::!
RB-223
RB-18
Th
3,21
0,28
0
�
;j
Mulevik
lgn.matrix
68,52
0,24
15,37
4,89
0,03
RS-103
Mulevik
lgn.matrix
70,92
0.23
14,61
3,45
0.03
1.95
0,65
4,70
1,18
0,02
1,55
99,30
27
4
13
155
45
174
4
50
49
24
18
RS-104
Mulevik
lgn.matrix
70,87
0,23
14.53
3,52
0,03
2.00
0,64
3,96
1,18
0,02
1,49
98,47
28
1
10
5
101
43
146
62
383
5
42
46
719
17
11
RB-105
Mulevik
lgn.matrix
70,05
0,22
14,14
4,91
0,02
3,16
0,28
2,76
1,24
0,02
2,08
98,87
30
2
18
4
161
35
215
115
390
10
70
58
841
22
R9-84
Mulevik
lgn.matrix
71.70
0,20
12,73
4,84
0,02
3,24
0,97
4,81
0,67
0,02
1,83
101,40
29
8
18
6
179
16
378
81
348
5
25
12
61
37
48
217
RB-102
R9-82
Mulevik
Rhy. core
RB-40
Mulevik
Rhyolite
RB-98
Mulevik
Ash-layer
R9-85
Mulevik
Ash-layer
83,54
0,13
7,98
2,56
0,02
0,41
0,14
4,20
0,02
0,01
0,56
99,57
73,18
0,23
14.18
2,02
0,02
0,31
0,46
7,59
0,04
0,02
0,65
�8.70
66,58
0,22
15,07
6,14
0,05
3,63
0,35
1,40
2,87
0,02
2,56 - 98,88
65,47
0,24
16,36
6,04
0,05
4.07
0,29
1,12
3,74
0,00
2,75
100,20
4
47
353
49
48
395
16
19
40
20
11 27
65
73
352
68
66
16
28·-
24
155
80
101
57
392
54
34
3319
10
28
23
159
106
87
72
414
46
42
4180
4
7
30
86
LOI, loss on ignition; F�o; , total iron as Fe203 + 1 . 1 x FeO; Listed values of major oxides of mafic volcanic rocks are recalculated values to 100% at LOI
=
"" � ;::s !:) ;::s
�
11
0.
�
250
R. Ravnas and H. Fumes
A
100
10
Q)/>;�
/'/-
. .
.
•
�� c.< '< c�
COMENOITE
B
Olislolilh Slenesel- Alkaline
PANTELLERITE
I i
\PHONOLITE
\
'
' '............
0.1
- ... ... .. ZrfTI02
0.1
100
®
0,01
Olistolith Kalvegjerdsneset
•
A NDESITE
Subalkaline MORB-Iike (KO)
--�
I BASALT
-- -,t- --
� ., ,� �
ALKALI· BASALT
SUB-ALKALINE BASALT 10
0.001
+------+---1---i
O.Ql
10
0.1 Nb/Y
®
1000 0.1
100
®
,/'"'
I
Olislolilh B otnanesesl- Subalkallne MORB-Iike (BO)
, , , ,
10 ppmCr
I I I
\I
100
� •
.t.6
1
' '
m a: 0 :::E
100
I ' I I I I
'
0.1
-" ()
0 a:
I
G)
10
IAT
f, 1: h
WPB
I\
I\ • I \
\ I I
I
I I ', I \ :I
+-------��---'�-�--�---i '
100
10 10
ppmY
®
10
----- ......
100
'
\ I
I
0.1
® Olistolilhs of IAT metabasahs
TI02(%)
\ \ \ \ I
1
'
0.1 100
10
ppmZr
•
0
01. alkaline metabasalts 01. subalkaline/MORB-Iike metabasalts (KO) 01. subalkaline/MORB-Iike metabasalts (80)
0.1
O.Ql +--+---+--+--1----'< Ti y Cr p Zr Rb Be Nb Ce Sr K
o
01. transitional metabasalts
•
01. IAT metabasalts
1000
2S1
Geochemical data for determining provenance
Table 4. Classification of metabasalts of mafic volcanic rock olistoliths according to the discriminant plots of Fig. 58
Discriminant plot
B
Alkaline
Subalkaline basalt
MORB/WPB WPB (MORB)
MORBIWPB MORB/WPB
WPB
MORB?
Nb/Y- Zr/Ti02 Y - Cr Zr-Ti02 (%)
c
A
Total
D
E
Subalkaline andesite/ Subalkaline andesite/ Subalkaline andesite/ basalt basalt basalt !AT MORB/WPB MORB/WPB !AT (MORB) All MORB (!AT) MORB?
!AT
Transitional
A, Large olistolith of alkaline metabasalts exposed at Sleneset; B, large olistolith of subalkaline MORB-Iike metabasalts exposed at Kalvegjerdsneset (KO); C, large olistolith of subalkaline MORB-like metabasalts exposed at Botnaneset (BO); D , smaller olistoliths of transitional metabasalts; E , smaller olistoliths of !AT metabasalts. Table 5. Isotope data from volcanic-rock olistoliths
470 Ma Sample AI kaline basalt MORB-like (80) basalt MORB-like (BO) basalt andesite Ignimbrite matrix
450Ma
430 Ma
ENd
Esr
EN d
Esr
ENd
Esr
7.08 8.13 4.14 7.08
71.26 1 87 . 8 6 69.43 -50.81
6.82 7.99 4.06 6.9 6
72.38 1 87 . 87 70. 65 -45.55
6.5 6 7.86 3.98 6.74
73.49 1 87 . 89 7 1 . 88 -40.3
BO, large olistolith of subalkaline MORB-like metabasalts exposed at Botnaneset.
character, more akin to om or CFB. The lack of positive Nb anomalies and in som e cases slight relative depletion ofNb (Fig. SA), and the relatively low values of Ti02, P205, Zr and Y jointly favour a CFB-type rather than an OIB (Thompson et al., 1984; Wilson, 1989 ) . Low values of Ti, Nb, Zr and Y are generally attributed to crustal contamination or reflect subduction-related magma genesis (Perfit et al., 1980; Saunders et al., 1980; Cox & Hawkes warth, 1984; Thompson etal. , 1984). Moreover, the subalkaline MORB-like metabasalts also display geo chemical characteristics similar to subduction-related potassic and high-K calc-alkaline basalts (Foley et al. , 1987; Wheller et al. , 1987; Stolz et al. , 1990). The isotope data from a basaltic pahoehoe (ENct = 8) and an andesitic aa (ENd = 4) flow-unit are indicative of MORB/back-arc basin basalts and oceanic island arc/some continental rift-related (northern Basin
and Range) basalts respectively (Faure, 1986; Wilson, 1989) . The lower ENd value of the latter may, how ever, indicate that the more felsic volcanics have experienced higher degrees of crustal contamination compared with the basaltic types (Faure, 1986 ) . Transitional basalts chiefly consist of scoriaceous and subordinate massive, aphyric or porphyritic aa metabasalts and basaltic meta-andesites. Rare phenocrysts are saussuritisized plagioclase and FeTi oxides. Calcite- and minor quartz-filled amygdules are common, and locally small sandstone xenoliths are present. Similar to the alkaline and subalkaline MORB-like basalts, the transitional types also display some geochemical signatures that are characteristic of alkaline OIB and/or MORB (Fig. S). However, the pronounced negative Nb anomalies (Fig. SA) strongly argue for a subduction-related setting. Although most geochemical evidence favours an
Fig. 5. (Opposite. ) (A) MORB normalized trace-element diagrams for volcanic rock olistoliths. See Fig. 2 for location and
stratigraphical position of different olistolith types of mafic volcanic rocks. (1) alkaline metabasalts, (2) subalkaline MORE like metabasalts of olistolith located at Kalvegjerdsneset (higher stratigraphical position compared to subalkaline MORB-like metabasalts of olistolith located at Botnaneset), (3) subalkaline MORB-like metabasalts of olistolith located at Botnaneset, (4) transitional type metabasalts, (5) !AT metabasalts, and ( 6) acidic volcanic rocks. The order of the elements and the values of the normalizing constants are from Pearce ( 1983) , except for Cr, which is from Pearce (1980) . (B) Discriminant plots for olistoliths of ( 1 ) volcanic rocks, and (2) and (3) mafic volcanic rocks. See Fig. 2 for location and stratigraphical position of different olistolith types of mafic volcanic rocks. Discriminant plot (8 . 1 ) is from Winchester & Floyd ( 1977) , whereas discriminant plots (B .2) and (8.3) are from Pearce ( 1980).
2S2
R . Ravnas and H. Fumes
original back-arc setting, some element abundances and element ratios (e.g. high AI, low Ni, low FeO/ MgO) are more characteristic of calc-alkaline basalts and island arc tholeiites. Basalts associated with incipient back-arc rifting or initial stages of back-arc spreading are thus thought to be the best analogues of these m etavolcanics (Bruhn et at., 1978; Saunders etal., 1979; Weaver et at., 1979; Tarney eta/., 1982). Island arc basalts (IAT) comprise scoriaceous, porphyritic aa flows with phenocrysts of saussuritis ized plagioclase and minor FeTi-oxides. The MORB normalized, multi-element variation diagrams (Fig. SA) show typical IAT character with negative Nb anomalies and relative depletion of Ti, Zr, Y and Cr, thus testifying the subduction-related (i.e. island arc or active continental margin) affinity of these metabasalts (see also Fig. SB). Acidic volcanic rocks comprise massive, por phyritic rhyolite and rhyolitic ignimbrite, with an inferred co-ignimbrite ash layer. Phenocrysts or crystaloclasts are quartz (commonly blue), zoned, twinned and 'chessboard' sodic plagioclase, sanidine, myrmekite, quartz/alkali-feldspar spherulites and inferred glass shards. In addition there are pumice and lithic fragments present in the ignimbrite. Geo chemical analyses of the ignimbrite matrix, rhyolitic core of individual ignimbrite flows , ash-layer and massive rhyolite, suggest that these acidic volcanic rocks are co-magmatic (Fig. S; Ravnas, 1991). The geochemical and isotope composition of these silicic rocks favour a subduction-related original tec tonic setting (Tarney et a/., 1977; Saunders et a/., 1979; Pearce et a/., 1984; Faure, 1986; Wilson, 1989). The pronounced depletion of Nb, P and Ti (Fig. SA) indicate derivation from a depleted IAT source (Pearce et a/., 1984; Holm, 198S), whereas positive anomalies of Zr and Y are more typical of silicic rocks from tholeiitic associations (Pearce et a/., 1984, and references therein). The isotope com position, which is close to that of MORB (Table S), excludes magma generation by partial melting of continental crust and suggests little influence of melted sediments in the magma source region (Faure, 1986). The acidic volcanic rock olistoliths are, accordingly, interpreted to represent island arc/ active continental margin volcanics erupted at a considerable distance from the trench, i. e. behind the arc with respect to the trench. Alternatively, these volcanics may have been derived from a slightly contaminated MORB source, representing oceanic back-arc, rift-related acidic volcanics (Gill et a/., 1984).
As noted above, the mineralogical and geo chemical similarities between the granitic and rhyolitic conglomerate debris and the olistoliths of acidic volcanic rocks suggest a common protolith. The higher K, Rb, Zr, Nb, Y, and Th content of the rhyolitic rocks and the higher plagioclase content of the granitic rocks may accordingly reflect fractional crystallization, and particularly plagioclase fraction ation (Taylor et a/., 1968; Saunders et al., 1979). Notably, the different volcanic-rock olistolith types are present at certain stratigraphical levels. Olistoliths of IAT lava-flow basalts and acidic volcanic rocks are found at a lower stratigraphical level, although generally above 'mega-olistoliths' of terrigenous sediments, whereas transitional and subalkaline MORB-like metabasalts of increasingly alkaline affinity are found at successively higher stratigraphical levels (Fig. 2). However, the olistolith consisting of alkaline metabasalt makes an important exception. Whereas volcanic-rock olistoliths gener ally are found in the upper part of the melange succession, the olistolith of alkaline metabasalt occurs in the lower part of the melange succession in an interval containing olistoliths of rhyolite, bedded chert, pebbly mudstones and shallow-marine sediments. Most major and trace elements of the mafic vol canic rock olistoliths seem to display 'gradational' variations from alkaline through subalkaline MORE like to transitional types, and finally to !AT-type basalts, particularly demonstrated by their Fe, Mg, K, Ti, Nb , Zr and Y content. All metabasalt types have high AI and similar Rb, Ba and Sr contents (Fig. SA & Table 3). Moreover, all volcanic rock olistoliths, including the alkaline metabasalts, show evidence of a primary subduction-related setting. These points suggest that the protoliths of the mafic volcanic rock olistoliths may be genetically related and represent parts of a continuous volcanic series. Below it is argued that the volcanic rock olistoliths represent parts of a volcanic succession in which the alkaline and the IAT types represent the recorded end-members. Olistoliths of sedimentary rocks
Geochemical analyses have been performed on samples from all types of sedimentary-rock olisto liths. However, only the lithic-volcanic arenites from the olistoliths of shallow-marine sediments and deep marine sandy turbidites will be discussed in some detail here.
Geochemical data for determining provenance The lithic-volcanic arenites consist of angular to subrounded quartz (blue or colourless), plagioclase (saussuritisized, zoned, twinned or 'chessboard'), microline, sanidine, lithic-volcanic grains, lithic sedimentary grains, minor iron oxides, ilmenite, calcite (locally dolomite), apatite, epidote and accessory sphene and zircon. The lithic-volcanic grains represent massive or foliated metabasalts, m eta-andesites, acidic volcanics and minor myrme kite. The lithic-sedimentary grains represent chert or quartz schist and minor inferred intrabasinal sand stones and mudstones. Interbedded pebbly sand stones and polymictic conglomerates show a similar gravel admixture to the channel-fill conglomerates. Bhatia ( 1 983) and Bhatia & Crook ( 1986) sug gested a simplified plate-tectonic classification of continental m argin and oceanic basins based on the nature of the crust from which the sediments were derived. Four types of plate tectonic setting were recognized: oceanic island arcs, continental island arcs, active continental margins and passive margins. The geochemical signatures of the lithic-volcanic arenites (Table 6) clearly suggest derivation from island arcs, either continental or oceanic (Fig. 6). According to the classification schemes of Bhatia ( 1983) and Bhatia & Crook ( 1986) most m ajor elements favour a continental ist'and arc as prov enance, whereas most trace elements and ratios argue in favour of an oceanic arc. The 'oceanic island arc' setting includes sedimentary basins adjacent to intra-oceanic island arcs or island arcs partly formed on thin continental crust, whereas the 'continental island arc' type includes sedimentary basins adjacent to island arcs formed on well developed continental crust or on thin continental margins (Bhatia, 1983 ) . A comprehensive data set on the geochemistry of detritus from an intra-oceanic arc was provided by Hiscott & Gill ( 1992) . The lithic-volcanic arenites of the sedimentary rock olistoliths of the Kalvag Melange differ from the volcaniclastic deposits discussed by Hiscott & Gill ( 1 992) by their higher content of Nb, and, although less pronounced, also Zr, La and Th, which is here interpreted as the signature of the addition of alkaline and/or continental source rocks (see below). A m agmatic arc (Dickinson, 1985 ; Valloni, 1985) is also suggested as the source for these arenites by the modal analyses, although such analysis is in this case hampered by post-depositional alteration effects, e.g. metamorphic recrystallization and subsequent partial obliteration of primary grain boundaries. Further constraints on the provenance are
253
provided by the framework grains and the gravel population of the interbedded pebbly sandstones and conglomerates. The type of framework grains and the polymodal distribution of the gravel popu lation of most debris types strongly suggest that volcanic and sedimentary rocks now present as vol canic rock olistoliths and extrabasinal conglomerate debris of the channel-fill deposits formed a m ajor source for these lithic-volcanic arenites. The geo chemical composition of the lithic-volcanic arenites in fact can be explained by mixing of detritus derived by erosion of such rocks alone (Fig. 7). Moreover, the deviations in element and element ratio ranges compared with those suggested for basins of island arc settings (Bhatia, 1983; Bhatia & Crook, 1986; see Table 6) should be expected according to this interpretation (Fig. 7). Alternatively the apparent andesitic composition of the lithic-volcanic arenites (Fig 7A), may suggest that andesitic igneous rocks originally formed a larger volumetric proportion of the provenance region. Notably, there appears to be no systematic stratigraphical variations in mineral ogical and/or geochemical composition, suggesting that the different source rock types were all present in the provenance region(s) throughout the life-span of the depositional basin. Groundmass deposits
The m atrix of the olistostromal groundmass deposits is geochemically similar to the m udstones in the olistoliths of shallow-marine deposits. The m atrix is, however, enriched in K, Fe, Mg, V, Cr, and to a lesser degree Zr and Nb, and depleted in Si, Ca and Na compared with the lithic-volcanic arenites. These effects are attributed to a larger volumetric pro portion of original m ud-sized p articles (clays and m icas) in the groundmass deposits (Roser & Korsch, 1986; Argast & Donnelly, 1987). This supports the interpretation that the finer grained, more distal mem bers of the shallow-marine deposits provided most of the m aterial of the melange' s groundmass (Ravnas, 1 99 1 ) . The enrichment of fines in the olisto stromal matrix can be ascribed to remoulding, during which most of the sandier deposits remained as clasts/olistoliths, or, perhaps more likely, to incor poration of shelf and slope mud into the ambient mud slurries or debris-flows from which the olisto stromal 'groundmass' was deposited. The 'matrix' of the channel-fill conglomerates and the gully- and channel-fill turb idites is, on the other hand, mineralogically and geochemically similar to
N Ul .,.
Table 6. Representative geochemical analyses of sedimentary-rock olistoliths Sample no.
location
Si02
Olistolith type
Ti02
Al203 Fe203"
MnO
MgO
CaO
Na20
1<20
P205
LOI
Total
v
Cr
Co
Ni
Cu
Zn
Rb
Sr
y
Zr
Nb
Ce
Nd
Ba
La
Th
109
144
30
34
36
91
87
144
35
242
13
89
45
361
37
10
1 04
154
25
35
10
79
52
166
28
203
11
49
46
458
24
151
15
42
27
78
0
220
41
203
10
24
50
27
36
13
135
27
39
23
90
65
211
29
228
13
56
48
305
29
19
12
91
R8-1 1 0
Annevika
Shallow marine sed.
63,41
0,84
14,92
6,99
0,09
3,81
4,16
2,90
1 ,84
0,11
1 ,73
1 00,81
R8-124
Annevika
Shallow marine sed.
65,97
0,87
1 1 ,93
6,19
0,09
4,62
6,09
3,13
1,13
0,12
0
1 00,84
R8-1 1 5
Annevika
Shallow marine sed.
54,64
0,91
9,60
5,21
0,21
3,80
18,09
2,16
0,04
0,17
6,65
101 ,47
6,58
0,08
4,83
5,17
2,83
1 ,53
0,12
0,82
1 00,96
5,12
10,17
1 ,59
2,36
R8-1 1 9
Annevika
Shallow marine sed.
64,55
0,88
R8-1 1 7
Annevika
Shallow marine sed.
R8-122
Annevika
Shallow marine
R8-127
Annevika
R8-129 R9-10
13,56
)1
94
105
3
56,85
0,80
12,63
5,81
1 00,87
96
137
20
46
103
345
29
209
13
54
42
328
38
11
64,39
0,91
1 3,07
6,41
0,08
4,67
4,97
2,83
1 .88
0,13
0,80
100,14
1 04
143
26
36
22
85
81
202
32
282
14
70
39
373
39
6
65,76
0,96
1 1 ,83
7,54
0,10
4,86
4,57
2,35
1 ,70
0,12
0,73
100,51
134
193
31
37
20
79
84
199
32
211
7
50
27
,244
29
8
Annevika
sed. Shallow marine sed. Shallow marine sed.
67,43
0,89
12,07
6,55
0,09
4,03
3,92
3,44
1 , 34
0,13
0,75
100,64
107
153
26
35
25
78
67
200
29
262
12
84
44
259
44
10
Varenesvika
Shallow marine sed.
64,63
0,88
14,27
6,88
0,07
3,76
2,10
3,24
2,35
0,12
1 ,50
99,82
116
140
30
30
21
88
109
117
31
252
14
83
33
376
39
16
19
93
107
117
36
11
5,28
0,11
0,14
R9-11
Varenesvika
Shallow marine sed.
63,40
0,96
14,65
7,01
0,08
4,01
2,80
3,51
2,23
0,13
1 ,48
1 00,27
114
136
30
30
300
15
106
51
439
45
R8-213
Varenesvika
Shallow marine sed.
64,42
0,86
12,95
6,16
0,09
3,93
6,57
2,62
1 ,84
0,13
0,90
1 00,48
101
112
23
29
23
83
85
138
34
242
13
75
50
369
48
16
R8-21 1
Varenesvika
Shallow marine sed.
63,71
0,94
1 5,26
6,98
0,08
3,88
2,25
3,03
2,42
0,15
1 ,64
1 00,33
117
139
31
34
20
95
118
111
38
296
17
101
50
439
41
15
R9-7
Varenesvika
Shallow marine sed.
68,01
0,86
12,90
6,73
0,08
3,67
2,17
1,60
1 00,23
117
145
29
29
22
81
79
108
31
252
14
86
46
214
35
8
R9-4
Varenesvika
Shallow marine sed.
67,63
0,76
1 3,02
6,54
0,09
3,78
2,71
2,95
1 ,43
0,10
1 .46
1 00,48
112
143
27
33
33
83
70
127
27
184
8
80
43
255
41
9
4,15
4,75
0,50
0,10
0,97
1 00,51
132
185
31
35
18
71
20
179
25
126
7
42
30
123
12
13
99,44
147
40
135
28
2,46
1 ,63
0,11
R8-31
Langeneset
Deep marine turb.
68,45
0,86
1 1 ,13
7,95
0,13
3,85
R8-32
Langeneset
Deep marine turb.
64,95
0,96
1 1 ,93
7,95
0,13
4,15
4,75
2,82
0,12
0,88
201
34
39
22
75
5
38
20
168
15
12
R8-33
langeneset
Deep marine turb.
68,06
0,71
12,76
6,32
0,07
3,54
2,13
2,93
2,11
0,13
0,81
99,57
113
145
28
35
19
76
97
125
27
170
9
40
31
301
12
16
7,04
0,09
3,81
2,40
3,59
2,28
0,14
1 ,03
99,99
121
158
31
39
22
93
1 02
121
38
326
15
76
38
385
35
18
7,27
0,08
3,83
2,44
4,29
2,25
0,14
1 ,21
99,04
126
159
31
37
28
94
99
122
36
303
16
72
45
455
31
4
0,07
3,44
1 00,08
112
147
26
29
21
74
92
129
24
160
8
38
34
299
20
11
111
R8-34
Langeneset
Deep marine turb.
64,36
0,99
14,27
R9-61
langeneset
Deep marine turb.
62,04
1,00
14,49
Deep marine turb.
69,27
0,82
137
1 ,96
0,10
1 ,06
R9-63
langeneset
Deep marine turb.
70,57
0,80
1 1 ,79
6,18
0,08
3,33
2,66
2,69
1 , 54
0,11
0,92
1 00,68
150
26
33
18
74
80
127
24
215
9
48
283
28
7
R9-64
Langeneset
Deep marine turb.
69,69
0,86
1 1 ,94
6,04
0,09
3,10
2,18
5,10
1 ,46
0,12
1,31
101 ,90
107
156
25
31
28
77
62
110
30
298
12
72
51
274
37
7
R9-65
Langeneset
Deep marine turb.
67,37
0,86
13,31
6,57
0,08
3,52
2,00
3,38
1 ,99
0,12
1 ,19
1 00,39
114
161
29
33
20
83
89
117
29
256
14
83
39
322
30
R9-66
Langeneset
Deep marine turb.
70,07
0,85
1 1 ,27
6,29
0,10
3,69
3,46
2,91
0,99
0,11
0,86
1 00,60
118
169
27
34
21
71
50
157
28
1 92
11
72
47
289
31
R9-62
LOI,
Langeneset
loss on ignition;
F�O�',
0,69
total iron as
12,50
5,98
F�03 + 1 . 1 x FeO.
2,18
2,82
70
:;.;;, :;.;;, .,
"<: ;:s !':lo "'
.,
;:s "'-
� � ..., ;:s
�
(b)
(a)
1 ,4 1,2
� N'
0
i=
,'
0,8
.... ... _
0,6
I
0
/,- ,_'- -'--- - - - - "1
0,4 "
"
0,2
60
_
,_ ....
,
/D "/
/
"
,
0
,
"
:_ _
c
... ... ...
""
,
/
"
_
..
-
--
.. ... ..
....
Do • •- - - - - - .oD
'\ I
-'
- -
.,. ..... - --
0,3
A;n��. c::.". ' J ,' . . .... D� a -
,/ "
-- -'- - -: B ,.,..
0
,"
-
�
•'
/
N
Q
� 0 N
A
0,2
/
,' C
) j, .. ..;
0,1
/"
/ /� "" D ./
, , ..._ _ ..
.... ... -
"' .,..
:- - '
...0,
r"
•
"'ib",,- ' ,' B ',, ....( - -...� ..... �·
,'-'\
<
I
/,.
/
/" -.... ....
./
6 -.-
. .
/
I
�·t
�oo,_-·.., � c;, o o oaD
,"
� 3 \:!.
g2 ..,
I I I I I I I
D
s+
/
�
, ,'
(c)
I I I I I
I
I
�7 :
/ c
('-
<
' ..... _ _ /,"
I I I I I I r-- -
-
....
•• ',
�-:.. ... • / B : - -� --·------,
-- --
\
',,
/
--� �
!(
0
o: _,.,.
-''),'-•'/ \',�
,, ,�
9::-. oa 0" A ' o• ....... -•- - o- - - " .
C'l "' 0
g..
"'
4
2
10 8 6 Fe203 MgO (%) +
12
14
16
0
+---�--�---+--��-�
0
2
8 10 6 Fe203 MgO (%)
4
+
(e)
(d)
14
16
0 +---�--�---+--� 0 2 10 12 14 16 4 8 6 Fe203 MgO +
!} s
C)> ...,
Th
»' � � v
50
12
;:! r;· �
�
� s· �·
40
"<:::
(3
'<:
E
15: 30
.. .....
, ....... ..,...
/./
�
0 !"' I 0 5
I
10
I
15
I
20
I
25
I
30
( Co
.....
_
__
__
A
__ _
_
1-
-
/ ·r... _, ... .. ""'!. ', ,' __ D '\ 0. • "'� 0 \ • B \'-'... o · , .o. . ... ' \ / ',_ __
20 10
c /:
...
•
_ __ _
:c• � D-��- _ _ ... ...
..... 0
•
Shallow marine sediments
o Deep marine turbidites
"' ;:: ., ;::
�
"'
,'
Zr/ 1 0
Th ppm
Fig. 6. Discriminant plots of major and trace elements from lithic-volcanic arenites and wackes from sedimentary rock, olistoliths. The outlined fields are: A, oceanic island arcs, B, continental island arcs; C, active continental margins, D , passive margins (Bhatia, 1983; Bhatia & Crook, 1986). The 'shallow-marine sediments' and 'deep-marine turbidites' are lithic-volcanic arenites and wackes from olistoliths of shallow-marine deposits and deep-marine turbidites respectively. Discriminant plots (a) to (d) are from Bhatia ( 1983), whereas (e) is from Bhatia & Crook ( 1986).
N U\ U\
R. Ravnas and H. Fumes
256
COMENDITE
A
B
\ I
PA NTELLERIT
10
\ PHONOLITE '
•
.• •
0.1
.
RHYOLITE
'
' , .._
--
<•
- - - - - .......
....
0
Ti02 (%) 0 -/ - - -r - -
Q}/"
- -%s-�-��slmg>
· �����<0� -v""c·:·�.:�: '
BASALT
0
BONINITES
0.1
I \
•
....··
••
,
•
•
·
\ · /" \". ••
•
•
I
':
.
\
\ \ I
PLATE
\
\
\
WITHIN
\
I
\
LAVAS
\
\
\
1•: :r '\\
lo I
�------�--+---+--; 10
10
0.1
••
ARC LAVAS
0
A LKALI·
0.001
�--.\ .
1
NEPHELINITE
SUB·ALKA LINE BASALT
0.01
\
0
ZrfTi02
A NDESITE I BASALT
'
0
1000
100 ppm Zr
Nb/Y
-
c
•
1 000
sediments
-1
o mafic vole. ol.
D
• acidic vole. ol. o
400
•
350
mafic debris acidic debris
J
300
0 ppm Cr
250
100
o
o
•
IAT o
10
\\
'
I\ 1: /1
00
WPB
1 'I
•
g
150
I\ \ 0 "I I I \
/
0
ppm V 200
\ I I
0
100
\ 1 : I
50
· --� ·'----�'�� --'--� +-----------------� 10
1 00
ppm Y
10
15
20
ppm Ti/1000
Fig. 7. Discriminant plots of lithic-volcanic arenites from sedimentary rock olistoliths and suggested source rocks (i.e. conglomerate debris and volcanic rock olistoliths). The filled squares labelled 'sediments' represent lithic-volcanic arenites of the sedimentary-rock olistoliths, matrLx of the channel-fill conglomerates, gully- and channel-fill turbidites and the olistostromal groundmass. Note how the lithic-volcanic arenites and the groundmass deposits in all discriminant diagrams plot in an intermediate position between the suggested source rocks. Discriminant plot (A) is from Winchester & Floyd ( 1 977), (B) and (C) are from Pearce ( 1980), and (D) is from Shervais ( 1982). See text for further discussion.
the lithic-volcanic arenites contained in olistoliths of shallow-marine sediments and of deep-marine turbidites. The channel- and gully-fill de posits are accordingly interpreted to represent resedimented, originally shallow-marine sands and grave ls.
PROVENANCE AND TECTONIC SETTING OF THE KALVAG MELANGE
On the basis of geochemical d ata, most volcanic and sedimentary-rock olistoliths unequivocally favour a subduction-related setting for their original depositional environment. The great thickness of
Geochemical data for determining provenance
the melange succession (> 2500 m) and its facies assemblage and their distribution, all suggest depo sition in the lower part or at the base of a fault controlled, steep submarine slope that was subjected to persistent mass-failure processes (Ravnas, 199 1 ) . The type, abundance and geochemical signature of olistoliths of volcanic rocks (basaltic, andesitic and acidic) and first-cycle , marine, volcaniclastic rocks indicates that the resedimentation zone was on the submarine slope of a volcanic arc. The stratigraphical distribution of the various types of olistolith lithology with the shallow-marine material in the lower part of the melange, and the deep-marine turbidites, cherts and volcanics in the upper part (Figs 2 & 3 ) , suggest that the melange was formed by progressive 'unroofing' of the volcanic terrane (Fig. S). This interpretation is further supported by the vertical changes in the geochemical affinity of the volcanic rock olistoliths. The general lack of pre-emplacement tectonic deformation of olistoliths argues against an origin of the melange as multiple slides tapping a terrain of tectonically juxtaposed blocks. The occurrence of clasts and olistoliths of sedimentary and volcanic rocks within the same debris-flow units suggests that the volcanic debris represents volcanic rocks that either were interbedded with the sedimentary deposits and/or constituted the substratum upon which the sediments were originally deposited. Progressively older deposits of the volcanic arc margin are inferred to have been uplifted by tectonic faulting (thrusting) and resedimented during sub sequent gravitational mass movement (Fig. SA). The resulting 'olistolith-composition stratigraphy' in the melange succession (Fig. 2) would thus be an inversion of the original stratigraphy of the volcanic/ clastic sedimentary apron of the volcanic arc. Accordingly, the olistoliths at the lowest strati graphical level in the melange succession are thought to represent the resedimented youngest portion of the volcanic-arc edifice, whereas those present at successively higher stratigraphical levels would represent progressively older portion of the arc margin. The apparently systematic geochemical variations of the mafic volcanic-rock olistoliths from alkaline through subalkaline MORB-like to transitional types, and finally to IAT-type basalts (Fig. 5 & Table 3) are interpreted to reflect (parts of) the gradual magmatic evolution of the melange's provenance region. The IAT-type basalts, occurring at a lower stratigraphical level than the transitional and subalkaline MORB-like
257
basalts (Fig. 2) , are interpreted to represent the latest stage(s) of this inferred continuous magmatic evolution. Even though the single olistolith of potassic, alkaline basalt is found at a lower strati graphical level, the alkaline volcanics are thought to represent the oldest record of this assumed volcanic rock series (Ravnas, 199 1 ) . The magmatic evolution suggested b y the volcanic rock olistoliths can best be explained by a gradual development from continental intraplate, rift-related potassic magmatism, possibly associated with con tinental flood-basalts, through incipient stages of rifting to initial back-arc spreading, and finally to island arc calc-alkaline and tholeiitic magmatism. The progressively increasing influence of subduction related processes in the magma source region during the final stages of the recorded magmatic evolution is demonstrated by the more pronounced negative Nb anomalies. Magmatic development is therefore related to an area undergoing extension and rifting of continental crust behind an active continental margin. The rifting is thought to finally have separated a volcanic arc, detached from the mainland by a back-arc basin or marginal sea (see discussion below) . A comparable system would be the present Bransfield Strait, which is a narrow basin separating the South Shetland Islands from the Antarctic Peninsula, and which is thought to have formed by recent back-arc extension behind the South Shetland volcanic arc (Weaver et al. , 1979; Garrett & Storey, 19S7 ) . The formation and evolution of the Middle Jurassic to Lower Cre taceous 'rocas verdes' of southern Chile (Bruhn et al., 197S; Saunders et al., 1979) may constitute another analogue. Shortly after the formation of the back-arc basin, a thick pile of first-cycle, lithic volcanic terrigenous clastic materials and possible intercalated volcanics were deposited in inferred fault-controlled basin(s) (Fig. SA; Ravnas, 199 1 ) . The most favourable setting for the Kalvag Melange is thought to be along the volcanic arc margin of the back-arc basin. The rhyolite and rhyolitic ignimbrite olistoliths are interpreted to have been derived from acidic volcanics associated either with the incipient rifting to initial back-arc spreading stages, or alternatively to reflect the inferred final island arc magmatism. Based on the assumption that the granitic pebbles and boulders represent the plutonic equivalents of these acidic volcanics, the presence of such material as conglomerate debris suggests that the volcanic arc itself was deeply eroded.
258
R. Ravnas and H . Furnes
A
B
Sea level
Retrogressive slumping (7) bottleneck
s l u m p scars
� �
Sandy
�
turbidites
�
Extensive
�
s l u m ping/sliding of shallow· marine deposits
� � 0
•
Intraoceanic
/ /inactive
Active
a rc-related magmatism
Active spreading MORB affinity
Active spreading B o n i n itic affinity Obducted slices of 1. 2 and 3 Subduction related magmatism
Laurentia
Baltica Toquima - Table Head faunas Mixed T·TH/Baltic faunas Hirnantian faunas
Continued thrusting and tectonic uplift causes further steepening of the slope a n d renewed mass-failure. Slidi ng/slumping of: Large olistoliths of shallow· marine deposits
Deep-marine turbidites Bedded chert Volcanic rocks Deposition of channe l-fill conglomerates and gully·fill turbidites
Fig. 8. (A) Suggested depositional model for the Kalvag Melange and geological evolution of its provenance and tectonic
setting. 1, Deposition of the back-arc basin, volcanic arc margin along which the Kalvag Melange originated, 2, Initial destabilization and gravitational resedimentation of the volcanic arc apron. 3-4, Deposition of the lower (3) and upper (4) part of the melange. See text for further discussion. (B) Evolutionary model for the Lower Ordovician ophiolite complexes and island arc sequences, and the Upper Ordovician to Lower Silurian rift-related sequences. 1 - 2 , Initiation of subduction and development of immature island arc systems. 3, Build-up of mature island arcs that became colonized by shallow marine, Toquima- Table Head (Laurentian) faunas. 4, Accretion of the island arc system to the Laurentian continental margin. 5-6, Rifting of the active continental margin and formation of a late Ordovician to early Silurian marginal basin inhabited by Hirnantian and Holorhynchus (linked Laurentian and Baltic) faunas. From Pedersen et al. (1992).
Despite some slight variations in geochemical composition , the IAT- and MORB-type metabasaltic debris, in addition to the calcareous material, may have been derived from intercalated mafic volcanics of the proposed volcanic arc apron, representing eroded IAT- and transitional-type basalts, respect ively (compare multi-element variation diagrams of
Figs 4A & 5B). However, metabasalts of boninitic affinity have not been observed as olistoliths in the melange. Boninitic rocks within the Scandinavian Caledonides have so far been reported only from the Lower Ordovician ophiolite complexes/island arc sequences (e.g. see Pedersen et at., 1988; Pedersen & Fumes, 199 1 ) , which thus provide the only likely
Geochemical data for determining provenance
source for this debris type among the terranes exposed along the orogen today. Moreover, ongoing research has revealed the presence of large accumu lations of gabbroic rocks of boninitic affinity associ ated with one of these complexes (Fumes & Pedersen, 1994) . Characteristically, the Lower Ordovician ophiolite complexes/island arc sequences show a long-lived magmatic development (c. 500-470 Ma) , ranging from typical MORB to IAT, boninites, calc-alkaline and alkaline metabasalts and differentiates (Pedersen et al., 1988; Pedersen & Fumes, 199 1 ) . It is note worthy that the relatively flat MORB normalized pattern displayed by the MORB metabasaltic conglom erate debris (Fig. 6a) is typical for the metabasalts from the axis sequence of the Lower Ordovician ophiolite complexes, thereby suggesting an alter native source for this debris type . Based on mineral ogical and geochemical data the trachyandesite/ dacite debris are also assumed to have been derived from the Lower Ordovician island arc sequences. The sedimentary cover to these ophiolite complexes, consisting of interbedded metamorphosed volcanics, phyllites and chert, may thus represent a probable source for the quartzite and possibly also the sub arkosic conglomerate debris (Ravnas, 199 1 ) . The back-arc basin within which the volcanic arc apron originally formed received sediments from two different source rock associations or provenance regions: the newly formed volcanic arc and an ?older ophiolite complex/island arc sequence. As no systematic vertical variations in mineralogical and/ or geochemical composition have been observed , either in the sedimentary-rock olistoliths or in the olistostromal 'groundmass', sediments were ap parently shed continuously from both source rock associations into the original receiving basin(s). This suggests that sediments were derived either (i) con tinuously from two areally distinct provenances; (ii) from a volcanic arc formed by arc splitting of the Lower Ordovician ophiolite complexes/island arc sequences; or (iii) from a deeply eroded volcanic arc separated from and formed by extension of an active continental margin onto which the Lower Ordovician ophiolite complexes/island arc sequences had already been emplaced. Based on the isotope character of the volcanic-rock olistoliths combined with the present knowledge of the development of marginal basins within the outboard terranes of the Scandinavian Caledonides, the latter is preferred (see discussion below) . The accreted or obducted ophiolite complexes/island arc sequences are inter-
259
preted to have formed part of the crust upon which the volcanic arc was founded, thus suggesting that the volcanic arc was built on a sliver of continental crust.
REGIONAL CORRELATIONS AND INTERPRETATIONS
The Scandinavian Caledonides, being dominated by thrust-emplaced tectonic units, is made up of accreted suspect and exotic terranes (sensu Coney et al., 1980), representing respectively the outer part of the continent margin (suspect terranes) and 'oceanic' or island arc/ophiolite complexes, meta morphic complexes and continental lithosphere con sidered to have been derived from outboard of the early Palaeozoic continent Baltica (exotic terranes) (Stephens & Gee, 1985 , 1989; Roberts , 1988). The development of marginal basins and associated magmatism within the outboard terranes of the Scandinavian Caledonides are summarized in a number of papers (e.g. Pedersen & Fumes, 1991 ; Pedersen et al., 1992). Two generations of ophiolite complexes and spreading related mafic magmatism are now established within the Caledonian Appalachian orogenic belt; these are of early Ordovician and late Ordovician to early Silurian age, respectively (e.g. see Pedersen & Fumes, 199 1 ; Pedersen e t al., 1992, and references therein) . The Lower Ordovician ophiolite complexes and island arc sequences of the Scandinavian Caledonides are thought to reflect the build-up of an island-arc- marginal-basin system within the Early Palaeozoic Iapetus Ocean (Pedersen & Fumes, 199 1 ; Pedersen e t al., 1992, Pedersen & Dunning, 1995). Pedersen & Dunning ( 1995) suggest a continuous development from an immature island arc system comparable to the present Palau- Kyushu ridge of the western Pacific into a mature island arc similar to the present Sunda- Banda arc. The Upper Ordovician-Lower Silurian ophiolite and associated igneous complexes are related to rifting of an active continental margin and the development of small, fault-controlled marginal basins (Tucker et al., 1990; Pedersen & Fumes, 199 1 ; Pedersen et al., 199 1 , 1992) . Of these Upper Ordovician t o Lower Silurian marginal basins, only that represented by the Solund- Stavfjord Ophiolite Complex developed into an oceanic basin (Fumes et al., 1 990) . The relationship between the Lower Ordovician ophiolite complexes/island arc sequences and the
260
R . Ravnas and H. Fumes
Upper Ordovician to Lower Silurian magmatic complexes is poorly known. However, a line of evidence suggests that the Upper Ordovician to Lower Silurian marginal basins formed by extension along a continental margin onto which Lower Ordovician ophiolite complexes/island arc sequences had already been accreted or emplaced (Fig. 8B) (Pedersen & Fumes, 199 1 ; Pedersen et a!., 199 1 , 1992; and references therein). Based on faunal and isotopic data, the Lower Ordovician ophiolite complexes/island arc sequences are interpreted to have been accreted onto the Laurentian margin of the Iapetus Ocean in the early to mid-Ordovician (Bruton & Bockelie , 1980; Pedersen et a!., 1988, 199 1 , 1992; Pedersen & Fumes, 199 1 ) . A major problem in the reconstruction of the origin and evolution of the Kalvag Melange is the lack of age constraints and rather poorly defined tectono-stratigraphical relationship to other early Palaeozoic marginal basin complexes/sequences of the Scandinavian Caledonides. The geochemical characteristics of volcanic-rock olistoliths from the Kalvag Melange compare well with various types of volcanic rocks from the Lower Ordovician ophiolite complexes/island arc sequences reported by Fumes et a! . ( 1986) and Sivertsen ( 1992) . These volcanics are now regarded as representing the late, evolved stages of the build-up of a mature island arc (Sivertsen, 1992; Pedersen & Dunning, 1995 ) . However, the combined geochemical and isotope signature of the volcanic-rock olistoliths, and par ticularly that from the rhyolitic ignimbrite olistolith , do not compare favourably with any known magmatic rocks from the Lower Ordovician com plexes. Moreover, based on geochemical data and the relatively undeformed nature of the alkaline and subalkaline MORB-like volcanic-rock olistoliths, these olistolith types are not thought to have been derived from a hypothetical oceanic island or seamount that shed material into the trench along the subduction zone (e.g. see Ogawa, 1985; Pautot et a!., 1987) above which the Lower Ordovician ophiolite com plexes/island arc sequences formed. The suggested setting and evolution of the back arc, depositional basin of the Kalvag Melange resembles closely that proposed for the late Ordo vician to early Silurian marginal basins, e.g. the Solund-Stavfjord Ophiolite Complex and the Sulitjelma Gabbro (Fumes et al., 1990; Pedersen & Fumes, 199 1 ; Pedersen et a!., 199 1 , 1992 ) . A cor relation of the proposed volcanic arc apron along which the Kalvag Melange originated with the Upper
Ordovician to Lower Silurian rift-related igneous complexes and their cover, is further strengthened by the close spatial relationships with the Solund Stavfjord Ophiolite Complex and the suggested tectono-stratigraphical position of the Kalv�lg Melange (Fig. 1) (Fumes et al., 1990; Hartz et a!., 1 994) . On the basis of isotope data obtained from the sedimentary cover (single zircon dates) , the marginal basin represented by the Solund-Stavfjord Ophiolite Complex also received sediments from a composite terrane, which included the Lower Ordo vician ophiolite complexes/island arc sequences (Pedersen & Dunning, 1993 ) . Moreover, the inferred rift-related alkaline and subalkaline MORB-like vol canic rocks present as olistoliths in the melange are geochemically similar to metabasaltic dykes in the Hs;;yvik Group of the middle tectonic unit on Atls;;y further south (Ravnas, 1991 ) . Although the origin of these dykes is still uncertain (Fumes eta!., 1990) , they may possibly be related to the same rifting event inferred from the volcanic-rock olistoliths of the Kalvag Melange above . The lower part of the melange is interpreted to reflect the initial destabilization of the volcanic arc slope and gravitational failure of the sedimentary wedge, whereas the upper part reflects repeated mass-failure processes that finally lead to resedimen tation of the deeper levels of the original arc apron (Fig. SA) . The spatial distribution of facies assem blages and olistolith types , together with the large thickness of the melange succession , thus suggest major fault-related uplift of the volcanic arc margin. Tectonic uplift and steepening of the basin margin may be attributed to at least two mechanisms, which may not be mutually exclusive. The Andaman Sea has been advocated as a modern analogue for the formation and evolution of the Solund- Stavfjord Ophiolite Complex (Fumes et a!., 1990). Tectonic faulting may, accordingly, be related to transpression due to large-scale strike-slip faulting along the island arc, similar to the transcurrent movements along the present Andaman-Nicobar Ridge (Curray et a!., 1979; Hla Maung, 1987) . Alternatively, the tectonic uplift may be attributed to thrusting during the closing stages of the back-arc basin history. In the latter case the Kalvag Melange may be coeval with the Sunnfjord Melange, which is interpreted to have been formed during obduction of the Solund · Stavfjord Ophiolite Complex, i . e . during the closing stage of this Late Ordovician to Early Silurian marginal basin (Berg, 1988; Andersen et a!., 1990; Alsaker & Fumes, 1994, Osmundsen & Andersen,
Geochemical data for determining provenance
1994). In either case, continued uplift led to oversteepening of the basin margin , gravitational instability and deposition of the Kalvag Melange as a series of successive gravitational resedimentation events which tapped progressively deeper levels of the volcanic arc apron. Final emplacement of the melange onto the continental margin probably occurred during the closure of the Iapetus Ocean as a result of the continent- continent collision of the Laurentian and Baltic shields in the late Silurian (Bryhni & Sturt, 1985; Ziegler, 1985) .
AC KNOWLEDGEME N T S
Financial support for this study has been provided through grants from the Norwegian Research Council for Science and Humanities (NAVF) , project code D .4 1 .3 1 . 147 , and the University of Bergen. We are grateful to Drs W. Nemec and R . B . Pedersen for their encouragement, helpful criticism and discussions during the course of this study. We thank Drs J. Winchester, R.M. Hiscott and A . G . Plint for thorough reviews o f a n earlier draft of this manuscript, and J. Ellingsen for help in preparing the illustrations. This work represents publication no. 155 in the International Lithosphere Project (ILP) .
REFERENCES
E . & FURNES, H. (1994) The geochemistry of the Sunnfjord Melange: sediment mixing from different sources during obduction of the Solund-Stavfjord Ophiolite Complex, Norwegian Caledonides. Ceo/. Mag., 131, 105 - 12 1 . ANDERSE N , T . B . & JAMTVEIT, B . (1990) Uplift o f deep crust during orogenic extensional collapse: a model based on field studies in Sogn-Sunnfjord region of Western Norway. Tectonics, 9, 1097 - 1 1 1 1 . ANDERSEN , T . B . , SKJERLIE, K . P . & FURNES, H. (1990) The Sunnfjord Melange, evidence of Silurian ophiolite accretion in the West Norwegian Caledonides. J. geol. Soc. London, 147, 59-68. ARGAST, S . & DONNELLY, T.W. (1987) The chemical discrimination of clastic sedimentary components. J. sediment. Petrol., 57 , 813-823. BASALTIC VOLCANISM STUDY PROJECT ( 1981) Basaltic Volcanism on the Terrestrial Planets. Pergamon Press, New York, 1286 pp. BASU, A. ( 1985) Influence of climate and relief on compositions of sands released at source areas. In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 1 - 18. NATO ASI Series, Series C : Mathematical and Physical Sciences, Vol. 148. Reidel, Dordrecht.
ALSAKER,
261
R.L. & JACKSON , J.A. ( 1987) Glossary of Geology. American Geological Institute, Alexandria, Virginia, 788 pp. BERG, T. ( 1988) Sedimentology of the Her/and Group (Lower Silurian) on Atl¢y, Sunnfjord, western Norway. Unpublished Cand. Scient. thesis, University of Bergen, 217 pp. BHATIA, M.R. (1983) Plate tectonics and geochemical com position of sandstones. J. Geol., 91, 6 1 1 -627. BHATIA, M.R. & CROOK, K.A.W. ( 1986) Trace element characteristics of graywackes and tectonic setting dis crimination of sedimentary basins. Contrib. Mineral. Petrol. 92, 181- 193. BHATIA, M.R. & TAYLOR, S.R. ( 1981) Trace-element geochemistry and sedimentary provinces: a study from the Tasman Geosyncline, Australia. Chern Geol., 33, 1 15 - 125. B LAT!", H . , MIDDLETON, G.V. & MuRRAY, R. (1980) Origin of Sedimentary Rocks. Prentice-Hall, Englewood Cliffs, NJ , 782 pp. BREKKE, H. & SoLBERG, P.O. (1987) The geology of Atl!'ly, Sunnfjord, western Norway. Bull. Norg. geol. Unders., 410, 677-690. BRIGGS, R . M . , UrriNG , A.J. & GIBSON, I.L. (1990) The origin of alkaline magmas in an intraplate setting near a subduction zone: the Ngatutura basalts, North Island, New Zealand. J. Volcano/. Geotherm. Res., 40, 55-70. BRUHN, R.L., STERN , C.R. & DE Wn, M.J. (1978) Field and geochemical data bearing on the development of a Mesozoic volcano-tectonic rift zone and back-arc basin in southernmost South America. Earth planet. Sci. Lett., 41, 32-46. BRUTON, D . L. & BocKELIE, J . F . ( 1980) Geology and paleontology of the H!'llonda area, western Norway - a fragment of north America? In: Proceedings of the Inter national Geological Program - Caledonide Orogen Project 27 (Ed. Wones, D . R . ) , pp. 41 -47. Department of Geological Sciences, Virginia Polytechnic Institute and State University, Blacksburg, V.A . , Memoir 2. BRYHNI, I. & LYSE, K. (1985) The Kalvag Melange Norwegian Caledonides. In: The Caledonide Orogen Scandinavia and Related Areas (Eds Gee, D . G . & Sturt, B . A . ) , pp. 417-427. Wiley, Chichester. BRYHNI, I. & STURT, B.A. ( 1985) Caledonides of southwestern Norway. In: The Caledonide Orogen Scandinavia and Related Areas (Eds Gee, D . G. & Sturt, B . A . ) , pp. 89- 109. Wiley, Chichester. BRYHNI, I . , BOCKELIE, J . F . & N YSTU E N J . P . ( 1981) Excursions in the Scandinavian Caledonides. Uppsala Caledonide Symposium (UCS), 1 5 1 pp. CAMERON, W.E., McCuLLOCH, M.T. & WALKER, D . A . (1983) Boninite petrogenesis: chemical and Nd-Sr isotopic constraints. Earth planet. Sci. Lett., 65, 75-89. CAN N , J.R. (1970) Rb, Sr, Y , Zr and Nb in some ocean floor basaltic rocks. Earth planet. Sci. Lett., 10, 7 - 1 1 . CAs, R.A.F. & WRJGHT, J.V. (1988) Volcanic Successions, Allen & Unwin, London, 528 pp . CoiS H , R.A. (1977) Ocean floor metamorphism in the Betts Cove Ophiolite, Newfoundland. Contrib. Mineral. Petrol., 60, 277-302. CoNEY, P . J . , JoNES, D . L. & MoNGER, J.W.H. (1980) Cordilleran suspect terranes. Nature, 288, 329-333. Cox, K.G. & HAwKESWORTH , C.J (1984) Relative contriBATES,
,
262
R . Ravnas and H . Fumes
butions of crust and mantle to flood basalt magmatism, Mahabaleshwar area, Deccan Traps. Philos. Trans. R. Soc. London, Ser. A, 310, 627-64 1 . CRAWFORD, A . J . (1989) Boninites and Related Rocks. Unwin Hyman Ltd, London, 465 pp. CROOK, K.A.W. ( 1974) Lithogenesis and geotectonics, the significance of compositional variations in flysch arenites (graywackes) In: Modern and Ancient Geosynclinal Sedimentation (Eds Dott, R.H. & Shaver, R.H.) Spec. Pub!. Soc. econ. Paleontol. Mineral. , Tulsa, 19, 304-310. CURRAY, J . R . , MOORE, D . G . , LAWVER, L . A . , EMMEL, F . J . , RAITI, R . W . , HENRY, M . , e t al. ( 1 979) Tectonics o f the Andaman Sea and Burma. In: Geological and Geo physical Investigations of Continental Margins (Eds Watkins, J . , Montadert, L. & Dickinson, P . ) , Mem. Am. Assoc. petrol, Geol . , Tulsa, 29, 189- 198. DICKINSON , W . R. (1985) Interpreting provenance relations from detrital modes of sandstones. In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 333-361 . NATO ASI Series, Series C: Mathematical and Physical Sciences Vol. 148. Reidel, Dordrecht. DICKJNSON , W.R. & SuczEK, C.A. (1979) Plate tectonics and sandstone compositions. Bull. Am. Assoc. petrol. Geol. , 63, 2-3 1 . DICKJNSON , W.R. & VALLONI , R . ( 1980) Plate settings and provenance of sands in modern ocean basins. Geology, 8, 82-86. DoRSEY, R.J. (1988) Provenance evolution and unroofing history of a modern arc-continent collision: evidence from petrography of Plio-Pleistocene sandstones, Eastern Taiwan . J. sediment. Petrol. , 58, 208-218. D u N N I N G , G.R. & PEDERSEN , R.B. ( 1988) U/Pb ages of ophiolites and arc-related plutons of the Norwegian Caledonides: implications for the development of Iapetus. Contrib. Mineral. Petrol., 98, 13-23. EDGAR, A.D. (1980) Role of subduction in the genesis of leucite-bearing rocks: discussion. Contrib. Mineral. Petrol. , 73, 429-43 1 . EWART, A. (1982) The mineralogy and petrology of Tertiary- Recent orogenic volcanic rocks with special reference to the andesite-basaltic compositional range. In: Andesites: Orogenic Andesites and Related Rocks (Ed. Thorpe, R.S.), pp. 26-87. Wiley, Chichester, 697 pp. FAURE, G. ( 1986) Principles of Isotope Geology. Wiley, New York, 589 pp. FoLEY, S . F . , VENTURLLI, G . , GREEN , D . H . & TosCANI , L. (1987) The ultrapotassic rocks: characteristics, classifi cation, and constraints for petrogenic models. Earth Sci. Rev. , 24, 8 1 - 134. FURNES, H. & PEDERSEN , R.B. (1994) The Lyngen Magmatic Complex: geology and geochemistry. Abstracts 21 :a Nordiska Geologiska Vintermdtet, Lulea 1994, p. 58. FURNES, H . , BREKKE, H . , NORDAS, J. & HERTOGE N , J. ( 1986) Lower Palaeozoic convergent plate margin vol canism on Bl!lmlo, southwest Norwegian Caledonides: geochemistry and petrogenesis. Geol. Mag., 123, 123 - 142. FURNES, H . , PEDERSEN , R.B . , SUNDVOLL, B . , TYSSELAN D, M. & TuMYR, 0. (1989) The age, petrography, geo chemistry and tectonic setting of the late Caledonian
Gasl!ly Intrusion, West Norway. Norsk geol. Tidsskr. ,
69, 209-224.
H . , SKJERLIE, K . P . , PEDERSE N , R.B . , ANDERSEN, T.B . , STILLMAN, C.J . , SuTHREN , R.J . , eta/. (1990) The Solund-Stavfjord Ophiolite Complex and associated rocks, west Norwegian Caledonides: geology, geo chemistry and tectonic environment. Geol. Mag. , 127, 209-224. GARRETI, S .W. & STOREY, B.C. ( 1987) Lithospheric exten sion on the Antarctic Peninsula during Cenozoic sub duction. In: Continental Extensional Tectonics (Eds Coward, M . P . , Dewey, J.F. & Hancock, P.L.), Spec. Pub!. Geol. Soc. London , 28, 419-43 1 . Geological Society of London, Bath. GILL, J . B . , STORK, A.L. & WHELA N , P.M. (1984) Vol canism accompanying back-arc development in the southwest Pacific. Tectonophysics, 102, 207-224. GoviNDARAJU , K. (1984) 1984 compilation of working values and sample description for 170 international reference samples of mainly silicate rocks and minerals. Geostand. News/. , 8, 1 -9 1 . HARTZ, E . , ANDRESEN, A. & ANDERSEN, T . B . ( 1994) Structural observations adjacent to a large scale extensional detachment zone in the hinterland of the Norwegian Caledonides. Tectonophysics, 231 , 123- 127. HAuGHTON , P.D.W. & HALLIDAY, A.N. (1991) Signifi cance of a late Caledonian igneous complex revealed by clasts in Lower Old Red Sandstone conglomerates, central Scotland. Geol. Soc. Am. Bull. , 103, 1476- 1492. HELMOLD, K.P. ( 1985) Provenance of feldspathic sandstones - the effect of diagenesis on provenance interpretations. In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 139- 163. NATO ASI Series, Series C : Mathematical and Physical Sciences Vol. 148. Reidel, Dordrecht. HlcKEY, R.L. & FREY, F.A.F. (1982) Geochemical characteristics of boninite series volcanics: implications for their source. Geochim. Cosmochim. Acta , 46, 2099-2115. H1scorr. R.N. ( 1984) Ophiolitic source rocks for Taconic age flysch: trace-element evidence. Geol Soc. Am. Bull., 95, 126 1 - 1267. Hlscorr, R.N. & GILL, J.B. (1992) Major and trace element geochemistry of Oligocene to Quaternary volcaniclastic sands and sandstones from the Izu- Bonin arc. In: Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 126. (Eds Taylor, B . , Fujioka, K . , et a/.), pp. 467-485 . Ocean Drilling Program, College Station. TX. HLA MAUNG (1987) Transcurrent movements in the Burma-Andaman Sea region. Geology, 15, 9 1 1 -912. HoLM, P.E. (1985) The geochemical fingerprints of differ ent tectonomagmatic environments using hygromag; matophile element abundances of tholeiitic basalts and basaltic andesites. Chemical Geology, 5 1 , 303-323. HoSSACK , 1 .R. (1984) The geometry of listric faults in the Devonian basins of Sunnfjord, W. Norway. J. Geol. Soc. London, 141 , 629-637. INGERSOLL, V.R. & SucZEK, C.A. (1979) Petrology and provenance of Neogene sand from Nicobar and Bengal fans, DSDP sites 2 1 1 and 18. J. sediment. Petrol. , 49, 1217- 1228. JoHNSSON, M.J. & STALLARD, R.F. (1989) Physiographic FURNES,
Geochemical data for determining provenance
controls on the composition of sediments derived from volcanic and sedimentary terrains on Barro Colorado Island, Panama. J. sediment. Petrol. , 59, 768-78 1 . KoLDERUP, N.H. ( 1921) Der Mangeritsyenit und umgebende Gesteine zwischen Dalsfjord in Spndfjord im westlischen Norwegen. Bergen Museum A. rbok , 1920-21 (5) 71pp. KoLDERUP, N.H. (1928) Fjellbygningen i kyststrpket mellem Nordfjord og Sognefjord. Bergen Museum A. rbok, Naturvitenskapelig Rcekke, I , 222 pp. LARU E , D . K . & SAMPAYO, M . M . ( 1990) Lithic-volcanic sandstones derived from oceanic crust in the Franciscan Complex of California: 'sedimental memories' of source rock geochemistry. Sedimentology, 37 , 879-889. MARSAGLIA, K . M . , INGERSOLL, R.V. & PACKER, B .M . (1992) Tectonic evolution o f the Japanese islands as reflected in modal compositions of Cenozoic forearc and backarc sand and sandstone. Tectonics, 1 1 , 1028 - 1044. McBRIDE, E .F. (1985) Diagenetic processes that affect provenance determinations in sandstone. In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 95- 1 13 . NATO ASI Series, Series C: Mathematical and Physical Sciences Vol. 148. Reidel, Dordrecht. McLENNAN , S . M . , TAYLOR, S . R . , McCuLLOCH, M.T. & MAYNARD, J . B . ( 1990) Geochemical and Nd- Sr isotopic composition of deep-sea turbidites: crustal evolution and plate tectonic associations. Geochim Cosmochim. Acta , 54, 2015-2050. MoRRIS, J . D . & HART, S . R. ( 1983) Isotopic and incom patible element constraints on the genesis of island arc volcanics from Cold Bay and Amak Island, Aleutians, and implications for mantle structure. Geochim. Cosmochim. Acta , 47, 2015-2033. MuTn, E . (1985) Turbidite systems and their relations to depositional sequences. In: Provenance of Arenites (Ed . Zuffa, G . G . ) , pp. 65 -93. NATO ASI Series, Series C : Mathematical and Physical Sciences Vol. 148. Reidel , Dordrecht. NORMARK, W.R. (1985) Local morphologic controls and effects of basin geometry on flow processes in deep marine basins. In: Provenance of Arenites (Ed. Zuffa, G.G.), pp. 47-64. NATO ASI Series, Series C: Mathematical and Physical Sciences Vol. 148. Reidel, Dordrecht. NORRISH, K. & HUTTO N , T.T. (1969) An accurate X-ray spectroscopic method for the analysis of a wide range of geological samples. Geochim. Cosmochim. Acta, 33, 431 -453. NORTON, M.G. (1986) Late Caledonian extension in western Norway: a response to extreme crustal thickening. Tectonics, 5, 195-204. OGAWA, Y. (1985) Variety of subduction and accretion processes in Cretaceous to Recent plate boundaries around southwest and central Japan. Tectonophysics, 112, 493 -518. 0SMUNDSEN , P.T. & ANDERSEN, T.B. ( 1994) Caledonian compressional and late-orogenic extensional defor mation in the Staveneset area, Sunnfjord, Western Norway. J. struct. Geol. , 16, 1385- 1401 . PAUTOT, G . , NAKAMURA, K . , HUCHON, P . , ANGELlER, J . , BouRGOIS , J . , FuJIOKA, K . , et al. (1987) Deep-sea submersible survey in the Suruga, Sagami and Japan trenches: preliminary results of the 1985 Kaiko cruise,
263
Leg 2. Earth planet. Sci. Lett. , 83 , 300-312. J.A. (1980) Geochemical evidence for the genesis and eruptive setting of lavas from Tethyan ophiolites. In: Ophiolites (Ed. Panayitou, A . ) pp. 261 - 272. Geo logical Survey Department, Cyprus, 781 pp. PEARCE, J.A. (1983) The role of sub-continental lithosphere in magma genesis at destructive plate boundaries. In: Continental Basalts and Mantle Xenoliths (Eds Hawkesworth, C.J. & Norry, M.J.), pp. 230-249. Shiva Geology Series, Nantwich. PEARCE, J . A . , HARRIS, N.B.W. & TINDLE, A.G. (1984) Trace element discrimination diagrams for the tectonic interpretation of granitic rocks. J. Petrol. , 25, 956-983. PEDERSE N , R . B . & DUNNING, G.R. (1993) Provenance of turbiditic cover to the Caledonian Solund-Stavfjord ophiolite from U - Pb single zircon dating. J. geol. Soc. London, 150, 673-676. PEDERSE N . R.B. & DUNNING, G.R. (1995) Age relations between ophiolite and island arc complexes: new U/Pb ages from the ophiolitic terrane of SW Norway. Contrib. Mineral. Petrology, (in press) . PEDERSE N , R . B . & FURNES, H. ( 1991) Geology, magmatic affinity and geotectonic environment of some Caledonian ophiolites in Norway. J. Geodyn. , 1 3, 183-203. PEDERSEN , R . B . , FURNES, H . & DUNNING, G.R. (1988) Some Norwegian ophiolite complexes reconsidered. Norges geol. Unders. Spec. Pub/. , 3, 80-85. PEDERSE N , R.B . , FURNES, H. & DUNNING, G.R. (1991) A U/Pb age for the Sulitjelma Gabbro, North Norway: further evidence for the development of a Caledonian marginal basin in Ashgill-Llandovery time. Geol. Mag. , 128, 141- 153. PEDERSE N , R.B . , BRUTON , D.L. & FURNES, H . ( 1992) Ordovician ophiolites, island arcs and faunas in the Scandinavian Caledonides. Terra Nova , 4, 217-222. PERFIT, M . R . , GusT, D . A . , BENCE, A . E . , ARcuLus, R.J. & TAYLOR, S.R. (1980) Chemical characteristics of island arc basalts: implications for mantle sources. Chern. Geol. , 30, 256-277. PETTIJOHN, F.J . , POTTER, P.E. & SIEVER, R. (1987) Sedimentary Rocks. Springer-Verlag, New York, 553 pp. RAvNAS, R. ( 1991) A sedimentological, geochemical and provenance study of the Kalvag Melange, western Norwegian Caledonides. Unpublished Can d . Scient. thesis, University of Bergen, 2 12 pp. RAYMOND, L.A. ( 1984) Classification of melanges. In: Melanges: their Nature, Origin and Significance (Ed . Raymond , L.A.), Geol. Soc. Am. Spec. Paper, Boulders, 198, pp. 7-20. REUSCH, H . ( 1903) Forsteninger i fjeldet paa Frpyen. Naturen, 7 , 160. Rlcc1 LuccHI, F. (1985) Influence of transport processes and basin geometry on sand composition . In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 19-45. NATO ASI Series, Series C: Mathematical and Physical Sciences Vol. 148, Reidel, Dordrecht. RoBERTS, D . (1988) The terrane concept and the Scandinavian Caledonides: a synthesis. Bull. Norges geol. Unders. , 41 3 , 93-99. ROBERTSON, A.H.F. & HEN DERSON , W.G. ( 1984) Geo chemical evidence for the origins of igneous and sedi mentary rocks of the Highland Border, Scotland. Trans. PEARCE,
264
R. Ravnas and H. Fumes
R. Soc. Edinburgh: Earth Sci. , 75, 135- 150. ROSER, B.P. & KORSCH, R.J. ( 1986) Determination of tectonic setting of sandstone- mudstone suites using Si02 content and K20/Na20 ratio. J. Geol. , 94, 635-650. SAUNDERS, A.D. & TARNEY, J. (1979) The geochemistry of basalts from a back-arc spreading center in the East Scotia Sea. Geochim. Cosmochim. Acta, 43, 555-572. SAUNDERS, A . D . , TARNEY, J . , STERN , C.R. & DALZIEL, I.W . D . (1979) Geochemistry of Mesozoic marginal basin floor igneous rocks from southern Chile. Geol. Soc. Am. Bull. , 90, 237-258. SAUNDERS, A.D . , TARNEY, J. & WEAVER, S . D . ( 1980) Transverse geochemical variations across the Antarctic Peninsula: implications for the genesis of calc-alkaline magmas. Earth planet. Sci. Lett. , 46, 344-360. ScHWAB, F.L. (1975) Framework mineralogy and chemical composition of continental-margin type sandstone. Geology, 3, 487-490. SERANNE, M. & SEGURET, M. ( 1987) The Devonian basins of western Norway: tectonics and kinematics of an extending crust. In: Continental Extensional Tectonics (Eds Coward, M . P . , Dewey J.F. & Hancock, P.L.) Spec. Publ. Geol. Soc. London, 28, pp. 537-548. Geo logical Society of London, Bath. SHERVA!S , J.W. (1982) Ti-V plots and the petrogenesis of modern and ophiolitic lava. Earth Planet. Sci. Lett. , 59, 101 - 1 18. SIVERTSEN , J.-E. ( 1992) Stratigraphy and geochemistry of the extrusives of the Torvastad Group: marginal basin deposits associated with the Karm¢y Ophiolite Complex, SW Norway. Unpublished Cand. Scient. thesis, Univer sity of Bergen, 1 1 8 pp. SKAR, 0 . , Fu R N ES, H. & CLAESSON, S . (1994) Middle Proterozoic orogenic magmatism within the Western Gneiss Region, Sunnfjord, Norway. Norsk. geol. Tidjskrift. , 74, (in press) . SMITH, G.A. , CAMPBELL, N . P . , DEACON , M.W. & SHAF!QULLAH, M. (1988) Eruptive style and location of volcanic centers in the Miocene Washington Cascade Range: Reconstructions from the sedimentary record . Geology, 16, 337-340. STAUD!GEL, H. & HART, S.R. (1983) Alteration of basaltic glass: mechanism and significance for the oceanic crust seawater budget. Geochim. Cosmochim. Acta, 47, 37-50. STEPHENS, M . B . & GEE, D . G . (1985) A tectonic model for the evolution of the eugeoclinal terranes in the central Scandinavian Caledonides. In: The Caledonide Orogen Scandinavia and Related Areas (Eds Gee, D . G . & Sturt, B . A . ) , pp. 953 -978. Wiley, Chichester. STEPHENS, M . B . & GEE, D . G . (1989) Terranes and polyphase accretionary history in the Scandinavian Caledonides. In: Terranes in the Circum-Atlantic Paleozoic Orogens (Ed. Dallmeyer, R.D.). Geol. Soc. Am . Spec. Pap . , Boulder, 230, 17-30. STOLZ, A . J . , VARN E, R . , DAVIES, G . R . , WHELLER, G.E. & FoDEN , J . D . (1990) Magma source components in an arc-continent collision zone: the Flores-Lembata sector, Sunda arc, Indonesia. Contrib. Mineral. Petrol. , 105, 585-60 1 . TARNEY, J . , SAUNDERS, A.D. & WEAVER, S . D . (1977) Geochemistry of volcanic rocks from the island arcs and marginal basins of Scotia Sea region . In: Island Arcs,
Deep. Sea Trenches and Back-arc Basins (Eds Talwani, M. & Pitman, W.C., III), pp. 367-378. Maurice Ewing Series, Vol. 1 , American Geophysical Union, Washington, DC. TARNEY, J . , WEAVER, S . D . , SAUNDERS, A . D . , PANKHURST, R.J. & BARKER, P.F. ( 1982) Volcanic evolution of the northern Antarctic Peninsula and Scotia arc. In: Andesites: Orogenic Andesites and Related Rocks (Ed. Thorpe, R.S.), pp. 371-402. Wiley, Chichester. TAYLOR, S . R . , EwART, A. & CAPP, A.C. (1968) Leuco granites and rhyoHtes: trace element evidence for frac tional crystallization and partial melting. Lithos, 1 , 179 - 186. THOMPSON, R.N. (1977) Primary basalts and magma genesis. III. Alban Hills, Roman Comagmatic Province, Central Italy. Contrib. Mineral. Petrol. , 60, 9 1 - 108. THOMPSON, R.N. & FOWLER, M . B . (1986) Subduction related shoshonitic and ultrapotassic magmatism : a study of Siluro-Ordovician syenites from the Scottish Caledonides. Contrib. Mineral. Petrol., 94, 507-522. THOMPSON , R . N . , MORRISON , M . A . , HENDRY, G.L. & PARRY, S.J. ( 1984) An assessment of the relative roles of crust and mantle in magma genesis: an elemental approach. Philos. Trans. R. Soc. London, Series A, 297 , 409-455. THORPE, R . S . , FRANCIS, P.W. & O'CALLAGHAN, L. (1984) Relative role of source composition, fractional crystal lization and crustal contamination in the petrogenesis of Andean volcanic rocks. Philos. Trans. R. Soc., London, Ser. A, 310, 675-692. TucKER, R . D . , BoYD, R. & BARNES, S.-J. ( 1990) A U - Pb age for the Rana intrusion, N. Norway: new evidence of basic magmatism in the Scandinavian Caledonides in Early Silurian time. Norsk geol. Tidsskr. , 70, 229-239. VALLONl, R. (1985) Reading provenance from modern marine sands. In: Provenance of Arenites (Ed. Zuffa, G . G . ) , pp. 309-332. NATO ASI Series, Series C : Mathematical and Physical Sciences Vol. 148. Reidel , Dordrecht. VALLON!, R. & MEZZARDl, G . ( 1984) Compositional suites of terrigenous deep-sea sands of the present continental margins. Sedimentology, 31, 353-364. WEAVER, S . D . , SAUNDERS, A . D . , PANKHURST, R.J. & TARNEY, J. (1979) A geochemical study of magmatism associated with the initial stages of back-arc spreading. Contrib. Mineral. Petrol., 68, 1 5 1 - 169. WHELLER, G . E . , VARN E, R . , FODEN , J . D . & ABBOTT, M.J. (1987) Geochemistry of Quaternary volcanism in the Sunda-Banda arc, Indonesia, and three-component: genesis of island-arc basaltic magmas. In: Tectonic Controls on Magma Chemistry. (Eds Weaver, S . D . & Johnson, R.W.) J. volcano/. Geotherm. Res. , 32, 137160. WILSON, M. (1989) Igneous Petrogenesis: a Global Tectonic Approach . Unwin Hyman, London, 466 pp. WINCHESTER, J.A. & FLOYD , P.A. ( 1977) Geochemical discrimination of different magma series and their dif ferentiation products using immobile elements. Chern. Geol., 20, 325-343. ZIEGLER, P.A. (1985) Late Caledonian framework of western and central Europe . In: The Caledonide Orogen - Scandinavia and Related Areas (Ed. Gee, D . G . & Sturt, B . A . ) , pp. 3 - 19. Wiley, Chichester.
Spec. Pubis int. Ass. Sediment. (1995) 22, 265-281
Differential subsidence and preservation potential of shallow-water Tertiary sequences, northern Gulf Coast Basin, USA MARC B . EDWARDS 5430
D umfr e i s, H o uston, Texas77096-4020, USA
ABSTRACT
Growth faulting, which accompanied shelf-edge progradation and filling of the northern Gulf Coast Basin, resulted in partitioning of the basin margin into fault blocks with contrasting subsidence rates. Study of correlative sections in juxtaposed fault blocks reveals that contrasting subsidence rates can result in strongly differing facies patterns in neighbouring areas. This complicates the task of predicting sandstone reservoir occurrence and properties. Shallow-water clastic sections from Eocene to Miocene in age were investigated using extensive well-log observations, supplemented with micropalaeontology and seismic profiles. All depositional environments involve an ongoing, complex interplay between sedimentation and erosion at different time- and physical scales. In certain settings, a greater subsidence rate causes the preservation of certain facies that would otherwise have been eroded at lower subsidence rates by processes inherent to the environment. The critical subsidence rate that separates preservation from non-preservation is termed the preservation potential threshold for a particular depositional facies. Examples are provided for progradational mouth-bar facies in a deltaic setting (Wilcox), and storm deposited shoreface-shelf muds in a prograding shoreline setting (Frio). Where rates of subsidence are even greater, the growth fault may produce a topographical scarp at the surface, which will influence the disposition of depositional environments. Here, the concept of preservation thresholds is not adequate to account for the observed facies changes. Rather, the presence of the surficial scarp as the surface manifestation of the subsurface fault causes the preferential development and preservation of channel activity in the topographic lows, and progradational environ ments with channel bypass in the topographical highs . An example is provided for a series of prograding stacked deltas (Miocene). These concepts may help to focus attention on the role of subsidence in constraining the appearance of the sedimentary record .
INTRODUCTION
In growth-faulted regions, such as the northern Gulf Coast Basin of Texas and Louisiana, contempor aneous faulting structurally offset stratigraphical surfaces shortly after their formation. Through time, the sediment on the upthrown block (footwall) sub sided at a lower rate than sediment on the down thrown block (hangingwall). This setting provides the opportunity to study the relationship between the preserved stratigraphical record and changing subsidence rates, while other important variables, such as eustatic sea-level , sediment supply and depositional environment, remain comparatively unchanged.
Exploring in such regions, It IS common to drill into fault blocks in which strata of a particular age have not been penetrated previously. Typically, little is known about how the downthrown section will differ from the comparatively well-known equivalent section on the upthrown block. This paper describes several examples of characteristic changes that take place across growth faults , and attempts to explain the observed changes. The most obvious manifestation of the effect of growth faulting on sedimentation is an increase in thickness of a genetic unit from the upthrown to the downthrown blocks. Expansion ratios (downthrown
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
265
M.B. Edwards
266
thickness divided by upthrown thickness; Thorsen, 1963) as great as 10 have been documented. I n addition t o thickness, other properties that vary across the faults include net sandstone, percentage sandstone and log facies development. The way in which these properties change across growth faults will be described in the examples below. The interpretation of three stratigraphical units, the Palaeocene -Eocene Wilcox Group, the Oligo cene Frio Formation, and the Lower Miocene (no formal nomenclature) of the Texas-Louisiana Gulf Coast all illustrate contrasting relationships between growth-faulting and sedimentation patterns. One important relationship concerns the relative rates of subsidence and sediment supply in shallow-water clastic depositional systems, in which sea-level is the major constraint on base level (Wheeler, 1964). If sediment supply is much greater than subsidence, then significant relief is unable to develop where the fault trace emerges at the depositional surface. How ever, where differential subsidence rates are suf ficiently high, significant topographical relief can be created, which can then influence sedimentation patterns. Sedimentologists and stratigraphers commonly invoke mechanisms such as changes in sea-level, subsidence rate and sediment supply to explain the distribution of sedimentary facies and the character istics of vertical profiles. In ancient sediments it is generally impossible to validate these interpretations independently or to isolate their respective effects. However, growth faults that rise to the depositional surface cause contrasting subsidence rates in juxta posed areas, while other conditions are relatively
TEXAS
unchanged. Hence it is possible to identify those changes that most likely relate to subsidence rate. Sequence stratigraphy has heightened interest in the temporal significance of certain types of strati-· graphical surfaces. Peculiarly, emphasis has been given to surfaces that are admittedly not isochronous (e .g. sequence boundaries; see Posamentier & Weimer, 1993) as opposed to surfaces that are (e .g. flooding surfaces; see Galloway, 1989). The require-· ment that sequence boundaries everywhere separate older from younger sediments adds to the demands placed on the stratigrapher to resolve the age of a stratigraphical section. Thus, the lessons learned while studying the effects of growth faulting can be applied to the problems of attempting to distinguish surfaces formed by normal environmental processes, such as channel migration (termed 'source diastems'; Swift et al., 1991), from those formed by imposition of external relative sea-level controls, such as 'incised valleys' (which are components of sequence boundaries).
STUDY AREA AND SCOPE
Examples from diverse geographical locations andl geological ages (Fig. 1) have been chosen to illus trate the principles set forth in this paper: the Palaeocene-Eocene Upper Wilcox Group of South Texas, the Oligocene Frio Formation of South Texas, and the Lower Miocene of southwest Louisiana. The data set consists primarily of well logs at a scale of 1 in. =100ft, supplemented by seismic and micropalaeontological data. With cali-
LOUISIANA
San Patricio County Live Oak County
GULF OF MEXICO
100 Miles 100 Km
Fig. 1. Index map to the location of the four examples presented in this paper : 1, Upper Wilcox in Zapata County; 2, Upper Wilcox in Live Oak County, 3, Frio in Nueces and San Patricio Counties; 4, Lower Miocene in Vermilion Parish. General location of regional sandy shelf-edges are indicated in stippled patterns for Wilcox, Middle Frio and Lower Miocene trends. (From fig. 2, Winker & Edwards, 1983.)
267
D ff i erent a i l subsd i ence bration from whole cores, it is possible to make fairly reliable inferences about lithology from the electric logs (Fig. 2). In this paper, diagrams show only the spontaneous potential (SP) curves, but all logs were examined and interpreted by integrating the SP with the induction logs. The Upper Wilcox of South Texas (areas 1 and 2, Fig. 1) was deposited in a series of shelf-edge delta lobes as part of the Rosita delta system (Edwards, 1980, 1981). A variety of sand geometries reflecting a varied depositional environment were identified. Strike continuous, upward-coarsening sand bodies, appear to represent wave-dominated shorelines, such as strandplains, whereas sections with scattered, blocky and fining upward sandstones suggest depo sition along wave-influenced deltas with pronounced mouth-bar development. The Frio of South Texas includes an interdeltaic embayment dominated by stacked and prograding barrier bar and strandplain sandstones that have pronounced strike continuity (Boyd & Dyer, 1964; Galloway et al., 1982a, b). Sandstones pinch out into lagoonal and continental mudstones and siltstones up-dip, and grade down-dip into storm-deposited interbedded siltstones and mudstones of the shelf. The Lower Miocene of southwest Louisiana was deposited as a series of stacked delta lobes at the initiation of the major sediment influx during the Miocene (Curtis, 1970). Intense contemporaneous structural activity focused sediment into a series of
HC suppression of SP
Sandstone
Shale
SP
Fig. 2.
1
......
:::::::
Hydrocarbons
�
I
�""""
�
Increasing silt content upwards
INDUCTION
Sample electric log showing response of SP and induction resistivity logs to lithology. In this paper, only the SP curves are shown, but in all cases both sets of curves were utilized in the stratigraphical studies upon which this paper is based. (Modified from Edwards, 1984.)
structural basins (Sloane, 1971), probably largely controlled by subsurface salt movement. Detailed mapping indicates that deltas were supplied by large distributary channels and incised valleys. Numerous subsurface studies in this region have shown that the major sand-bearing stratigraphical units on the contemporaneous shelf and upper slope can be divided into regressive packages, or cycles, that can be correlated from tens to hundreds of miles along strike (e .g. Curtis & Picou, 1978). These packages are bounded by transgressive shales that are associated with marine flooding surfaces or maxi mum flooding surfaces. Recent studies tying planktic foraminiferal assemblages to a chronostratigraphical framework suggest that the cycles have durations of about 100-200 ka (Edwards, 1990; Mitchum & van Wagoner, 1991). A recent attempt to relate Lower Miocene oxygen-isotope cycles to well-log cycles suggested a dominant periodicity of 100 ka (Ye et al., 1993). In the examples that follow, most of the correlation markers shown are thought to bound cycles of this order (see Figs 5, 9, 10, 11 & 14).
EFFECTS OF GROWTH FAULTING ON SEDIMENTATION
A considerable literature covers many aspects of the setting and effects of growth faulting on sedimen tation . Curtis & Picou (1978) placed the major growth-faulted trends of the Gulf Coast Basin into an offlapping delta model. Winker & Edwards (1983) examined the delta model in a shelf margin setting and pointed out some of the ways in which this setting differs from a platform setting. We can envision growth faults as part of the extensional head region of a large gravity driven slope failure that also includes a contractional toe region with folds and thrusts in deep water. Study of salt tec tonics in the past few years has focused on the role of salt in terms of both lateral and vertical flowage. Growth faults in salt-dominated areas can form in response to: the evacuation of deep salt into extrud ing diapirs, the evacuation of salt sheets, or combi nations of salt-driven and slope-driven gravity systems (VendeviUe & Jackson, 1992). The most obvious effect of growth faulting on sedimentation is the change in thickness (Thorsen, 1963). Aside from such thickness changes, the character of a stratigraphical unit can appear unchanged, or can show significant changes across a growth fault. In either case, it is important to attempt
268
M.B. Edwards
to predict sandstone properties for the purpose of hydrocarbon exploration and production. The fol lowing sections present examples of various depo sitional responses across growth faults.
2900m
Jonnell gas
1
Ramos
1--
9500'
J onnell gas
--11
Yzaguirre
Expansion only
Stratigraphical sections with large expansion ratios usually display significant stratigraphical changes across a growth fault. A noteworthy exception occurs in the Upper Wilcox in Zapata Country. A fault with an expansion ratio of greater than three, shows no obvious changes in percentage sandstone or log facies (Fig. 3). These sandstones were illustrated in stratigraphical and structural sections (see Edwards, 1981, Figs 10 & 11). Comparison between the upthrown and down thrown blocks is facilitated by changing the vertical scale of one of the logs, in order to make correlation markers subparallel (Fig. 3). This procedure of 'double datuming' attempts to reconstruct the strati graphy as if all of the locations had been subject to the same subsidence rate. In this and the following examples, no attempt has been made to decompact the sections in order to compensate for greater compaction of shale relative to sandstone. Con trasting amounts of compaction could influence present-day thickness where lithology changes sig nificantly across the fault. However, these sections are overpressured, which has resulted in compara tively small amounts of shale compaction. Inspection of the two logs (Fig. 3) indicates that percentage sandstone and log facies are largely unaf fected by the fault, except for the presence of a high frequency signal in the middle part of the down-dip well. In this example, the proportion of net sand in the two wells is the same as the expansion ratio . Facies changes and channel erosion
Extensive mapping in the South Texas Upper Wilcox (Fig. 1) reveals the presence of stacked delta lobes with mappable distributary channels (Edwards, 1980, 1981; Winker & Edwards, 1983). A set of eight wells arranged in a dip section (Fig. 4) has been selected to show thickness and log facies changes (Fig. 5). In an unfaulted area, with a uniformly increasing subsidence rate down-dip (flexure), a gradual change in log facies would be expected, with a change from distributary channels up-dip, through mouth bar to distal mouth bar down-dip. However, in this example
"
.><
£ E ro
"
.><
£ E "'
(") � 0 0
c: 0
·�
Q) !/) a.
:0
a. ::J
"'
�
0
t
0 ro
(") c: 0
'£
Q) !/) a.
:0 c: ::
0 0
Fig. 3. Stratigraphical dip section across a large growth fault in Zapata County, South Texas. The vertical scales of stratigraphically equivalent section both upthrown and downthrown to the fault have been adjusted to compensate for differential subsidence. Note the similarity of the SP logs, despite the expansion ratio of 3.45. Similar 'double datumed' sections are also shown in Figs 5, 1 1 & 13. (Modified from Edwards, 1984.)
(Fig. 5), a striking change in log facies occurs at the up-dip growth fault (between wells 2 and 3), with more subtle changes at the other faults. The effects of growth faults on stratigraphical preservation can be observed by comparing and contrasting adjacent
Dfi ferent a i l subsd i ence Texas
Louisiana
INDEX MAP
I
5ooo· 1500 m
I
Live Oak County
4. Index map showing the location of wells depicted in Fig. 5, Live Oak County, Texas (see Fig. 1 for context).
Fig.
subsidence-normalized wells that are in the same versus different fault blocks. For example, in the Luling regressive cycle (Fig. 5B), wells 1 and 2 are very similar, as are wells 3 and 4. However, wells 2 and 3 are clearly different. The preserved components of each progradational deltaic cycle (Fig. 5) include mouth bar and distal mouth bar deposits, typically arranged in upward coarsening successions; distributary channel deposits represented by blocky sandstones and upward-fining successions; and coastal plain and interdistributary bay deposits, which consist of mixed sandstones and shales in a variety of patterns, usually thinner than the overall regressive succession . Subsidence creates space for the preservation of facies beneath erosion surfaces. In this example, major fluvial erosion is represented by the base of the blocky and upward-fining sandstones. Subsid ence allows the deposition and burial of pro gradational facies beneath the depth of erosion attained by channels. This relationship becomes critical in up-dip areas, where the decreasing thick ness of the space available for preservation of progradational facies approaches the depth of dis tributary channels. Up-dip of this point, the Slick and Luling genetic regressive successions are charac terized by amalgamated sandy channel-fill deposits. Subsidence rate also affects regional facies trends. High subsidence rates favour stacking of delta lobes
269
with 'foreshortening' of dipwise facies gradients (the distance along dip from proximal to progressively distal facies) rather than down-dip translation of facies belts by continued progradation into the basin. A schematic facies preservation diagram (Fig. 6) enables prediction of the depositional sequences that can be juxtaposed across a growth fault in a prograding delta. At the top (Fig. 6A) is a subsidence rate graph . The duration of the progradational phase is suggested to have lasted about 1000 yr, although it could have been much shorter or longer. The growth fault curve (solid line) shows subsidence rates that reflect the thickness changes observed due to growth faulting and roll-over (up-dip thickening on the downthrown block toward the fault). The vertical steps show the locations of individual growth faults. Shown for reference is the flexure curve (dashed line), which assumes a linear increase in subsid ence rate down-dip. Arrows labelled U and D are explained below. Figure 6B illustrates some of the facies relation ships and inferred environments of deposition. After transgression and abandonment of the previous delta lobe, the cycle begins by progradation of the delta front, with superimposed higher frequency trans gressive -regressive cycles. The progressive basin ward shift of environments down-dip has been shown by the down-dip termination of channel erosion surfaces and mouth bars. Coastal-plain facies are depicted as blocky sandstones, although they are in reality complex intercalations of various channel and bay deposits. There is a total of 5 0-200ft (15 -60m) of subsidence at the up-dip end, whereas at the down-dip end there is 25 0-2000ft (75 -600 m) of subsidence . It is assumed that distributary chan nels are 5 0-200ft deep regardless of location, although it is likely that they would decrease in depth down-dip due to channel bifurcation and dis charge through crevasses. The subsidence curves can be used to estimate preservation of vertical sections on either side of a growth fault. At a growth fault, the regionally aver aged subsidence rate does not occur. Instead, much higher subsidence rates occur on the downthrown block, and much lower rates occur on the upthrown block. On the upthrown block (see Fig. 6A), the lower subsidence rates resulted in a vertical section that resembles that developed up-dip. This is shown by the arrow U, which is projected to the left until it intersects the flexure curve, where the appropriate lower subsidence rate would have occurred in a non growth-faulted setting (depicted by flexure curve in
M.B. Edwards
270
Downdip
Updip 3
2
4
5
6
8
7
- -100m - -500'
91m
- -1000'
\F
-
-1500'
A - -500m u L T -2000' ' -;-5 614m
�
19 579m
A Paleostructure B Subsidence normalized 3
2
4
PD
5
7
6
8
PD F L T
Fig. 5.
Thickness and facies changes in two stacked delta complexes, the older Luling (A and B) and the younger Slick (C and D) units in the Upper Wilcox of South Texas (see Figs 1 & 4 for locations) . For both deltas the upper panels (A and C) show a palaeostructural cross-section with the top to the delta complex as the upper datum, and growth faults that were active during delta formation. The lower panels (B and D) are 'double-datumed' sections in which vertical scales were adjusted to normalize for differential subsidence. Generalized deltaic facies are identified using electric log characteristics. CH, channel; MB, mouth bar; PD, prodelta. No horizontal scale.
Fig. 6A and facies distribution in Fig. 6B). At the intersection point, the arrow is then extended down into the facies diagram, in order to determine the vertical section that corresponds to that subsidence rate. A similar procedure can be carried out for the downthrown block by projecting the downthrown subsidence rate to the right (arrow D, Fig. 6A) until it intersects the flexure curve, and then extending it
down into the facies diagram below. This procedure predicts that the vertical sections on either side of the growth fault would show considerable differences in facies profiles that could not be explained solely as a function of changing palaeogeographical lo cation. The thickness of the upthrown and down thrown profiles could then be expanded or contracted to restore their present-day relative thicknesses. However, this method may predict excessive lateral
Different a i l subsd i ence
271 Downdip
Updip 2
3
4
5
6
7
8
- -100' - -200' - -300' - -400' - -500' - -600' C Paleostructure
D Subsidence normalized 2
4
5
6
7
8
PO
Fig. 5.
:><: u ::::;
(f)
(Continued.)
facies changes because the distance between the two schematic wells (below arrows U and D, Fig. 6) will be greater than the actual distance between two wells on either side of a growth fault. Facies changes and storm/wave erosion
The South Texas Frio Formation (Fig. 7) illustrates shorelines that prograde as a line source (i . e . wave dominated deltas, barrier islands and strandplains), rather than point sources (such as river-dominated deltas). It is based on complex stratigraphical and structural relationships determined from approxi mately 600 well-logs (Weise et a!., 1981; Bebout et a!., 1982; Edwards, 1986), coupled with micro palaeontological studies that indicate neritic to coastal environments throughout the section, with overall shallowing upwards (Martin, 1969, 1970). The sandy, wave-dominated Frio shoreline was bordered by muddy coastal plains up-dip, muddy slope deposits down-dip, and by major deltaic depo centres along strike (Galloway et a!., 1982a, b). These palaeogeographical relationships strongly
suggest that sand was supplied to the Frio shoreface by wave-driven alongshore transport in the fore shore/upper shoreface, rather than by prograding deltas (Martin, 1969). A typical well-log in the Frio shows an upward transition from shelf mudstone to strandplain sand stone over a depth range of about 5 000 ft. Subdivision of the section into component cycles reveals that the thickness of the cycles increases markedly with greater depth, from less than 5 0 ft to almost 5 00ft (Fig. 8). This suggests that structurally deeper sedi ments were deposited at much greater subsidence rates than shallower Frio sediments. Rigorous cor relation of hundreds of well-logs in the area allows the tracing of these varied cycles across growth faults. Palaeostructural cross-sections (Figs 9 & 10) show the position of growth faults and their effect on thickness and log facies. The analysis of log facies changes is presented in a set of four well-logs, each in different fault blocks (Fig. 11). The log correlations look questionable at their original depth-scales, but when normalized for differential subsidence, the correlation of the indi-
M.B. Edwards
272 A Subsidence rates (thickness/arbitrary unit of time)
0' 1-��;===�-----�� ��---- ------------------------�_0' ----1�� Upthrown . _1 U - -Growth faults \ = ---I= � ----J / Downthrown . _ F lexure � / .= rr:- I= D 1= 50'= 250'r 200' 15-60m \"-0II0"e 2000' u D 75-600m -
-
-
�
-
---� --
B Facies Updip
Downdip
a
o 0
E
0
� � b � � ....
Fig. 6. Schematic diagram of facies preservation relationships in a deltaic environment characterized by significant channel erosion. (A) Two subsidence rate relationships are shown, a straight dashed line indicating uniformly increasing subsidence rate with distance as in a flexure, and a complex solid line indicating abrupt increases in subsidence rate down-dip across growth faults and decreasing subsidence rate up-dip within some fault blocks due to roll-over. (B) Facies preservation trends and characteristic SP logs. Proximal to distal facies trends have been exaggerated. The figure can be used as a kind of nomograph to estimate facies changes across growth faults in this setting. See text for additional explanation. (Modified from Edwards, 1984.)
vidual cycles is much more obvious. Figure 11 illus trates the change in log facies from thick sand bodies up-dip, to thin interbedded sandstones (siltstones) and mudstones down-dip. The palaeostructural sec tions emphasize that the facies changes take place abruptly across the growth fault. However, the cor relatability of the individual cycles across the fault suggests that significant sea-floor relief was not main tained at the location of the fault. Core study of the down-dip interbedded facies (Berg & Powell, 1976) suggested that the sandstones were deposited by turbidity currents. Studies of the modern Texas shelf, however, suggested that shelf sands are deposited from storm-generated geo strophic flows that produce turbidite-like beds (Morton, 1981). Transport is oblique to the shore line, but the resulting thin beds have enormous strike continuity (Snedden & Nummedal, 1991). The continuity is consistent with the excellent gas
production that can be achieved despite the thin, ratty character of the sands. The large amount of interbedded mudstone was presumably derived from the adjacent deltaic depocentres. A schematic diagram (Fig. 12) suggests how abrupt facies changes can develop across growth faults, without the necessity of an abrupt change in depo sitional environment. The model proposed in Fig. 12 relies upon differential subsidence to produce a con trast in the sediments preserved, even though the depositional processes on both sides of the fault are virtually identical. Unfortunately, the available data do not allow many aspects of the model to be resolved until additional core data are obtained . The dynamics of shoreface -shelf systems have been discussed by many authors (e .g. Swift & Thorne, 1991), however, a considerable number of issues remain contentious. For the present dis cussion, it suffices to emphasize the key points. The
Dff i erent ai l subsd i ence Louisiana Texas
N
t
Nueces
I
10 Miles 10 Km
Kleberg
Fig. 7.
Index map showing location of cross-sections A-A' and B-B' in Figs 9 & 10, and general location of Frio investigation around Corpus Christi Bay.
precise roles of fair-weather versus storm conditions in transporting sediment on the upper shoreface is often unclear and may vary from one shoreline to another. The Frio shoreline prograded steadily through time due to the high sediment supply, and significant wave transport along shore and storm transport to the inner shelf. Logs and whole cores indicate a progression from relatively coarse-grained homogeneous sandstones up-dip, through interbedded thick sandstones and shales, to thinly interbedded 'ratty' siltstones and
273
shales down-dip. It seems likely that the coarse grained shelly sandstones up-dip were deposited in a foreshore and upper shoreface setting and are com posed of wave-reworked and amalgamated storm beds. The ratty interbedded section down-dip was deposited on the lowermost shoreface to shelf and consists of interbedded storm-deposited sands and muds. The abrupt facies change at the fault remains to be explained. The important distinction is that mud layers are preferentially preserved on the more rapidly subsiding down-dip block. The mud could have been deposited from suspension either by turbid nearshore, semi-permanent currents, or by waning flow following a storm event. In either case, it is important that the mud was not eroded subsequently, either by storms or fair-weather wave activity. The Frio shelf system can be regarded in terms of profiles of equilibrium and changing base levels. Areas of greater subsidence rate permitted less sedi ment to be reincorporated by storm currents and bypassed on to the shelf, and hence more sediment was allowed to accumulate. If eustasy and sediment supply are held constant, then the high subsidence rates on the downthrown block effectively raise the local base level (increase accommodation ) , increasing the probability that storm-deposited muds would be preserved. At some critical subsidence rate, sand beds stop being amalgamated, and instead are separated by thin mud layers. At this point, the reservoir characteristics of the sand body change drastically. The most rapidly subsiding areas immediately down-dip of the growth fault would probably have
A
8
7000' -
�
8000'-
�
J 0 Fig. 8.
Example of Frio log from Corpus Christi Bay showing vertical changes in cycle thickness, suggesting increasing subsidence rate with depth. Small portions of the log have been enlarged to illustrate cycle styles and thickness. The numbers next to logs A, B and C refer to the thickness of the log segments illustrated.
a: u..
A
9000' -
3000m 10 000' -
]
8
11 000' -
12 000' -
]
c
�
�
1
E "'
t:.
0
"' N
� j
�
�
c
1
1
E "'
E "' (")
0
0
l
j
0 :::. "' (")
"' ...
M. B. Edwards
274
NW A
32
31
30 GULf 2 ST TR
47 2
-1> (8560')
GULf ST TR
53
-1> (8678')
33
CITIES SER !c SUNRAY 1 ST TR 52 1 ST TR 51
-1> (8670')
-1> (8740')
34 1
ATlANTIC ST TR 81
-1> (8885')
38
37
35 CHERRYVILLE 1 ST TR 81
6
-1> (8983')
ATlANTIC ST 470
�
(10360')
1
TENNECO w IA TANG I
SE
�
(10038')
�;�o���o����o���o��!�O·�=-+- ----.!i�:':n='-------t---.,2��0�
A
'
OAlUM HORIZON 10
/
/
INDEX MAP
E
I
oo oo I{) PALEOSTRUCTURAL DIP SECTION IN FRIO FORMATION, NUECES CO., TX. DAlUM: LOCAL MARKER IN FRIO V. E.
=
32 X
Fig. 9.
Frio palaeostructural section A-A' (see Fig. 6 for location and Fig. 1 for setting) showing an overall regressive stratigraphical interval bounded by marine flooding surfaces expanding dramatically across two growth faults, with concomitant facies changes. (Modified from Edwards, 1986.)
served as an effective sediment trap, especially as the local onshore-directed subsidence gradient was opposed to the obliquely offshore-directed sediment transport gradient. Storm-deposited sands typically pinch out down-dip, leading to the development of combination traps in off-structure positions. Facies changes at faults that develop topographical relief
The balance between sediment supply and subsid-
ence occasionally results in a significant topographi cal scarp being developed where the fault intersects the sea-floor. This is illustrated by the Lower Miocene of southwestern Louisiana (Fig. 1), where a large depositional basin formed as a result of the sudden and rapid removal of subsurface salt, and then stabilized when the salt body was fully evacu ated (Edwards, unpublished). Thin stratigraphical units in this area were mapped using almost 2000 well-logs, micropalaeontological data, and seismic data. The maps show the devel--
275
Different ial subsd i ence 39
40
41
BRITlSH-AMERICAN
CITlES SERVlCE
HAMON
#1
12
13
i} (8549')
(8288')
(8185')
ST TR
ST TR
f/1
(9400')
10 #1
ST TR
i} (9121')
42 15 #3
ST
44
3
TENNECO
RENWAR I< PHILLIPS
43
HAMON HAMON
8#2
ST
8#2
ST TR
5
(9029') (9021') (9220')
#1
HOGG
f/1
(8905')
sE B'
HORIZON 20
B
E
1
*-......
oo oo 1[)..--
INDEX MAP
2
*
\ 39
N
1
40
*-0
\ 41
*
....
10,000' 3000
m
42 43
PALEOSTRUCTURAL DIP SECTION
44
*·-¢-.0
'----
NUECES AND SAN PATRICIO COS., TX
3
0
DATUM: LOCAL MARKER IN FRIO
B'
V. E.
=
32 X
Fig. 10. Frio palaeostructural section B-B' (see Fig. 6 for location and Fig. 1 for setting) showing dramatic expansion of two sandy regressive cycles, each bounded by marine flooding surfaces, across a growth fault, with concomitant facies changes. Note excellent roll-over structure with expansion and improved development of sandstones up-dip to the northwest into the growth fault. (Modified from Edwards, 1986.)
opment of channel sandstone bodies that traverse the up-dip stable shelf and trend toward the down dip basin, where large quantities of sand were depo sited on the downthrown side of the fault (Fig. 13). The pattern is repeated in several successive units (Fig. 14). The coincidence of the structural boundary and the facies boundary indicates that the basin was a topographical as well as a structural low, which not only collected sediment but also attracted distribu tary channels and incised valleys from the adjacent highs (Fig. 13B). Similar relationships between struc tures, relief and facies have been recognized in
other settings (e.g. Hopkins, 1987; Leeder & Alexander, 1987). The up-dip areas (Fig. 13B) appear to be com prised largely of shallow-water mudstones and local ized mouth-bar sandstones, which were deposited while sea-level was relatively high . The development of narrow incised valleys allowed the preservation of these fine-grained deposits, and contrasts with the Wilcox example, described above, in which there was extensive scouring at the bases of distributary channels. The apparently stable channels suggests a relative fall in base level, due either to eustasy or to local structural uplift beneath the upthrown block
276
M.B. Edwards
Updip
Down dip 2
3
4
Down dip
Updip
Datum
2
4
3
Datum Top
Top J
J
Paleostructure
2000' A
610m
Base J B
230'
405'
500'
2000'
Fig. 11. (A) Frio schematic palaeostructural section of one stratigraphical unit ('J') with well-log segments selected from four wells. (B) Frio stratigraphical section with vertical scales adjusted to normalize for differential subsidence across growth faults. The correlations suggest that thin digitated SP facies up-dip have closely time-equivalent ratty/serrated SP facies down-dip. Figures 9 & 10 demonstrate that the facies changes occur at growth faults.
due to movement of deep salt. The down-dip basin fill attests to the huge sediment volumes that were bypassed through the relatively small valleys.
PRESERVATION POTENTIAL THRESHOLDS
The above examples raise the question as to why some of those faults that are not associated with depositional topography show dramatic facies changes, whereas others do not. It is customary to express the significance of a growth fault in terms of its growth ratio, as this can be measured directly in the subsurface data. However, this ratio gives no information about the absolute rates of subsidence on either side of the fault. It seems likely that a major factor controlling facies development is the absolute subsidence rate. Along a continuum of subsidence rates, will be rates that separate domains that are associated with the preservation or non-preservation of a particular facies component, or sedimentary feature (Fig. 15). These specific subsidence rates can be referred to as preservation potential thresholds for a particular sedimentary feature in a particular depositional
environment. As noted above, absolute subsidence rate has to be evaluated in the context of the other variables that affect base level: eustatic fluctuations, and the amount and calibre of sediment supply (e.g. Swift & Thorne, 1991). For example, in the deltaic setting discussed above, distributary channels are a significant source of erosion. With sufficiently low subsidence rates, migrating channels will remove all or most of the progradational facies, resulting in amalgamated multistory channel sand bodies. When the threshold subsidence rate is exceeded, progra dational facies will be preserved . Another example is the balance of erosion and deposition on the shoreface . A threshold value of subsidence, in the context of the other controlling variables, separates regimes in which mud layers will be eroded from those in which they are preserved. Analogous thres holds could be postulated for other environments and sedimentary features.
CONCLUSIONS
1 The interplay between erosional and depositional processes in sedimentary environments is controlled to a large extent by subsidence rates. The latter
277
Dff i erent ial subsd i ence
l
FORESHORE Bay/
Barrier/
lagoon
strandplain
SHOREFACE
Storm erosion
SHELF
_
A Stable setting
B Unstable setting
Prograding wave-dominated shoreline
Fig. 12.
Schematic cross-sections showing wave-dominated shoreline progradation in (A) stable and (B) unstable settings. Characteristic well-logs (SP only) are shown. (A) This section indicates the presence of wave grading shaping the foreshore and upper shoreface, and storm events resulting in erosion of the upper shoreface and deposition of a storm couplet on the shelf. (B) The enhanced subsidence rates in the unstable setting allows for the preservation of the muddy portion of the storm couplet from erosion, whether by wave grading or a subsequent storm event. The enhanced preservation seems to have resulted from greater subsidence rates, which modified local base levels. (Modified from Edwards, 1984.) A
Topography
B
healed by sedimentation
Topography controls sedimentation
Fig. 13.
Schematic block diagram contrasting: (A) normal sedimentation across growth fault without development of topographical relief, versus (B) growth fault with associated topographical scarp and bypassing of coarse sediment across up-dip block to rapidly subsiding down-dip block. Based on a study of the Lower Miocene. (Modified from Edwards & Tuttle, 1993.)
Amalgamated channel complex
Expansion with drastic facies changes across growth fault Expansion without facies changes across growth fault
M. B. Edwards
278 - DowndipHawthorne 1 Delcambre E 135-04E-001 (12281')
A
Texaco lle Blanc B 135-04E-Q49 (12036')
B
- Downdip143'/44m
137'/42m
Jllt lllf Jllt�JIIr �II).: (IIi ill}.} ill[
Fault/channel relationships
1 Fault buried; no channels in area
,04
168'/51m
108'/33m
153'/47m
132'/40m
316'/96m
2 Fault active but channels not closely related to fault traces
283'/86m
3 Fault active with tight control on channel location. Channel updip and downdip
743'/226m
4 Fault active with tight control on channel location. Channel in downdip well only
780'/238m
5 Fault active with tight control on channel location. Channel in downdip well only
411'/125m
-1700'/520m
111r� �111
6 Fault active with tight control on channel location. Channel in downdip well only
14. (A) Palaeostructural cross-section showing change of Lower Miocene section across a prominent growth fault that had topographical relief developed during part of its history. The section was subdivided into six units for mapping and correlation, of which the deepest unit has no clear lower boundary in this well. (B) The six units are shown with vertical scales adjusted to normalize for differential subsidence across the fault. For each layer, the original thicknesses are shown, and the expansion ratio is shown between the logs. Fig.
279
Dff i erent ial subsd i ence
Examples of different fault/subsidence relationships 0)(1) (1) c·- :J ro � V>0 "' Vl "' "' (.) � .0 c (.) "' "' .!: "0
Fault A
Fault
1
Fault 2
Fault 4
Fault 3
'iii
D
.0 ::J Vl
Preservation threshold: storm-deposited mud
-
Preservation threshold: progradational facies
-
Channel inc;sion threshold:
Preservation of storm-deposited mud Preservation of progradational facies Updip bypass due to channel incision
u
D -----------
------- -
---
---------
--
-- -----
D - ---
-------
--
- - - - - --- -
D
+ -0
-
u
\\
�--
-I
--
D
J _i��_]__ --
-
--
----
-
-
--
----
-
-
--
----
\\
-
u
]\�1: : 1\�
Presence of features
:I I I I I I I I I I I •
I§!
•
I§!
•
•
I§!
•
I§!
Fig. 15.
Representation of preservation potential thresholds in growth-fault settings. The vertical axis shows increasing absolute subsidence rate, but is schematic with no scale. Most of the stratigraphical units depicted in this study are considered to be 'fourth order', and hence have a duration of the order of approximately 100 ka. Preservation thresholds are depicted for progradational facies at a lower rate than that for storm-deposited mud in the transitional lower shoreface to shelf environment. Five different fault situations indicate how various combinations of depositional environment and subsidence rates determine whether particular facies will or will not be preserved across specific growth faults. Note that expansion ratio is related to relative rather than absolute subsidence rate. If blocks on either side of a growth fault are on the same side of the preservation threshold, then facies patterns will be similar with regard to the particular process threshold, although thickness will be different. Settings that represent examples described in this paper are fault A, Lower Miocene of southwestern Louisiana; fault 1, amalgamated sandstones on both sides of fault common in up-dip areas; fault 2, Wilcox in Live Oak County, Texas; fault 3, Frio in South Texas; fault 4, thick sections with similar facies on both sides of the growth fault, common in down-dip areas.
affects preservation potential via thresholds that separate domains of facies preservation versus erosion. 2 Subsidence rates, through their effect on preser vation potential, influence facies composition, including geometry, bedding characteristics, fabric, palaeontology and seismic response. The effect of subsidence rates is readily demonstrated in the Gulf Coast Basin due to the presence of growth faulting. In tectonically stable basins, the effect of subsidence rate on facies composition may not be determined readily. 3 Where subsidence rates were sufficient to create topographical scarps at the depositional surface, the effect of faulting on sedimentary facies is greater, as there is then a feedback effect on the location of
major facies, and not just the preservation potential of sensitive component facies. 4 The documented effects of preservation thresholds on reservoir architecture indicate that it would be useful to be able to predict preservation patterns on unexplored fault blocks in growth-faulted basins. However, the present study indicates that expansion ratios are generally inadequate to make such predic tions: absolute subsidence rates are required, but are often difficult to obtain or predict.
ACKNOWLEDGEMENTS
The Wilcox and Frio work was initiated while I was with the Bureau of Economic Geology of the
M. B. Edwards
280
University of Texas. I thank my colleagues there for their generous support, especially Don Bebout, Bonnie Weise and Bill Galloway. Competent assist ance was provided by Rick Schatzinger, Jim Lockley, Down Downey, Susan Hallam, Steve Mann and Doug Wilson. Contract funding which led to some of these results was provided by the US Department of Energy and the Gas Research Institute. This early and much of my subsequent work on Gulf Coast onshore stratigraphy has been supported by many oil and gas companies, for which I express my appreciation. The Miocene work benefited from the micropalaeontological expertise of J . Loyd Tuttle and seismic data from Geophysical Pursuit, Inc. Yvonne Bowlin ably assisted in all parts of this study. This manuscript has been improved by the reviews of William R. Dupre, William C. Ross and the Editor Guy Plint. In addition, many colleagues , too numerous to mention, have discussed the concepts discussed herein, offering anything from constructive criticism to disbelief. Finally, I thank Harold Reading for picking me up at Gatwick after my first trans-Atlantic journey, and for attempting to teach me to drive on the correct side of the road.
REFERENCES BEBOUT,
D.G., WEISE, B.R., GREGORY, A.R. & EDWARDS, M.B. ( 1982) Wilcox sandstone reservoirs in the deep subsurface along the Texas Gulf Coast. Tex. Univ. Bur. econ. Geol. Rep. Invest., 117, 125 pp. BERG, R . R . & PowELL, R . R. (1976) Density-flow origin for Frio reservoir sandstones, Nine Mile Point Field, Aransas County, Texas. Trans. Gulf Coast Assoc. Geol. Soc. , 26, 310-319. BoYD, D.R. & DYER, B.F. ( 1964) Frio barrier bar system of south Texas. Trans. Gulf Coast Assoc. Geol. Soc., 14, 309-321. CuRTIS, D.M. ( 1970) Miocene deltaic sedimentation, Louisiana Gulf Coast. In : Deltaic Sedimentation Modern and Ancient (Eds Morgan, J.P. & Shaver, R.H.), Spec . Pub!. Soc. econ. Paleontol. Mineral, Tulsa, 15, 293-308. CuRTIS, D .M. & Picou , E.B., JR. (1978) Gulf Coast Cenozoic; model for application of stratigraphic concepts to exploration on passive margins. Trans Gulf Coast Assoc. Geol. Soc. , 28, 103- 120. EDWARDS, M.B. (1980) The Live Oak delta complex : an unstable shelf-edge delta in the deep Wilcox trend of South Texas. Trans Gulf Coast Assoc. Geol. Soc., 30, 71-79. EDWARDS, M.B. (1981) Upper Wilcox Rosita delta system of South Texas: growth-faulted shelf-edge deltas. Bull. Am. Assoc. petrol. Geol. , 65, 54-73.
EDWARDS,
M.B. (1984) Stratigraphic and structural analysis of growth-faulted regions using well logs: a workshop. Houston, Texas. Unpublished lecture notes and problems. EDWARDS, M.B. (1986) Sedimentary effects of differential subsidence in Frio shoreface-shelf sediments, Gulf Coast Tertiary. Houston geol. Soc. Bull., 28, 10-14. EDWARDS, M . B . (1990) Stratigraphic analysis and reservoir prediction in the Eocene Yegua and Cook Mountain Formations of Texas and Louisiana. In: Sequence Stra tigraphy as an Exploration Tool (Ed. Armentrout, J .M.), pp. 151-164. Gulf Coast Section, Soc. econ. Paleontol. Mineral. 1 1th Ann. Res. Conf. EDWARDS, M.B . & TuTTLE, J.L. (1993) Regional Sequence Stratigraphy and Exploration Potential of the Lower Miocene of Southwest L ouisiana. Proprietary industry study (unpublished) 121 p. GALLOWAY, W . E . (1989) Genetic stratigraphic sequences in basin analysis I: architecture and genesis of flooding; surface-bounded depositional units. Bull. Am. Assoc. petrol. Geol. , 73, 125-142. GALLOWAY , W.E., HoBDAY , D . K. & MAGARA, K. (1982a) Frio Formation of the Texas Gulf Coast Basin - depo sitional systems, structural framework, and hydrocarbon origin, migration, distribution, and exploration potential. Tex. Univ. Bur. econ, Geol. Rep. Invest., 122, 78 p. GALLOWAY, W.E., HOBDAY, D . K. & MAGARA, K., (1982b) Frio Formation of the Texas Gulf Coast Plain: depo sitional systems, structural framework, and hydrocar bon distribution. Bull. Am. Assoc. Petrol. Geol. , 6�>, 649-688. HoPKINS, J.C. (1987) Contemporaneous subsidence and fluvial channel sedimentation: Upper Mannville C Pool, Berry Field, Lower Cretaceous of Alberta. Bull. Am. Ass. petrol. Geol. , 71, 334-345. LEEDER, M.R. & ALEXANDER, J. ( 1987) The origin and tectonic significance of asynunetrical meander-belts. Sedimentology, 34, 217-226 . MARTIN , G.B. (1969) The subsurface Frio of South Texas: stratigraphy and depositional environments as related to the occurrence of hydrocarbons. Trans. Gulf Coast Assoc. Geol. Soc. , 19, 489-499. MARTIN , G.B. (1970) Depositional history: key to explo ration. Oil Gas J. , January 12, 98-106. MiTCHUM, R.M. & VAN WAGONER, J.C. (1991) High frequency sequences, and their stacking patterns: sequence-stratigraphic evidence of high-frequency eustatic cycles. Sediment. Geol. , 70, 131-160. MoRTON, R . A . (1981) Formation of storm deposits by wind-forced currents in the Gulf of Mexico and the North sea. In: Holocene Marine Sedimentation in the North Sea Basin (Ed. Nio, S.D.), Spec. Pubis int. Ass. Sediment., No. 5, pp. 385-396. Blackwell Scientific Publications, Oxford. PoSAMENTIER, H . W . & WEIMER, P. ( 1993) Siliclastic sequence stratigraphy and petroleum geology - where to from here? Bull. Am. Assoc. petrol. Geol. , 77 , 731-742. SLOANE, B.J. (1971) Recent developments in the Miocene Planulina gas trend of south Louisiana. Trans Gulf Coast Assoc. Geol. Soc. , 21, 199-210. SNEDDEN , J.W. & NU MMEDAL, D. (1991) Origin and geometry of storm-deposited sand beds in modern sedi ments of the Texas continental shelf. In: Shelf Sand and
Dff i erent ial subsd i ence Sandstone Bodies: Geometry, Facies and Sequence Stra tigraphy (Eds Swift, D.J.P., Oertel, G.F., Tillman, R . W. & Thorne, J.A.), Spec. Pubis int. Ass. Sediment., No. 14, pp. 283-308. Blackwell Scientific Publications, Oxford. SwtFT, D.J.P. , PHILLIPS , S. & THORN E , J .A. (1991) Sedi mentation on continental margins, IV: lithofacies and depositional systems. In: Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (Eds Swift, D.J.P. , Oertel, G.F., Tillman, R.W. & Thorne, J.A.), Spec. Pubis int. Ass. Sediment., No. 14, pp. 89- 152. Blackwell Scientific Publications, Oxford. SwtFT, D.J.P. & THORNE , J.A . (1991) Sedimentation on continental margins, I: a general model for shelf sedimentation. In: Shelf Sand and Sandstone Bodies: Geometry, Facies and Sequence Stratigraphy (Eds Swift, D.J.P., Oertel, G.F., Tillman, R.W . & Thorne, J.A.), Spec. Pubis int. Ass. Sediment . , No. 14, pp. 3-3 1 . Blackwell Scientific Publications, Oxford. THORSE N , C. E. (1963) Age of growth faulting in southeast Louisiana . Trans. Gulf Coast Assoc. Geol. Soc., 13, 103- 1 10. VENDEVILLE, B.C. & JACKSON , M.P.A. (1992) The rise and fall of diapirs during thin-skinned extension. Tex. Univ.
281
Bur. econ. Geol. Rep. Invest. , 209, 60 p. B.R., EDWARDS, M.B . , GREGORY , A.R . , HAMLIN, H.S. , JIRIK, L . A. & MORTON , R.A. (1981) Geologic Studies of Geopressured and Hydropressured Zones in Texas: Test-well Site Selection, Final Rep ort. Texas University Bureau of Economic Geology, unpublished contract report prepared for Gas Research Institute, 308 pp. WHEELER, H.E . ( 1964) Baselevel, lithosphere surface, and time-stratigraphy. Geol. Soc. A mer. Bull., 75, 599-6 10. WINKER, C. D. & EDWARDS, M.B. (1983) Unstable progra dational clastic shelf margins. In: The Shelfbreak: Critical Jnte1Jace on Continental Margins (Eds Stanley, D.J. & Moore, G.T.), Spec. Publ. Soc. econ. Paleontol. Min eral . , Tulsa, 33, 139-157. YE, Q . , MAn-HEWS, R.K., GALLOWAY, W. E . , FROHLICH, C. & GAN , S. (1993) High-frequency glacioeustatic cyclicity in the Early Miocene and its influence on coastal and shelf depositional systems, NW Gulf of Mexico Basin. In: Rates of Geologic Processes: Tectonics, Sedimen tation, Eustasy and Climate (Eds Armentrout, J.M., Bloch, R. & Olson, H.C.), pp . 287-298. Gulf Coast Section, Soc. econ. Paleontol. Mineral. 14th Ann. Res. Conf.
WEISE,
Sequence and Seismic Stratigraphy in Facies Analysis
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
Spec. Pubis int. Ass. Sediment. (1995) 22, 285-303
Seismic-stratigraphical analysis of large-scale ridge-trough sedimentary structures in the Late Miocene to Early Pliocene of the central North Sea J O E CAR T WRI G HT Department of Geology, Imperial College of Science, Technology and Medicine, Prince Consort Road, London SW7 2BP, UK
ABSTRACT
This paper describes the geometry, seismic-stratigraphical characteristics and palaeogeographical setting of a system of ridges and troughs measuring approximately 5-20km long, by 1-3km wide by 100-300 m high, developed in the mudstone-dominated Upper Miocene and Lower Pliocene of the central North Sea. The ridge-trough system occupied a base-of-slope to basin floor setting centred on UK Quadrant 22 at the time of deposition. The ridge crests are oriented NNW-SSE and are almost straight in plan view. The troughs are characterized by incision at the base, followed by drape, and subsequent onlap fill. The ridges and troughs have a distinctly aggradational character. Two possible mechanisms are proposed to explain the development of these structures. The geometry and scale of the ridge-trough structures is similar to that described for mud waves formed on deep-marine sediment drifts by the action of oceanic bottom currents. By analogy, it is possible that the ridge-trough structures could have formed in response to bottom currents flowing along the axis of the central North Sea, linking the Norwegian-Greenland Sea with the Bay of Biscay. This mechanism requires a marine connection in the southern North Sea, and direct evidence for this possible connection has been removed by erosion during the Pleistocene. An alternative mechanism is that the ridge-trough system formed in response to downslope-directed currents of unspecified type developed on the pro delta slope. This mechanism is appealing in that it explains the orientation of the ridges and troughs (orthogonal to the nearest delta front) , and, most significantly, explains the significant incision observed along the troughs. The proximity of the ridge-trough structures to a rapidly prograding Neogene delta (the Skaggerak Delta) suggests a close link between pro-delta processes and the formation of these structures. The necessary confinement of the current system may have been provided by slope gullies, but conclusive evidence of this on the regional seismic data is lacking at present.
INTRODUCTION
described from the Cenozoic of the North Sea. They are on average 5-20 km long, and have a fairly regular spacing of 1-3 km. They have not been described in any previous publications on the North Sea Basin, and directly analogous structures have not, apparently, been described from any other basin. The lack of attention given to these extraordinary structures can probably be attributed to the fact that they are contained in an interval that has no economic potential from a petroleum perspective.
This paper describes a set of large-scale sedimentary structures that are developed in Upper Miocene to Lower Pliocene mudstones in the central North Sea. The structures are a series of ridges and troughs that are distributed throughout an elliptical area centred on UK Quadrant 22 (Fig. 1). The ridge-trough structures formed in a physiographic base-of-slope basin-floor setting, although the maximum water depths were relatively shallow, between 200 and 3 00m. The ridge-trough structures are by far the largest type of discrete sedimentary structure so far
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
285
286
J.
Cartwright
spacing of 25 km was used to map the distribution of the ridge-trough system (Fig. 1). The area of interest is located in the axis of the Cenozoic Basin, close to the region of maximum total Cenozoic: subsidence and sediment accumulation (Ziegler,, 1982; Joy, 1993).
The quality and quantity of petrophysical, litho logical and biostratigraphical data available for the Neogene interval in the North Sea is greatly inferior to that available for the more prospective Palaeogene and older formations. This contrast in the well data base explains why there have been so few studies on the Neogene of the North Sea, and contrasts markedly with the vast literature available on the Palaeogene (see Lovell (1990) for a comprehensive bibliography). The main aims of this paper are to describe the seismic facies characteristics of the ridge-trough structures with particular emphasis on their internal reflection configurations and external morphology, and to place these structures in a palaeogeographical context. The paper concludes with some preliminary ideas on possible depositional models.
Cenozoic Basin evolution
The Cenozoic North Sea Basin is widely regarded as a type example of a post-rift thermal sag basin (Dewey, 1982; White, 1989). Rifting ceased at the end of the Jurassic along most of the con stituent grabens of the North Sea Rift (Ziegler, 1982; Bertram & Milton, 1989; Cartwright, 1991). After a period of infilling of remnant depositional topography in the Early Cretaceous and early part of the Late Cretaceous, the rift axis and rift shoulders began a phase of regional sag-type subsidence that persists to the present-day (Joy, 1993). Uplift of the hinterlands bordering the North Sea Basin com-· menced at approximately the same time as regional subsidence (Watson, 1985; Rundberg, 1989), and
GEOLOGICAL SETTING
A grid of regional two-dimensional seismic data covering the central North Sea, with an average line
····
•
·
·····
·-
� �� !a ····· · ··
NORTH•
·· "············ · ···.: . ·
. . . . . !. . . . .
. -�-, .. .··········.
···············•
: :
,
·
�
········ ·········
.. .
•
··
.
.
� ·,("'.. """..
\
BlJ
Study Area
�·
.
.....
_,
..
.....
�.. .
,..•:
.......
····'·
Distribution of ridge/trough features
.
.
·
58°N
,.. ......... t ..
; ··•.······· ·········
.
.
r,,i-·::i:·''.
,
:.-"
l
. ... ....... --····
: :
\• • • �·· ·
15
i • \ \ ', l \� \
: :
�
·
. . .. . .
;
.
. . .. ...
.
.. .
\
.
. .
:
..
. .. . .... . ..
r--.
16
Quad Numbers
Fig_ 1. Location map of the study area for the detailed mapping of the ridge-trough system in the central North Sea. Also
located are the two line drawings of regional seismic lines illustrated in Fig. 2.
Ridge structures, central North Sea
sediment supply changed from dominantly carbonate material during the Late Cretaceous, to dominantly clastic material throughout the Palaeogene and Neogene (Ziegler, 1982; Lovell, 1990). Basin-margin deltaic systems were established along certain sectors of the basin at different times, presumably in response to changes in the location of the main centres of marginal uplift. Uplift of the Scottish Massif in the Late Cretaceous to Early Cenozoic (Watson, 1985) is generally considered to pre-date the main phase of uplift of Fennoscandia, which probably dates from the Oligo-Miocene (Holtedahl, 1960; Spjeldnaes, 1975). One result of this diachroneity in uplift of the eastern and western marginal areas is that the pattern of infill of the Cenozoic is asymmetric. The main progradational clinoforms built out from the western margin for the Palaeocene to Oligocene interval, with sediment mainly supplied from the Scottish Massif and the Shetland Platform (Parker, 1975; Rochow, 1981; Lovell, 1990). A major reorganization of the clastic supply into the basin occurred at approximately the beginning of the Middle Miocene, and progradation switched to a dominantly easterly source. Over large areas of the Norwegian, Danish, and German sectors of the North Sea, the Middle Miocene to Late
287
Pleistocene interval is dominated by a large deltaic system, which prograded from the eastern basin margin in Denmark, crossing the Danish Sector en route to the central North Sea Basin (Fig. 2). This large deltaic complex is referred to in this paper as the Skaggerak Delta. This name was chosen because the main axis of sediment supply has been mapped along a NE-SW course running through the present day Skaggerak (Morgan, 1992). The scale and influence of the Skaggerak Delta on the Neogene fill of the basin is illustrated in two schematic sections based on interpret ations of regional seismic lines calibrated with the released well-data (Fig. 2). The Skaggerak Delta is represented on these two sections by large, westward-building sigmoidal and tangential oblique clinoforms above the regional middle Miocene marker (MM). The magnitude of the clastic input from the east is particularly apparent in the way that the clinoforms cross the basin axis and downlap or onlap the western margin of the basin. These two sections also illustrate the asymmetric pattern of basin-fill in the Cenozoic, with contrasting westerly and easterly sources of sediment supply in the Palaeogene and Neogene, respectively. The Late Miocene to Early Pliocene ridge-trough
(a) ENE
WSW
(b)
Q._
_.7pm
4km
_____
�
Late Miocene - Early Pliocene Ridge - Trough System
Fig. 2. Line drawings of two regional profiles across the central North Sea, showing the elements of the Mesozoic rift, and
the configurations of the post-rift succession. The Late Miocene to Early Pliocene ridge-trough system is located at the base of slope of large-scale prograding clinoforms associated with the Skaggerak Delta. MM, middle Miocene marker; uM, upper Miocene; eP, lower Pliocene; Ku, upper Cretaceous.
J.
288
Cartwright
system occupies a position in the basin axis in the two sections illustrated in Fig. 2. The ridge-trough system is associated with the toesets of major west ward prograding clinoforms of the Skaggerak Delta. Progradation and aggradation of the clinoforms is matched in part by aggradation and southwesterly migration of the ridge-trough system.
EXTERNAL MORPHOLOGY AND INTERNAL CONFIGURATION OF THE RIDGE-TROUGH SYSTEM
Database and methodology
The descriptive parts of this paper are based largely on interpretations of reflection seismic data. Litho logical data used in this study are restricted to cut tings descriptions compiled in completion logs for over 100 released well-data in the central North Sea, and a limited amount of petrophysical data. Unfor tunately, the biostratigraphical control of many of these released well-data is very poor, in part due to lack of sampling and in part due to drilling methods (Gradstein et at., 1992). In addition, the lack of velocity information means that accurate well-to seismic calibration has not been possible. Conse quently, detailed lithostratigraphical correlations have not been successful. The analysis is therefore based on a seismic-stratigraphical approach, which
although limited by the vertical seismic resolution of the order of 1 0-15 m in the Neogene interval, is none the less sufficient to give some preliminary indications of the processes responsible for the devel opment of the ridge-trough structures. Detailed interpretation and mapping of the ridge trough structures was carried out on a study area of approximately 400 km2 in Quadrant 22, of the UK Sector of the North Sea. This study area is located in the central portion of the larger region in which the ridge-trough structures are developed (Fig. 1). The area was selected for detailed study because well and seismic data were readily available. Mapping was based on a seismic grid with an average line spacing of 1 km, which was sufficient to allow indi vidual ridges and troughs to be correlated without difficulty. The seismic data is a late 1980s vintage, and the quality and resolution of the Late Miocene to Early Pliocene interval are excellent. External morphology
The external morphology of the ridge-trough struc tures is illustrated with reference to a representative seismic section from the study area (Fig. 3). The section is oriented NE-SW, approximately orthog onal to the long axes of the ridge-trough structures in the study area. The base of the main ridge-trough system in the study area is picked at a low-frequency reflection that appears to mark the upper limit of a
Fig. 3. Seismic profile from the study area showing the cross-sectional geometry of the ridge-trough system. Note the diachroneity evident in the abandonment of active ridges from right to left. Also note the asymmetry of the post abandonment infill of the troughs. Section oriented SW-NE. Datum at top of section is 600ms TWTT (two-way travel time).
289
Ridge structures, central North Sea
precursor set of ridges and troughs (reflection A on Fig. 3). The upper boundary is picked wherever a clear onlap can be seen in a trough (e.g. reflection B on Fig. 3). The ridges have an external form that is best described as mounded (Mitchum et at., 1977). The intervening troughs have a channel-like profile. The ridge-trough structures have crest-to-crest spacing of between 1 and 3 km, and vary from almost perfectly symmetric to slightly asymmetric in form. The crest-to-trough amplitude of the structures defined using the upper and lower bounding reflec tions varies between 1 00 and 200m. The ridges have broad crests with fairly steep flanks, with dips of up to 7° (uncorrected for compaction). Throughout the study area there is clear evidence that the system of ridges and troughs is diachronous from northeast to southwest. This diachroneity can be seen, for example, in Fig. 3 in the contrasting topography and external form of the reflections comprising the ridge-trough system across the section. The relief on the shallow, high amplitude reflections that can be seen draping the ridges on the northeastern side of the section (e.g. reflection C) increases to the southwestern side of the section. Reflection C changes from a post-abandonment drape of the ridge in the northeast, into an internal reflection of the ridge-trough structure in the southwest. This indicates a younging of the active growth of the ridges to the southwest.
Internal reflection configurations
The typical internal configurations of the ridges and troughs are illustrated in Fig. 4, a seismic section parallel to that shown in Fig. 3, but reproduced at an enlarged scale to enhance the resolution. The ridges are composed of low- to moderate-amplitude reflections with moderate continuity. The lowermost reflections are concordant with the base, but there is increasing convexity exhibited by successively shallower reflections, such that the maximum relief between ridges and adjacent troughs is at the upper bounding reflection (e.g. at reflection A). Numerous reflection terminations can be observed within the ridges illustrated in Fig. 4, particularly towards the margins of individual ridges. Some of these reflection terminations may be indicative of genuine stratal truncation, pinch-out or downlap. However, considering the probable distortive effect on ray-paths of the troughs, and the possibility of multiples being generated in the overlying high amplitude reflections, it is possible that many of the reflection terminations are artefacts produced by interference or tuning. Genuine downlap is observed only rarely within the ridges, and consequently the most common diagnostic feature used to define a constructional sedimentary mound (Mitchum, 1985) cannot be confidently applied to the ridges illustrated in Figs 3 & 4. 1:
0
. .
....
.
1.
�Ni:-
,,-"''IHI•I�����
1km
Fig. 4. Enlarged scale seismic section oriented at right angles to the axes of the ridge-trough structures showing truncation
of events and downlap at the margins of the ridges and the aggradational configurations of stratal reflections within both the ridges and the troughs. Datum at top of section is 600 ms TWIT.
290
J.
Cartwright
The troughs imaged on the section in Fig. 4 are characterized by an almost concordant set of high amplitude reflections whose lower contact is difficult to define. The overall configuration of the troughs is almost equivalent to an upside-down version of the ridges. No obvious incision can be seen at the base of the lower reflections defining the troughs, and direct correlation between reflections from the troughs to adjacent ridges is hampered by changes in reflection character and continuity. However, at least some of the low-amplitude internal reflections of the ridges pass directly into the concave, high amplitude reflections of the troughs. It is possible that the contrast in reflection amplitude between reflections in the troughs and ridges is due to lateral variations in lithofacies. For example, some form of lag deposit in the troughs could explain the higher amplitudes. However, it is also possible that the variation is due to tuning and the concave shape of the reflectors. At the upper levels of the ridge trough system, the relief across the troughs is gradually diminished by the development of onlap fill (e.g. Fig. 4, point X). Thus, although the troughs seen in Fig. 4 have the general morphology sugges tive of some type of channel, there is little conclusive evidence that the troughs on this section formed by an erosive process. A clearer indication of a possible erosional origin of the troughs can be seen on another NE - SW
oriented seismic section (Fig. 5), located several kilometres to the northwest of Fig. 4. Figure 5 shows an example of a trough that has a clearly recognizable basal reflection, with a classical channel profile (trough A). Erosional truncation can be seen on both margins of the trough, and there are sufficient genuine reflection terminations to define the position of the basal erosion surface. This basal surface (reflection X) passes laterally into the neighbouring ridges, and can be correlated with the erosional bases of the two neighbouring troughs (points B and C). Above the erosional base in trough A, the overlying reflection is concordant with reflection X, and above this there is a clear onlap of the trough margin by three subhorizontal reflections. The evol ution implied by these reflection configurations is: (i) incision, (ii) drape, and (iii) onlap fill. Taking the truncation and onlap relationships together, trough A can only reasonably be interpreted as an erosional feature. A similar pattern of erosion, drape and onlap can be seen on the other two troughs in Fig. 5. The variability in the detailed configurations of the three troughs seen on the section in Fig. 5 is typical of that observed throughout the study area. Incision is invariably observed along at least part of the axial extent of every trough. However, the degree of incision, and the clarity with which the incision is expressed, varies widely, both from trough to trough, and along any individual trough. The
Fig. 5. Seismic section oriented at right angles to long axes of ridge-trough structures, showing contrast in reflection configurations of three troughs, and evidence of erosional truncation at their bases . Datum at top of section is 600ms TWTI.
Ridge structures, central North Sea
troughs in Fig. 4 are typical of those in which basal incision is difficult to define from the cross-sectional reflection configurations, whereas trough A in Fig. 5 is a type example of a trough whose seismic expression most clearly conforms to that of an ideal incised profile. Figure 6 is a comparison of two representative seismic sections located along the axes (NW-SE) of a ridge and trough pair. The internal configurations of the ridge and trough are comparable, with a consistent parallel form with respect to the lower and upper bounding reflections. There is no evidence of any significant thickness variation along the axes of either the· ridges or the troughs, and no evi dence of internal downlap, progradation, onlap or erosional truncation in the along-axis direction of either the ridge or the trough. Interpretation
The reflection configurations of the ridge-trough system observed in cross-section are summarized in a schematic line drawing in Fig. 7. One of the most
291
puzzling aspects of the ridge- trough system illus trated in this figure is that there is a combination of features that are diagnostic of both constructive and destructive process elements. Included for com parison in Fig. 7 are two idealized line drawings through a definitive constructional mound and an incised channel separating interchannel ridges based on criteria outlined in Mitchum et al. ( 1977) and Mitchum (1985). The incision observed at the base of at least a portion of every trough clearly points to the activity of an erosive current regime for at least part of the time represented by the growth of the ridge-trough structures. In contrast, the increasingly convex upwards configurations of the reflections within the ridges is interpreted as evidence that the ridge struc ture is at least partly constructional. The common development of drape configurations in both ridges and troughs indicates that the constructional and erosive components of the system alternated with a dominantly pelagic or hemi-pelagic mode of deposition. Applying the criteria outlined by Mitchum (1985)
NW
�:: :;:�: :::--�:::=�::::::;:;::_,,>M,�I"
'II''.!I!INI'"'I�-1!::!.;:;¥:· -;;;'P!!'P:O
0 I
"'"
SE :·'
·�,;,::::�:;:;;;;;2::1:=:==���-��-�....;_,�,
'"'"'t•
I tiMrll:....... .
..Ill
1km
I
Fig. 6. Comparison of two longitudinal seismic sections along a ridge (A) and adjacent trough (B). Note the absence of erosional truncation, onlap, downlap or progradation in the internal reflections of either the ridge or the trough.
J.
292
Cartwright
-
0
EROSIONAL CHANNEL
1 km
--- ---
BASE MIDDLE MIOCENE MARKER
INTER-MOUND TROUGH (NON-EROSIONAL)
Fig. 7. Summary diagram showing the main stratigraphical features of the ridge-trough system observed from the study
area. Idealized sections across an erosional channel and a pair of constructional mounds shown beneath main figure for comparison. Note that the ridge-trough structures have affinities with both these contrasting idealized cases. Note also that with low-resolution seismic data it may be difficult to distinguish the two cases, particularly If multiples obscure the true stratal configurations at the ridge margins.
for the recognition of different types of construc tional mounds, it is difficult to make a conclusive argument that the ridges are primary constructional features. Neither is it reasonable to interpret the ridges as purely remnant relief produced by erosion along the troughs. Without the benefit of detailed lithological calibration of the seismic data it is impossible to determine the extent to which erosive and depositional processes alternated or dominated throughout the evolution of the ridge-trough system.
MAPPING THE RIDGE-TROUGH SYSTEM
Mapping of the ridge-trough system was under taken in order to determine the spatial relationship between ridges and troughs and to determine their orientation relative to the major physiographic features of the basin. Detailed mapping of indi vidual ridges and troughs was restricted to the small study area (Fig. 1) because it was only there that
the relatively closely-spaced seismic grid ( 1-2 km) allowed correlation of individual ridges and troughs. Structural contour mapping of specific reflections within the Late Mioc�ne to Early Pliocene interval was attempted initially, but correlating horizons across the grid proved to be too difficult, mainly due to loss of reflection continuity in the troughs. Consequently it was decided to map the individual trough axes because these are narrower than the ridges and are therefore less likely to be miscorrelated. The map of trough axes in the study area is presented in Fig. 8. The most important features of this map are all well constrained by the seismic grid. These are: 1 the orientation of the trough axes is confined to a narrow range from 320°N to 340°N; 2 the troughs are linear to gently sinuous, and die out abruptly in both directions along their axes; 3 there is no interconnection between individual troughs - they neither bifurcate nor anastomose in either direction; 4 the length of those troughs that die out along
293
Ridge structures, central North Sea
··············
Seismic lines
-�- Axis of Trough • --·•
Termination of Trough
-
(decreasing relief in direction shown)
0
N
+
2
Fig. 8. Map of the trough axes in the study area. Also shown is the seismic grid used for the mapping. Note the
discontinuous array of trough axes in both an upslope and downslope direction (regional palaeoslope to the northwest).
their axes in both a northwesterly and southeasterly direction within the confines of the mapped area varies from 2 to 10 km - one trough is over 20 km long, and continues outside the limits of the seismic grid; 5 the trough axes are fairly consistently spaced over the area, with the intertrough spacing varying between 1 and 3 km. One of the most surprising aspects of Fig. 8 is the lack of any interconnection between adjacent troughs. The incision observed in the troughs (e.g.
Fig. 5) clearly indicates that currents strong enough to erode muddy sediments flowed along the trough axes. Because it has not been possible to correlate the erosional bases of individual troughs across the study area, the extent to which erosional activity in the troughs was synchronous is not known. For example, did the basal incision along the troughs take place at approximately the same time through out a narrow zone, a broad zone, or was it confined to a single trough at any one time? The overall diachroneity in the growth of the
J.
294
Cartwright
system from northeast to southwest evident on the seismic data (e.g. Fig. 3) and on the regional sections (Fig. 2) suggests that only a few of the trough axes indicated on Fig. 8 were active at any one time. This interpretation is based on the assumption that, because the depositional unit containing the ridge trough structures is diachronous as a whole, then it is probable that the time of initial incision of troughs was also diachronous. If this is correct, then Fig. 8 should not be viewed as a palaeogeographical reconstruction, but as an overall progression estab lished during several million years of deposition and erosion. Another important feature of the organization of the troughs is the observation that the troughs die out in both a northwesterly and southeasterly direction. Of the 20 troughs partly or wholly mapped within the study area, 13 die out along their axes at one end of the trough within the mapped area, and five die out at both ends along the trough axis (Fig. 8). A series of line drawing cross-sections constructed at intervals along one of the troughs
I I
� I
:
�
l
(A-B of Fig. 8) with terminations in both directions along-axis are presented in Fig. 9. This diagram illustrates the subtle changes in internal configuration of the trough, with variable depth of incision and a general decrease in relief towards both ends of the trough. At the limit of the trough, only a vague indication of an erosive feature can be discerned, and beyond this point, the trace of the trough axis projects directly into a ridge structure.
PALAEOGEOGRAPHICAL SETTING OF THE RIDGE-TROUGH STRUCTURES
A palaeogeographical reconstruction of the Early Pliocene North Sea Basin is presented in Fig. 10. The reconstruction is based on a basin-wide cor relation of a seismic horizon marking the end of the period in which the ridge-trough system formed . This horizon is dated as approximately 4 Ma, i.e . Early Pliocene. The depositional shoreline break
UNE1
13------------ ---------- -------------0 LINE1 Be
i
i i ! :
f i
I
G----------L·-·---r·+------�------o UNE 2
\I
I !
I I
I
�
=
: :
:
1 i
!
.. . . .
i
it ti ��:: I
.
:
/ i
/
!
:I
iI
..
u
I • : :t --{J UNE 5 G- --- -- - --- ----- ------- ------
1 : ;
� X
1 I
o
•
/i
Ii
JJJ<
./ii
/ /
� ,:
I
i c�: I
.... .
�
t
: ;
1;
I
·
,
·
\i I
p �••
\ i
G ··- ----1- ---·-----··-···--·-··1--
\
I
d UNE7
Fig. 9. Line drawings of seismic sections oriented SW-NE along the trough A-B (location in Fig. 8) , showing the change in reflection configurations of the trough along its axis. Note that the trough dies out towards the north-northwest and merges into a neighbouring ridge structure.
Ridge structures, central North Sea
was defined at the break of slope between topsets and foresets for clinoforms along the progradational sections of the basin margin (Posamentier & Vail, 1988). Progradational shorelines were differentiated from retrogradational or static shorelines by defining the extent of marine onlap for the Late Miocene to Early Pliocene interval. Major depocentres were
295
identified by mapping regions dominated by pro gradational clinoforms. The palaeobathymetry of the basin was estimated by measuring the relief on clinoforms using topset surfaces as mean sea-level projection datums. Figure 10 shows that the North Sea Basin was a narrow marine gulf or seaway in the Early Pliocene,
Actively prograding shoreline at 4ma
\9ma
\
Passively onlapped shoreline at 4ma
I
Position of depositional shoreline break for Skaggerak Delta at various times Progradation route of Skaggerak Delta
Bathymetric contour in metres I
�istribution of Ridge-Trough System �bandoned at 4ma
Other major distributary systems
Active at 4ma
t �·
. ,, ,: ,,
i :
i :
...... ................\� ........
Fig. 10. Palaeogeographical reconstruction for the Late Miocene to Early Pliocene. The North Sea is shown as a semi
enclosed marine seaway. Marine connection to the Norwegian and Greenland Sea was open via a narrow strain between two deltas. The marine connection to the south is uncertain (see text). The position of the depositional shoreline break of the Skaggerak Delta is shown for various time intervals, and illustrates the rapid advance of this major depositional system.
296
J.
Cartwright
with a NNE-SSW trending elliptical central basin narrowing to the north and south. The bathymetry estimated from clinoform relief indicates that maxi mum water depths of 300-400 m were developed in a basin-plain region centred on Quadrant 22 of the UK Sector (ct. Figs 1 & 10). Palaeobathymetric constraints from the limited sampling of microfauna in wells in the central North Sea are imprecise, but are broadly consistent with the magnitude of water depth interpreted from the clinoform geometry. Water depths derived from an analysis of subsidence of the central North Sea are in close agreement with the values presented here (Joy, 1993). The region of active ridge-trough structures is shown to be confined to the southern end of the basin plain in the Early Pliocene. The full extent of the ridge-trough structures (active and abandoned) is also indicated in Fig. 10, in order to show the overall southwestward migration of the ridges and troughs from the Late Miocene to the Early Pliocene. The most probable marine connection into the North Sea was at the northern end of the basin, with the Norwegian-Greenland Sea (Ziegler, 1982). Whether there was an additional connection to the Bay of Biscay to the south is not clear. Neogene strata are absent throughout much of the south western part of the North Sea (south of this study area), and a major erosional unconformity is developed near the base of the Late Pleistocene (Cameron et al., 1987). Hence it has not been possible to extend the palaeogeographical recon struction into the southern North Sea and clarify the nature of any possible marine connection. Four main depocentres prograded into the North Sea in the Early Pliocene. The courses of distributary systems supplying these depocentres are sche matically illustrated in Fig. 10, to give some indi cation of probable source areas, although it should be noted that there is little direct evidence concerning the exact positions of the distributary systems. In the extreme north of the basin, two large deltaic systems prograded towards each other simultaneously, narrowing the marine connection to the Norwegian Greenland Sea. In the south (south of 57°30'), clino forms of the Skaggerak Delta dominated the infilling of the basin. The rapid westward progradation of the Skaggerak Delta is indicated on Fig. 10 by the mapped positions of the depositional shoreline break through the Late Miocene to Early Pliocene interval. A relatively minor depocentre was confined to an embayment in the otherwise passively onlapped southwestern margin of the basin.
In summary, the ridge-trough system in the Early Pliocene occupied a base-of-slope-basin-fioo:r position, in water depths of 300-400 m, along the southern portion of a narrow marine basin that was being rapidly infilled by a major delta prograding east to west and either narrowing or closing any possible marine connection to the south.
DISCUSSION
Any explanation of the genesis of the ridge-trough system needs to take account of the following key observations. 1 The ridge-trough system developed during a short interval of time (c. 1-2 Myr) spanning the Late Miocene to Early Pliocene. No remotely similar features have been documented from comparable base-of-slope depositional systems in the North Sea at any other period in the Cenozoic. 2 The dimensions of the ridges are large in com parison to most constructional bedforms. They are of the order of 10 km long, 1-3 km wide, and 100-200 m thick. 3 The orientation of the long axes of the ridges and troughs range from 320°N to 340°N. 4 Ridges and troughs exhibit a variety of construe-· tiona) and erosional characteristics, from stratal thickening towards the ridge axes (constructional), to truncation of ridge flanks and incision along trough axes (erosional). 5 There is no evidence of lateral accretion of indi vidual ridges. The internal reflection configurations indicate that ridges grew by aggradation: abandon- ment was followed by a migration of active growth to new ridges to the southwest (of the inactive ridges). 6 The troughs do not interconnect and, where mapped to their full extent, die out along both axial directions. 7 The zone of ridge growth migrated from northeast to southwest with time. Depositional models Mudwaves formed by the action of bottom currents
Large-scale, constructional muddy waveforms formed by the long-term action of bottom currents have been described only from slope and abyssal plain settings in oceanic basins (see Pickering et al. (1989) for a comprehensive review), but there is no
Ridge structures, central North Sea
reason why similar structures should not be found in locations other than oceanic basins, provided that the necessary current regime was operative for a sufficient period. The dimensions and internal stratification of the ridge-trough system are com pared with a compilation of line drawings of mud waves deposited by bottom currents in various oceanic localities (Fig. 11). The relief of these mud wave examples is of the order of 10-20 m, i.e. considerably less than the c. 100-200 m relief of the ridges in the central North Sea. However, the shape, internal reflection configurations and spacing of some of the mudwaves are comparable with the ridge trough structures. Better grounds for comparison can be made with large-scale bedforms associated with a sediment drift in the Northeast Atlantic, described by Kidd & Hill (1986). The mudwaves developed on the Feni and Gardar Drifts were studied by these authors using a combination of sparker and GLORIA sidescan sonar data. The resulting map of the mudwaves (Fig. 12) shows a pattern with some similarity in scale and organization to that exhibited by the ridges and troughs from the central North Sea study area, although there is less regularity to the mudwave organization than to the ridge-trough system (Fig. 8). An interesting feature of the map shown in Fig. 12 is that the mudwave crests close at both ends of their long axes. This is in contrast to the ridge trough system where the mapping showed that it is the troughs that terminate in both directions along axis. It is not clear whether this attribute is diagnostic of mode of origin. In common with the ridges in the North Sea, the mudwaves on the Feni and Gardar Drifts show no evidence for lateral migration on an individual basis. The long axes of the mudwaves are thought to be oriented obliquely to the bottom current flow direction. The limited basis for comparison means that it is hard to make an unequivocal case that the ridge trough system in the central North Sea was the result of bottom-current activity. The similarity with the published descriptions of mudwaves associated with sediment drifts is sufficient to at least consider bottom currents as a possible driving mechanism for the development of the ridge-trough system. However, although a bottom-current mechanism may offer a reasonable explanation for both the distribution of the ridge-trough system in the basin axis· and the long-term migration of the zone of active ridge-trough structures towards the south west (i.e. along the likeliest route for currents linking
297
the two oceanic basins), it is difficult to reconcile this mechanism with the often significant incision observed along many of the troughs. In particular, it is difficult to see how bottom currents could have produced the incision of tens of metres of fine grained sediments along the troughs in a north northwest orientation (i.e. transverse to the long axis of the basin and hence the likely flow route for bottom currents), and on a longer term lead to a general southwesterly migration of the entire ridge trough system. In conclusion, the interpretation of the ridges as some form of bottom-current deposit cannot be discounted, particularly in view of the uncertainties regarding a southerly marine connection, but the observation of incision along the troughs argues against this depositional model. Pro-delta current regime
An alternative model for the development of the ridge-trough system is suggested by two observations: 1 the approximately orthogonal relationship be tween the axes of the ridges and troughs (NNW SSE to NW-SE) and the strike of the nearest depositional shoreline break (NE-SW); 2 the skewed distribution of the ridge-trough struc tures to that sector of the basin plain immediately adjacent to the toe of the Skaggerak Delta, which was by far the most significant Mio-Pliocene pro gradational system in the North Sea. The evidence of erosional truncation and incision in the troughs (Fig. 5) implies that currents flowing along the trough axes were of sufficient strength to result in the cumulative erosion of tens of metres of muddy sediment. If the troughs were foci for strong currents flowing along the trough axes, it is reason able to infer that their origin was directly upslope, either at the delta front or on the upper reaches of the pro-delta. Because the nearest delta front to the ridge-trough system was along the southeastern margin of the basin, it is reasonable to argue that the ridge-trough system developed in response to a current regime associated with the rapidly prograding Skaggerak Delta. If some form of pro-delta current regime was responsible for the development of the ridge-trough system, several questions need to be addressed. 1 What was the specific nature and flow regime of the current system? 2 How and where did the currents originate?
298
J.
Cartwright
0
1km
Fig. 1 1 . Line drawings of high-resolution shallow seismic images of mud waves from the Central and Southern Atlantic
showing typical dimensions and internal configurations. Taken from Pickering eta/. (1989). Line drawing of ridge-trough structures displayed at the bottom for comparison.
299
Ridge structures, central North Sea
N
1
0
KILOMETRES
5
r-------7'"--.:--� so·N
Fig. 12. Detailed bathymetric map
over a field of sediment waves located on the Feni Drift, Rockall Trough, North Atlantic. Note that the sediment wave crests (stippled) are discontinuous. Taken from Kidd & Hill (1986).
so•
30'
How did these currents result in the stratal patterns observed on the seismic data? 4 In the absence of any evidence linking similar ridge-trough structures to the other deltas along the basin margin, why was this current system uniquely developed in connection with the Skaggerak Delta? 5 Why was this current regime operative only during
3
the Late Miocene to Early Pliocene, given that the Skaggerak Delta was first established in the Early Eocene, and continued prograding until the Early Pleistocene? 6 Why did the locus of current activity migrate towards the southwest from the Late Miocene to the Early Pliocene? The postulated downslope current system must
300
J.
Cartwright
have been organized in linear, confined flow paths with lateral dimensions of 1-2 km in order to produce the regular, isolated and linear troughs mapped in the study area. To explain the lack of interconnection between the troughs and the obser vation that the troughs die out at both ends of their axes (Figs 8 & 9), it is necessary to invoke downslope currents that were either of a finite extent in the flow direction, or currents whose strength varied along the flow direction. Both of these conditions would be expected to apply to a downslope pro-delta current system. Any downslope currents would reach their maximum velocity at the base-of-slope, and would gradually lose velocity along the horizontal basin floor. The finite longitudinal extent of the incision observed along the troughs could, therefore, be interpreted as a result of the current exceeding a critical erosive velocity threshold for the muddy sediments as the currents accelerated downslope in a northwesterly direction. Then, as the power of the current decreased as flow proceeded along the basin floor, the degree of incision would also decrease until the trough was no longer recognizable as such (Fig. 9). One of the problems with this pro-delta current model is the large distance from the likely point of origin of the currents at the head of. the slope, to the toe region where the ridge-trough structures formed. A question that cannot be resolved as yet is how sufficient flow confinement would have been maintained from the delta front, where the currents may have originated, for example, as some form of turbid underflow (Bates, 1953), over a distance of tens of kilometres to the northwest at the base of slope, where incision commenced. The extent to which the Coriolis effect would have deflected downslope flowing currents originating at or close to the delta front is also a problem with this depositional model. If a modern system, such as the Huanghe (Wright et al., 1990), represents a reason able analogue, then along-slope deflection should be expected (Hill, 1984). It is also difficult to envisage a mechanism whereby the downslope flow would maintain its confined flow pattern once it began to decelerate at the lower break of slope. One method of providing the necessary flow confinement would be if there was a pro-delta channel or gully system linking the distributary channels at the delta front with the troughs at the base of slope (Fig. 13A). Linear gully systems have been observed and mapped on the present-day Scotian slope of eastern Canada (Piper & Sparkes, 1987). The scale and
spacing of these slope gullies is closely comparable to those exhibited by the troughs in the central North Sea, but the erosional relief of the Scotian gullies is only of the order of 10 m. It is possible, however, that even shallow erosional gullies could provide the necessary confinement and guide the flow paths from the upper reaches of the pro-delta to the toe region. Minor erosional features have been observed on the pro-deltaic foresets of the Skaggerak Delta on regional seismic data (Morgan, 1992), but the wide spacing of the seismic grid available at this time makes any direct correlation of these features with the ridges and troughs in Quadrant 22 extremely tenuous. A direct relationship between the distributary channels on the delta front and confined axes of flow at the base of slope is appealing in that it may explain the overall southwesterly migration of the ridge-trough system from the Late Miocene to the Early Pliocene. If downslope currents originated at the outlets of the major distributary channels on the delta front, possibly during one of the glacially forced relative lowstands of sea-level (Haq et al., 1987), then the locus of trough development would prob ably have been confined to the sector of the basin floor directly downslope from the major channels. If for any reason the channels avulsed at the delta front, then the locus of active ridge-trough growth would be expected to migrate accordingly (Fig. 13B) . A depositional model linking the evolution of the Skaggerak Delta directly to the development of the ridge-trough system also provides a satisfactory explanation for the aggradational nature of the stratal patterns interpreted on the seismic data (Fig. 7). The Late Miocene to Early Pliocene is known to have been a period of rapid basin sub sidence (Kooi & Cloetingh, 1989; Joy, 1993), and this led to increased aggradation of the basin-margin deltas (Morgan, 1992). The aggradational character istics observed for the troughs and ridges can be viewed as a response to the aggradation of the deltaic depositional system as a whole, including the foreset and bottomset depositional units, and any downslope current system originating at the delta front. The pro-delta current mechanism can thus explain many of the observations documented in this paper, but a major unresolved question is why the ridge- trough structures were restricted to the Late Miocene to Early Pliocene in particular? What was so unique about this period that either a pro-deltaic downslope current system or a bottom-current system should
301
Ridge structures, central North Sea
NE
A Pro-delta
"'
-
Topset
} aggradation Trough
' Bottomset
500m
�
C u r rent flow path
ll----.,--
'�
SW
--,
0
10
20km
Approximate scale
B
Palaeogeography time t1
Palaeogeography time t2
N
'
Avulsion of major distributary c h a n nel
Progradation of depositional shoreline break
Fig. 13. (A) Evolutionary reconstruction of the development of the ridge-trough system based on a pro-delta current
regime. Flow confinement from the delta front is provided by a system of slope gullies. Maximum flow velocity is reached at the base of slope, and the trough incision begins at this position. Deceleration along the basin floor results in the elimination of the troughs in a north-northwest direction. (B) Migration of the ridge-trough system is achieved by avulsion and channel switching at the delta front.
have been operative for this specific period, and not throughout the entire Cenozoic? The major climatic change that took place with the onset of the glacial interglacial cycles in the Late Miocene to Early
Pliocene (Lidmar-Bergstrom, 1982) may be part of the solution. The resolution of this question will require considerable additional research, not least of which would be a comprehensive study of the
J.
302
Cartwright
morphological evolution and palaeogeographical setting of the Skaggerak Delta throughout the Cenozoic. The impetus for additional work on the commer cially unprospective Neogene interval of the North Sea Basin may be provided by the enigmatic linear sandbodies of the Eocene in the North Sea. The troughs described in this paper may be examples of large sedimentary features that resulted from a process involving the transport and confinement of reservoir-prone units from the delta-top into deep water environments. If the troughs were a locus for sand transport and deposition, then the resultant reservoir geometry would be a linear body compar able in scale and thickness to reservoir sands devel oped on the Eocene slope and basin floor.
CONCLUSIONS
1 The ridge-trough structures of the central North Sea are unique features in the context of the Cenozoic stratigraphy of the North Sea, with apparently no direct analogues in the published literature. 2 The linearity of the structures and the regularity of their spacing suggest that they either formed in a current regime that was highly oblique or even transverse to their long axes, or in one that was parallel to their long axes. 3 Two contrasting current regimes are regarded as the basis for a genetic model: (i) bottom currents flowing along the NNE-SSW basin axis, or (ii) pro deltaic currents flowing downslope in a NW-SE direction from the delta front of the actively pro grading Skaggerak Delta. 4 Although both current regimes can provide satis factory explanations of the dimensions and mor phology of the ridge-trough structures, the often pronounced incision observed in the troughs is dif ficult to reconcile with bottom currents flowing obliquely across the long axes of the troughs. The incision is compatible with currents flowing along the trough axes. Hence the preferred model is the pro-delta current regime. 5 A weakness in the pro-delta current model is the lack of a conclusive explanation as to how the necessary confinement in the current system was initiated and maintained down the slope and across the basin floor so as to produce the linearity observed for the troughs. Slope gullies observed on foresets of the Skaggerak Delta may have provided the
necessary confinement, but much additional map ping is required to confirm this suggestion.
ACKNOWLEDGEMENTS ·
I would like to thank Fina Exploration Ltd for financial support and Fina and partners for access to well and seismic data. Thanks also to colleagues at Fina for valuable discussions, namely Rod Laver, Dave Fassom, Mick Cope and Joe Staffurth. The manuscript was greatly improved as a result of the many suggestions made by the reviewers, Chris Baldwin and Phil Hill, and the Haroldesque editorial skill of Guy Plint. Thanks also to Howard Johnson and Dick Selley for suggestions for improving the manuscript. Finally, my sincerest thanks to Harold for putting me on the right track over the last 15 years (derailments are my responsibility !). Tony Brown and Jubrail produced the figures.
REFERENCES
C.C. ( 1953) Rational theory of delta formation. Bull. Am. Assoc. petrol. Geol. , 37, 2 1 19-2 162. BERTRAM, G.T. & MILTON, N.J. ( 1989) Reconstructing BATES,
basin evolution from sedimentary thickness; the import ance of palaeobathymetric control, with reference to the North Sea. Basin Res. , I, 247-257. CAMERON , T.D.J. , STOKER, M.S. & LoNG , D. ( 1987) The history of Quaternary sedimentation in the UK sector of the North Sea Basin. I. geol. Soc. London, 144, 43 -58. J.A. ( 199 1 ) The kinematic evolution of the Coffee Soil Fault. In: The Geometry of Normal Faults (Eds Roberts, A.M. , Yielding, G . & Freeman, B . ) , Spec. Pub!. geol. Soc. London, No. 56, pp. 29-40. Geological Society of London, Bath. DEWEY, J.F. ( 1982) Plate tectonics and the evolution of the British Isles. I. geol. Soc. London, 139, 371 -412. GRADSTEIN , F.M . , KRISTIANSEN, l . L. , LOEMO, L . & KAMINSKI, M. ( 1992) Cenozoic foraminiferal and dino flagellate cyst biostratigraphy of the central North Sea. CARTWRIGHT,
Micropalaeontology , 38, 101 - 137 . HAQ, B.U. , HARDENBOL, J. & VAIL, P. R . ( 1 987) Chron ology of fluctuating sea levels since the Triassic. Science, 235, 1 156- 1 167. HILL, P.R. ( 1984) Facies and sequence analysis of Nova
Scotian slope muds: turbidite vs 'hemipelagic' deposition. In: Fine-grained Sediments: Deep Water Processes and Facies (Eds Stow, D. A . V. & Piper, D.J.W. ) , Spec . Pub!. geol. Soc. London, No. 1 5 , pp. 3 1 1 -318. Geo logical Society of London, Bath. HoLTEDAHL, 0. ( 1960) On supposed marginal faults and the oblique uplift of the land mass in Cenozoic time. Norges geol. Unders. , 208, 35 1 -357. ( 1993) Post-rift subsidence of the North Sea
JoY, A.M.
Ridge structures, central North Sea
Basin. Unpublished PhD thesis, University of London, 248 pp. Kmo, R.B. & HI LL, P.R. ( 1 986) Sedimentation on mid ocean sediment drifts. In: North Atlantic Palaeocean ography (Eds Summerhayes, C.P. & Shackleton, N.J.), Spec. Publ. geol. Soc. London, No. 2 1 , pp. 87 - 102 .
Geological Society o f London, Bath. KoOI, H. & CLOETINGH, S. ( 1989) Intra-plate stresses and the stratigraphic evolution of the North Sea Central Graben. In: Extensional Tectonics and Stratigraphy of the North A tlantic Margins (Eds Tankard, A.J. & Balkwill, H.), Mem. Am. Assoc. petrol. Geol. , Tulsa, 46, 541 -558.
K. ( 1982) Pre-Quaternary geo morphological evolution in southern Fennoscandia. Sver. geol. Unders. Ser. C. , 785, Arsbok 75 , 6. LovELL, J.P.B. ( 1990) Cenozoic. In: An Introduction to the Petroleum Geology ofthe North Sea (Ed. Glennie, K.W), pp. 273-293. Blackwell Scientific Publications, Oxford. MITCHU M , R.M. ( 1985) Seismic stratigraphic expression of submarine fans. In: Seismic Stratigraphy, II: an Integrated Approach to Hydrocarbon Exploration (Eds Berg, O.R. & Woolverton, D.G . ), Mem. Am. Assoc. petrol. Geol. , Tulsa, 39 , 1 17 - 138. MrrcH U M , R.M. , VAIL, P.R. & SANGREE, J.B. ( 1 977) Strati graphic interpretation of seismic reflection patterns in depositional sequences. In: Seismic Stratigraphy Applications to Hydrocarbon Exploration (Ed. Payton, C.E.), Mem. Am. Assoc. petrol. Geol. , Tulsa, 26, LIDMAR-BERGSTRO M ,
107 - 134.
MoRGAN, C. ( 1992) The advance of the Baltic Delta. Unpublished MSc thesis, University of London, 86 pp. PARKER, 1 . R. ( 1975) Lower Tertiary sand development in the central North Sea. In: Petroleum and the Continental Shelf of N. W. Europe (Ed. Woodland, A.W.), pp. 447-452. Applied Science Publishers, London.
303
K.T . , HISCOTr, R.N. & H EIN , F.J. ( 1989) Deep Marine Environments, Clastic Sedimentation and Tectonics . Unwin & Hyman, London, 416 pp. PIPER, D.J.W. & SPARKES, R. ( 1987) Proglacial sediment instability features on the Scotian slope at 63 degrees W. Mar. Geol. , 76, 1 5 - 3 1 . PoSAMENTIER, H.W. & VAI L , P.R. ( 1988) Eustatic controls
PICKERING,
on clastic deposition I I : sequence and system tracts models. In: Sea-level Changes: an Integrated Approach (Eds Wilgus, C.A. , Hastings, B.A. , Kendall, C. G.St C. , Posamentier, H.W. , Ross, C.A. & Van Wagoner, J.C.), Spec. Publ. Soc . econ. Paleontol. Mineral. , Tulsa, 42, 125 - 154. RocHow, K.A. ( 198 1) Seismic stratigraphy of the North Sea 'Palaeocene' deposits. In: Petroleum Geology of the Continental Shelfof North- West Europe (Eds Illing, L.V. & Hobson, G. D.) , pp. 255-266. Institute of Petroleum,
London. Y. ( 1989) Tertiary sedimentary history and basin evolution of the Norwegian North Sea: an integrated approach . Dr. Ing. thesis, University of Trondheim. SPJELDNAES, N. ( 1975) Palaeogeography and facies dis RuNDBERG,
tribution in the Tertiary of Denmark and surrounding areas. Norges Geol. Unders. , 316, 289- 3 1 1 . WATSON , 1 . ( 1985) Northern Scotland as an Atlantic-North Sea divide. J. geol. Soc. London , 142, 22 1 -243. WHITE, N.J. ( 1989) Nature of lithospheric extension in the North Sea. Geology, 17, 1 1 1 - 1 14. WRJGHT, L . D. , WISEMAN, W .J . , YAN G , Z.S. , BORHOLD, B . D . , KELLER, G.H. , PRJOR, D.B. eta!. ( 1990) Processes of marine dispersal and deposition of suspended silts off the modern Huanghe (Yellow River). Continent. Shelf Res. , 10, 1 - 40 .
ZIEGLER, P. A. ( 1982) Geological A tlas of Western and Central Europe . Shell International Petroleum BY, The Hague, 130 pp.
Spec. Pubis int. Ass. Sediment. (1995) 22, 305-327
Millstone Grit cyclicity revisited, II: sequence stratigraphy and sedimentary responses to changes of relative sea-level O L E J . M A R T I N S E N* J OH N D . C O L L I N S ONt and B R I A N K. H O L DS W O RTH:j: *Norsk Hydro Research Centre, N-5020 Bergen, Norway; tCollinson Jones Consulting, Westgate House, Hill's Lane, Shrewsbury SYI 1 QU, UK; and
:j:
Department of Geology, Keele University, Keele, Staffordshire ST5 5BG, UK
ABSTRACT The Namurian (Upper Carboniferous) Millstone Grit cyclothems of northern England were formerly interpreted as the results of deltaic progradational episodes formed during lowstands of glacial-eustatic sea-level, punctuated by widespread fossil-bearing marine bands formed during sea-level highstands. New research, based on detailed fieldwork in the Askrigg B lock and Craven Basin area, has led to a reinterpretation of Millstone Grit cyclothem stratigraphy. Although the broad framework of a cyclothem bounded by goniatite-bearing marine bands is retained, a more complex internal archi tecture is recognized . The tectonic framework of northern England in the Carboniferous evolved from an active extensional phase in the Dinantian to a thermal sag phase in the Namurian. As a response, two different marine band-bounded sequence types developed in the Askrigg Block and Craven Basin area: (i) a 'block-edge' sequence, formed when there was significant differential bathymetry between earlier active fault blocks and adjacent deep basins, and (ii) a ramp sequence, formed when clastic fill and decaying differential subsidence had reduced the bathymetric contrast between the Askrigg Block and the Craven Basin to a gently inclined ramp. The main differences between these two sequence types are: (i) the formation and preservation of fine-grained highstand strata, which is a particular featu re in the 'block-edge' sequence; (ii) formation of a falling-stage systems tract, which occurred only in the ramp sequence; (iii) the formation of a lowstand turbiditic wedge in the 'block-edge' sequence; and (iv) the greater thickness of the 'block-edge' sequence. These differences can be attributed to varying basin bathymetry and available accommodation as a response to the evolving tectonic and palaeogeographical situation.
INTRODUCTION
1992; Maynard, 1992). This area is particularly favourable for an appraisal of the consequences of sequence stratigraphical analysis particularly because of the highly refined Namurian biostratigraphical framework (Fig. 2; Holdsworth & Collinson, 1988). In particular, based on a detailed data set from the Askrigg Block and Craven Basin area in northern England (Martinsen, 1990), we discuss: (i) the goniatite-bearing marine bands and their relation to sea-level curves; (ii) the architectural arrangement of Millstone Grit cyclothems; (iii) the sedimentary responses to relative sea-level change; and (iv) the
Recent advances in stratigraphical methods, par ticularly in sequence stratigraphy (e.g. Embry & Podruski, 1988; Van Wagoner et a/., 1988; Galloway, 1989; Cross & Gardner, 1991), have caused new light to be shed upon the interrelationships of stra tigraphy and sedimentology (ct. Walker (1990, 1992) for a critical discussion). One basinal succession for which new interpretations based on sequence stra tigraphical analysis are being presented and debated is the classic Namurian (Upper Carboniferous) Millstone Grit of northern England (Fig. 1; cf. Martinsen, 1990, 1993; Read, 1991; Collinson et al.,
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
305
O.J. Martinsen, J.D. Collinson and B. K. Holdsworth
306
ALSTON BLOCK BLF STAINMORE BASIN
N
(Cross-section in Fig. 6a) ASKRIGG BLOCK
0
4, 6
2
(Cross-se . Ctton Lancaster CRAVEN BASIN
in F'
tg. 6b) Cross-section in Figs 6c and 8)
Fig. 1. Location map of northern England and simplified structural map of the study area. For a detailed structural map of northern England, see Gawthorpe eta/. (1989). The location of cross sections in Figs 6 & 8 are shown, and the location of localities mentioned in the text are indicated by numbers: 1, Lad Gill, Stonesdale; 2, Kettlesbeck, near Settle; 3, Cononley Beck, Airedale; 4, Backstean Gill, Nidderdale; 5 , Horn Crag Q uarry, Silsden; 6, general area of several localities on Askrigg Block used in Fig. 8. For grid references, see Figs 6 & 8.
PF, Pennine Fault; BLF, Bu tterknowle-Lunedale Fault.
existence of a Namurian major-scale cyclicity. The aim is to extend existing stratigraphical models for the transition from the Askrigg Block into the Craven Basin (Martinsen, 1993) and to provide the impetus for a thorough re-evaluation of Millstone Grit cyclicity.
NAMURIAN GEOLOGICAL FRAMEWORK
Stratigraphy
The concept of sedimentary cyclicity is not new in Carboniferous geology and dates back to Phillips (1836), who recognized that the Carboniferous Yoredale cyclothems of northern England consist of repeated lithological successions and that there are lateral variations within these. For the Namurian, the detailed stratigraphy is dependent largely on the detailed goniatite bio stratigraphy established by the work of Bisat (1914, 1924) and Hudson (1934, 1939, 1941, 1944). Because of inadequate exposures, complex structure, and local, inconsistent stratigraphical naming of coarser grained lithologies, the goniatite-bearing marine bands are vital for stratigraphical analysis. The later work of Ramsbottom and co-workers (Ramsbottom et al., 1962, 1978; Ramsbottom, 1 977, 1979) has been instrumental in consolidating and expanding Bisat's and Hudson's work. Holdsworth & Collinson (1988) presented a goniatite-based biostratigraphical scheme. It presents a refined subdivision of Namurian stratigraphy and will be used in the following. Other
schemes include those of Riley (1985) and Riley et al. (1987). There are about 60 marine bands in the Namurian succession, most containing their own goniatite species (see Fig. 2; Holdsworth & Collinson, 1988). The average duration of a Millstone Grit cycle is about 180 000 yr, although significant variability probably existed (Maynard & Leeder, 1992). Most marine bands have an interbasinal extent (Ramsbottom et al., 1978). Some marine bands are particularly widespread and extend over much of Europe, in fact, a few can be traced from western Ireland to the Urals (Wagner et al., 1979; Martinsen, 1990). Other essential work for establishing a reliable Namurian stratigraphical framework was carried out in numerous mapping projects organized by the British Geological Survey. Perhaps the most sig nificant of these is the classic Rossendale Anticline Memoir (Wright et al., 1927), where the authors not only presented a reliable stratigraphical scheme based on goniatites, but also showed a sophisticated understanding of how a particular Millstone Grit cyclothem operated in terms of depositional setting and relative sea-level change. Structure and regional depositional patterns
The central and northern England area (the Central Province) was in early Carboniferous time domi nated by a pronounced differential bathymet1ry related to carbonate build-ups on tilted fault blocks (Gutteridge, 1989), some of which were granite cored (Miller & Grayson, 1982; Fig. 3A). The
Millstone Grit cyclicity MINOR CYCLE H2c(ii) z
<X: f=
a: 0 "�
<X:
::J
e
�c 0
H. smithi-undulatum
H2a(i)
H. smithi/Hd. proteus
H1b(iv)
H. aff. beyrichianum
H. beyrichianum H. beyrichianum
Q
H1a(iv)
Homoceras sp. nov.
H1a(iii)
I.
subglobosum
H1a(ii)
I.
subglobosum
H1a(i)
I.
subglobosum
E2c2(iii)
N.
nuculum
E2c2(ii)
N.
nuculum
E2c2(i)
N.
nuculum
E2c1
N.
stellarum
E2b3
Ct. nititoides
E2b2(ii)
F.
E2b2(i)
Ct. nitidus
E2b1(iii)
C. subplicatum Grp.
E2b1(ii)
Ct. edalensis/Ct. bisati
E2b1(i)
C. subplicatum Grp.
E2a(iii)
E.
yatesae
E2a(ii)
E.
ferrimontanum
E2a(i)
C. cowlingense
E1c
C. malhamense
UJ
0 :I: u
Q. Q. ::J c co ·c: ::J
E
C, Cravenoceras; Ct, Cravenoceratoides; E, Eumorphoceras; F, Fayettevillae; H, Homoceras; Ht, Homoceratoides; N, Nuculoceras; V, Valites.
. H2a(ii)
eostriolatus
H1b(i)
Q;
Fig. 2. Excerpt of the goniatite band stratigraphy established for the Namurian. Based on Holdsworth & Collinson (1988).
V.
H. undulatum
<X: a:
co
co
H2c(i) H2b
H. aff. beyrichianum
u
"-
Ht. prereticulatus
H1b(iii)
-2
co z ..... 0 t
GONIATITE SPECIES
H1b(ii)
z
(j)
307
z
<X: 0
a: UJ C!l (f) z a:
<X:
Pendleian (part)
shallow-water limestone platforms were separated from adjacent deeper water basins by steep sub marine slopes (Fig. 3A; Collinson, 1988; Gawthorpe et al., 1989). Most authors agree that this configur ation resulted from N -S extension north of the northwards migrating Variscan front (Leeder, 1982, 1988; Leeder & McMahon, 1988). Active extension probably ceased in early Namurian time, causing a change from mainly fault-controlled subsidence to thermal sag, which progressively decreased during the Namurian (Fig. 3B; e.g. Leeder & McMahon, 1988). The area discussed in this paper, the Askrigg Block to Craven Basin transition, lies at the northern margin of the central Province, and the two components are separated by the North Craven
holmesi
Fault (Fig. 1). The Askrigg Block is cored by the Wensleydale Granite (Fig. 3A), which contributed significantly to maintaining the Askrigg Block as a structural high throughout Carboniferous time. The Askrigg Block was tilted to the north during the main extensional phase, and a half-graben basin, the Stainmore Basin, was formed along its northern edge. This basin is bounded to the north by the Butterknowle- Lunedale Fault, which also marks the southern end of the Alston Block. The Stainmore Basin ceased to subside differentially by early Namurian time (based on uniform thickness data of overlying strata across basin), and throughout the rest of the Namurian, had no significant impact on depositional patterns (Martinsen, 1990). Therefore, for the cyclothems investigated in the present paper,
308
0.1. Martinsen, J.D. Collinson and B . K. Holdsworth
A END-DINANTIAN BLOCK-AND-BASIN BATHYMETRY CENTRAL PROVINCE N-S CROSS-SECTION
N 30km
Rossendale Block
25km
s
30km
Wensleydale Granite
Only relative vertical scale TRANSITION OF SOUTHERN END OF ASKRIGG BLOCK
B N
S
s
N
E1c-E2a(iii); pronouced 'block-edge' with turbidite-fronted deltas
E2b2(i)-Kinderscoutian: gently-inclined ramp with shallow-water deltas
Inactive Craven Fault system Only relative vertical scale
Ca. 40km
Fig_ 3. (A) North-south cross-section across the Central Province to show the end-Dinantian block-and-basin physiography. Modified from Collin son (1988). (B) Cartoon illustrating the change of the southern end of the Askrigg Block from a pronounced 'block-edge' to a ramp in the Namurian. Based on Martinsen (1993).
the Stainmore area is considered part of the Askrigg Block. Although the Namurian basin-filling sediments of the Central Province were partially deposited in deep water and therefore are expected to be largely marine, it is important to realize that the basin complex formed in an intracratonic setting far removed from the nearest ocean. In fact, the contemporaneous Mid-European Ocean probably lay as far south as the present position of Portugal (Anderton et at., 1979; Ziegler, 1986, 1988), possibly connected to the Central Province through the Variscan foredeep in central Europe. The Central Province may also have had a connection to the Canadian Maritime Provinces (Ziegler, 1988). This intracratonic setting and isolation from the ocean is important for explaining marine bands and Millstone Grit cycle salinity variations (Collinson, 1988; Holdsworth & Collinson, 1988). The initial Namurian depositional systems pro graded from the north and northeast into the Central Province. At the southern end of the Askrigg Block, across the Craven Faults at the transition to the Craven Basin, a pronounced
bathymetric break occurred, which was a first-order control on early Namurian depositional patterns. To the south, turbidite sedimentation was import ant in the earliest Namurian cycles (Pendleian and earliest Arnsbergian; Baines, 1977; Collinson, 1988; Martinsen, 1993). With time (from latest Arnsbergian), this bathymetric break transformed into a gently inclined, southerly dipping ramp (Fig. 3B), resulting from decreased differential sub sidence rates due to transition to thermal sag and progressive fill of the Craven Basin. Consequently, the potential for significant turbidite sedimentation immediately south of the Craven Faults declined, and shallow-water deltaic sediments dominated deposition in the later Namurian cycles of this area (cf. Martinsen, 1993). Further south in the Central Province, in the Pennine Basin, basinal shales dominated deposition until the mid-Namurian (Kinderscoutian). These shales were succeeded first by turbidite-fronted deltas and then by shallow-water deltas (Walker, 1966; Collinson, 1969, 1988). Although this tran sition occurred at a much later stage compared with the northern margin of the Craven Basin, it was
Millstone Grit cyclicity
perhaps not a result of ongoing, prolonged differ ential subsidence in the south. Rather, because the clastic supply was mainly from the north, the feeding systems did not manage to prograde into the area until the middle part of the Namurian. Basin-filling was therefore delayed in the Pennine Basin compared with the Craven Basin, probably as a result of the clastic supply being absorbed by depocentres beneath the southern North Sea. In the Staffordshire Basin, a slightly different situation existed. There, protoquartzitic turbidites were supplied from the southern Midland land mass in several minor cycles from as early as E1b (Holdsworth, 1963; Trewin & Holdsworth, 1973). In the Widmerpool Gulf to the east, the same situation seems to have taken place (Aitkenhead, 1977). Although sand was supplied to these areas at this early stage, the supply was insufficient to fill the available accommodation space, and complete basin filling did not occur until the Marsdenian R2b(ii) Ashover/Roaches Grit was supplied from the north (Jones, 1980; Collinson, 1988).
GONIATITE-BEARING MARINE BANDS AND RELATIONSHIP TO SEA-LEVEL
The marine bands, which allow such a refined bio stratigraphical breakdown of the Namurian, are virtually the only fossiliferous sediments in Namurian basinal successions. Fossiliferous intervals range from a few centimetres up to around 2 m in thickness and generally constitute only a small proportion of the total Namurian succession (Collinson, 1988; Holdsworth & Collinson, 1988). Sediments between marine bands are virtually devoid of body fossils, irrespective of whether the sediments are fine grained, deep-basinal mudstones, thick sequences of turbidite sandstones or progradational deltaic suc cessions (Fig. 4). Where the intervening sediments are deltaic progradational successions, the marine bands occur closely above an abandonment surface or just below the base of the next upward-coarsening pro-delta unit, leaving a clearly asymmetric cyclo them motif (Fig. 4). These observations, coupled with the widespread nature of many marine bands (see above) suggests an external control on the occurrence of goniatite-bearing marine bands. Most Namurian researchers have believed this mechanism was glacial-eustasy, a theory which is increasingly supported by growing evidence for widespread
309
glaciations in Gondwana during the Carboniferous (Caputo & Crowell , 1985; Veevers & Powell, 1987). Although goniatites constitute the stratigraphically most useful faunal phase of marine bands, it is not uncommon for various other faunal elements to occur in distinct faunal phases stacked one above the other. Based on detailed analysis of a cored condensed basinal succession at Ashover, Derbyshire, an ideal succession of faunal phases has been constructed (Ramsbottom et al., 1962; Holdsworth & Collinson, 1988; Fig. 4). In the Ashover boreholes the total succession from base Namurian to the Marsdenian is of basinal mudstones with no coarse clastics. The Ashover succession was deposited on part of the Derbyshire Massif, a mudstone-draped Dinantian carbonate platform, isolated from clastic supply by a partial moat of deep-water gulfs. The isolation from clastic supply means that there is virtually no change of lithofacies between marine bands and intervening barren mud stones. In such circumstances it is very difficult to conclude that marine bands are purely reflections of sediment starvation and condensation. The presence or absence and the changes of fauna in a succession like Ashover almost certainly result from changes in the overlying water column, in particular its salinity. Of the faunal elements present in the sequence, only the thick-shelled goniatites are thought to reflect fully marine salinity (Fig. 4). The other phases are thought to record progressively less saline conditions. The implication is that the marine bands do not reflect faunal concentration but represent the arrival of fauna that had been absent from the basin during intervening times. From the succession of faunal phases, the arrival of a fully marine fauna appears to be associated with pro gressively increasing salinity (Fig. 4)·. Because many of the faunal elements are nektonic or pelagic, it is not clear whether the salinity changes applied to the whole water column or whether they affected only the photic zone, where most of the fauna swam. Whatever the extent of desalination, it is apparent that the time of maximum salinity in the basin is most likely to coincide with or occur very slightly after the peak of the eustatic sea-level curve. If one considers the wider palaeogeographical context of the Namurian basins of the British Isles (see above), it is clear that they occupied an enclosed intra continental setting, removed by considerable dis tance and topographical complexity from the nearest ocean. Connections between the ocean and the British basins were therefore probably long and
O.J. Martinsen, J.D. Collinson and B . K. Holdsworth
310
Eustatic curve F-R
E1c minor cycle in Craven Basin (Ca. 800 m)
E2a(i)
E 1 c minor cycle in Staffordshire Basin (Ca. 25-90 m)
6
Delta top sands
Delta slope heterolithics
Ideal faunal/salinity cycle (based on Ashover cycles)
6 5
E2a(i) Barren shale
Turbidite sands and ironstones
E_j_
Barren s ale Calcareous siltstone 6
E1c
SIMPLE, MINOR CYCLE
Goniatite AID 4 Spat 3 Lingula 2 Planolites 1 Fish BARREN 1 Fish 2 Planolites 3 Lingula 4 Spat 5 AID 6 Goniatite 5
4
3
L
AEN
T Ideal
faunal/salinity oscillation
1
Fig. 4. Minor cycle stratigraphy showing variations of the basic cyclicity, from the ideal faunal/salinity cycle erected from Ashover borehole data (cf. Ramsbottom et al. , 1962; Holdsworth & Collinson, 1988), to thicker, sand-prone cycles from the Staffordshire Basin and the Craven Basin. The probable relation to the eustatic curve is also shown. (Modified from Holdsworth & Collinson, 1988.)
A, Anthracoceras; D, Dimorphoceras (thin-shelled goniati tes) .
tortuous. That, coupled with the fact that the British basins were supplied by very large volumes of fresh water, which delivered an abundant supply of clastic sediment, makes it easy to envisage how fluctuations in eustatic sea-level would have induced changes in the salinity of basinal waters. During lowstands, the cross-sectional areas of the connecting channels would be much reduced, especially if they were of low relief. Their main function would then be to accommodate the discharge of fresh water from the basins to the ocean. During highstands, expanded cross-sectional areas would have allowed oceanic waters to enter the basins and thereby increase salinity. Salinity changes are recorded from the Red Sea as a result of Pleistocene/Holocene changes in sea-level, although in this case lowstand led to hyper salinity because of the climate and Jack of river discharge (Thunell et al., 1988 ). The fact that sal inities appear to have oscillated throughout the Namurian within rather constant limits suggests that the straits which connected the basins with the ocean were rather stable and not subjected to major tectonic change or sedimentary choking. Partial analogues at the present day, such as the Skagerak Kattegat, the Hudson Straits and the Straits of
Hormuz are all relatively sediment free. Simple but conservative calculations based on likely river discharges (approximately 1 x 10 4m3 s-1, based on Kinderscout Grit channels; Collinson, 1969 ) and basin water volumes (around 1 x 10 13 m3, based on an assumed areal extent of the basins of 5 x 10 4km2 and a depth of 200 m) suggest that river water could have replaced basinal water in a matter of 30-35 yrs if the process was 100 % efficient. Given the average duration of a minor cycle was 180 000 yrs, desalination seems likely even at much lower efficiencies. In order for basinal waters to have become fully saline after a period of lowstand, and thereby allowing the deposition of a marine band, it is clear that some threshold of eustatic level must have been exceeded and that sea-level must have remained above the threshold for a period of time. The marine bands, therefore, on this basis, are thought to record periods of time at and around eustatic maxima. Given that there was probably some lag in the response of salinity to eustatic level, it is likely that the interval of marine band deposition may be skewed slightly later about the eustatic peak (Fig. 4). From their positions close to the bases of pro-
Millstone Grit cyclicity
gradational deltaic cyclothems it is likely that marine bands are associated with periods of relative sedi ment starvation and of rapid deepening. If marine bands are considered only in this context then it could be suggested that they record the time of maximum rate of rise in relative sea-level, i.e. around the R inflection point of the Exxon sequence stratigraphical model (Jervey, 1988; Posamentier et at., 1988). However, we believe this to be unlikely. If we are correct in believing that full marine salinity was necessary to support a fauna including thick-shelled goniatites, a view supported by the common occur rence of Radiolaria in this faunal phase and (as far as is known) no other, then, given the ocean-distant setting of the Millstone Grit basins, it seems likely that such salinities could have been achieved only close to the eustatic maxima. This would normally be expected to significantly post-date the R inflection point. It could even be inferred that, in some cases, parts of the curve between the R inflection point and the eustatic maximum may be represented by the inferred lower salinity phases of fauna that occur within marine bands below the thick-shelled goniatites. In other cases, there is compelling evi dence to place the response to the R inflection point clearly below the marine band. In the E1c- E2a cycles in Staffordshire (Trewin & Holdsworth, 1973), units of 'transitional shale', unfossiliferous and lacking carbonate, invariably separate a sideritic protoquartzitic turbidite unit from calcitic marine siltstones below and goniatite-bearing marine bands above (Fig. 4). Such shales can readily be interpreted as recording periods of rapid adjustment of environ ment when conditions favoured neither marine fauna or mineralogy nor influx of delta-derived clastics. The transitional shales below a marine band probably closely approximate the R infection point, whereas the overlying transitional shales mark the earliest stages of eustatic sea-level fall. In cyclic deltaic sequences, maximum conden sation, predicted as a response to the R inflection point, is most likely recorded by heavily bioturbated flooding surfaces or coal seams, which commonly terminate sandstone sequences. Such surfaces commonly are separated from the overlying marine band by a thin interval of mudstone. This mudstone is therefore thought to be the record between the R inflection point and the sea-level maximum of the marine band. Whilst these modifications regarding marine band
311
occurrence in relation to eustatic oscillation differ only slightly from those of conventional sequence stratigraphical theory, they illustrate ways in which the principles of sequence stratigraphy must be adapted in settings other than those in which they were formulated initially.
M ILLSTONE GRIT CYCLOTHEM ARCHITECTURE: RESPONSE TO RELATIVE SEA-LEVEL CHANGE
Introduction
Since Wright et al. (192 7) published the classic Rossendale Anticline memoir, the established view has been that Namurian cyclothems represent more or less continuous progradational episodes formed during falling or low relative sea-level, punctuated by marine transgressions, highstands of sea-level and deposition of goniatite-bearing marine bands (Fig. 5). In the following, two Millstone Grit cyclo thems are discussed in order to illustrate what has been interpreted as a different and more complex internal architecture. The two cyclothems were selected based on outcrop quality, and because they show different lateral facies and thickness develop ments. It is focused on the cyclothem development from the Askrigg Block and into the northern part of the Craven Basin. The depositional history of each cyclothem is reviewed briefly (for a full dis cussion of this aspect for the lower Namurian (Pendleian -Chokierian) cyclothems, see Martinsen, 1990, 1993). In the following, we take a less rigid view concern ing the use of the word 'sequence' than some existing definitions. In the American Geological Institute definition, for example (Bates & Jackson, 1987), 'sequence' is constrained to a depositional unit bounded by unconformities. However, as there is in existence more than one sequence stratigraphical model (Embry & Podruski, 1988; Van Wagoner et a/., 1988; Galloway, 1989; Cross & Gardner, 1991), and the models differ in what type of strati graphical surface is chosen as a boundary, we use the term 'sequence' for the stratigraphical unit that is practical to use in a given basin. The most practical sequence type is chosen based on an objective evaluation of the nature and extent of various strati graphical surfaces. The most easily defined, most extensive surfaces, based on detailed facies analysis,
0.1. Martins�n, J.D. Collinson and B . K. Holdsworth
312
,_ Fluvio-deltaic distributary
Goniatite band Barren mudstones
channel systems
�
Goniatite band
Slope channels feeding turbidites
Continuous progradation with successive clinoforms downlapping base-of-cycle goniatite band
Base-of-slope turbidites related to
Adjacent goniatite bands
gravitational instability or underflows
only separated by thin shale
during progradation
in basinal areas
Coarsening-up siltstone m
t#
Goniatite band
Fig. 5. The traditional Mill stone Grit cycle interpretation as a result of d elta progradation (turbidite-fronted delta example). See text for discussion.
are chosen as sequence boundaries. To avoid conflict with allostratigraphy, the same type of surface is picked throughout the succession studied. Goniatite-bearing marine bands are considered to represent the most significant stratigraphical levels in the Namurian of the Askrigg Block to Craven Basin succession, because of their lateral extent and distinctive faunal characteristics (Martinsen, 1993). They represent maximum transgression between successive progradational episodes and compare with the genetic stratigraphical sequence boundaries of Galloway (1989). Marine bands have a thickness and are not surfaces in a strict sense. However, when we describe them in the following, we see their mid-point as a 'surface', representing peak marine band conditions. Detailed facies analysis and regional understand ing of marine bands suggest that they represent the units across which the most significant palaeogeo graphical changes occurred in the Namurian suc cession. The marine bands are therefore chosen as sequence boundaries. Within a marine-band bounded sequence, however, the idea is to establish a detailed stratigraphical architecture, using both other stratigraphically significant surfaces (e.g. lowstand erosion surfaces or unconformities) and systems tracts.
The E2a(i) cyclothem Boundaries and thickness
The Cravenoceras cowlingense goniatite band forms the base of the E2a(i) cyclothem (Fig. 2), and can be found across most of the Askrigg Block, as well as in the western and easternmost parts of the Craven Basin (cf. Martinsen, 1990). The top of the cyclothem is marked by the base of the Eumorphoceras bisul catum ferrimontanum band, which forms the base of the overlying E2a(ii) cyclothem. This band is found throughout most of the Craven Basin, and on the northwestern side of the Askrigg Block. On the Askrigg Block, the cyclothem is approxi mately 70 m thick in the northwest, whereas it thins to around 40 m in the southeast (Fig. 6A-C). In the Craven Basin it is highly variable in thickness. In the northeast of the basin it is apparently less than 1 m thick due to condensation (further discussion below), whereas in the Lancaster Fells in the west, the E2a(i) cyclothem approaches 150 rn in thickness (Fig. 6A; cf. also Moseley, 1954, 1956). Depositional model and tectonism
Detailed studies (cf. Martinsen, 1993) suggest the following depositional model for the E2a(i) cyclo-
Millstone Grit cyclicity
them (Fig. 7). A phase of mudstone deposition occurred on the Askrigg Block. A zone of barren mudstones were succeeded by a phase of normal marine conditions represented by a fossiliferous interval with the Cravenoceras cowlingense fauna. A slight coarsening upward trend within the mud stones overlying the fossiliferous level suggests progradation of a clastic shoreline from the north. A drop of relative sea-level probably followed, and a rapid basinward shift of facies is interpreted to have taken place, causing establishment of the shoreline at the southwestern margin of the Askrigg Block. From this shoreline, sediment was supplied down into the Craven Basin as turbidity currents and suspended silts, leading to the establishment of a slope apron (Fig. 7). During the ensuing rise of relative sea-level, fluvial aggradation took place within the incised palaeo valleys (Upper Howgate Edge Grit) and eventually led to the deposition of a fluvial sheet sandstone (the Tan Hill Grit; Figs 6A & 7; Martinsen, 1993). Deposition corresponded with continued pro gradation and fill of the basin, with the development of a deltaic plain in the northwestern parts of the Craven Basin. The deltaic plain was subsequently flooded and transgression took place. In the initial phase of the transgression, shallow-water conditions persisted on the northwestern parts of the Askrigg Block, allowing for sporadic, high-frequency pro gradations of small delta lobes (Figs 6A & 7). Finally, the relative sea-level rise overwhelmed sediment supply, causing transgression and establish ment of offshore shelf conditions where mud stone sedimentation dominated. The conditions were shallow enough for occasional storms to rework benthic fauna into a shell bed, the Lad Gill Limestone. Fluvial channels were confined largely to the western parts of the block, leaving the eastern part relatively starved of sediment (Fig. 7). The cause for this was probably increased subsidence along the Dent Line, a long-lived lineament bounding the Askrigg Block in the west (cf. Underhill et al. , 1988). The Dent Line probably experienced minor reactivation during the early Namurian (Martinsen, 1993). The lack of sediment supply to the eastern parts of the block caused sediment starvation in the eastern parts of the Craven Basin (Figs 6C & 7), because most sediment was supplied to the basin across the western part of the Askrigg Block. This continued sediment supply to the western parts of the Craven Basin caused the slope apron or turbiditic
313
wedge to develop there (see above), because there was still a significant bathymetric contrast between the block and the basin. E2a(i) cyclothem architecture and sequence stratigraphy
The Cravenoceras cowlingense marine band itself is interpreted to represent the eustatic maximum or shortly thereafter (discussion above; Fig. 6A- C), whereas the thin, barren mudstones below the marine band probably represent maximum trans gression (maximum flooding). The fine-grained deposits above the marine band (eustatic maximum), and directly below the major fluvial complexes, but separated from these by a major erosion surface across which a major basin ward shift of facies occurs, are interpreted as late highstand deposits (Fig. 6A & C). The mudstones are interpreted to conform to the upper part of the falling limb on the eustatic curve, and on the relative sea-level curve prior to initiation of the fall, but at a time when the rate of rise had decreased to the point where clastic systems could prograde back into the area from the north. The erosion surface underneath the Upper Howgate Edge Grit fluvial complex (Fig. 6A) is interpreted as a lowstand erosion surface. Based on the interpretation that a prominent basinward shift of facies takes place across the surface, it is also interpreted as an Exxon-type sequence boundary. The surface is unlikely to correlate with the base of the turbiditic slope sediments in the Craven Basin. This is because incision probably evolved with time, and the final position of the incision surface most likely corresponds to some cryptic level within the upper parts of the turbidite slope succession (Fig. 6A). Therefore, the downlap surface below the turbidite deposits in the Craven Basin is more likely to record initiation of sea-level fall rather than cor relate directly with the final position of the lowstand erosion surface below the Upper Howgate Edge Grit on the Askrigg Block. The stratigraphical level corresponding to the ultimate lowstand of relative sea-level is considered as the boundary between a lower and upper part of the lowstand systems tract (Fig. 6A). Architecturally, the unit below this surface is a wedge in the Craven Basin, and the correlative of this lowstand wedge on the Askrigg Block is the lowstand erosion surface. When relative sea-level started to rise, and the incised valleys started to fill, turbidite sedimentation
314
O.J. Martinsen, J.D. Collinson and B . K. Holdsworth
NW Askrigg Block Lad Gill (SE 888046)
Sea-level curves (Lad Gill section)
Ca. 50km
-
3 -
j, 1 1
NW Craven Basin Kettlesbeck (SD 7 43632)
Wave-influenced shallow-water deltaic sediments Braided, fluvial channel complexes and overbank fines
-- -
IFS
Upper LST
---
E
"' c. �
Q)
-
�
g_
Lower LST
R
F Relative
c"O 0 Q) ·- "' � 0 "' c.
EX �
F
Q)
��
Eustatic
.,-o -"' co.o "0-
Distal part of highstand delta complex
E
:J .0 Q) 0 a:
Lower sequence boundary (Cravenoceras cowfingense marine band)
ms
A
Sea-level curves Relative
NW Craven Basin +----- 4okm Kettlesbeck (SD 743632)
-----•
NE Craven Basin Cononley Beck (SD 986468)
Eustatic River mouth bar
E
"' c.
Upper sequence boundary and superimposed initial flooding surface
�
Q)
� .2-
--o c Q) 0 "' 0 �c.
·-
co X E., 0�
u..-o Q) "' -.0 "' "
E
:J .0 Q) 0 a:
Upper LST --Lowstand, turbiditic slope wedge
--
Lower LST
Cravenoceras cowfingense
--
10
-
_
�
Edge Marine Band (E2a (ii))
0
m
m
fs
Lower sequence boundary (base E2a(i)) and conformable lowstand surface
marine band (E2a(i)) F ..
R ..
F ..
R ..
B
315
Millstone Grit cyclicity
Sea-level curves Relative F
R
,._____..
Eustatic F
SE Askrigg Block +----- 25km Backstean Gill (SE 058723)
.,.
____
NE Craven Basin Cononley Beck (SD 986468)
R
,._____..
40 Upper sequence boundary
Lower sequence boundary
c
Cravenoceras cowlingense marine band
All surfaces merge into condensed section in basin
(base E2a(i) cycle) on Askrigg Block
Fig. 6. (Opposite. ) (A) Stratigraphical correlation and proposed sea-level curves for the cross-section of the E2a ( i ) sequence from the northwestern part of the Askrigg Block into the northwestern part of the Craven Basin. (Opposite.) (B) Like (A), but cross-section from the northwestern part of the Craven Basin to the northeastern part of the basin. (Above. ) (C) Like (A) , but correlation from the southeastern part of the A skrigg Block into the northeastern part of the Craven Basin. For section locations, see Fig. 1.
MFP, Mirk Fell palaeosol ( Elc) ; UHEG, Upper Howgate Edge Grit; THG, Tan Hill Grit, LG Lst, Lad Gill Limestone; MFS, maximum fl ooding surface; LES, lowstand erosion surface; IFS, initial flooding surface; SB, sequence boundary. Systems tracts; HST, highstand; LST, lowstand; TST, transgressive.
may have gone on for some time. The time for cessation of turbidite sedimentation within the E2a(i) cycle may have been a result of elimination of the differential bathymetry (and accommodation space) at the southern margin of the Askrigg Block. How ever, during times of rising sea-level, rivers are more likely to aggrade and trap sediment, particularly in their downstream reaches (Schumm et al., 1987; Schumm, 1993), which in turn may reduce the poten tial for voluminous turbidity current generation. Several factors complicate the correlation of major stratigraphical surfaces from the basinal part of a succession upstream to a significant distance into fluvial systems. Firstly, there is likely to be a lag in the time the relative sea-level behaviour is recorded in upstream reaches. For instance, it is known from field data as well as flume experiments that upstream reaches may incise while downstream reaches aggrade (Shepherd, 1979; Schumm, 1993). This suggests that sequence boundaries commonly may have very complex histories and are not strictly correlative over large distances in upstream downstream profiles.
Secondly, the effects of a particular sea-level change may reach only a surprisingly short dis tance upstream from the coast. Saucier (1981; cf. review in Schumm, 1993) suggested that the effects of the approximately 120-m sea-level lowering during the late Wisconsin lowstand extended only 370 km upstream in the Mississippi river valley, in contrast to the 1530 km estimated by Fisk (1944). Also Blum (1994) showed similar ·effects for the Colorado River of Texas. Both points are important for the E2a(i) cyclo them. If there was a lag time between the initial sea level rise and the fluvial response by aggradation, the fluvial deposits on the northwestern part of the Askrigg Block may not be coeval, but perhaps be later than those in the basin or at the edge of the Askrigg Block, 60 km to the south. This is similar to back-filling during relative sea-level rise, where deposition starts at sea-level and with time progresses back up the alluvial tract. Such a relationship cannot be verified from the existing data set. An even more refined chronostratigraphy is needed to work out such details.
316
O.J. Martinsen, 1. D. Collinson and B. K. Holdsworth
A
B
Condensation in NE part of basin (lateral setting, no supply from block)
c
D
Bay deltas on SE block, condensation in NW part of basin
Distributary channels in basin, braided fluvial style on block leaving sheet sandstone (Tan Hill Grit)
� t N
I
Rapid coastal transgression, but minor, shallow-water deltas briefly interrupted retrogradation Salinity increase and faunal reintroduction
Start of prolonged period of slow deposition and condensation in basin and on the block
..,.
..a �
9
Fig. 7. Palaeogeographical summary diagram of the E2a(i) cyclothem related to different stages of sea-level. A, Eustatic maximum; B, lowstand; C, rising sea-level; D, rapidly rising sea-level.
It is fairly certain that the effects of the E2a(i) sea level fall extended as far upstream as the north western part of the Askrigg Block (provided the equivalent shoreline during the sea-level minimum lay at the southern end of the Askrigg Block), because the facies interpretation clearly indicates that a sea-level fall and basinward shift of facies occurred (Fig. 6A; Martinsen, 1993). On the southeastern part of the Askrigg Block, the abrupt transition from offshore mudstones to burrowed and rooted, thin crevasse splay deposits probably records the lowstand phase (Figs 6C & 7). Overbank deposition was probably limited during the early parts of the lowstand period when most sediment was transferred directly to the block-edge. During later aggradation, more sediment may have been supplied to floodplain or interchannel areas, but the proposed tectonic tilting of the Askrigg Block to the west would have prevented substantial volumes from reaching the southeastern parts of the block (Fig. 7). On the Askrigg Block, the incised Upper Howgate
Edge Grit lies directly on offshore mudstones. The sheet sandstone occurs above (the Tan Hill Grit), within the part of the cyclothem where relative sea level probably was rising steadily and significant: alluvial aggradation occurred (Fig. 6A). During very rapid rates of relative sea-level fall, the only possi bility for the fluvial system is to incise, particularly when the substrate is fine-grained sediments with pronounced cohesive stability and little possibility to compensate by channel-belt widening and increased bed roughness (Schumm, 1993). The resulting fluvial morphology of the Upper Howgate Edge Grit is therefore likely to be different from the braided, amalgamated model of Shanley & McCabe (1993). This may have been the case for the Upper Howgate Edge Grit of the E2a(i) cyclothem. Only when relative sea-level had risen further, perhaps to where the channels were no longer confined by an incised valley did the braided, sheet-like morphology of the Tan Hill Grit develop (Fig. 7). The thin deltaic coarsening upward succession on top of the Tan Hill Grit in the northwestern part of
Millstone Grit cyclicity
the Askrigg Block suggests that a flooding surface occurs between the two units (Fig. 6A). This is the first indication of inundation and transgression within the E2a(i) cyclothem. It also suggests that the boundary between a lowstand and transgressive systems tract should be put at this surface (see above). Moreover, the mudstones and goniatite bearing Lad Gill Limestone on top of the deltaic succession indicate further transgression and delta abandonment. These data support a transgressive systems tract interpretation of the 8-m deltaic suc cession, perhaps including the 4-5 m of mudstones below the Lad Gill Limestone. In the basin, there is no evidence for coarser grained, transgressive sediments. Instead, the thin deltaic succession is inferred to thin basinwards and merge with the mudstones on top of the distributary channel complex (Fig. 6A). In effect, the trans gressive systems tract in the basin is represented by a condensed section, which probably merges with the condensed shales at the base of the overlying E2a(ii) cyclothem (Fig. 6A). In the southeastern parts of the block and in the eastern parts of the Craven Basin, there is little evidence for a transgressive phase (Fig. 6C). Only, the intense burrowing on top of the crevasse sand stones indicate delta abandonment and trans gression. The burrowed level is sharply overlain by dark, offshore mudstones (Fig. 6C). The E2a(ii) marine band, the E. bisulcatum ferrimontanum band caps the E2a(i) cyclothem. Like its E2a(i) counterpart, the C. co wlingense band, it is interpreted to represent the peak of the eustatic curve. Dark mudstones underlie the marine band. It is considered likely that maximum transgression occurred at some point during deposition of these barren mudstones. The E2c2 cycl othem Boundaries and thickness
One of three upper Arnsbergian cyclothems with at the base (i.e. E2c2(i, ii or iu); cf. Fig. 2) is described as a counterpart to the E2a(i) cyclothem. There is uncertainty concerning the exact age of both the interpreted basal marine band and the cyclothem itself, because all the three nuculum bands are not developed everywhere. On the Askrigg Block, no N. nuculum bands are found , probably as a result of occlusion or erosion (see below). In the northeastern part of the Craven
Nuculoceras nuculum
317
Basin, only one or two bands are developed at any one location (e.g. Wilson, 1977), leaving a great deal of ambiguity as to which exact cyclothem the described rocks belong. For the present purpose, it is described as 'the E2c2 cyclothem'. Based on facies analysis (Martinsen, 1990), map ping (Stephens et al. , 1953), and local stratigraphical studies (Wilson, 1960, 1977), there is a high degree of confidence in correlating the described cyclothem between the southeastern part of the Askrigg Block and the northeastern parts of the Craven Basin, over a N-S distance of approximately 25 km. Therefore, the exact age and correlation of the described cyclo them with the minor cycle chart (Fig. 2), is of no importance for the present purpose. The E2c2 cyclothem varies considerably in thick ness. It is 5 -36 m in the southeastern corner of the block, around 34 m in the northeastern part of the Craven Basin, only 12 m in the north-central parts of the basin, and around 40 m thick in the northwestern parts of the basin (Fig. 8). On the Askrigg Block, the cyclothem base is interpreted to be the erosive base of the Lower Follifoot Grit, whereas in the basin, the base of the cyclothem is one of the three N. nuculum bands (Fig. 8). Martinsen (1993) interpreted the underlying E2cl cyclothem to be entirely eroded on the block by the Lower Follifoot Grit, which also erodes into the E2b3 cyclothem (Fig. 8). The cyclothem top is interpreted to coincide with the base of a multi cyclothem condensed interval, which encompasses at least nine minor cycles (E2c2(i or ii)- H1b(iii)), and which can be traced across the entire Craven Askrigg region (Fig. 8; ct. Martinsen, 1990). Depositional model and tectonism
In the southeast part of the Askrigg Block, the Lower Follifoot Grit is a trough cross-stratified sandstone, which has been interpreted to represent deposition in low-sinuosity, braided river channels (Martinsen, 1993; Figs 8 & 9). Geometrically, it has two modes of occurrence. For the most part, it is a tabular sheet sandstone recognizable over several tens of square kilometres (Fig. 8; Wilson, 1960). However, locally, two well-defined storeys of sand stone are observed. The lower one is very lenticular, forming ribbon-shaped bodies up to 15 m thick and hundreds of metres wide (Fig. 8). These sandstone bodies are clearly incised into the underlying depo sits. The upper storey is much more extensive lateraLly, forming a pronounced sheet. The Lower
318
0.1. Martinsen, J.D. Collinson and B . K. Holdsworth
SE Askrigg Block ---- 20-25krn ---+--• NE Craven Basin S
N Great Whernside West Scrafton Beldin G i l l S l i pstone Crag Thrope Edge SE 0074 SE 082824 SE 095799 SE 1 38821 SE 1 06752
Sea-level curves Relative
Eustatic
!Horn Crag Quarry, Silsden F R F R SE 053480 ---- ----
Scar House Beds (E2b3) 35m Mu lti-cycle condensed section (contains maxi m u m flooding s u rface and top seq u e n ce boundary of this cyclothem)
Lowstan d erosion su rface (coincides with sequence boundary)
Highstand systems tract
0
Fig, 8, Cross-section showing stratigraphical correlation and proposed sea-level curves of the studied E2c2 cyclothem from the southeastern part of the Askrigg Block to the northeastern part of the Craven Basin. See Fig. 1 for section location.
FSST, falling stage systems tract; LST, lowstand systems tract; MFS, maximum flooding surface; LES, lowstand erosion surface; FRS, forced regression surface.
Follifoot Grit is abruptly overlain by a multi cyclothem condensed section, which contains gonia tites of the lsohomoceras subglobosuml Homoceras beyrichianum marine bands of H1a- H1b age (Fig. 2; Wilson, 1960). In the basinal section above Silsden in Airedale (Fig. 8), to the south, offshore mudstones occur directly above the Nuculoceras nuculum marine band at the base of the cyclothem (Stephens et al. , 1953). These mudstones are only a few metres thick, and are then abruptly overlain by a 26-m-thick, entirely sandy succession (Fig. 8). This sandstone succession has been interpreted as friction-dominated and wave-influenced delta-mouth-bar deposits (Fig. 9; Martinsen, 1990). The mouth bar sandstone is growth-faulted to the depth of the base of the sand stone, suggesting that the fault soles out within the underlying mudstones. The mouth bar sands are overlain by an erosively based, very coarse-grained pebbly sandstone, which is entirely trough cross stratified (Fig. 8). The sandstone is interpreted as a braided fluvial complex, and is eventually overlain by fossiliferous mudstones of the overlying multi cyclothem condensed section (cf. Stephens et al. , 1953; Wilson, 1977).
E2c2 cyclothem architecture and sequence stratigraphy
The Nuculoceras nuculum goniatite band probably represents maximum eustatic sea-level (discussion above). Consequently, the dark mudstones that overlie the goniatite band in the basin are interpreted to represent late highstand deposits (Fig. 8), laid down when the nearest shoreline was still far to the north. In the basin , the abrupt contact between mud stones and the delta sandstones, and the close juxtaposition of the goniatite band and the delta sandstones (only separated by a few metres; Fig. 8), suggest that a relative drop of sea-level took place, represented by the lower boundary of the sandstone. One interpretation of the lower boundary is as a lowstand surface of erosion or Exxon-type sequence boundary formed during the most rapid rate of sea level fall. The delta sandstones above the lowstand surface would further be interpreted as lowstand sediments, suggesting that the entire delta is a lowstand delta. An alternative interpretation is that the sharp transition from the mudstone to the deltaic sandstone was produced by accretionary forced regression during relative fall of sea-level, making the delta
Millstone Grit cyclicity
A
319
B
Falling-stage (forced regression) delta with growth faults c N
N
r
coastal transgression
Start of prolonged period of slow deposition and condensation in basin and on the block
Salinity increase and faunal reintroduction
Fig. 9. Palaeogeographical summary diagram of the studied E2c2 cyclothem related to different stages of sea-level .
A, E ustatic maximum; B, falling-stage; C, lowest/ slowly rising sea-level; D , rapidly rising sea-level.
sandstones part of a falling stage systems tract (Fig. 8; e.g. Plint, 1988; Hunt & Tucker, 1992; Posamentier et al. , 1992). This interpretation is supported by the fact that the braided, fluvial sand stone overlying the deltaic rocks is so much coarser than the deltaic sandstones, shows entirely different palaeocurrents and therefore does not seem to be related to the deltaic sandstones (i.e. it is not their distributary channel). The erosion surface below the fluvial sandstones could represent the relative sea level lowstand and thus equate with an Exxon-type sequence boundary (Fig. 8). On the Askrigg Block, the erosion surface below the lenticular channels of the Lower Follifoot Grit probably marks an extensive erosion surface related to falling and low relative sea-level (Fig. 8). Martinsen (1993) suggested that extensive incision took place on the block at this time, and the under lying E2c l cyclothem was largely eroded. The incision phase may correlate with the deposition of
the inferred falling-stage delta in the basin (Fig. 8), but perhaps more likely, it represents the incision related to relative lowstand and correlates with the erosion surface between the deltaic sandstones and the very coarse fluvial sandstones in the basin. This interpretation implies that erosion below the Lower Follifoot Grit on the Askrigg Block involved both initial formation during the falling stage and further development and modification during the relative lowstand. If this interpretation is correct, the fluvial sand stones on the block and in the basin were deposited during a relative rise of sea-level (Fig. 8). Questions as to whether the shoreline prograded during filling of the incised fluvial channels, or if enough sediments were trapped on the alluvial plain to enhance trans gression at the contemporaneous shoreline cannot be answered due to inadequate exposure. In the latter case, the fluvial interval should be termed part of a transgressive systems tract. However, the lack
320
O.J. Martinsen, J. D. Collinson and B. K. Holdsworth
of evidence for transgressive deposits, and analogy with the E2a(i) cyclothem suggests that it is more likely that progradation continued up until or very close to cyclothem abandonment. In that case, the fluvial rocks should be termed part of the lowstand systems tract (Fig. 8). It is likely that sea-level fall across a slowly sub siding area like the Askrigg Block would enhance incision and stabilize single channels or 'valleys', particularly if incision was rapid and occurred in a fine-grained substratum (see Schumm, 1993). Only during sea-level rise and consequent increasing accommodation would the incised channels be 'overflooded' and fluvial deposition take place out side the incised channels. If grain size, discharge and gradient were favourable for a braided fluvial style to develop at this time, the resulting fluvial sediments would probably be sheet-like. Provided the fluvial rocks of the E2c2 cyclothem form part of a lowstand systems tract, no trans gressive systems tract seems to have developed above, because there is an abrupt transition to marine mudstones with goniatites, both in the Craven Basin and on the Askrigg Block (cf. Wilson , 1960, 1977). This transition suggests a rapid rise of relative sea-level, flooding and cyclothem abandon ment with no time for development of a 'late' progradational interval, like that of the E2a(i) cyclo them, at least in the study area. Therefore, in summary, within the E2c2 cyclo them, falling-stage sediments are interpreted to predominate in the basin , whereas on the block, lowstand systems tract sediments dominate above a complex lower erosional surface. Highstand sediments were probably eroded on the block, whereas in the basin they are restricted to a thin interval immediately overlying the Nuculoceras nuculum marine band (Fig. 8). Transgressive sys tems tract sediments are condensed, and probably occur as a thin zone between the fluvial sediments and the overlying multi-cyclothem, goniatite-bearing condensed section (Fig. 8).
post-extensional subsidence but also of decreasing overall subsidence (cf. Leeder & McMahon , 1988). One similarity between the two cyclothems is the proposed prolonged absence of clastic supply from the time of maximum transgression (thin shales below marine bands), until maximum eustatic highstand (marine bands; see Figs 6 & 8). This encompasses one-quarter of a eustatic curve, and, depending on the rate of subsidence, probably corresponds to the middle part of the rising, extended limb on a relative sea-level curve. Therefore, it seems that clastic supply was absent for a prolonged period compared with conventional sequence strati graphical models (cf. Van Wagoner et al. , 1990). The decreasing overall subsidence is important for explaining why late highstand deposits are not pre served on the Askrigg Block in the E2c2 cyclothem in contrast to the E2a(i) cyclothem within which highstand deposits are relatively thick. A lower rate of subsidence at this time probably enhanced and amplified the fall of relative sea-level on the Askrigg Block causing deeper incision and erosion. In addition, a lower subsidence rate during the relative highstand generated less new space for sediment accumulation, which gave less accommodation and a thinner highstand succession to start with. Another contrast is the interpreted development of a lowstand, turbiditic wedge during E2a(i) time whereas a falling stage, forced regression delta evolved during E2c2 time. These depositional patterns follow closely the shelf-edge and ramp models of van Wagoner et al. (1988). However, in addition , in the higher subsidence situation of the E2a(i), the response time of the sedimentary system to falling relative sea-level was longer. This longer response time may have slowed regression during E2a(i) time. The opposite was probably the case for the E2c2 delta. A similar conclusion was reached by Holdsworth & Collinson (1988, p. 138), where the response time of sand deposition to lowered sea level was related to bathymetry, the latter being a combined effect of compaction of previously deposited sediments, eustasy and subsidence.
Sequence model Choice of sequence stratigraphical model Introduction
The E2a(i) and E2c2 cyclothems illustrate different stratigraphical architectures as responses to con tJ·asting physiographic relationships between the Askrigg Block and the Craven Basin. The contrast evolved as a result not only of decreasing differential,
Local sequence stratigraphical models should fulfil the same requirements as facies models (Reading, 1986; Walker, 1992). Thus, a sequence strati-· graphical model for the E2a(i) and E2c2 cyclothems will have to serve several purposes: (i) act as a norm so that the model can be compared with sequence
Millstone Grit cyclicity
models from other areas; (ii) the model should be rigorous enough to be a framework for future studies; (iii) it should aid prediction of sequence stratigraphical architecture in adjacent areas; and (iv) it should be a basis for interpretation of Namurian sequence stratigraphy. The models described here are considered to fulfil these requirements, and they are modified versions of the general models of Galloway (1989) and Exxon (e.g. Jervey, 1988). Several local factors prevent the general models from being applied as templates in the present situation, and modified versions are required. First, erosion surfaces at the bases of the incised fluvial complexes are laterally discontinuous and cannot be correlated with confidence between individual outcrops (e.g. through palaeosols), nor in a proximal- distal direction (cf. Martinsen, 1993). Secondly, goniatite-bearing marine bands are the most laterally continuous stratigraphical horizons (see above), and are interpreted to be the levels across which the most important palaeogeographical reorganizations and stratigraphical breaks occur (cf. Figs 7 & 9). The erosion surfaces are recognized as lowstand surfaces of erosion and compare with Exxon-type sequence boundaries, but they are not considered as the most prominent breaks in the succession, nor are they laterally continuous enough to be reliable for correlation (Martinsen, 1993). For example, the lowstand erosion surfaces are not re cognizable in the Craven Basin (Figs 6A-C & 8). Thus, the goniatite bands are the most sensible stratigraphical levels to pick as sequence boundaries. Thirdly, since lowstand wedges of sediment occur both in the E2a(i) and E2c2 cyclothem, the Galloway (1989) model cannot be applied in a straightforward way because such units are not emphasized in that model. However, Xue & Galloway (1993) described low stand prograding complexes from both shelf-edge margins and ramp margins in a study from the Songliao Basin in China, thus extending the original Galloway (1989) model. The Xue & Galloway (1993) model is therefore a reasonable comparison for the models presented herein for the E2a(i) and E2c2 cyclothems, with the added complexity of a forced regressive unit/falling stage systems tract for the E2c2 cyclothem. Cyclothem asymmetry, glacio-eustasy and spatial development of systems tracts
In the E2a(i) and the E2c2 cyclothems, as well as in
321
most other Millstone Grit cyclothems deposited near the basin margin (Wright et al. , 1927; Stephens et al. , 1953; Wilson, 1960; Collinson, 1988; Holdsworth & Collinson, 1988; Martinsen, 1993), there is a pronounced lithological asymmetry (see Figs 6 & 8; cf. also modelling by Collier et al. , 1990). The asymmetry developed in the thicker Millstone Grit cyclothems contrasts with the symmetry of the basinal cycles at Ashover (Rams bottom et al. , 1962; Holdsworth & Collinson, 1988; Fig. 4). The Ashover faunal cycles were probably controlled by salinity (see above; Holdsworth & Collinson, 1988). Although salinity probably fluctuated symmetrically with time (Fig. 4; discussion above), the supply of coarser sediments did not. Consequently, rather than salinity, sand deposition must have been related to other mechanisms, such as supply variations and subsidence. Therefore, relative sea-level rather than eustasy controlled the succession in the Askrigg Block and Craven Basin area. Because of the added influence from subsidence, relative sea-level curves are characterized by longer periods of rise (e.g. Fig. 6A). The sedimentary expression and character of the succession at Ashover would, therefore, be expected to be different from the Askrigg Block to Craven Basin succession, even if sediment input was similar. Given the inherent nature of a glacial world, with major glaciations in Gondwana (Caputo & Crowell, 1985 ; Veevers & Powell, 1987), glacigenic sea-level falls tend to be much slower than corresponding sea level rises due to slow ice-cap build-ups and rapid melt-downs (e.g. Williams, 1988). As pointed out by Read (1991), although periods of eustatic sea-level fall were relatively short, the rate at which sea-level fell must have been high enough to overcome sub sidence rate and produce a relative sea-level fall. During these periods, sandy systems filled the basins. The rate of relative sea-level rise must have been extremely rapid, particularly in basinal areas, where it was amplified by high subsidence rates. Consequently, in low-lying coastal plains, which probably formed an important part of the Namurian systems (cf. Holdsworth & Colbnson, 1988), the rate of coastal transgression must have been high, with little possibility for the sedimentary systems to leave significant sandy deposits behind. Also, the generally observed lack of significant wave or tidal action decreased the possibility for significant reworking of the previously deposited sediments. These relationships suggest that the asymmetric lithological response to relative sea-level change was
322
0 .J. Martinsen, J. D. Collinson and B. K. Holdsworth
inversely related to the time needed to deposit the facies. The thick, coarsening upward successions in the basins were probably deposited in a relatively short time, whereas the thin condensed mudstones and marine bands record most of the time rep resented by each Millstone Grit cyclothem. This asymmetry contributed to the areal distri bution of systems tracts for both the 'block-edge' and ramp situation. The rapid transgression resulted in a very thin transgressive systems tract, with, at the most, only a thin shallow-water delta on top as a last sedimentary response to the accelerating sea-level rise. The transgressive systems tract developed a recognizable thickness only on the Askrigg Block. The highstand systems tract of the two cyclothems, which is mudstone-dominated in the study area, is probably better developed to the north (discuss ion above), closer to the palaeoshoreline. The ex pectedly slower rate of glacio-eustatic sea-level fall resulted in a slower rate of relative sea-level fall, and a well-developed lowstand systems tract in the Craven Basin. However, even in an ideal situation, with no along-strike displacement and variability of systems tracts, no one vertical section is expected to show a complete development of a marine-band bounded sequence. The systems tracts are displaced either down-dip or up-dip with relation to each other (Figs 6A & 8), and their complete presence in any one vertical section is unlikely. With the added complexity of a strike-variable development of systems tracts (see Fig. 6A -C; compare cross sections), a common situation along any shoreline (Martinsen, 1994), it is self-evident that only very few two-dimensional cross-sections would show a complete development of systems tracts.
MULTI-CYCLOTHEM CONDENSED SECTIONS AND MAJOR SCALE CYCLICITY
Background
Although the basic cyclothem pattern ('minor cycles' of Holdsworth & Collinson, 1988) of the Millstone Grit has been established since Wright et a!. 's (1927) classic Rossendale Anticline Memoir, the issue of a cyclicity at a larger scale has been more actively debated recently (cf. also Read (1991) for a dis cussion of cyclicity at several orders in the Namurian of the Midland Valley of Scotland). In 1977, Ramsbottom argued that Millstone Grit cyclothems could be grouped into larger scale cycles, termed
'mesothems'. The mesothem theory has been subject to severe criticism and its validity questioned (see Holdsworth & Collinson (1988) and Martinsen (1990) for a thorough discussion). It is proposed that major cyclicity exists, but in a different form than mesothems. Nam urian maj or-scale cyclicity
Within the Namurian E1c- H2c succession of the Craven-Askrigg region, three regionally correlative stratigraphical intervals can be defined. In each are several minor cycles, which are condensed, represent non-deposition and/or are entirely fine-grained (Fig. 10). Interval 1 extends from cyclothem E2b1(i) to cycle E2b2(ii) and contains five minor cycles, interval 2 extends from cyclothem E2c2(ii) to cyclothem H1b(iii) and contains nine minor cycles, and interval 3 extends from cyclothem H2a(i) into the Kinderscoutian (R1) and contains five minor cycles to the base of the Kinderscoutian. Common to these three, multi-cyclothem con densed intervals, is that they are sand-free and only small thicknesses of fine-grained sediments were deposited. Only locally within one section can evi dence for subaerial exposure be documented, and erosion has not been proven. Therefore, these fine grained intervals are mainly of subaqueous origin. There are ambiguities to some of the correlations, particularly on the Askrigg Block where few goniatite bands occur. Nevertheless, the general framework seems to hold, not least because the intervening cycles are dominated by sandstones, which them selves can be correlated regionally both because of adjacent goniatite faunas, classic mapping and their distinctive depositional environments (cf. Martinsen, 1990). Based on these multi-cyclothem condensed mud stone intervals, it is suggested that three major cycles of deposition can be defined within the E1c H2c succession (Fig. 10). Major cyclothem 1 (MC1) comprises the E1c- E2a(iii) minor cycles, bounded below by the sub-E1c unconformity on the Askrigg Block (ct. Dunham & Wilson, 1985) and the base of the Cravenoceras malhamense goniatite band in the basin, and above by the base of the E2b1(i) Cravenoceras subplicatum group marine band. This band forms the base of the lowest multi-cyclothem mudstone interval. All the cycles in MC1 contain sandstones. Major cyclothem 2 (MC2) encompasses minor cycles E2b1(i)- E2c2, and is bounded below by the base of the Cravenoceras subplicatum group
Millstone Grit cyclicity
323 H l bliv)
SE Askrigg Block .----- 1 5-20km --------. NE Craven Basin
Base E2c2(i?) unconformity MC2
Base E 1 c unconformity
\
E1c
Craven Fault System
� t
\ 1 50
\
MC1
om
• Multi-cycle mudstones III:t!Il Dominantly sandstones D Other mudstones and siltstones Minor cycles
Major cycles
Fig. 10. Proposed division of the Namurian Elc-H2b stratigraphy into major cyclothems. See text for discussion.
marine band, and above by the base of the second Nuculoceras nuculum marine band at the base of the E2c2(ii) cyclothem, which also forms the base of the second multi-cyclothem condensed interval. Whereas the lower five cycles in MC2 are sandstone free, the upper three contain sandstones (see below). Major cyclothem 3 (MC3) contains ten minor cycles comprising E2c2(ii) to H1b(iv). The lower boundary is the base of the second Nuculoceras nuculum marine band at the base of the E2c2(ii) cyclothem, whereas the top boundary is defined at the base of the Homoceras smithi/ Hudsonoceras proteus goniatite band at the base of the third multi cyclothem condensed interval. Only the uppermost cyclothem, H1b(iv), contains sandstones, whereas the underlying nine make up the second multi cyclothem condensed interval.
Maj or c ycl othem architecture and genesis
In only one major cyclothem (MC2) is there indi cation of a systematic organization of the contained minor cycles. The E2b3, E2c1 and the one sandy E2c2 cyclothem show an increasingly progradational character (cf. Martinsen, 1993). In addition, Martinsen (1993) argued that the lowstand erosion surface below the E2c2 Lower Follifoot Grit on the Askrigg Block was, in addition to the unconformity below the E1c Grassington Grit, the most prominent within the E1- H2 part of the Millstone Grit in this area. Therefore, it is probable that the sequential evolution of the MC2 cycles reflects progressively lower sea-level with time. Such a pattern is compar able to a progradational sequence-set architecture, as described by Mitchum & Van Wagoner (1990), and may reflect longer term sea-level changes than
324
0.1. Martinsen, J.D. Collinson and B . K. Holdsworth
those at minor cycle scale. However, none of the other proposed major cycles show a systematic stacking pattern. Although there is no direct reason to argue against sea-level changes as the main control on the major cycles, in most of the multi-cyclothem condensed sections, the basic cyclicity is still present, with faunal bands being interbedded with barren mud stones (e.g. Wilson & Thompson, 1959; Arthurton et al. , 1988). This suggests that lowstands were still reducing salinities below the critical marine-band threshold. In addition , there is evidence for subaerial exposure in the E2b1(i)- E2b2(ii) condensed section in the Colsterdale Marine Beds on the eastern side of the Askrigg Block (Wilson & Thompson, 1959). Therefore, it appears that there are no major changes to the sedimentation and cyclicity patterns beside the decreased thickness and the lack of sand. Possibly, one would expect a change of the cyclicity patterns, particularly in the condensed sections (the assumption then being that these would represent a particularly high sea-level), if major-scale sea-level changes caused the major cyclicity. Major changes of sediment supply are more likely to have caused the major cyclicity. Within the E2c2(ii)- H1b(iii) cyclothems, which form a con densed section, particularly in the eastern part of the Craven-Askrigg area, the mudstones thicken dramatically towards the west, include turbidite sandstones and reach 165 m in thickness (Wilson et al. , 1989). This relationship suggests that a major switch of sediment supply to the west occurred within the E2c2(ii)- H1b(iii) time interval, and probably explains the condensation in the Craven -Askrigg area. It is suggested also that the other multi cyclothem condensed sections were formed by large scale switches of sediment supply. To the east, over the southern North Sea, large Namurian depocentres are known to have been present. However, the stratigraphical subdivision in this area is yet not refined enough to test the hypothesis that those areas were alternative sites of sediment accumu lation . Therefore, a sequence stratigraphical scheme at major cyclothem scale is not considered to be possible at present.
CONCLUSIONS 1 To illustrate the lateral and temporal develop ment of Namurian Millstone Grit cyclothems , the Arnsbergian E2a(i) and E2c2 cyclothems in the
Askrigg Block and Craven Basin area have been re-evaluated in terms of architectural and strati graphical response to relative sea-level change. 2 Goniatite-bearing marine bands are thought to represent the highest eustatic sea-level evidenced by salinity data. This interpretation contrasts with the conventional sequence stratigraphical models where faunal concentrated bands are considered to represent maximum transgression and flooding. Instead, thin , condensed shale below the marine bands probably represent maximum rate of rise of sea-level. 3 The E2a(i) cyclothem developed when there was a significant bathymetric contrast between the Askrigg Block and the Craven Basin. Prolonged condensation took place, and clastic deposition did not occur until the late part of the highstand of sea level. Subsequent lowstand sediments are developed both on the Askrigg Block as fluvial systems and in the basin as a turbiditic slope apron. The trans gressive systems tract is very thin due to rapid trans gression and condensation in the basin. 4 The E2c2 cyclothem developed at a time when there was a gently inclined ramp basinward of the Askrigg Block. Prolonged condensation and/or non deposition took place until sea-level started to fall and a forced regression delta formed. This delta was incised during the subsequent sea-level lowstand , and fluvial deposition took place o n the Askrigg Block. No underlying highstand or overlying trans gressive systems tract sediments are preserved or were deposited. 5 Major cycles of deposition are interpreted to exist in the Namurian of the Askrigg Block and Craven Basin area. The cycles are bounded by condensed, multi-cyclothem mudstone levels. Except in one case, the major cycles show no systematic architec ture, suggesting they were not related to major changes of relative sea-level. Instead, they prob ably represent major avulsive shifts of sediment depocentres.
ACKNOWLEDGEMENTS
We wish to thank numerous colleagues who have inspired and guided our thinking about the Millstone Grit both in the UK and in Ireland, particularly Harold Reading, Trevor Elliott, Roger Walker and Colin Jones. Ole Martinsen wishes to thank Total Norway for sponsoring 3 years of doctoral research, Norsk Hydro for sponsorship of 3 years of post-
Millstone Grit cyclicity doctoral
research ,
and
Frank
Heggernes
and
Karianne Krag Kjos for providing new ideas about the northwest quarter of the Askrigg Block. The referees Charles Bristow, Robert
L.
Gawthorpe and
Peter J. McCabe are thanked for reviewing the manuscript.
REFERENCES AITKENHEAD, N. ( 1977) The Institute of Geol ogical Sciences borehole at Duffield, Derbyshire. Bull. geol. Surv. G. B. , 59, 1-38. ANDERTON , R., BRlDGES, P.H., LEEDER, M.R. & SELLWOOD, B.W. ( 1979) A Dynamic Stratigraphy of the British Isles: a Study in Crustal Evolution . George Allen & Unwin, London, 301 pp. ARTHURTON, R.S . , JoHNSON, E.W. & MuNDY, D.J.C. ( 1988) Geology of the country around Settle. Mem. geol. Surv. G.B., Eng. Wales, Explan. Sheet, 60, 147 pp. BAINES, J.G. ( 1977) The stratigraphy and sedimentology of the Skipton Moor Grits (Namurian Elc) and their lateral equivalents. Unpubli shed PhD thesi s, Keele University. BATES, R.L. & JACKSON, J . A. (1987) Glossary of Geology, 3rd edn. American Geological Institute, Alexandria, Va, 788 pp. BISAT, W.S. ( 19 14) The Millstone Grit Series between Ma sham and Great Whernside. Proc. Yorks. geol. Soc. , 19, 20- 24. BISAT, W.S. ( 1924) The Carboniferous goniatites of the north of England and their zones. Proc. Yorks. geol. Soc. , 20, 40- 124. BLUM, M.D. ( 1994) Genesis and architecture of alluvial stratigraphic sequences: a late Quaternary example from the Colorado River, Gulf Coastal Plain of Texas. In Siliciclastic Sequence Stratigraphy: Recent Develop ments and Applications (Eds Weimer, P. & Posamentier, H.W.) Mem. Am. Assoc. Petrol. Geol., 58, 259-283. CAPUTO, M.V. & CROWELL, J.C. ( 1985) Migration of glacial centers across Gondwana during the Paleozoic Era. Geol. Soc. Am. Bull. , 96, 1020 - 1036. COLLIER, R.E. LL., LEEDER, M . R. & MAYNARD, J . R. (1990) Transgressions and regressions: a model for the influence of tectonic subsidence, deposition and eu stasy, with application to Quaternary and Carboni ferous examples. Geol. Mag. , 127, 117-128. CoLLINSON, J.D. ( 1969) The sedimentology of the Grindslow Shales and the Kinderscout Grit: a del taic complex in the Namurian of northern England. J. sediment. Petrol. , 39, 194-221. COLLINSON , J . D. ( 1988) Controls on Namurian sedimen tation in the Central Province basins of northern England. In: Sedimentation in a Synorogenic Basin Complex: the Upper Carboniferous of Northwest Europe (Eds Besly, B.M. & Kelling, G.), pp. 85 -101. Blackie, Glasgow. COLLINSON, J.D., HOLDSWORTH, B.K., J oNES, C.M. & MARTINSEN, O.J. (1992) Discussion of 'The Millstone Grit (Namurian) of the southern Pennines viewed in the light of e ustatically controlled sequence stratigraphy by
325
W . A. Read. Geol. J. , 27, 173 - 180. CRoss, T.H. & GARDNER, M . H . ( 1991) Base level, accommodation space and high-resolution sequence stratigraphy of the Ferron Sandstone (Cretaceou s). Utah. In: Proceedings of the NUNA Conference on High-Resolution Sequence Stratigraphy (Eds Leckie, D . A . , Posamentier, H.W. & Lovell, R.V.V.), pp. lO l l . Geological Association of Canada, St J ohn's. DuNHAM, K . C. & WILSON , A . A . ( 1985) Geology of the Northern Pennine Orefield : Volume 2, Stainrnore to Craven. Econ. Mem. geol. Surv. G.B., Eng. Wales, 247 pp. EMBRY, A.F. & PODRUSKl, J.A. (1988) Third-order depo sitional sequences of the Mesozoic succession of Sverdrup Basin. In: Sequences, Stratigraphy, Sedimentology: Surface and Subsurface (Eds James, D . P. & Leckie, D.A . ) , Mem. Can. Soc. Petrol . Geol . , Calgary, 15, 73 -84. FISK, H.N. ( 1944) Geological Investigation of the Alluvial Valley of the Lower Mississippi River. Missi ssippi River Commission, Vicksburg, MS, 78 pp. GALLOWAY, W.E. (1989) Genetic stratigraphic sequences in basin analysis. I : architecture and genesis of flooding surface bounded depositional units. Bull. Am. Assoc. petrol. Geol. , 73, 125 - 142. GAWTHORPE, R.L., GuTIERIDGE, P. & LEEDER, M . R. (1989) Late Devonian and Dinantian basin evolu tion in northern England and North Wales. In: The Role of Tectonics in Devonian and Carboniferous Sedimentation in the British Isles (Eds Arthurton, R.S., Gutteridge, P. & Nolan, S.C.), Occas. Pubis Yorks. geol. Soc., 6, 1-23 . GuTIERlDGE, P. (1989) Control s on sedimentation in a Brigantian intrashelf basin (Derbyshire). In: The Role of Tectonics in Devonian and Carboniferous Sedimentation in the British Isles (Eds Arthurton, R.S., Gu tteridge, P. & Nolan, S.C.), Occ. Pubis Yorks. geol. Soc., 6, 171- 187. H oLDSWORTH, B.K. ( 1963) Pre-fluvial, autogeosynclinal sedimentation in the Namurian of the Sou thern Central Province. Nature (London) , 199, 133-135. HOLDSWORTH, B.K. & COLLINSON , J.D. (1988) Millstone Grit cyclicity revisited. In: Sedimentation in a Synoro genic Basin Complex: the Upper Carboniferous of Northwest Europe (Eds Besly, B.M. & Kelling, G.), pp. 132-152. Blackie, Glasgow. HuDSON, R. G.S. (1934) The Mill stone Grit succession south of Harrogate. Trans. Leeds geol. Assoc. , 5, 118-124. HuDSON, R.G.S. ( 1939) The Millstone Grit succession of the Simonseat Anticline. Proc. Yorks. geol. Soc. , 23, 3 19. HuDSON, R.G .S. (1941) The Mirk Fell Beds (Namurian E2) of Tan Hill, Yorkshire. Proc. Yorks. geol. Soc. , 24, 259-289. HuDSON, R.G.S. ( 1944) The faunal succession in the Ct. nitidus zone in the mid-Pennines. Proc. Leeds Philos. Lit. Soc. , Sci. Sect. , 4, 233-242. HuNT, D. & TucKER, M . E. ( 1992) Stranded parasequences and the forced regressive wedge systems tract: deposition during base-level fall. Sediment. Geol. , 8 1 , 1-9. JERVEY, M.T. (1988) Quanti tative geological modeling of siliciclastic rock sequences and their seismic expression. In: Sea-Level Changes: an Integrated Approach (Eds
326
0.1. Martinsen, J.D. Collinson and B. K. Holdsworth
Wilgus, C. K . , Hastings, B.S. , Kendall, C.G.StC., Posamentier, H.W. , Ross, C.A. & Van Wagoner, J.C.), Spec. Pubis Soc. econ. Paleontol . Mineral . , Tulsa, 42, 47-69. JONES, C.M. (1980) Deltaic sedimentation in the Roaches Grit and associated sediments (Namurian R2b) in the southwest Pennines. Proc. Yorks. geol. Soc. , 43, 39-67. LEEDER, M . R. (1982) Upper Palaeozoic basins of the British I sles: Caledonide inheritance versus Hercynian plate margin processe s. J. geol. Soc., London, 1 39 , 479-491. LEEDER, M . R. (1988) Recent developments in Carbonifer ous geology: a critical review with implications for the British I sles and N.W. Europe. Proc. Geo/. Assoc. , 99, 73 - 100. LEEDER, M . R. & McMA HON , A . H. (1988) Upper Carbon iferou s (Silesian) basin subsidence in northern Britain. In: Sedimentation in a Synorogenic Basin Complex: the Upper Carboniferous of Northwest Europe (Eds Besly, B.M. & Kelling, G . ) , pp. 43-52. Blackie, Glasgow. MARTINSEN, O.J. (1990) Interaction between eustasy, tectonics and sedimentation with particular reference to the Namurian Elc-H2c of the Craven-Askrigg area, northern England. Unpubl. Dr. Sci. thesis, University of Bergen, Norway, 495 pp. MARTINSEN, O.J. (1993) Namurian (late Carboniferous) depositional systems of the Craven-Askrigg area, northern England: implications for sequence strati graphic models. In: Stratigraphy and Facies Associ . ations in a Sequence Stratigraphic Framework (Eds Posamentier, H.W. , Summerhayes, C . P . , Haq, B. U . & Allen, G.P.), Spec. Pubis int. Ass. Sediment. , 18, 247-281. MARTINSEN, O.J. (1994) Sequence stratigraphy, three dimensions and philosophy. In: Sequence Stratigraphy: Advances and Applications for Exploration and Pro duction in Northwest Europe (Eds Steel, R.J. et a/.) , Spec. Pubis Nor. petrol. Soc . , 5 , 23 - 29. MAYNARD, J. R. (1992) Sequence stratigraphy of the Upper Yeadonian of northern England. Mar. petrol. Geol. , 9 , 197 - 207. MAYNARD, J . R. & LEEDER, M . R. (1992) On the periodicity and magnitude of Late Carboniferou s glacio-eustatic sea-level changes. J. geol. Soc., London, 149, 303-311 . MILLER, J. & GRAYSO N , R.F. p 982) The regional context of Waulsortian facies in northern England. In: Symposium on the Palaeoenvironmental Setting and Distribution of Waulsortian Facies (Eds Bolton, K. , Lane, H . R. & Lemone, D . V . ) , pp. 17-33. El Paso Geological Society/University of Texas at El Paso. MITCHU M , R.M. JR. & VAN WAGONER, J.C. (1990) High frequency sequences and e u static cycles in the Gulf of Mexico Basin. In: Sequence Stratigraphy as an Explo ration Tool: Concepts and Practices in the Gulf Coast. Programs and Abstracts, 11th Annual Research Confer ence, Gulf Coast Section, Society of Economic Paleon tologists and Mineralogists, Houston, pp. 257-267. MoSELEY , F. (1954) The Namurian of the Lancaster Fells. Q. J. geol. Soc. , London, 109, 423-454. MoSELEY, F. (1956) The geology of the Keasden area, west of Settle, Yorkshire. Proc. Yorks. geol. Soc. , 30 , 331-352. PHILLIPS, J. (1836) Illustrations ofthe Geology of Yorkshire,
Part II: the Mountain Limestone District. John Murray, London, 253 pp. PuNT, A.G. (1988) Sharp-based shoreface sequences an d 'offshore bars' in the Cardium Formation of Alberta: their relationship to relative changes in sea-level. I n : Sea-Level Changes: an Integrated Approach (Eds Wilgus, C . K . , Hastings, B. S . , Kendall , C.G.StC . , Posamentier, H.W. , Ross, C.A. & Van Wagoner, J.C.), Spec. Pubis Soc. econ. Paleontol. Mineral. , Tulsa, 42, 357-370. PosAMENTIER, H.W. , JERVEY, M.T. & VAIL, P . R. (1988) Eu static controls on clastic deposition I - conceptual framework. In: Sea-Level Changes: an Integrated Approach (Eds Wilgus, C.K. , Hastings, B.S . , Kendall, C.G .StC . , Posamentier, H.W. , Ross, C.A. & Van Wagoner, J . C . ) , Spec. Pubis Soc. econ. Paleontol. Mineral . , Tulsa, 42, 1 09-124. POSAMENTIER, H.W. , ALLEN , G.P. , JAMES, D . P . & TESSON , M . (1992) Forced regressions i n a sequence stratigraphic framework: concepts, examples, and exploration signifi cance. Bull. Am. Assoc. petrol. Geol. , 76, 1687-1709. RAMSBOITOM, W . H.C. (1977) Major cycles of transgression and regression (mesothems) in the Namurian. Proc. Yorks. geol. Soc. , 41, 261-291. RAMSBOITOM, W.H .C. (1979) Rates of transgression and regre ssion in the Carboniferous of NW Europe. J. geol. Soc. , London, 1 36 , 147-154. RAMSBO'ITOM, W.H.C. , RHYS, G . H . & S M ITH , E . G . (1962) Boreholes in the Carboniferou s of the Ashover district, Derbyshire. Bull. geol. Surv. G. B. , 19, 75 - 168. RAMSBOlTOM, W.H.C. , CALVER, M . A . , EAGAR, R.M.C. , HODSON , F. , HOLLIDAY, D.W. , STUBBLEFIELD, C. et a/. (1978) A correlation of Silesian rocks in the British I sles. Spec. Rep. geol. Soc. , London, 10, 81 pp. READ, W.A. (1991) The Millstone Grit (Namurian) of the sou thern Pennines viewed in the light of e ustatically controlled sequence stratigraphy. Geol. J. , 26, 157-165. READING, H.G. (1986) Facies. In: Sedimentary Environ ments and Facies (Ed. Reading, H.G . ) . Blackwell Scientific Publications. Oxford, pp. 4 - 19 . RILEY, N.J. (1985) Asturoceras and other dimorphoceratid ammonoids from the Namurian (E2) of Lancashire. Proc. Yorks. geol. Soc. , 45, 219-224. RILEY, N.J . , VARKER, W . J . , OWENS, B . , HIGGINS, A . C & RAMSBOTTOM, W.H.C. (1987) Stonehead Beck, Cowling, North Yorkshire, England: a British proposal for the Mid-Carboniferous boundary stratotype. Cour. Forsch. Inst. Senckenberg, 98, 159-177. SAUCIER, R.T. (198 1 ) Current thinking on riverine processes and geologic hi story are related to human settlement in the southeast. Geosci. Man, 22, 7-18. ScHUMM, S.A. (1993) River response to baselevel change: implications for sequence stratigraphy. J. Geol. , 101 , 279-294. SCHUMM, S . A . , MOSLEY, M . P. & WEAVER, W.E. (1987) Experimental Fluvial Geomorphology. Wiley, New York, 413 pp. SHANLEY, K.W. & Mc CABE , P . J . (1993) Alluvial archi tec ture in a sequence stratigraphic framework: a case hi story from the Upper Cretaceous of southern Utah, USA. In: The Geological Modelling of Hydrocarbon Reservoirs and Outcrop Analogues (Eds Flint, S.S. & Bryant, I . D . ) , Spec. Pubis int. Ass. Sediment. , No. 15, p p . 21-56. Blackwell Scientific Publications, Oxford.
Millstone Grit cyclicity
R.G. ( 1979) River channel responses to bedrock lithology and stream capture, Sandy Creek Drainage, central Texas. In: Adjustments of the Fluvial System (Eds Rhodes, D . D . & Williams, G . P . ) , pp. 255-276. Kendall/Hunt, Dubuque, Ia. STEPHENS, J.V., MITCHELL, G . H . & EDWARDS, W. ( 1953) Geology of the country between Bradford and Skipton. Mem. geol. Surv. G. B. , Eng. Wales, Explan. Sheet, 69, 180 pp. THUNELL, R . C . , LOCKE, S . M . & WILLIAMS, D . F. ( 1 988) Glacio-eustatic sea-level control on Red Sea salinity. Nature, 334, 601 -604. TREWIN , N.H. & HoLDSWORTH, B . K . ( 1 973) Sedimentation in the lower Namurian rocks of the North Staffordshire Basin. Proc. Yorks. geol. Soc. , 39, 3 7 1 - 408. UNDERHILL, J . R . , GAYER, R . A . , WOODCOCK, N . H . , DoNNELLY, R . , JOLLEY , E . J . & STIMPSON, I . G . ( 1 988) The Dent Fault System, northern England reinterpreted as a major oblique-slip fault zone. J. geol. Soc. London, 145, 303 - 3 1 6 . VAN WAGONER, J.C. , POSAMENTIER, H.W., MITCHU M , R.M. JR. , VAIL, P.R., SARG , J.F., LouTIT, T.S., et al. ( 1 988) An overview of the fundamentals of sequence stra tigraphy and key definitions. In: Sea-Level Changes: an Integrated Approach (Eds Wilgus, C . K . , Hastings, B . S . , Kendall, C.G .StC . , Posamentier, H . W . , Ross, C . A . & Van Wagoner, J.C.), Spec. Pubis Soc. econ. Paleontol. Mineral . , Tulsa, 42, 39-45. VAN WAGONER, J . C . , MITCHU M , R.M. JR . , CAMPION , K . M . & RAHMANIAN , V . D . ( 1 990) Siliciclastic sequence stra tigraphy in well-logs, cores and outcrops. Am. Assoc. petrol. Geol. Meth. Explor. Ser. , 7, 55 pp. VEEVERS, J.J. & POWELL, C.McA. ( 1 987) Late Paleozoic glaciations in Gondwanaland reflected in transgressive regressive depositional sequences in Euramerica. Geol. Soc. Am. Bull. , 98, 475-487. WAGNER, R . H . , HIGGINS, A.C. & MEYEN, S . V . (eds) (1979) The Carboniferous of the USSR. Occas. Pubis. Yorks. geol. Soc . , 4 , 247. WALKER, R . G . ( 1966) Shale Grit and Grindslow Shales: transition from turbidites to shallow-water sediments in the Upper Carboniferous of northern England. SHEPHERD,
J. sediment. Petrol. , 36,
327 90- 1 14 .
R.G. ( 1990) Facies modeling and sequence stra tigraphy. J. sediment. Petrol. , 60, 777-786. WALKER, R . G . ( 1 992) Facies, facies models and modern stratigraphic concepts. In: Facies Models - Response to Sea-Level Change (Eds Walker, R . G . & James, N . P . ) , p p . 1 - 14 . Geological Association o f Canada, St John's. WILLIAMS, D .F. ( 1 988) Evidence for and against sea-level changes from the stable isotopic record of the Cenozoic. In: Sea-Level Changes: an Integrated Approach (Eds Wilgus, C.K . , Hastings, B . S . , Kendall , C.G .StC . , Posamentier, H.W. , Ross, C . A . & Van Wagoner, J . C . ) , Spec. Pubis Soc. econ. Paleontol. Mineral. , Tulsa, 42, WALKER,
31 -36.
A.A. ( 1960) The Millstone Grit Series of Colsterdale and neighbourhood, Yorkshire. Proc. Yorks. geol. Soc. , 32, 429-452. WILSON , A.A. ( 1977) The Namurian rocks of the Fewston area. Trans. Leeds geol. Assoc. , 9, 1 -42. WILSON , A . A . & THOMPSON, A.T. ( 1 959) Marine bands of Arnsbergian age (Namurian) in the south-eastern portion of the Askrigg Block, Yorkshire. Proc. Yorks. geol. Soc. , 32, 45-67. WILSON, A . A . , BRANDON , A . & JOHNSON, E . W . ( 1 989) Geological Context of the Wyresdale Tunnel Methane Explosion . Geological Survey of Great Britain, Technical Report, WA/89/30, HMSO, London. WRIGHT, W . B . , SHERLOCK, R.L. , WRAY , D . A . , LLOYD, w. & ToNKS, L.H. (1927) The geology of the Rossendale Anticline. Mem. geol. Surv. G. B . , Eng. Wales, 182 pp. XuE, L. & GALLOWAY , W.E. (1993) Genetic sequence stratigraphic framework , depositional style, and hydro carbon occurrence of the Upper Cretaceous QYN formations in the Songliao lacustrine basin, northeastern China. Bull. Am. Assoc. petrol. Geol. , 76 , 1792- 1808. ZIEGLER, P . A . ( 1 986) Geodynamic model for Palaeozoic crustal consolidation of western and central Europe. Tectonophysics, 126, 303-328. ZIEGLER, P .A. ( 1 988) Evolution of the Arctic-North Atlantic and the Western Tethys. Mem. Am. Assoc. petrol. Geol. , 43, 196 pp. WILSON ,
Facies Analysis in Reservoir Sedimentology
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
Spec. Pubis int. Ass. Sediment. (1995) 22, 331-354
Productive Middle East clastic oil and gas reservoirs: their depositional settings and origins of their hydrocarbons Z I A D R. B E Y D O U N American University of Beirut, Lebanon
ABSTRACT The Middle East region contains 67% of the world's reported recoverable oil reserves and 30% of its reported recoverable gas reserves. The greater bulk of these are housed in carbonate reservoirs of Late Permian-Miocene age, with sandstone reservoirs of that age range being minor and restricted to specific areas of the region. The predominance of carbonates since the latest Permian is related to the plate tectonic position of the region, located principally in low latitudes and warm climatic belts. By contrast, from the Infra-Cambrian to the Early Permian, the Middle East region was located principally in middle to high southern latitudes in cool-wet to cold-dry climatic belts, when clastic sedimentation predominated and carbonates and evaporites were deposited mainly during brief excursions into lower latitudes and/or during short episodes of aridity in higher latitudes. Only fairly recently has the petroleum potential of the Palaeozoic clastic reservoirs begun to be properly appreciated after discover ies in Oman demonstrated ample sourcing from extensive organic-rich Infra-Cambrian to Early Cambrian sediments, with both vertical and lateral migration into Cambro-Ordovician and Permo Carboniferous sandstone reservoirs. More recently still, important discoveries in Central Arabia have confirmed sourcing of Permo-Carboniferous sandstone reservoirs from regionally deposited organic-rich Lower Silurian shales related to transgressive global sea-level rise. Either of these sources, where overmature, have most probably also provided most of the region's vast Permian Khuff carbonate reservoir gas reserves. The depositional environments and relevant petrophysical properties of the productive sandstone reservoirs of the region are reviewed by stratigraphical position and age and the source rocks charging these are briefly discussed. It is concluded that tremendous petroleum potential exists in the Middle East for substantial new Palaeozoic sourced and reservoired hydrocarbons. However, unlike the extensive lateral facies continuity of the post-Palaeozoic source-reservoir-seal relationships and the giant traps of the very wide Arabian shelf region (resulting in almost predictable accumulation trends), the Palaeozoic clastic reservoirs and seals generally exhibit rapid lateral and vertical facies changes coupled with the generally smaller sizes of local fault-block traps and/or low-amplitude fold traps. Consequently, the distribution of hydrocarbon accumulations in Palaeozoic sequences is less easily predicted, and in some areas is critically dependent upon seal distributions. Provided source and 'kitchen' distributions can be better understood and migration paths to traps better constrained, major additional reserves in Palaeozoic sandstone reservoirs will be established. Additional discoveries in post-Palaeozoic sandstone reservoirs are likely to be mainly restricted to the identification of hidden rift plays similar to those of Yemen and the Euphrates area of Syria.
I N T RO D U C T IO N west Iran-southeast Turkey, lying, respectively, southwest and south of the Zagros and Bitlis (Taurus) sutures, together with the Levant states and including the Sinai Peninsula but not Egypt proper (Fig. 1).
The geological boundaries of the Middle East region, as discussed in this survey, are those of the Arabian Plate and the adjacent Levant-Sinai fragment to its northwest (Fig. 1). Geographically, this encompasses the countries of the Arabian Peninsula and south-
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
331
Z. R.
332 Jo• LEGEND
40° oo· CJ t1lJlll !:::::>:I /
Beydoun
oo·
Quaternary-Recent Cenozoic Mesozoic
C ©
@
r::o-o-ol Palaeozoic @ � (Infra-Cambrian in
k;:���::�·�l
Oman)
PreCambrian
f"V\'VVl Mesozoic-Cenozoic � (Ophiolites & Basalts} Igneous Rocks
Approximate extent of currently hydrocarbon-productive basins with predominance of sandstone reservoirs and their ages P.M. C. (see legend for key)
N.B. Small lener (e.g. pl denotes only minor production or potential.
--::::::;::;
35° 00'
Main fractures or thrusts New Oceanic Crust
Jo• oo·
20° 00' SUDAN
( �
·· .r
.
1s•oo·
)
118506923
40° 00'
Fig. I. Geological sketch map of the Middle East.
Middle East clastic oil and gas reservoirs
The Middle East region, as outlined above (includ ing the Gulf of Suez rift basin), at the start of 1992 contained a reported 665.6 billion (B) barrels of recoverable oil and 1325.8 trillion (T) cubic feet (tcf) of recoverable gas (88.51 B tonnes and 3 37.54 T m respectively) (Oil and Gas Journal, 30 December 1991). This is equivalent to 67.2% of the World's reported recoverable oil reserves and 30.3% of the reported recoverable gas reserves. In 1991, the region accounted for 28.5% of the World's annual oil production (Oil and Gas Journal. , 30 December 1991); this figure, however, is determined more by economic and engineering constraints than by limitations on reservoir potential and productive capacity. Oil initially was discovered in the region in 1908, at Masjid-i-Suleiman in southwest Iran in the Oligo-Miocene carbonate reservoir of the Asmari Formation, and that field and others went on pro duction shortly afterwards. Commercial oil was found subsequently in Iraq, at Kirkuk in 1927 in a similar carbonate reservoir. Oil initially was found in Egypt at Gemsa field, located onshore on the western side of the Gulf of Suez in 1909, again in a carbonate reservoir of Miocene age. This was fol lowed by the discovery of the first clastic production in the Gulf of Suez at Hurghada field near to Gemsa in 1915, from Cretaceous 'Nubia' sandstone reser voirs as well as from a Miocene carbonate reservoir. On the eastern side of the Arabian Peninsula, the next discovery was made in Bahrain in 1932 in the Cretaceous sandstone and carbonate reservoirs of the Awali field, and in 1938 in eastern Saudi Arabia at Dammam in Upper Jurassic carbonate reservoirs and also in Kuwait at Burgan in Cretaceous sand stone reservoirs. World War II interrupted further exploration activity, but shortly after it was over a remarkably rapidly expanding exploration pro gramme commenced in many different parts of the region and continues to the present; discoveries have been made in every single country of the region except Lebanon, where the protracted phase of civil strife commencing in the early -mid- 1970s has cur tailed all exploration activities. The last country to join the producers club was Yemen in 1984. The magnitude of the many discoveries and devel opments varies considerably, from the 'giant' (0.5 -4.5 B barrels estimated ultimate recoverable (EUR) reserves of oil or oil equivalent) and 'super giant' (> 5 B barrels) fields of Iran, Iraq, Kuwait, Saudi Arabia, Qatar, the UAE and Oman, to the major fields of Syria and Yemen and the Gulf of Suez region, to the relatively modest fields of Bahrain
333
and southeast Turkey and to the minor production of Jordan and Israel (Beydoun, 1988). Cumulative Middle East oil production from all reservoirs since the first Iranian discovery has totalled about 175.3 B barrels, making the ultimate recoverable oil reserves for the region around 841 B barrels of oil (Oil and Gas Journal, 30 December 1991). Cumulative gas production figures are not discussed here because all gas production in the region is at present for local usage (desalination, power generation, fertilizers, feedstock, light industry) or for reservoir pressure maintenance by reinjection; some gas is still being flared during oil and gas separation, pipeline oper ations and refining. No gas as yet is being exported, although plans to export are well advanced in several countries of the region and exploration for non associated gas to establish additional reserves has been stepped up. Most of these immense reserves are housed in reservoirs predominantly of Permian to Oligo Miocene age, of which 80% are carbonate; the bulk of these reserves (98.7% of the oil and 98.8% of the gas) are located along the prolific northeast Arabian shelf. This region had formed the Arabian passive Tethyan margin of Gondwana (and subsequently of Afro-Arabia) from the initial cratonization of the Arabian Shield in the late Proterozoic until the collision and suturing with Eurasia. Initial collision commenced in the medial Tertiary and continued to the present, overprinting the Zagros foreland basin on the outer edge of this margin; the northeast margin extends from northwest Iraq through south west Iran and the eastern Arabia-Gulf region to northern central Oman (Fig. 1). The current hydrocarbon-producing reservoirs of the region and those being developed for production are shown by country in Table 1. The clastic reser voirs are shown in Fig. 2 and the principal areas of clastic reservoir production are indicated on Fig. 1. Apart from the relatively long established major production areas of the Gulf of Suez ('Nubia' Palaeozoic-Mesozoic, and Miocene, constituting Egypt's most prolific hydrocarbon province) and the Basra -Kuwait-northeast Saudi Arabia-Bahrain Cretaceous reservoir areas near the head of the Arabian-Persian Gulf, the remainder of the import ant producing areas constitute new additions to the list, with the Oman Palaeozoic discoveries, many dating from the mid- 1970s, the Syria and Yemen Mesozoic discoveries, from the mid-1980s, and the Central (Saudi) Arabian Upper Palaeozoic discover ies, from the late 1980s. Other minor clastic reservoir
Z. R.
334
Beydoun
Table 1. Petroleum producing formations of the Middle East. Expanded and updated from Beydoun (1991), reproduced by permission of AAPG Periods, Epochs, Ages
(Modified fonn Dunnington MIQ. PLIOCENE
1967)
GULF OF
YEMEN
SUEZ
ISRAEL/
PA L ES T INE
JORDAN
S.E. TURKEY
SYRIA
MIQ.PLIOCENE M · U MIOCENE L-M MIOCENE OGLIQ.
OLIGO. MIOCENE
U • M EOCENE L.EOCENE·PALAEDCENE
EOCENE/ PALAEOCENE
Habshiya : 0. (Gulf of Aden)
Belayim: 0 Kareem: 0 Rudeis: 0 Nukhul: 0
JOJTibe: O.G. Chilou: O.G. Jaddala : 0. Sinan : 0.
MAESTRICHTIAN U.CAMPANIAN L.CAMPAN-/SANTDNIAN CONIACIAN
u c
Sudr: 0
Garzan: 0
Matulla: 0
Beloka : 0. Karabogaz : 0.
Shiranish: O.G. Amah: O.G. ''Massive"
R E T A
TURONIAN M
c
APnAN L
BARREMIAN HAUnERIVIAN
VALANGINIAN BERRIASIAN
J
u
Karababa : 0.
PORTLANDIANM. KIMMERIDGIAN
Limestone : 0.
+ Derdere: o. (Mardin)
"Judea" : 0.
Nubia "A"
Qishn Clastics: 0
("Malha"): 0 Heletz: 0
Rutbah : 0.
Naifa: O.G. Alit: O.G.
u L. KIMMERIDGIAN
R A
s s I c
Ajlun Gp: 0. (Hummar & & Na'ur)
Raha: 0
ALBIAN
E
0 u s
CENOMANIAN
OXFORDIAN/CALLOVIAN M L
BATHONIAN BAJOCIAN
Amran : G.
Nir Am: 0
Kohlan: 0
Brur: O.G (Zohar: G)
LIASSIC L. LIASSIC
Butmah : 0. Camurlu: G--+Kurachine : O.G.
TRIASSIC
(Mulussa) : G.O. PERMIAN PERMQ.CARBONIFEROUS
p A L A
Dubaidib:
G
Katin : G.O.
CAMBRQ.ORDOVICIAN
I
CAMBRIAN I INFRA· CAMBRIAN
c 0 =Oil
Markadah : G.
Nubia "C"
CARBONIFEROUS SILURIAN-DEVONIAN
E 0 z 0
G
=
Gas, as predominant production
G after 0
=
Gas alone in some fields
Condensate in many of the non-Permian Gasfields
hydrocarbon production in the region occurs in Turkey (Palaeozoic), Jordan (Palaeozoic) and Israel (Cretaceous). The identified or inferred source rocks that charge these Middle East reservoirs, both car bonate and clastic, are indicated in Fig. 2 and those that have a known relevance to sourcing the clastic reservoirs are discussed briefly in the text, together with the environmental characteristics and some relevant petrophysical properties of the reservoirs. In a survey and review of this scale, no attempt has been made to describe the clastic sediments
concerned in the interpretive detail reached by some of the other contributors or, indeed, that which might be considered adequate by the dedicatee of this volume. Nevertheless, the review is offered with the intention of drawing out the important distinctions necessary in viewing the hydrocarbon productive clastic sediments of the prolific Middle East region in terms of age, plate tectonics, palaeolatitude, climate and broader depositional environments. Numerous oil and gas field names are mentioned
335
Middle East clastic oil and gas reservoirs
Table 1. (Continued. )
N. IRAQ
S.W. IRAN
OFFSHORE IRAN
CENTRAL & S. IRAQ
KUWAIT & DIVIDED ZONE
SAUDI ARABIA
QATARBAHRAIN + OFFSHOR E
UNITED ARAB EMIRATES
OMAN
Fars: G. "Main" O.G. limestone :
Asmari: O.G. (Kalho<)
AvanaiJaddala /Aaliji: 0.
Ahwaz/Asmar i
"Ghar": 0.
G.O. Jahnum: 0.
Shiranishl "Pilsner" :0.
Gurpi : 0.
Belayim/Rudois : G.O. (Rod Sea basin)
Ghar: 0
Pabdeh:
Dammam : O.G. U. Radhuma : 0.
G.
Hartha : 0.
llam : 0. Magwa: 0. (Mishri� Komotan/ Dokan I 0. "Qamchuqa" :
L. Qamchuqa and 0. Garagu
Mishrif: 0.
Mishrif: 0.
Nahr Umr : 0.
Rumaila : 0. Ahmadi: 0. Wara : 0. Mauddud: 0. Burgan : 0.
Servak : 0. Ahmadi : 0. Wara: 0. Mauddud: 0. Burgan: 0.
Wara: 0. Mauddud: 0. Burgan: 0.
Dariyan: G.
Shuaiba: 0. Biyadh :
Zubair: 0.
Zubair : 0.
(Zubair) :
Yamama : 0.
Minagish :0. (Ratawi)
Yamama : 0. Sulaiy: 0.
0. Fahliyan : G.
Fahliyan : 0. (Yamama)
0. Bubwaib: 0.
Hith: 0. Arab A�: 0. Arab D : 0 Jubaila : 0. Hanila : 0. Tuwaiq: 0.
Arab : 0. Arab D : 0.
Dhruma: 0.
Najmah : 0.
Mishril: 0. A hmadi/ Khatiyah: 0 Wara: 0 Maud d ud: 0 Nahr Umr: 0. Shuaiba : 0 Kharaib: 0
G. Mauddud : 0. Thamama (Shuaiba) : 0.
o.
Thamama (Shuaibo) : 0.
Thamama: 0.
Arab A�: G.O. Arab D I Darb: 0.
Arab I Darb. 0.
Araej: 0.
Araej : 0.
Marra!: 0.
Natih:
Mafraq: 0
Marrat: 0.
Kurachino : 0. Kangan: G. Dalan : G.
Dalan: G.
' Khuff: G.
Khuff: G. Unayza: O,G.
Sara: 0 Sajir: 0
Khuff: G.
Khuff: G.
Pre-Khuff :
G.
Khuff: O,G. Haushi : 0. (G harif, AI Khlata) Haima: O,G.
Haima: 0, G (Haradh, Amin, M ahawis/Andam) Ara: 0. Buah: G.
in the text but not given precise location on Fig. 1 because of the scale involved. A better overview of field locations (with the exception of those in the Levant-Sinai fragment) can be obtained from the enclosure map in Beydoun ( 199 1), which was also issued with Bull. Am. Assoc. petrol. Geol., 75(7) for July 1991.
P A L A EO G EO G R APH I C A L A N D P A L A EO C L I M A T I C S Y NOP S I S From the latest Proterozoic to the latest Palaeozoic
(Permian) the Arabian Plate, including the Levant Sinai fragment, was attached to the African Plate, which together formed part of Gondwana and was located in the southern hemisphere; it was part of the long and very wide shelf adjacent to the 'northern' margin of Gondwana and bordered the Palaeo-Tethys ocean. This wide shelf expanse was covered intermittently by shallow epeiric seas along side lowlands consisting of the cratonized Arabo Nubian and other shields; the areal extent of these shelf seas changed in response to succeeding trans gressions and regressions. Following plate tectonic reconstructions by Scotese et at. ( 1979) and Scotese
336
Z. R.
<<< ffi < << ��
,i
z « :!; 0
II
l :�
��!,�
IIU
-�
"
:0 z
-=�
.... :z:
:0 0
<
< � , �< ! ��� < < :<6& < <<;;_ <�&I, <
!
�I t
Beydoun
;
<<<<< <
I
Q)
m
""
li
! <-<�1?
I�
�
tt
jl<: IJI ' li �: II'\ ' I� << I< ' ' ' ' 1 �I<� !f ·� IJ: iI ll ' <<<< << I< 1!11 !l � l·.-:1 I < l � i � � i I l� �< ' .1 1�:1 J? hjt �� ! i .i·. !X\ I e� < i l : .,:\ � 1111 1<< 1_1 a:<<< li Ji f!� llI ; � I K! I � ·� i
�.... w
.,;
!"'
, ,
« ii'
�
>
�
I "'
I
� �.
·•
-�,
N w :0
�
-� " r-:0
·············••d
1: > : :
:�
.
>-
\1<
J�
�
!�
Q:
1-:-::.i!:·:·:·:·
�
z
Q) > .., u ::l 'C 0
,F
If.·�:-:
�
��
t::
:0
w
ffi :l:
ili
u
0
�
ii
3N3�03N 13N3�03'o'l'o'd
:0
w �
0 0
�
a:
�
g
sno3::J'<>'l3�-.
a:
� :0
::JISSvl:lnr
a:
�
g
l
:o
l
�
87 :o
l
�
Q) u
:;
0 "' 'C Q) :;:: .., c: Q) :!:!
Middle East clastic oil and gas reservoirs
& McKerrow (1990), the Arabian Plate (within the borders defined above) from latest Precambrian to Early Silurian had its 'long axis' (Fig. 1 inset) oriented) approximately NE-SW, with its present northern extremity in the southwest (Fig. 3). Palaeolatitudinal positions during this time changed through the following steps (Fig. 3):
Latest Precambrian Late Cambrian Middle- Late Ordovician Late Ordovician Early Silurian
between 7°S (Oman) and 2SOS (Sinai) between 30°S (Yemen) and 55°S (Sinai) between 28°S (Oman) and 52°S (Sinai) between 35°S (Oman) and 58°S (Sinai) between 29°S (Oman) and 50°S (Sinai)
The 'long axis' orientation changed with the advent of the Devonian and continued to shift orientation for the remainder of the Palaeozoic, as tabulated below, with the position of the present northern extremity ( PNE) indicated for each new orientation. (Fig. 3): Devonian Early Carboniferous Permian
orientation E-W, PNE west orientation NW-SE, PNE northwest orientation N-S, PNE north
Palaeolatitudinal positions during this time span changed as follows (Fig. 3): Early Devonian Early Carboniferous
Late early Permian
Late Permian
between 22°S (Zagros suture) and 39°S (Yemen) between 14°S (southeast Turkey/northern Iraq) and 37°S (Yemen) between 28°S (southeast Turkey/western Iraq) and 47°S (Yemen) between l2°S (southeast Turkey/northern Iraq) and 36°S (Yemen)
This sequence of changes would have placed the Middle East region throughout most of the Palaeo zoic mainly in temperate middle to high latitudes, where a humid continental climate with large con trasts in seasonal temperatures would have prevailed and where clastic sedimentation would have pre dominated. The region drifted into the edge of the warm and dry (arid) climate belt towards the end of the Palaeozoic. Only sporadic migration into or
337
influence of warmer climatic belts occurred before the Late Permian, allowing warm water carbonate shelf sedimentation, with or without accompanying aridity and deposition of evaporites, for relatively limited spans of time. For example, during latest Proterozoic to Early-Middle Cambrian time, evap orites and carbonates were deposited in the Gulf region and Oman (latitudes 5° -15°S) (Huqf and Hormuz Groups) and carbonates (Middle Cambrian Burj Formation) in the Levant (25°S); during the Devonian and Early Carboniferous, Jauf Formation limestones were laid down in Saudi Arabia (latitude 22°-20°S) and Markada and Harur Formations car bonates in Syria and Iraq (latitudes 15°-28°S). Finally, in the Late Permian, the Khuff cyclic car bonates and evaporites, which were laid down over most of eastern Arabia (S0-30°S) during a rise in sea-level following the end of Gondwanan Permo Carboniferous glaciation heralded the Mesozoic predominance of low-latitude carbonates (Fig. 3). During the Late Ordovician, the present part of central and western Arabia formed the northernmost extent of the Ordovician (Saharan) glaciation, lying at about 50°S (Zarqa and Sarah Formations of Saudi Arabia), and during Late Carboniferous to Early Permian, the present southern part of Arabia (Oman-Yemen and southwest Saudi Arabia Ethiopia) again underwent an episode ofglaciation (the AI Khlata Formation of the Haushi Group of Oman and its correlatives in Yemen and Ethiopia). However, considering that this glaciated strip of Arabia (and adjacent Afghanistan-India to its south) was located in middle latitudes during that phase (average 42°-45°S) the glaciation there was probably of uplifted (upwarped) areas prior to Permo-Triassic rifting of this part of Gondwana. During the Mesozoic, the Arabian Plate region moved to tropical latitudes, and its orientation ('long axis') was similar to that of the present. In the Early Triassic it was still wholly in the southern hemisphere (0°-25°S), but from the Early Jurassic onwards it straddled the equator, extending to no more than 20°S but generally averaging about equal areas north and south of the equator (Beydoun, 1991). In the Cenozoic the move northwards accelerated and from the late Palaeocene onwards it lay totally in the northern hemisphere, its northern tip not extending northwards beyond about 35°N. In consequence, warm-climate carbonate sedimentation has been predominant from the Late Permian onwards, accompanied from time to time by episodes of aridity; the deposition of associated clastics was
Z. R.
338
Beydoun
3o• s
30° s
so• s
c
so•
D
o•
o•
NUB - Nubia
IR - Iran
AFG - Afghanistan TUR - Turkey
IND - India
A.
LATE CAMBRIAN
MAD - Madagascar
B.
MIDDLE-LATE ORDOVICIAN (LLANDEILIAN - CARADOCIAN)
SOM - Somalia
c.
LATE ORDOVICIAN (ASHGILLIAN)
D.
EARLY SILURIAN (LLANDOVERIAN)
E.
LATE EARLY DEVONIAN (EMSIAN)
F.
EARLY CARBONIFEROUS (VISEAN)
G
LATE EARLY PERMIAN (ARTINSKIAN)
H.
LATE PERMIAN (KAZANIAN)
Fig. 3. Palaeozoic plate tectonic reconstructions. Arabia and surroundings. Modified with permission from Scotese eta/.
(1979); Scotese
&
McKerrow (1990).
Middle East clastic oil and gas reservoirs
principally in the form of transgressive shales or pelagic marls, some of which form major source rocks and/or seals. Episodes of coarse clastic depo sition were restricted to either arid conditions in specific climatic rain-shadow belts (Triassic) or to major fluvio-deltaic regimes draining uplifted regions and confined to specific time slices and areas. Examples are the Lower and Middle Cretaceous systems of the southern Iraq-Kuwait and north eastern Saudi Arabian region, the Gulf of Suez area and the Levant; the Oligo-Miocene systems of the same Kuwait-Basra area; the clastic deposits associated with initiation of rifting and subsequent marginal infill (Marib-Jauf in Yemen, Gulf of Suez): the regressive/basal transgressive coarse clastic deposition following regional uplift (end Jurassic early Cretaceous in various sectors): or finally, the Mio-Pliocene molasse associated with the rising Zagros-Taurus orogens of southwest Iran, Iraq and southeast Turkey. It is these varied palaeolatitudinal settings of the Middle East Arabian Plate region that have localized the development of clastic (reservoir) facies from the Permian onwards but permitted their extensive regional development prior to the Late Permian. Post-Palaeozoic deposition of carbonates over the wide Arabian shelf region gave rise to very areally extensive reservoir facies, periodically alternating with widespread source-seal units, thus enabling the generation and trapping of the very large hydro carbon reserves that are located principally in the northeastern part of the region, in large to 'giant' structural traps with predictably identifiable 'kitchen' areas and migration pathways. By contrast, the gen erally quite rapid lateral facies changes of the predominantly clastic Palaeozoic sequences and particularly the relative paucity of thick efficient (unbreached) seals, precludes short migration path ways from source to reservoir into trap, and often permits considerable vertical migration until an effective (regional) seal is encountered. Moreover, the distribution of Palaeozoic source sediments has been dependent largely on deposition in tectonically controlled basins or during major transgressive episodes, which subsequent tectonic events have obscured or stripped off. So far, only two major regional source facies have been identified: the Infra Cambrian to Lower Cambrian Huqf Group of Oman and the Gulf region and its correlatives, which are now generally deeply buried beneath later sediments and are mainly overmature, except in Oman, and the Lower Silurian Qusaiba/Safiq shales and their
339
correlatives; other potential Palaeozoic source rocks are of local significance only. Because Palaeozoic formations, and especially the pre-Devonian ones, are stratigraphically much older and, generally, deeper as exploration targets in a region fortunate in its profusion of prolific Tertiary and Mesozoic carbonate-reservoired oil accumu lations and in Late Permian carbonate-reservoired gas accumulations, the widespread search for pre Permian targets had to await changes in global economic and energy demands before being initiated. The first such exploration was carried out in Oman, where the Palaeozoic sediments are not deeply buried in the southeastern part of the country, and later in other parts of the region lacking younger reserves ( Jordan), before the latest successful ven tures in Saudi Arabia. Based on these recent very major discoveries in central Arabia in Upper Car boniferous to Lower Permian sandstone reservoirs and the identification of the source rock for these substantial accumulations as being the Lower Silurian shales (see next section), the potential for finding considerable new Palaeozoic reserves else where in the Middle East region is particularly promising. Potentially, overall the Middle East reserves could well double in the coming two to three decades of intensive exploration of the stratigraphically deeper levels, provided due understanding of the palaeogeographical/subcrop distribution of regional source sediments, source 'kitchens' and maturation levels are arrived at to direct the search (Beydoun, 1991).
P A L A EO ZO I C C L A STI C R E S E RVO I R S At present, the most important Palaeozoic clastic reservoir production in the Middle East region is that in central and southern Oman (Fig. 1), which first commenced regular production in 1975. Heavy oil of 19°-22° API gravity had been discovered in southern Oman (Dhofar Province) as early as 1956, by the Cities Service Company at Marmul, but it was never developed because of the heavy and viscous nature of the oil and the low oil prices prevail ing prior to 1973, which made the venture non commercial; the age of the reservoir had not then been established precisely nor had the significance of the oil sourcing been appreciated or properly understood. The sharp rise in oil prices in 1973 coincided with the early phase of serious exploration for older
340
Z.R.
deeper reservoirs in central and southern Oman (additional to the Cretaceous reservoirs of central Oman, which had been on production since the mid1960s), and this gave added impetus for accelerated exploration and development following the first dis coveries of light oil in Palaeozoic sandstone reser voirs. First commercial production began in 1975, from the fields in the Ghaba basin, and there are now at least 33 fields producing from Palaeozoic reservoirs in southern and central Oman, which had cumulatively produced some 890 million barrels of oil until the end of 1991. There are several billion barrels of recoverable oil in these reservoirs. Clastic reservoirs in Oman There are two major clastic reservoir groups in Oman. The older group comprises continental and shallow-marine sandstones of the Haima Group, which range in age from Early Cambrian for the oldest (Amin Formation) to the Late Cambrian to Early Ordovician (Andam Formation); the second group comprises glaciogenic sandstones and shallow marine and fluvial-channel sandstones of the Haushi Group, which ranges in age from Late Carboniferous (AI Khlata Formation) to Early Permian (Gharif Formation) (de Ia Grandville, 1982; Hughes Clarke, 1988; Sykes & Abu Risheh, 1989). Haima Group
The oldest oil-bearing Haima Group clastic sedi ments in Oman are those of the Karim and Haradh Formations (the Lower Haima of South Oman; Hughes Clarke, 1988). The Haradh is of greatest significance as a reservoir and comprises chert-rich cross-stratified sands with abundant synsedimentary deformation features. It probably represents a braided alluvial apron shed from highlands to the southeast, with the abundant chert being derived from the underlying formations of the Huqf Group (Heward, 1989). The Amin Formation reservoir is predominantly an intercalation of aeolian dune, sabkha and water lain sands. It is remarkable for the generally small size of the dunes that deposited it and reflects a period of arid continental conditions of widespread extent. Conglomerates derived from a deformation front to the west of Oman interfinger with these desert deposits. The Amin Formation probably cor relates, at least in part, with the widespread Saq (sandstone) Formation of Saudi Arabia.
Beydoun
The Mahwis Formation of South Oman comprises waterlain sheet-flood sands and conglomerates, probably again derived from the west (Heward, 1989). It passes northwards into the coastal andl storm-dominated shallow-marine sediments of the Andam Formation, which consists of fine sandstones, shales and thin ooidal and stromatolitic limestones in the middle part. The sediment package, of Late Cambrian to Early Ordovician age on the basis of its trilobite and other fauna, represents the first marine transgression in the Palaeozoic of Oman (Hughes Clarke, 1988). The Ghudun Formation reservoir (Haima Group), although an excellent reservoir, is only of minor regional significance, being of importance in only one South Oman field. Haushi Group
For the Haushi Group reservoirs, the lower or the AI Khlata Formation reservoir consists of a complex of clastic lithologies with considerable lateral and vertical changes in both facies and in thickness; this is particularly the case in South Oman, where these changes occur over very short distances (Levell et a/., 1988; Heward, 1990). Lithologies range from coarse-fine diamictites, conglomerates (particularly at the base), gravels, pebbly sands, sands, silts, silty shales and shales that are often varved. Boulders can reach up to 10 m in diameter and are of very varied composition, including granite, granophyre and volcanics, as well as Infra-Cambrian sedimentary rocks of the Huqf Group. The AI Khlata is a glacially generated sediment package and was deposited by a series of glacial advances and retreats from the southwest towards the northeast, as evidenced by striations and 'whale backs' in glacial pavements in outcrop (Levell et at. , 1988). The reservoirs are predominantly glaciofluvial and glaciolacustrine (Heward, 1990). The AI Khlata correlates with simi lar deposits present elsewhere in southern Arabia and Ethiopia (Beydoun, 1988; McClure, 1980; Levell et a/. , 1988). The overlying Gharif Formation reservoir of the Haushi Group consists of a series of fluvial deposits derived from the southeast. The arkosic character is due to recycling of granitic material from AI Khlata, presumably being stripped off from the uplift that finally rifted between Oman- Afghanistan-India to the southeast. The formation fines and becomes more shaley to the northwest; its deposits reflect an ameliorating climate, most being red-beds, locally
Middle East clastic oil and gas reservoirs
with calcrete and frequent occurrences of silicified wood in outcrop, with growth bands indicating a seasonal climate. Marine sands and limestones occur to the north and west and reflect the interplay between marginal-marine -coastal-plain deposits and continental red-bed facies (Hughes Clarke, 1988). The formation generally correlates with the upper Unayza Formation of Saudi Arabia discussed below. The considerable volumes of hydrocarbons in place, housed in the Palaeozoic clastic reservoirs of the Oman region, are demonstrated to have been abundantly sourced from Huqf Group source sedi ments at least 550 Ma in age (but also of older Upper Proterozoic age) straddling the Infra-Cambrian to Early Cambrian boundary and terminating with a widespread evaporite unit- mainly salt (Gorin et al., 1982; AI Marjeby & Nash, 1986; Grantham et al., 1988; Mattes & Conway Morris, 1990). A few fields in western Oman, however, adjacent to the eastern side of the Rub al Khali basin, have oil accumulations in Haushi Group clastic sediments that have been geochemically typed to Silurian (Safiq and equivalent) source rocks (Grantham et al., 1988; Beydoun, 1991). The intermittent uplift of the southeast Arabian plate margin since mid-Palaeozoic times, and par ticularly since the Silurian, when the region became relatively positive, is the principal reason why these very old source rocks in South Oman remain in the oil window and continue to generate oil; they lie, on the whole, at moderate depths and seemingly have not been more deeply buried. The relatively heavy nature of some of the oil is considered to be due to subsequent transformation, although some may be ascribed to generation from still relatively immature sediments at shallow depth. Hydrocarbon charge from these old source rocks into Palaeozoic clastic reservoirs occurs from beneath the terminal Ara Formation (Huqf Group) salt edge and takes place in conjunction with progressive subsurface salt removal and dissolution; this leads to the formation of residual cores composed of basin-fill bodies of the overlying post-salt Palaeozoic clastic deposits (AI Marjeby & Nash, 1986). Short- to medium-distance lateral migration into the Palaeozoic clastic deposits, and/or subsequent long-distance vertical migration within these clastic deposits, occurs until stopped by an effective seal in trapping geometry; these seals are predominantly shales or argillaceous sediments. In the clastic depositional facies patterns that charac terized the Palaeozoic of the region, shale seals
341
generally have limited lateral extent and are often thin and easily breached. Migration of Huqf sourced oil has, thus, often moved vertically up as high as the Carboniferous-Permian reservoirs of the Haushi Group, where sealing is provided by glaciogenic or shallow-marine-floodplain shales and mudstone deposits. Furthermore, where these have locally become breached or have proved inadequate, verti cal migration has continued in central Oman to charge the Mesozoic carbonate reservoirs of the largest Oman fields (Sykes & Abu Risheh, 1989). The subject and wider implications of Palaeozoic oils in Middle East exploration has been discussed and reviewed in more detail by Beydoun ( 1991). The AI Khlata reservoirs are determined by a number of facies belts whose distribution is partly influenced by local salt dissolution (of the underlying Huqf Group Ara Formation salt); where subsurface salt dissolution occurred during the glaciation episode, lakes were formed with generally shaley deposits. Along the margins of the lakes, glaciogenic deltas gave rise to mixed sequences of sands and shales (multiple reservoir-seal couplets) and consti tute the most prospective facies belt. Within the main South Oman salt basin, terrestrial deposition dominated, with much waste rock (diamictites) and erratic reservoir character (Levell et al., 1988). Deltaic deposits are occasionally incised by sand stone-filled channels and rapid lateral and vertical variations in lithology and stratigraphy make it diffi cult to estimate and indeed develop all recoverable oil reserves from the AI Khlata reservoir. The over lying transgressive Gharif Formation sands often provide a peripheral stratigraphical trap component reservoir to the structural trap formed by drape over Haima Group sediment pods left by salt withdrawals and dissolution (AI Marjeby & Nash, 1986). Porosities and permeabilities of the South Oman Palaeozoic clastic reservoirs are generally good and range from 15 to 35% and 100 mD to over 10 D in the friable and poorly cemented reservoir levels (de Ia Grandville, 1982; Levell et al., 1988). Clastic reservoirs in Saudi Arabia In central Saudi Arabia, several billion barrels of high gravity (c. 43° API), low-sulphur Palaeozoic oil (and gas) have been discovered recently in the region south of Riyadh (Fig. 1), principally in the Unayza Formation sandstone reservoirs of ?latest Carbon iferous to Early Permian age (AI Laboun, 1986; Husseini, 1992; McGillivray & Husseini, 1992), but
342
Z.R. Beydoun
also including some oil housed in Devonian and Ordovician sand reservoirs. Oil to potential source rock correlations have identified the principal source of these hydrocarbons to be the basal, organically rich shales of the Qusayba Member of the Lower Silurian Qalibah Formation (Abu-Ali et al., 1991; Husseini, 1992; McGillivray & Husseini, 1992; Mahmoud et at., 1992). The Qalibah Formation was previously unrecognized as a separate formation and was included in the middle part of the Tabuk Forma tion (Powers, 1968). The first significant Palaeozoic sandstone reservoir hydrocarbon discovery was made in 1979 in Qirdi (south of the Khurais field, then the westernmost established oilfield in Saudi Arabia) in an Upper Permian pre-Khuff (carbonate) Formation sandstone sequence and in the same year pre-Khuff gas was also tested in the 'supergiant' Ghawar field (Husseini, 1992). Other discoveries followed in deep tests on the Khurais trend south of that field, but at depths in excess of 4000 m. It was not until 1989 that the major discoveries in the previously unexplored areas were made, and at moderate depths. The Unayza Formation is a siliciclastic succession of fluvial, coastal-plain, deltaic and shallow-marine facies whose age may extend back to Late Carbon iferous (AI Laboun, 1986), but is principally of Early to Late Permian age; it appears broadly cor relatable with the Gharif Formation of Oman whose deposition is associated with the Late Carboniferous to Early Permian deglaciation episode that affected Gondwana, including the southern part of Arabia (see earlier). Distribution of the varied Unayza facies is controlled by palaeogeography, sea-level fluctuations and by local and regional structure (Ferguson & Chambers, 1991). The complex inter play of these factors coupled with erosion during sea-level lowstands, has resulted in a variety of reservoir types, varying through poorly sorted coarse silt-sand of tentative marine origin, accumulated in fault bound depressions; fine to pebbly cross-bedded sandstone of alluvial origin; a variety of paralic sandstone facies, including shoreface-foreshore, delta channel and coastal-plain channel sands (it is these paralic facies that have proved to be the most prolific reservoirs); the final facies is a transgressive lag of unpredictable distribution but consisting of clean fine- to very coarse-grained reservoir quality sandstone (Ferguson & Chambers, 1991). Porosities in these facies are highly variable but average 20%, although permeabilities of several darcies are not uncommon (Husseini, 1992).
Other reservoir sandstones in this region of Arabia include Ordovician glacially related sandstones and Devonian ?continental sandstones. The hydrocarbon charge to the trap from the Silurian Qusayba shales is controlled by juxtaposition of reservoir to source through faulting and/or fluid communication with the main Unayzah reservoir. Stratigraphical pinch out of the Unayzah reservoir against the basal Khuff seal unconformity may affect the trap form significantly. The Lower Silurian Qusayba shales (and the equivalent Safiq in Oman) are widespread across Arabia and reflect the first transgressive cycle related to global sea-level rise consequent on the deglaciation occurring at the end of the Late Ordo vician to Early Silurian 'Saharan' glaciation. Similar shales are widely distributed throughout the Middle East and North Africa, but subsequent post- Middle Devonian tectonism in the Middle East and uplift have removed them from elevated areas; conse quently, their subcrop palaeogeographical distri bution in the· region beneath the Late Permian Khuff seal needs to be better controlled. The Qusayba shales are, however, seemingly present everywhere in the basins of Saudi Arabia and at depths where they are sufficiently mature to generate oil and gas (Husseini, 1992); this includes those portions of the Arabian Peninsula adjacent to Jordan and southern Iraq (AI Laboun, 1986). Oil generation and up-dip migration from some of the Arabian depocentres commenced as long as 160 Ma (Abu-Ali et a!., 1991). Clastic reservoir in the Gulf Region In other areas of the Gulf region, gas had been discovered in pre-Khuff clastic reservoirs in Abu Dhabi and Bahrain and non-commercial gas found in onshore Qatar. These gas discoveries are as yet undeveloped although potentially producible should market forces warrant it; but although the gas is generally H2S free, appreciable amounts of associ·· ated nitrogen and C02 cast doubt on commerciality. The principal accumulations in these countries occur in Unyiza-equivalent clastic deposits, and the sequences consist of continental fluvio-deltaic sands, silts and shales with several reservoir sandstone intervals separated by shale bodies. The source of the gas is generally assumed to be the Silurian shales, which, however, have not been penetrated in any Abu Dhabi well (Ali & Sildawi, 1989; Loutfi & El Bishlawy, 1989), although they may have
Middle East clastic oil and gas reservoirs
been partly penetrated Chaube, 1987).
m
Bahrain (Samahiji &
Clastic reservoirs in Syria In Syria, a number of gas accumulations occur in anticlinal structures in the central part of the country (southern Palmyrid basin) (Fig. 1). These are cur rently undergoing delineation drilling in preparation for commercial production. The reservoir facies is sandstone, occurring in intervals separated by clay stones, thin shales and intercalated dolomites of the Upper Carboniferous Markada Formation; there are rapid vertical and lateral lithological variations. The depositional environments are probably shallow marine to ftuvio-deltaic, with provenance of clastic material from erosion of the adjacent uplifts (Rutbah High and ?Aleppo High) created by an 'Hercynian' tectonic event. Periodically reduced clastic supply allowed carbonate deposition, commensurate with palaeolatitudinal location in tropical waters. Log porosities in the clastic sediments are reported to range between 5% and 16% . The source of the gas (predominantly methane but with C2 and C3 fractions and varying small amounts of nitrogen) has been inferred to be from shales within the Markada itself; however, based on regional considerations, it seems more likely to be from unconformably underlying Silurian (Tanf Formation) shales in a similar manner to the Unayza accumulations of Central Arabia. Although not proven to underlie this area, these shales are extensively distributed in the region and have been encountered in a number of widely scattered wells (Lababidi & Hamdan, 1985 ; Beydoun, 1988, 1991). Clastic reservoirs in Jordan In adjacent Jordan, the Risha field, close to the Iraqi border (Fig. 1) , is producing gas from a clastic reservoir of Late Ordovician age; this is the Dubaidib Sandstone Formation of the Khreim Group lying directly beneath the organic-rich Lower Silurian (Mudawara Formation) shales, which appear to act both as source and seal for this reservoir (NRA Jordan, 1989). This area of northeast Jordan was an Early Palaeozoic basin that accumulated a thick succession of Palaeozoic sediments, several thousand metres of which (Early Cambrian-Silurian) are still preserved following inversion and erosion in the Late Palaeozoic. The Ordovician reservoir sequence here consists of shallow-marine cyclic sandstone
343
sequences with stringers and intercalations of shales, claystones and siltstones; lateral and vertical con tinuity varies rapidly, however, being the main reason why only three or four of the dozen or more wells drilled over the (?)imprecisely defined sub surface feature to date are producers (NRA Jordan, 1989). Good permeabilities, however, are developed locally, as evidenced by the reported frequent mud losses during drilling. Light but non-commercial oil has been found in the same reservoir in well Wadi Sirhan 4, some 250 km to the southwest (NRA Jordan, 1989). Clastic reservoirs in southeast Turkey In southeast Turkey, several small accumulations of gas and condensate have been discovered in Devonian age clastic sediments of the Katin Forma tion, below the southern front of the Taurus thrust belt (Fig. 1). The reservoir facies consists of clean shallow-marine sands, which are oil impregnated nearby at outcrop (Hazro sands, previously thought to be Permian in age; Beydoun, 1991). These reser voir sands are considered to have been sourced by the Ordovician-Silurian organic-rich Bedinan For mation shales (locally defined as the Dadas Forma tion) with TOCs up to 5% (Harput & Erturk, 1991); these also contain reservoir sand levels (Handof Member) that have contained strong but non commercial hydrocarbon shows in a number of wild cat wells in the southeast Turkey region. The cumulative production from these Palaeozoic reser voirs has been modest to date; however, geochemical work in the area by a number of operators confirms that the Bedinan Formation has been the source charging a number of oil-productive Cretaceous car bonate reservoirs in southeast Turkey in areas where the Middle Cretaceous often unconformably over lies the Palaeozoic (Mardin-Bozova High). It is thought likely that accumulations in older clastic Palaeozoic reservoirs, such as those found in the Barbes and Katin fields, could be discovered with better subsurface well control and seismic mapping. Clastic reservoirs in the Gulf of Suez In the petroliferous Gulf of Suez rift basin of Egypt, on the south western edge of the Levant-Sinai fragment of the Arabian Plate, the bulk of the oil reserves are housed in sandstone reservoirs, the majority being of Miocene age, but a number of fields also produce from 'Nubia' sandstone reservoirs
344
Z. R.
of Late Palaeozoic and Cretaceous ages. The Palaeo zoic reservoirs belong to the Nubia 'C' unit, which is of ?Devonian-Carboniferous age (the Nubia 'B' consists of Carboniferous black shales, which are regarded as a likely source for some of the oil by some authors- Ayouty, 1990). The Nubia 'C' depositional facies in the Gulf of Suez region is principally shallow marine with deltaic and fluvial intercalations over much of the area but with alluvial deposits in the southern part; the facies became predominantly continental to shallow marine towards the end of Nubia 'C' time in the Late Carboniferous to Permian (Klitzch, 1990). About six or seven oil fields produce from the Palaeozoic sandstone reservoir, including the now depleted Hurghada field in the west, where the bulk of the oil was housed in the Nubia 'C' reservoir and which produced during its life span (1915-1969) 41 million barrels of oil. Other Palaeozoic reservoir fields include the Ras Gharib, July, Ramadan, Ras Badran and Sidky fields, where reservoir porosities and permeabilities are fair to good and contain oils having gravities that range between 24° and 40° API (Ayouty, 1990). The oil in the Palaeozoic and younger reservoirs of the Gulf of Suez is considered by many to be sourced mainly from organic-rich Early Miocene shales and marls (Rudais-Karim Formations) brought into the right juxtaposition of source to reservoir in tilted fault-block traps or stratigraphical pinch outs. Additionally, and perhaps alternatively, Late Cretaceous and Eo-Palaeocene sediments also contain ample organic carbon, enough to generate the oil in Cretaceous and Neogene reservoirs (Ayouty, 1990), and the Carboniferous Nubia 'B' shales cannot be ignored as another possible source to be considered in the calculation of total oil charge volumetrics.
M E S O Z O I C C L AS T I C R E S E R V O I R S Commercial oil production from Mesozoic sandstone reservoirs was first established in the Middle East region on the western side of the Gulf of Suez, in what are now depleted fields (Hurghada, 1915; Abu Durba, 1918; etc.) in 'Nubian' sandstones of Cre taceous age (now known as the Nubia 'A') (Ayouty, 1990). Sustained and continuing commercial oil production from sandstone reservoirs in the Middle East region, however, was first made in Bahrain in 1932 with the discovery of the Awali field.
Beydoun
Clastic reservoirs in the Bahrain- Neutral Zone- Kuwait - Basra region The 'Bahrain zone' in the Awali field consists of several productive Middle Cretaceous sandstone- siltstone (Wara, Nahr Umr) and carbonate (Ahmadi, Mauddud) reservoirs. Subsequently, the discovery and development of large oil reserves from equiv alent age sandstone reservoirs occurred in Kuwait in 1938 (Wara and Burgan Formations), in Lower Cre taceous sandstone reservoirs in the Basra area o:f southern Iraq at Zubair in 1949 (Zubair Formation) and in Albian age sandstone reservoirs (Safaniya/ Khafji Members of the Wasia Formation) in the northeastern part of Saudi Arabia and the adjacent part of the Saudi-Kuwaiti Divided (Neutral) Zone in 1951 (among others, the Safaniya-Khafji field the largest offshore oil field in the world). The oil (and gas) fields of the southern Iraq Kuwait-Divided (Neutral) Zone-northeast Saudi Arabia onshore and offshore region, were discovered and developed since the late 1940s and earliest 1950s. They produce (or are being prepared to produce) from these Middle and Lower Cretaceous sandstone reservoirs, and although only 23 in number, they constitute (both in production and recoverable reserves) by far the most important clastic reservoir oil province of the whole Middle East region . The recoverable hydrocarbon volumetrics of this province alone involve a sustainable production of several million barrels of oil per day and many tens of billions of barrels of remaining recoverable reserves (Beydoun, 1988, 1991), well in excess of the oil totals for the entire North Sea Basin ! The sandstone reservoirs housing these vast reserves represent the deposits of a huge fluvio-· deltaic-littoral-marginal marine depositional system, which commenced in early Cretaceous (Hauterivian-Aptian) time, when this area of the Middle East lay athwart the humid equatorial belt , 2°-3° to north and south of the equator, and ter- minated in the Albian to early Cenomanian, when the area drifted to between so and l0°N (Dercourt et al., 1986; Beydoun, 1991). This fluvial system (a virtual 'Nile sized' river complex) transported immense quantities of clastic detritus northeastwards from the Arabian -Nubian shield highlands, and from the exposed Palaeozoic clastic sediments over lapping the shield, on to the shallow and very wide Arabian passive margin shelf. The delta complex prograded during sea-level lowstands but retreated during transgressive high sea-level stands when shelf
Middle East clastic oil and gas reservoirs
carbonates or fine clastic sediments were deposited between the coarser lowstand sandstones. A distribu tary channel system also extended the delta complex to the Baghdad area in Central Iraq (Jawad Ali & Aziz, 1993). The repetitive nature of the Zubair Formation deltaic cycle (Hauterivian-Aptian and correspond ing to the subsurface Biyadh Formation of Saudi Arabia), which is one of the major productive reser voirs of the area, recently has been documented qualitatively (Khaiwka, 1989): the cycle includes pro-delta, delta front and delta plain, with marine shelf facies concluding the cycle. Subaqueous levee, distributary channels , distributary mouths, distal bars, freshwater lake, mud-fiat, marsh, subarea) levee and lobe abandonment facies have all been recognized and have a strong repetitive vertical pattern. Rapid facies changes provide the lithological con trols on the variable hydrocarbon distribution in the lower regional (Zubair) clastic reservoir; it provides the major production in southern Iraq but is a relatively minor reservoir in Kuwait and the north east Saudi Arabian fields. Potentially important new oil production from the Zubair reservoir was reported as being established recently in the East Baghdad oilfield of central Iraq (Beydoun, 1988). The Zubair Formation is overlain by the trans gressive carbonate cycle of the Shuaiba Forma tion, which concludes with a terminal reefal build-up phase (Dunnington et al., 1959; Powers, 1968). The Albian (Nahr Umr and Burgan Formations) clastic sequences of the Basra-Kuwait area and their equivalents in northwest Saudi Arabia and the Neutral Zone (Khafji and overlying Safaniya Members of the Wasia Formation) initiate a new prograding fiuvio-deltaic-littoral-shallow-marine depositional cycle following the Shuaiba (transgress ive) carbonate cycle, similar to that of the preceding early Cretaceous Zubair cycle. This Nahr Umr to Burgan progradational episode grades up into the widespread shallow-shelf carbonate sedimentation of the Mauddud Formation, which terminates the Albian. The early Cenomanian Wara clastic depositional episode that followed the Mauddud was of more limited duration and areal extent, covering Kuwait and eastern Arabia to Bahrain, and representing a brief terrigenous interlude in an otherwise marine Middle Cretaceous depositional phase. The Wara consists of a sequence of sandstones with interbeds
345
of siltstone and shale and exhibits rapid lithological changes both laterally and vertically (Powers, 1968). In its lower part, in the Kuwait region, it includes a large lenticular sand body enclosed by grey, lignitic shales (Owen & Nasr, 1958; Dunnington et al . , 1959) which indicate a return t o fiuvio-deltaic depo sitional conditions. This sandstone body interfingers with littoral-shallow-marine facies to the southeast (Bahrain) and with deeper marine pro-delta facies to the north and northeast (southern Iraq); sub aerial delta deposits, tidal-fiat and coastal-plain deposits dominate to the west, with the submarine portion of the delta being located east of the present coastline. The Wara reservoir contains significant volumes of oil in Bahrain, in the Divided (Neutral) Zone and in northeast Saudi Arabia and southern Kuwait, where clean permeable sands are common. The Wara clastic cycle is terminated by a transgressive marine shale facies of the Ahmadi Formation, which forms the regional seal to the Wara reservoir (Brennan, 1990; Samahiji & Chaube, 1987). In the Basra area the Wara is not distinguishable as a sandstone sequence, the coeval interval being a facies of black silty shales and siltstones with a thin oil saturated sand (Dunnington et al., 1959) . Reservoir porosities and permeabilities for these different Lower and Middle Cretaceous sandstone reservoirs vary and data are generally sparse, but good reservoir properties are commonly present. (OAPEC, 1985 ; Brennan, 1990, 1991) . In southern Iraq, a petroliferous belt of Albian Nahr Umr clastic sediments coincides with a marginal-marine depo sitional area (Ibrahim, 1983) but production from this reservoir is subordinate to that from the Zubair reservoir. The hydrocarbons in the Lower and Middle Cre taceous clastic reservoirs of the 'head of the Gulf ' region of the Middle East are believed sourced from the Kazhdumi. Formation organic-rich sediments in the Kazhdumi basin immediately to the east (Brennan, 1990) and/or from a 'deeper' Lower Cre taceous source ( Ibrahim, 1981) within the Gotnia depositional basin of Murris (1980) (Sulaiy and Chia Gara/Sarmord and Garau source) with vertical migration and charging from the Zubair into the Nahr Umr reservoir of southern Iraq ( Ibrahim, 1981). This question of sourcing of the various Middle East reservoirs and the regional distribution of source rocks has been reviewed and discussed in more detail by Beydoun (1988, 1991) and by Beydoun et al. (1992).
346
Z. R.
Clastic reservoirs in Yemen Two new and potentially major hydrocarbon pro ductive Mesozoic clastic reservoir regions have been established in the Middle East since 1984. These are in the Marib -Jauf rift basin and the Wadi Hadhramaut basin of Yemen in southwest Arabia. The Marib-Jauf basin now has about seven oil fields on production, or being developed for pro duction, where the oil and gas/condensate reserves are housed in sandstone reservoirs of latest Jurassic age. The discovery field was Alit, which is also the biggest. In 1991, this basin's oil production averaged about 190 000 barrels per day and recoverable reserves were estimated at about 3-4 B barrels (Oil and Gas Journal, 30 December 1991); the gas (with condensate) reserves are estimated to run into sev eral trillion cubic feet and means for developing these reserves are under consideration. The Upper Jurassic subsurface succession within this rift basin differs from that outcropping nearby in that a thick synrift sequence of sediments was laid down ; this NW-SE trending basin started to subside rapidly in the post-Oxfordian, with in excess of 3500 m of sediments being deposited in about 6 Ma (Maycock, 1987). The Callovian-Oxfordian Amran (Shuqra) platform carbonates here are overlain by a thick and predominantly shaley sequence followed by carbon ate turbidites rapidly deposited following graben inception. This sequence, informally referred to as the 'Lam Formation', is correlative with the organic rich Madbi Formation of the outcrop area to the southeast (Beydoun, 1964; Beydoun & Greenwood, 1968) and is the principal generative source for the Marib-Jauf hydrocarbon accumulations ; the 'Lam' is overlain by the 'Alif Formation', which is the main reservoir and consists of a shallow-marine-deltaic fluviatile succession of clean sands that are of Tithonian age (Huurdeman et al . , 1989) and are succeeded by sandstones and shales of braided flu viatile and deltaic systems, in turn followed by the Sabatain (Main Salt) evaporite sequence, also of Tithonian age (Maycock, 1987; Huurdeman et al., 1989) ; these form an excellent regional hydro carbon seal. Excellent reservoir porosities and per meabilities have been recorded ranging from 16% to 18% and 500- 1200 mD, with the oil being of 39°-41 ° API gravity (Huurdeman et al . , 1989). Some 300km to the east of the Marib-Jauf rift basin (and about 225 km east-northeast of its Late Jurassic carbonate-reservoired Shabwa hydrocarbon basin extension) recent exploration has revealed
Beydoun
other Late Jurassic rifts on an E-W trend (Mills, 1992), with a deep lying pre-rift sediment package underlying a Kimmeridgian-Tithonian synrift se quence, which includes an analogue of the organic rich Madbi source sequence; a series of important discoveries have been made in this area, the first in December 1990; estimated recoverable reserves to date are conservatively estimated at about 250 million barrels. The succession in this Hadhramaut-Jeza rift basin area is that described by Beydoun ( 1964, 1966) and Beydoun & Greenwood ( 1968) for the typical non evaporitic Jurassic Amran Group of Hadhramaut overlying the Liassic to Middle Jurassic clastic Kohlan Sandstone Formation ; this consists of pre-rift Shuqra Formation platform carbonates and synrift Madbi and Naifa (Kimmeridgian Tithonian-Berriasian) organic-rich shales and marls and argillaceous limestones, respectively. No evap orites equivalent to the Sabatain Formation are developed here and hence there is no efficient seal on top of the Jurassic, as in the adjacent Marib - Jauf and Shabwa rift basins. Exploration wells have shown some unexpected subsurface differences in the post-rift Cretaceous stratigraphy from the out crop sequence described by Beydoun & Greenwood ( 1968). An early Cretaceous platform carbon ate unit occurs beneath the Lower Cretaceous (Barremian-Aptian) Qishn Formation consisting of carbonates and alternating shales unconformably resting on the terminal Jurassic Naifa Formation and separating it from the Qishn ; this sequence has been informally named the Saar Formation of Hauterivian to early Barremian age (Mills, 1 992) and appears to have been laid down in an E-W shallow marine embayment. Additionally, the Qishn itself exhibits a substantial development of clastics derived from bordering and still positive uplifts to the marine embayment, below its main Barremian-Aptian carbonate unit, which are not developed in the main reference surface section for the formation within the Tawilah Group (Beydoun & Greenwood, 1968) located near Mukalla along the coast about 150 km due south of the first discovery. These clastic sediments in the area of recent discoveries (Hadhramaut-Jeza rift) are not only well developed but have good reservoir qualities. In this area, the hydrocarbons were generated in the organic-rich and well-developed Madbi source rocks in the adjacent graben(s) and migrated into the suitably juxtaposed older Kohlan (Lower-Middle Jurassic) sandstone reservoir and also, in the absence of an
Middle East clastic oil and gas reservoirs
efficient seal (Tithonian evaporites of the Sabatain Formation being absent), through the overlying suc cession into the Qishn sands beneath the thick and tight main Qishn carbonate that forms the regional seal (Mills, 1992). Good quality oil of 36° API gravity is found in the Kohlan reservoir and 30° API oil in the Qishn clastic reservoir(s) and substantial flow rates were tested. Individual zones in the Qishn clastic deposits have flowed in excess of 4875 barrels per day of 29°-33° API gravity with little dissolved gas and low sulphur (Mills, 1992).
Clastic reservoirs in Syria In the Euphrates graben basin of eastern Syria, some 25 oilfields discovered since 1984 are now on production and collectively were producing about 330000 barrels per day of 34°-38° API gravity low sulphur oil in early January of 1992 ( Middle East Economic Survey, 35: 16, 20 January 1992). Recov erable reserves are reported to be well over 1 B barrels. The reservoir is the so-called Rutbah Formation sandstone of Early Cretaceous age (Beydoun, , 1988). The Rutbah type section was defined originally across the border in Iraq, south of the Euphrates River, where a 23-m section of vari coloured, white and ferruginous unfossiliferous sand stone is exposed, with the basal part probably being of continental origin and the uppermost part marine; the age was given as Late Triassic to early Ceno manian based on the relationship with the dated overlying and underlying formations (Dunnington et a/., 1959). The correlation of the Rutbah with the Nahr Umr of southern Iraq as a single diachronous sandstone deposit was not regarded favourably (see, however, Ibrahim ( 1981) who recorrelates the two). The informal ' Rutbah Sandstone' of Syria was for malized as the Cherrife Shale Formation of Late Jurassic to Early Cretaceous age (Daniel, 1963a). The type section was the Cherrife- 1 exploration well in the Palmyrid basin, where the formation comprises predominantly red-brown fet:ruginous sandy shale and green pyritic shale with intercalations of quartiz itic and calcareous sandstone and streaks of lime stone. Elsewhere in Syria, the succession includes decomposed basalt and ash and is more varicoloured, more sandy and, in certain levels, clayey and lignitic. Daniel ( 1963a) considered the Cherrife Forma tion an extension of the Hathira Formation of Jordan and the 'Gres de Base' of Lebanon, which are Upper Jurassic to Lower Cretaceous and
347
basal Cretaceous respectively (Daniel, 1963b; Dubertret, 1963). The 'Rutbah' reservoir of the Euphrates graben is sourced by the Upper Cretaceous, organically rich Shiranish-Soukhne Formation(s) marls; these are phosphatic as a consequence of upwelling, and were widely deposited pelagic sediments that extend in outcrop from Iraq across the Levant and northern Arabia through North Africa all along the southern Tethyan margin of North Africa. Their organic maturity level is low except where deeply buried in rapidly subsiding basins, such as the Euphrates graben, whose inception coincided with the start of the pelagic sedimentation regime (Beydoun, 1991). Juxtaposition of maturing source sediment with older reservoir sandstone facies in tilted extensional regime fault blocks in the subsiding grabens resulted in both oil charging and sealing of these generally good 'Rutbah' sandstone reservoirs (Beydoun, 1988, 1991). The presence of volcanics associated with the Euphrates rifting episode and the initial high palaeo heat-flow have together created locally 'dirty' sands, causing production problems that have necessitated remedial pressurization measures in some fields. Clastic reservoirs in other Middle East countries Elsewhere within the Middle East region, minor oil production or potential production from Mesozoic sandstone reservoirs occur in Oman, offshore Qatar and in southern Israel-Palestine. In the Oman area, producible oil has been found in a Lower Jurassic sandstone reservoir in (several) fields of the Ghaba basin of central Oman (Alsharhan & Kendall, 1986). The main production of the Saih Raw! field comes from the Upper Palaeozoic Haushi Group sandstone reservoirs and, more recently, considerable volumes of gas were estabbshed in the Lower Palaeozoic Andam Formation sandstone reservoirs. In 1978, commercial oil was discovered in the 'Marrat' reservoir (Beydoun, 1988). Hughes Clarke (1988) redefined the 'Marrat' of Oman and renamed it the Mafraq Formation (only partly equivalent to the Marrat of Saudi Arabia), with the lower Mafraq being continental and locally possessing the charac teristics of a palaeosol, whereas the upper Mafraq is shallow marine with variable thickness but including important clean reservoir sand levels (e.g. at Saih Raw!) and is regarded as being essentially a basal clastic unit to the Middle-Upper Jurassic Dhruma Formation; thus the upper Mafraq is younger than the Marrat of Saudi Arabia. Reservoir charge is
348
Z.R. Beydoun
probably from a deeply buried Huqf Group source that also charges the underlying Haushi Group reservoirs and not from organically rich units within the so called 'Marrat', as suggested by Alsharhan & Kendall (1986), nor from the overlying Middle Jurassic Dhruma Formation shales that provide the seal (Beydoun, 1988). It is appropriate here to note that Beydoun (1988), in discussing the 'Marrat' accumulation at the Saih Raw! field in Oman (p. 149) , made the passing remark that deeper exploration in producing fields in Kuwait (p. 185) had established substantial new reserves in lower Jurassic Marrat (assumed as clastic sediments in consequence of the Saih Raw I discovery in Oman and the Saudi Arabian surface section for the Marrat) and other formations. In reality, the Marrat facies in Kuwait, as in the adjacent por tion of Saudi Arabia, is predominantly carbonate (Powers, 1968), changing from the clastic facies of the surface section to the west, and the new Kuwait Marrat reserves are housed in carbonate reservoir(s). The minor offshore Qatar Mesozoic clastic reser voir accumulation occurs in an areally limited struc tural closure over a portion of the northern flank of the 'supergiant' Permian Khuff carbonate gas accumulation of the areally huge but very gentle North (Dome) field. The reservoir is a thin Nahr Umr (Albian) fine-grained sandstone, locally cemented, and tight to the southeast of the closure but shaling out westwards (Wells, 1987). Thus, the thin and areally limited reservoir sand facies, shaley at the base and top, represents the distal clastic regional facies of the Nahr Umr at this distance from the main fluvio-deltaic depocentres of the Basra Kuwait area. The hydrocarbon source for this sand stone reservoir (and also for the overlying and underlying Cretaceous carbonate permeability pinch-out reservoir accumulations on this same north flank of the North field) is the Upper Jurassic Hanifa bituminous carbonates, as discussed by Murris (1980). The migration route was complex ; initially up the flank of the Qatar Arch under the Upper Jurassic Hith anhydrite seal whence most of the oil moved towards the then culmination in southern Iran, but some was retained in palaeostructures or stratigraphical traps. Later, during late Tertiary times, when the North (Dome) structure was formed, this oil remigrated southwards, where the Hith anhydrite seal had been breached allowing vertical access into the Cretaceous reservoirs of the area (Wells , 1987).
Clastic reservoirs in the Levant- Sinai region The Levant-Sinai fragment constituted an integral part of the Arabian Plate until the Neogene. In this region two areas of Mesozoic clastic reservoir oil production occur. The first is of minor significance and is located in the coastal region of southern Israel-Palestine. Here the only sustained oil pro duction is from the Heletz-Brur-Kokhav field complex, where the predominant reservoir facies is a Lower Cretaceous sandstone with some carbon ate interbeds belonging to the Lower Cretaceous Kurnub Group. The discovery was made in 1955 on re-entering and deepening a well abandoned in Aptian sediments in 1948 by the Iraq Petroleum Co. subsidiary Petroleum Development (Palestine) because of security conditions. The trap is a faulted anticline with a major NNE-SSW boundary fault to the west and a strong stratigraphical trap component consisting of channel sands and pinch-outs. Facies range from marine in the west to fluvio-continental in the east. The source of the oil is not firmly established but appears to be 'basinal' Middle Jurassic Barnea Formation carbonates located off shore to the west, with TOC values of 0.5-2.6% of type II kerogen; other source candidates are laterally equivalent basinal Lower Cretaceous shales filling canyons cut in the Late Jurassic ( Gilboa et al., 1990). The principal reservoir intervals are sands, but a Lower Cretaceous dolomite also produces. The sandstone reservoirs vary in characteristics and environmental facies and include three main types : an offshore (marine) environment calcite-cemented sandstone; a tidal channel and/or lagoonal environ ment sand, calcite or dolomite cemented; a coastal aeolian sand, very fine-medium grained, well sorted, loosely packed, low cement content and with porosities reaching 32% (average 24%) and good permeabilities at times exceeding 2000mD; most of the recoverable oil is in the coastal environment facies and the oil gravity is 27-31° API. However, reserves are modest with 38 MM barrels in place and 19 MM barrels being the estimated recoverable; with cumulative production until 1988 standing at 16.5 MM barrels the field must be close to exhaustion (Gilboa et al., 1990). By far the most important Mesozoic clastic reser- voir oil region of the Levant-Sinai fragment of the Middle East is the Gulf of Suez. There, at least 10 fields produce oil from multiple sandstone reservoirs of Cretaceous age, including Lower Cretaceous Nubia 'A' sands, Cenomanian sands , Turonian sands
Middle East clastic oil and gas reservoirs
and lower Senonian sands (Ayouty, 1990). The depositional environments charactizing the different Cretaceous reservoirs vary. In Late Jurassic to pre Aptian time, the Gulf of Suez region was positive with erosion or non-deposition taking place. By the Aptian, the northern extremity was within a delta front domain and during the Albian, the entire region was receiving alluvial deposition, with a flood plain environment belt straddling the Gulf and extending across the adjacent Eastern Desert of Egypt and Sinai. During the Cenomanian, the Gulf of Suez, much of Sinai and the northern part of the Eastern Desert were located in an open-marine regime, the Cenomanian reservoir sands thus being mainly shallow marine in facies. This situation con tinued in the Turonian but with the southern Gulf of Suez area alternating between shallow marine and fluvial -deltaic. By the Coniacian, the alternat ing shallow-marine and fluvio-deltaic facies had expanded to include the Egyptian Red Sea and all of the Eastern Desert while the northern part of the Gulf of Suez became uplifted as a ridge, with open marine conditions to its east and north. In the Santonian, the whole of the Gulf of Suez region was within the open-marine-embayment domain bounded by positive uplifts to the north and south and southeast. In the late Campanian the Gulf of Suez region became part of the phosphatic 'flint' facies or 'shale with bone beds' facies covering much of Egypt and extending into the Sinai-Levant region and across North Africa (Said, 1990). Porosities and permeabilities in these reser voirs are variable but generally good to fair, and oil gravities vary greatly in the different reservoirs , from 20° to 39° API (Ayouty, 1990). These oils, as with those in the Upper Palaeozoic Nubia 'C' sandstone reservoir, are all thought to be sourced principally from Lower Miocene sediments with Eo-Palaeocene and Upper Cretaceous sediments being probable additional sources (Ayouty, 1990). However, as stated in the previous section, there is an alternative view postulating the Carboniferous Nubia 'B' black shales as the source, at least for the Palaeozoic and Cretaceous reservoirs. Sourcing from the younger units would require fault or uncon formity juxtaposition of source against reservoir and involve only short-distance migration.
CENOZOIC CLASTIC RESERVOIRS Only two areas o f the Middle East have established
349
Cenozoic commercial oil production from sandstone reservoirs. These are the Kuwait-Basra-Ahwaz (Iran) area at the head of the Gulf, and the Gulf of Suez rift of Egypt. Everywhere else, Cenozoic pro duction is from carbonate reservoirs, with the principal prolific producing reservoirs being the Oligo-Miocene Asmari Formation limestone of southwest Iran and its correlative of north Iraq. Towards the end of the Oligocene, the present head of the Gulf area was situated at about 25°N, a time coinciding with a global sea-level lowstand. Much of Arabia was positive and uplifted, with the exception of a narrow seaway along the line of the present Zagros range (Beydoun et al., 1992). A major river rising in the highlands of Central Arabia flowed northeast into the shallow shelf seas of the northern Gulf area, prograding a large delta in the Kuwait-Basra sector (Ghar Sandstone Formation); a pattern not unlike that initiating the deposits of the Cretaceous Zubair-Burgan-Wara Formations. From this delta, distal reservoir-quality sand tongues extended into the southwestern part of Khuzestan Province in Iran, to form the Ahwaz Sandstone Member of the Asmari (carbonate) Formation (Setudehnia, 1972; Beydoun et al . , 1992). The Ahwaz Member is considered stratigraphically equivalent to the Ghar Formation of Kuwait and southeast Iraq (Basra area) and is of imprecise Miocene age (Dunnington et al., 1959). During deposition of the lower part of the Asmari carbonates in Khuzestan, sands were being flushed from the west into this essentially carbonate shelf area; this situation continued until finally the carbonate facies of the Asmari prevailed (Setudehnia, 1972). The Ahwaz sands die out eastwards, but in a number of oil fields in southwest Khuzestan, such as Ahwaz, Mansuri and Marun, where the sands are of near shore, littoral -shoreline and aeolian environments, they constitute a dominant or significant part of the Asmari reservoir and are clean, loosely cemented, have porosities of 20-30% and permeabilities in the darcy range (Beydoun et al., 1992). The oil produced is of 30-33° API gravity and field reserves are sub stantial, with exceptional flow rates due to the excel lent reservoir properties. Clastic reservoirs in the Kuwait -Basra-Ahwaz area In the Kuwait-Basra area a few small shallow fields produce from the Miocene Ghar sandstone reservoir, but the oil is heavy and sulphurous, ranging between 18° and 20° API gravity. The fields with producible
350
Z. R.
Ghar oil are Zubair near Basra in Iraq, and Bahra and Ratga in Kuwait, and the reservoir facies is fluvio-deltaic; the heavy nature of the oil is either because of biodegradation or of early maturity of source, but the source itself remains debatable (Beydoun, 1988). Clastic reservoirs in the Gulf of Suez In the Gulf of Suez, the principal productive sand stone reservoirs of the region are of Miocene age and contain the bulk of the reserves in many fields. The main oil-bearing levels are sandstones (although occasionally some limestones too) in the Nukhul (oldest), Rudeis and Karim Formations of the Gharandal Group (Lower to Middle Miocene), representing the main Neogene clastic cycle, and in the overlying Middle Miocene Belayin Formation, which marks the beginning of the main Miocene evaporite cycle represented by the Ras Malaab Group (Ayouty, 1990; Hughes et al., 1992). A num ber of thin sand reservoir levels are also present in the overlying Middle -Upper Miocene South Gharib Formation and the succeeding mainly Upper Miocene Zeit Formation (the youngest of the Ras Malaab Group). The main oil-bearing reservoir units occur in the lower four formations (Ayouty, 1990). The main reservoirs of the Morgan, Belayim Land and Marine fields, and July, Shoab Ali and Zeit Bay fields are in the Rudeis-Karim Formations. Belayim Formation oil-producing reservoirs are found in Belayim Land and Marine fields, Morgan and Shoab Ali fields, with less important production coming from the same reservoirs in several other fields. The Nukhul sandstone reservoir is oil productive in the Rudeis, Sidri, Shoab Ali and GS-173 fields, but some production is also obtained from contempor aneous carbonate reservoirs deposited as reefs or platform carbonates laid down on submerged palaeohighs. The Nukhul Formation is the basal marine Miocene unit and reflects an irregula_r reljef in the basin floor, leading to a diversity of sediments represented by bioclastic limestones, sandstones, or shales with sabkha-type anhydrites present in the central and southern parts of the Gulf of Suez. The Rudeis consists of highly variable lithology, thickness and depositional environments, controlled by syn depositional fault-block topography and represented by marls, limestones or shales with minor sands, willch, however, dominate locally at the expense of the shales (Hughes et al., 1992). The overlying
Beydoun
Karim Formation is similarly varied in facies but broadly consists of a lower, characteristically evap oritic member and an upper, fine-coarse-grained clastic member; the fine clastic sediments are shales and marls, and the evaporite member also contains intercalated shales and marls. The evaporites are absent locally owing to lateral facies changes, so that the formation becomes mainly shale-marl in com position, with sand levels that in marginal areas become the dominant clastic deposits and hence good reservoirs; environments range from shelf edge through the sublittoral (Hughes et al., 1992). The Belayim is typically evaporite and shale, but in marginal areas and laterally it can become pre dominantly sand, with carbonates developing on submerged palaeorelief; it is mostly of a sublittoral environment, generally shallow to deeper inner sub littoral with localized lagoonal conditions (Hughes et al., 1992). The South Gharib Formation is typically a thick evaporite body with intercalations of shale and fine to very fine-grained, well-sorted sandstones. The Zeit Formation represents a shallow-marine regime with alternating restricted and open con ditions (Hughes et at., 1992). Porosities and permeabilities in these formations are variable but generally good to fair. The reservoirs are sourced by oils generated in the Lower Miocene Rudeis and Kareem Formations, as in the case of oils in Mesozoic and older reservoirs (Ayouty, 1990), although, as discussed earlier, additional input from older sources is not ruled out. The northernmost Red Sea has an identical Miocene succession to that of the Gulf of Suez (as also the remainder of the Red Sea with recognizable Gulf of Suez units; Hughes & Beydoun, 1992). So far, however, only one hydrocarbon field, Barqan, has been discovered in the northernmost sector of the Arabian side in Saudi Arabia, where light oil, gas and condensate were tested in four sand reser voirs, in Rudeis-Kareem and Belayim Formation equivalents, but the field remains undeveloped despite promising flow rates (Beydoun, 1988).
CONCLUSIONS The bulk o f oil and gas production i n the Middle East region comes from Permian to Miocene reser voirs, the overwhelming proportion of which are carbonates that were deposited when the region lay mainly in warm low latitudes where marine carbon ate sedimentation dominated; sandstone sedimen-
Middle East clastic oil and gas reservoirs
tation during that time was subordinate and was related to particular climatic or tectonic conditions. Arid or warmer, wet climatic episodes coupled with local or regional uplifts and/or sea-level lowstands resulted in substantial localized aeolian or fluvio deltaic deposition. Tectonic phases allowed detritus to inundate local, rapidly subsiding rift basins; clastic detritus from developing and rising collisonal compressional mountain belts along the northern and eastern peripheries rapidly infilled marginal seaways. During the pre- Late Permian Palaeozoic, by contrast, the Middle East region was located in the southern hemisphere in middle to high temperate (humid-cool) latitudes with only occasional influ ence from warmer belts , hence the dominance of clastic sedimentation in general and of sandstone facies in particula�;. The distribution of favourable (hydrocarbon) reservoir facies by itself is insufficient for the dis covery of commercial petroleum fields unless widely distributed organic-rich source facies are present to generate the hydrocarbons. Retention of hydro carbons in producible reservoirs also requires impervious seal facies that are suitably formed geometrically (by tectonics or stratigraphy) into a trap juxtaposed to the source. Moreover, the dis covery of these accumulations commences with the most easily accessible , invariably the shallowest (gen erally also the youngest) reservoirs. In time, older and deeper accumulations are explored for as the shallower ones are exhausted or become insufficient to meet rising demand. In the highly petroleum-rich Middle East region, the need to explore deeper than the Mesozoic reservoirs was for many years neglected because of the huge reserves established in the Mesozoic and Tertiary (predominantly carbonate) reservoirs. It is only in recent years that attention has turned to the (deeper) Palaeozoic sequences, where sandstone reservoirs predominate and recent successes in Oman and in Central Arabia have opened up a vast new exploration domain for the whole region. Exploration drilling results in Oman and in Cen tral Arabia have shown that the major new reserves of hydrocarbons housed in Cambro-Ordovician and Permo-Carboniferous sandstone reservoirs are sourced from two main regionally distributed source rock sequences of Infra-Cambrian/Early Cambrian and Early Silurian age. The former are thought to be present over much of the Gulf region and over Central Arabia and the
351
Rub al Khali basin, but in the Gulf region they are now too deeply buried and have become over mature; but the hydrocarbons they had generated in earlier geological time have probably charged Lower Palaeozoic reservoirs lying up-dip and awaiting dis covery. Elsewhere , exploration for these Infra Cambrian/Early Cambrian sourced hydrocarbons outside Oman has been minimal and their potential remains to be assessed by drilling, especially in the Rub at Khali basin. The Lower Silurian source sediments were depo sited regionally in response to global sea-level rise and they have been geochemically shown to have sourced the sandstone-hosted oils of central Arabia and, locally, of western Oman. Where they lie deeply buried under the Gulf, the consensus of specialist opinion is that they are responsible for the vast gas reserves housed in the Upper Permian Khuff car bonated reservoirs of the region (Beydoun, 1991). Post-Silurian sandstone reservoirs , located up-dip from the kitchen areas where these Lower Silurian shales are preserved , are highly prospective in an arc around the Arabian Shield, from the Rub al Khali to southwest Iraq and Jordan passing by central and northwest Saudi Arabia; similar areas in southeast Turkey and Syria up-dip from similar Lower Silurian kitchen areas are equally prospective. Many Palaeo zoic basins became inverted in the Late Palaeozoic and some of the sedimentary section was stripped off, so that better control of the palaeogeographical subcrop distributions of these source rocks is first required before exploratory drilling successes can follow routinely. The successful search for additional Mesozoic or younger sandstone reservoirs in the region depends mainly on the identification of 'disguised' rift basins, where source rocks have been buried to maturation and thus charge suitably juxtaposed sandstone reser voirs, as in the cases of Yemen and the Euphrates graben in Syria. The Jordan and other grabens merit more exploration with these models in mind. Unlike the wide areal continuity exhibited by Mesozoic-Cenozoic carbonate reservoir facies and evaporitic regional seal facies, the Palaeozoic clastic deposits generally show rapid lateral and vertical changes in continuity of the reservoir-seal couplets (although exceptions have been demonstrated in Oman and elsewhere); hence major vertical migration of hydrocarbons , either to dissipation at the surface or until stopped by younger more effec tive seals and diverted to accumulate in integral traps , can be expected. This factor will, in general,
Z. R. Beydoun
352
affect the size of many Palaeozoic accumulations, because the traps at this level will be subject to fault control and thin seals can be breached often. Nevertheless, the total hydrocarbons generated by regionally distributed source rocks are believed to be enormous and the exploration potential rewards are believed to be huge potentially.
ACKNOWLEDGEMENTS The author acknowledges with gratitude the typing and draughting support provided by the Exploration Department of Marathon International Petroleum (GB) Ltd in London during the preparation of this paper, and particularly to Mrs Wendy Milligan for drawing the figures and Miss Susan Hambrook for typing the manuscript. To Dr M.W. Hughes Clarke, sincere thanks are expressed for his continuing role as the authors' active 'sounding board' and for critically reading and discussing the contents of the paper and providing many constructive and helpful ideas and suggestions. Thanks are also expressed to Dr R. Crossley for reviewing the paper and for providing a number of helpful comments and suggestions. Gratitude is also expressed to Dr A.P. Heward for providing new insight into the Early Palaeozoic of Oman which has greatly improved that section of the paper, and for his other helpful remarks and suggestions. REFERENCES Asu-Au, M.A., FRANZ, U.A., SHEN , J . , MONNIER, F., MAHMOUD, M . D . & CHAMBERS, T.M. (1991) Hydrocar bon generation and migration in the Palaeozoic sequence of Saudi Arabia. Proceedings, 7th SPE Middle East Oil Technical Conference, November 1991, Bahrain, SPE 21376, pp. 345-356. Au, A. R. & SILDAWI, S.J. (1989) Hydrocarbon potential of the Palaeozoic pre-Khuff clastics in Abu Dhabi, U.A.E. Proceedings, 6th SPE Middle East Oil Technical Conference, March 1988, Bahrain , SPE 18009, pp. 819832. ALSHARHAN , A.S. & KENDALL, e.G. St C. (1986) Pre cambrian to Jurassic rocks of Arabian Gulf and adjacent areas: their facies, depositional setting and hydrocarbon habitat. Bull. Am. Assoc. petrol. Geol. , 70, 977 -1002. AYOUTY, M.K. EL- (1990) Petroleum geology. In: The Geology of Egypt (Ed. Said, R.) pp. 567-722. Balkema, Rotterdam. BEYDOUN, Z.R. (1964) The stratigraphy and structure of the eastern Aden Protectorate. Overseas Geology and Mineral Resources Bulletin , Supplement 5, HMSO, London, 107 pp.
Z.R. (1966) Eastern Aden Protectorate and part of Dhufar. In: Geology of the Arabian Peninsula. U.S. Geol. Surv. Prof. Pap., 560-H, 49 pp. BEYDOUN, Z. R. (1988) The Middle East: Regional Geology and Petroleum Resources . Scientific Press, Beaconsfield, 292 pp. BEYDOUN, Z. R. (1991) Arabian Plate hydrocarbon geology and potential - a plate tectonic approach. Am. Assoc. petrol geol. Stud. Geol. , 33, 77 pp. BEYDOUN, Z . R. & GREENWOOD, J.E.G.W. (1968) Aden Protectorate and Dhufar. In: Lexique Stratigraphique International, Vol. III, Asie, Protectorat d'Aden et Dhufar (Ed. Dubertret, L.). CNRS Paris, fasc. 10b2, 128 pp. BEYDOUN, Z.R., HUGHES CLARKE, M.W. & STONELEY, R. (1992) Petroleum in the Zagros Basin: a late Tertiary Foreland Basin overprinted onto the outer edge of a vast hydrocarbon-rich Palaeozoic-Mesozoic passive-margin shelf. In: Foreland Basins and Foldbelts (Eds Macqueen, R.M. & Leckie, D.A.), Mem. Am. Assoc. petrol. Geol ., Tulsa, 55, 309-339. BRENNAN, P. (1990) Greater Burgan field. In: Treatise of Petroleum Geology, Atlas of Oil and Gas Fields; Struc tural Traps I, Tectonic Fold Traps (Compilers Beaumont, E.A. & Foster, N . H.), pp. 103-128. American Associ ation of Petroleum Geologists, Tulsa. BRENNAN, P. (1991) Raudhatain field - Kuwait, Arabian Basin. In: Treatise of Petroleum Geology Atlas of Oil and Gas Fields; Structural Traps V, ( Compilers Foster, N . H. & Beaumont, E.A.), pp. 187-210. American Association of Petroleum Geologists, Tulsa. DANIEL, E.J. (1963a) Syrie Interieure. In: Lexique Stra tigraphique International, Vol. III, Asie, Liban, Syrie, Jordanie (Ed. Dubertret, L.), fasc. 10 cl, pp. 157-291. CNRS, Paris. DANIEL, E.J. (1963b) Jordanie (Ensemble). In: Lexique Stratigraphique International, Vol. III, Asie, Liban, Syrie, Jordanie (Ed. Dubertret, L.) fasc. 10 cl , pp. 293-399. CNRS, Paris. DERCOURT, J . , ZoNENSHAJN, C.P., Rtcou, L.-E., KAZMtN, V.G., LE PICHO N , X., KNIPPER, A.L., et al. (1986) Geologic evolution of the Tethys belt from the Atlantic to the Pamirs since the Lias. Tectonophysics, 1221, 241-315. DuBERTRET, L. (1963) Liban, Syrie. In: Lexique Stra tigraphique International, Vol. III, Asie, Liban, Syrie, Jordanie (Ed. Dubertret, L.) fasc. 10 cl, pp. 7-155, CNRS, Paris. D u N N INGTON , H.V., VAN BELLEN, R.C., WETZEL, R. & MoRTO N , D.M. (1959) Iraq. In: Lexique Stratigraphic International, Vol. III, Asie, Iraq (Ed. Dubertret, L.). CNRS, Paris, fasc. lOa, 333 pp. FERGUSON , G.S. & CHAMBERS, T.M. (1991) Subsurface stratigraphy, depositional history and reservoir devel opment of the early to late Premian Unayzah Formation in central Saudi Arabia. Proceedings, 7th SPE Middle East Oil Technical Conference, November 1991, Bahrain, SPE 21394, pp. 487-496. GILBOA, Y. , FLIGELMAN , H. & DERJN, B. (1990) Heletz- Brur-Kokhav field - Israel. In: Treatise of Petroleum Geology, Atlas of Oil and Gas Fields, Structural Traps IV, Tectonic and non-Tectonic Fold Traps (Compilers Beaumont, E.A. Foster, N.H.) pp. 319-345. American
BEYDOU N ,
Middle East clastic oil and gas reservoirs
Association of Petroleum Geologists, Tulsa. GORIN , G.E., RAcz, L.G. & WALTER, M.R. (1982) Late Precambrian-Cambrian sediments of Huqf Group, Sultanate of Oman. Bull. Am. Assoc. petrol. Geol. , 66, 2609-2627. G RANDVILLE, B.F. D E LA (1982) Appraisal and develop ment of a stratigraphic and structural trap oilfield with reservoirs in glacial and periglacial clastics. In: The Deliberate Search for the Subtle Trap (Ed. Halbouty, M.T.) Mem. Am. Assoc. petrol. Geol., Tulsa. 37, 267-286. GRANTHAM, P.J., LUMBACH, G.W.M., POSTHUMA, J., H uGHES CLARKE, M . W . & W !LLIN K , R.J . (1988) Origin of crude oils in Oman. I. petrol. Geol. , 1 1 , 61-80 . H ARPUT, O.B. & E RTURK, 0. (1991) The organic geo chemical evolution of the 9th and lOth Districts of SE Anatolia. J. Southeast Asian Earth Sci. 5, 421-428. H EWARD, A.P. (1989) Early Ordovician alluvial fan deposits of the Marmul oil field, South Oman. J. geol. Soc. London, 146, 557-565. H EWARD, A.P. (1990) Salt removal and sedimentation in southern Oman. In: The Geology and Tectonics of the Oman Region (Eds Robertson, A.H.F., Searle, M.P. & Ries, A.C.) Spec. Pub. Geol. Soc. London, No. 49. pp. 637-652, Geological Society of London, Bath. H uGHES, G . W. & BEYDO U N , Z.R. (1992) The Red Sea Gulf of Aden: biostratigraphy, lithostratigraphy and palaeoenvironments. J. petrol. Geol. , 15, 135-156. H uGHES, G.W., AB D I N E , S. & G IRGIS, M.H. (1992) Mio cene biofacies development and geological history of the Gulf of Suez, Egypt. Mar. petrol. Geol. , 9, 2-28. H uGHES C LARKE, M.W. (1988) Stratigraphy and rock-unit nomenclature in the oil-producing area of Interior Oman. I. petrol. Geol. , 11, 5-60. H usSEIN!, M.l. AL- (1992) Potential petroleum resources of the Palaeozoic rocks of Saudi Arabia. Proceed ings, 13th World Petroleum Congress, Buenos Aires, November 1991 , Vol. 2, pp. 3-13. H UURDEMA N , A.J.M., BREUNESE, J.N., ASBAHI, A.AL-, L UTGERT, J.E. & FLORIS, F.J.T. (1989) Assessment of halite-cemented reservoir zones with a combined geo logical/engineering approach; a case study (Yemen). Proceedings, 64th SPE Annual Technical Conference, San Antonio, TX, October, 1 989, SPE 19603, pp. 335343. IBRAHIM, M.W. (1981) Lithostratigraphy and subsurface geology of the Albian rocks of South Iraq. J. petrol. Geol. , 4, 147-162. IBRAHIM, M.W. (1983) Petroleum geology of southern Iraq. Bull. Am. Assoc. petrol. Geol. , 67, 97-130. J AWAD Au, A. & Az1z, Z.R. (1993) The Zubair Forma tion, East Baghdad oilfield. J. petrol. Geol. , 16, 353-364. KHAIWKA, M.H. (1989) Order and repetitive nature of the Zubair deltaic facies, central and southern Iraq and northern Kuwait. Mar. petrol. Geol. , 7, 38-43. KUTZCH, E. (1990) Palaeozoic. In: The Geology of Egypt (Ed. Said, R.), pp. 393-406. Balkema, Rotterdam. LABABIDI, M.M. & H AMDAN , A . M. (1985) Preliminary Lithostratigraphic Correlation Study in OAPEC Member Countries. Organization of Arab Petroleum Exporting Countries (OAPEC), Kuwait, 171 pp. LABOUN, A.A. AL- (1986) Stratigraphy and hydrocarbon
353
potential of the Palaeozoic succession in both Tabuk and Widyan Basins, Arabia. In: Future Petroleum Provinces of the World (Ed. Halbouty, M.T.) Mem. Am. Assoc. petrol. Geol., Tulsa, 40, 373-397. LEVELL, B.K., B RAAKMAN , J.H. & R UTTE N , K.W. (1988) Oil-bearing sediments of Gondwana glaciation in Oman. Bull. Am. Assoc. petrol. Geol. , 72, 775-796. LouTH, G. & BISHLAWY, S. EL- (1989) Geology and hydro carbon potential of the de.ep Palaeozoic sequence in Abu Dhabi. In: OAPECIADNOC Seminar on Deep Forma tions in Arab Countries: Hydrocarbon Potential and Exploration Techniques, Abu Dhabi, October 1 989, Proceedings Technical Papers, pp. 85-129. MAHMOUD, M.D., VASLET, D. & H u sSEIN !, M . l. (1992) The Lower Silurian Qalibah Formation of Saudi Arabia. Bull. Am. Assoc. petrol. Geol. , 76, 1491-1506. MARJEBY, A . AL- & NASH, D. (1986) A summary of the geology and oil habitat of the Eastern Flank Hydrocar bon Province of South Oman. Mar. petrol. Geol. , 3, 306-314. MATTES , B.W . & CoNWAY MoRRis, S. (1990) Carbonate/ evaporite deposition in the Late Precambrian-Early Cambrian Ara Formation of southern Oman. In: The Geology and Tectonics of the Oman Region (Eds Robertson, A.H.F., Searle, M.P. & Ries, A.C.) Spec. Publ. Geol. Soc. London, No. 49, pp. 617-636. Geo logical Society of London, Bath. MAYCOCK, I. (1987) Oil exploration and development in Marib/AI Jauf Basin, Yemen Arab Republic. Proceed ings, 5th SPE Middle East Oil Technical Conference, March 1987, Bahrain , SPE 15685 (unpublished). McCLURE, H.A. (1980) Permo-Carboniferous glaciation in the Arabian Peninsula. Geol. Soc. Am. Bull. , 91, 707-712. McGILLIVRAY, J.G. & H u ss EIN !, M.I. (1992) The Palaeo zoic petroleum geology of Central Arabia. Bull. Am. Assoc. petrol. Geol. , 76, 1473-1490. MILLS, S.J. (1992) Oil discoveries in the Hadramaut: how Canadian Oxy scored in Yemen. Oil Gas J. , 9 March, 49-52. MuRRIS, R.J. (1980) The Middle East: stratigraphic evol ution and oil habitat. Bull. Am. Assoc. petrol. Geol. , 64 , 597-618. NRA (NATURAL RESOURCES A UTHORITY) J ORDAN (1989) Petroleum Exploration Opportunities in Jordan , Amman, 39 pp. OAPEC (O RGAN IZATION OF A RAB P ETROLEUM E XPORTING Cou NTRIES) (1985) Register of Exploration Wells and Arab Petroleum Fields (in Arabic). OAPEC, Kuwait, 500 pp. OwEN , R.M . S . & N ASR, S.N. (1958) Stratigraphy of the Kuwait-Basra area. In : Habitat of Oil (Ed . Weeks, L.G.), AAPG Symposium , pp. 1252-1278. Tulsa, Oklahoma. PowERS, R.W. (1968) Arabie Saoudite. In: Lexique Stra tigraphique International, Vol. III, Asie (Ed. Dubertret, L.). CNRS, Pa1is, fasc. 10b1, 177 pp. SAID, R. (1990) Cretaceous palaeogeographic maps. In: The Geology of Egypt (Ed. Said, R.), pp. 439-449. Balkema, Rotterdam. SAMAHUI, J .A. & CHAUBE, A.N. (1987) Evolution of Bahrain Field in relation to western Arabian Gulf Basin. Proceedings, 5th SPE Middle East Oil Technical Confer-
354
Z.R. Beydoun
ence, March 1987, Bahrain, SPE 15697, pp. 97-107. C.R . & McKERROW, W.S . (1990) Revised world maps and introduction. In: Palaeogeography and Bio geography (Ed. McKerrow, W.S. & Scotese, C.R. ) , Mem. geol. Soc. London, No. 12, pp. 1-21 . Geological Society of London, Bath. SCOTESE, C.R., BAMBACH, R . K . , BARTO N , C., VAN DER Voo, R. & ZEIGLER, A.M. (1979) Palaeozoic base maps. J. Geol. , 87, 217-277. SETU DEHNIA, A.O. (1972) Iran du Sud-Ouest. In: Lexique Stratigraphique International, Vol. III, Asie, (Ed. Dubertret, L.) fasc. 9b, pp. 285-376. CNRS, Paris. STONELY, R. (1990) The Middle East basin: a summary overview. In: Classic Petroleum Provinces (Ed. Brooks,
ScOTESE,
J.) Spec. Pub!. Geol. Soc. London, No. 50, pp. 293298. Geological Society of London, Bath. R.M. & All u RISHEH, A . K. (1989) Exploration of deep Palaeozoic and Pre-Cambrian plays in the Sultanate of Oman. In: OAPECIADNOC Seminar on Deep For mations in the Arab Countries: Hydrocarbon Poten tial and Exploration Techniques, Abu Dhabi, October 1989, Proceedings, Country Reports and Case Studies, pp. E71 -113. WELLS, P.R.A. ( 1987) Hydrodynamic trapping of oil and gas in the Cretaceous Nahr Umr Lower Sand of the North Area, offshore Qatar. Proceedings, 5th SPE Middle East Oil Technical Conference, March 1 987, Bahrain, SPE 15683, pp. 17-26. SYKES,
Spec.
Pubis int. Ass. Sediment. (1995) 22, 355-380
The evolution of Oligo-Miocene fluvial sand-body geometries and the effect on hydrocarbon trapping: Widuri field, west Java Sea
R AY Y O U N G ,* W. E. H A RM O NYt
and
T H OM A S B U D I Y E N T Ot
*Maxus Ecuador Inc. , PO Box 650496, Dallas, Texas 75265-0496, USA; and tMaxus Southeast Sumatra Inc., Five Pillars Office Park, ]alan Letjen MT Haryono 58, PO Box 2759, Jakarta, Indonesia
ABS T R A C T
Widuri field is a faulted anticline located in a Tertiary back-arc basin at the western end o f the Java Sea in the Indonesian archipelago. The discovery well was drilled in early 1988 to a depth of 3735 ft subsea (1139 m SS) and penetrated 170ft (52 m) of net oil pay in late Oligocene and Early Miocene sandstones. Eight delineation wells, 47 development wells and five sidetracks confirmed a large oil field over an area of approximately 7500 acres. The oil is trapped in six separate reservoirs over a vertical interval of 325ft (99 m). The lowermost reservoir is a thick, immature, fluvial sandstone with multistorey and multilateral stacking, high permeabilities and a sheet geometry; the uppermost reservoir is a series of thin, lower permeability, 2000 ft (610 m) wide, distributary channel sandstones associated with a tide-influenced delta. The intermediate reservoirs show a complete gradation in size, geometry and reservoir quality between these two extremes. This gradual change from an immature fluvial system, through progressively more mature fluvial systems into deltaic sediments at the top is the result of a gradual relative rise in sea-level. The systematic upwards decrease in sand-body width is accompanied by a change in the trapping mechanism, from structural in the two sheet-like reservoirs at the base to stratigraphical-structural in the upper four reservoirs, where the existing structural closure is enhanced significantly through each reservoir having a restricted lateral extent and a favourable orientation with respect to structure. However, the Widuri reservoirs share a common oil-water contact defined by the spill point of the stratigraphically highest of the sheet-like reservoirs. The reservoirs must therefore be in communication and, without a stratigraphical component to trapping in every one of the overlying four reservoir intervals, the field oil-water contact would be significantly higher and the field size very much smaller. The understanding of sand body geometry and trapping has led to efficient field development (only two of the original holes needed to be side-tracked owing to lack of pay) and will form the basis for future reservoir management. From an exploration viewpoint , the Widuri work has provided analogue models in the search for other stratigraphically trapped oil in the contract area .
I N T R O DU C T I O N
500 million barrels of oil in late 1989 but virtually all of this oil came from the Sunda basin. It was not until late 1987 that the first oil was discovered in the Asri basin and this discovery subsequently led to the establishment of five new fields on the northwest flank of the basin (Fig. 2). Widuri field is by far the largest of these accumu-
Widuri field is located in the Southeast Sumatra Production Sharing Contract (PSC) area approxi mately 100 miles ( 161 km) north of Jakarta in the west Java Sea (Fig. 1). The PSC comprises approxi mately 5300 square miles ( 13 725 km2) and consists of two principal areas, the Sunda and Asri basins. Cumulative oil production from the PSC surpassed
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0
355
35 6
R. Young, W.E. Harmony and T. Budiyento
SOUTH CHINA SEA Natuna
JAVA SEA
0
500 Miles
Fig. 1. Location of the Pertamina-Maxus , Production Sharing contract area showing the relative position of the Sunda and Asri basins and the distribution of oil fields.
lations and was discovered in early 1988 when Widuri-1 was drilled to a total depth of 3735 ft subsea (1139m SS) and penetrated 1 70ft (52m) of net oil pay in multiple sandstones of late Oligocene and Early Miocene age. Eight delineation wells, 47 development wells and five sidetracks (as of mid1992) confirmed a giant oil field with over 500 million barrels of original oil in place spread over approxi mately 7500 acres. Recovery efficiency is expected t9 be excellent and the discovery was the largest in Indonesia since 1974. The oil is a high pour point, low sulphur crude with a gravity of 31 API. First oil was produced in December 1990 and there are currently five platforms (A-E, Fig. 3). To optimize reservoir management it was decided to complete wells in single zones only and this strategy meant that a detailed and predictive geological model was vital. The primary aim of this paper is to show the o
relationship between reservoir geometry and struc ture at Widuri field and to demonstrate the resulting hydrocarbon trapping configurations.
BA C K G R O U N D
G E O L O GY
The PSC area is situated in a Tertiary back-arc basin and the history of tectonics in the region is therefore essentially extensional but with some wrench modi fication (see also Hutchison, 198 6). Structurally, the Asri basin is an asymmetrical syncline or half-graben with a roughly N-S axial trend (Fig. 2) and a maximum sedimentary fill of approximately 15 OOO ft (4773 m) of Oligocene and younger sediments. The oil accumulations are trapped in upthrown closures related to NE-SW trending, down-to-the-northwest, normal faults on the gently sloping northwest flank of the basin. These faults probably originated in
Fluvial sand-body geometries
I i
______
_j
Cl 50 ms 10 Kms 10 Miles Fig. 2. Asri basin time-structure map on the top Talang
Akar Formation.
Fig. 3. Depth structure map on the top 34-2 reservoir showing (1) that Widuri field is a low-relief, faulted closure, (2) the location of the five platforms and 62 wells in the Widuri area and (3) the location of seismic line 88-196 (Fig. 4).
i i i i i I i i i I
357
early Oligocene times during basin formation but were subsequently reactivated at several times throughout the Tertiary. Widuri field is a low relief, three-way-dip, faulted closure (Figs 3 & 4), which formed approximately 19Ma (Early Miocene) and has not been altered significantly since. The stratigraphy in the Widuri field area is shown in Fig. 5. Lower Cretaceous, metamorphic basement was penetrated only over the A-platform area, which is the structurally highest part of the field. Basement is onlapped and overlain by upper Oligocene fluvial sandstones, shales and coals of the Zelda Member, Talang Akar Formation. The Lower Miocene Gita Member is a fluvio-deltaic succession of sandstones and shales with thin interbedded limestones and coals. The boundary between the Zelda and Gita members is the top of the Oligocene succession and is an unconformity over most of the Asri basin. At Widuri field, however, the boundary is probably a discontinuity that is associated with an important field-wide coal marker - Coal 'A' (see Figs 4 & 5). The Talang Akar sandstones are currently the only commercial reservoirs in the Asri basin. The top of the Talang Akar Formation is an unconformity that is generally marked by a relatively high gamma-ray (uranium) reading (Fig. 6). The overlying Batu Raja Formation has yielded non commercial oil shows from thin argillaceous lime stones and sandstones, but the predominant lithology is mudstone which acts as a local seal. The regional
358
R. Young, W.E. Harmony and T. Budiyento
NNW
Widuri Field (approx. 3 miles)
SSE TWT 0.7
.2
Fig. 4. Seismic line 88-196 over Widuri field (see Fig. 3 for location).
seal for both the Sunda and Asri basin is claystone of the overlying Gumai Formation (Bushnell & Temansja, 1986). The remainder of the succession consists of Miocene, shallow-marine shales, lime stones and sandstones, which are overlain by pre dominantly alluvial sediments and volcanics of the Plio-Pleistocene. An adequate source rock for the oil accumulation has not yet been penetrated in the Asri basin. However, by analogy with the Sunda basin, it is believed that lower Oligocene lacustrine shales are present in the central part of the basin and that these shales generated oil from Early Miocene until at least Late Miocene time.
R E S E RV O I R
OVE RVI E W
Figure 6 illustrates the full Talang Akar succession in the Widuri A-1 well, one of the few wells in the field that penetrated each of the main reservoirs.
The oil is trapped in six different sandstone units (from base to top the 36-1, 35-2, 35-1, 34-2, 34-1 and 33-series sandstones) over a vertical interval of approximately 325 ft (99 m). The reser voirs almost certainly share a common oil-water contact at 3680ftSS (1122 m SS - see below), but there is no evidence for direct, sand-on-sand com munication between any two reservoirs, nor any prediction that this might occur. However, the reservoirs must either be in communication or have been in communication at some time in the past through faulting, fractures or seepage. So far, however, the production history indicates that each reservoir is acting as a separate sealed unit. Reservoir descriptions are based on over 650ft (198 m) of core from seven wells. The sandstones vary from coarse- to very fine-grained and both conglomerate and granule layers are common . Compositionally, the sandstones are mature, with 75% being quartz arenites and the remainder split between sublitharenites and subarkoses. There does
Fluvial sand-body geometries
c::
� ::2: Q)
0
.<::: <.) 0
.0
c.
LL
c:: Q)
·a;
.<:::
c.
"'
Q; c. c. :::J
:J .0 :J
u
Parigi
Q) :0
..,.
.f
.
Q) c:: Q) <.)
2
"0
::2:
-
::2:
Q; c.
�
"' "'c::
c. :::J
-"
Q) co
Q;
;:0
<(
-.:::"'> UJ
--'
.f N c::
Q)
<.)
0
.2' 0
Fig. 5. Stratigraphy over Widuri field.
<.:)
�
"'
--'
Early to late Cretaceous
not appear to be any significant variation in detrital composition between the different reservoir units or between the different sedimentary facies. The maturity of the framework mineralogy probably reflects the provenance (see Folk, 1968) and this is supported by the preponderance of metasedimentary basement in the Sunda and Asri basins (see also Ben-Avraham & Emery, 1973). Clay minerals dominate the non-framework mineralogy but usually only account for less than 5% of the bulk rock volume in the majority of the reservoir quality sandstones. Kaolinite predominates with subordinate illite and chlorite, and minor smectite and illite-smectite mixed-layer clay. All
c:: "' (ij f-
"'
"0 (ij N
�
. �
-
�
\.
�� <::::;:!>
r=2. "'. ,,
� �'(
• •
":
· ·' • •
· ·.· . , . .
� :
.
..- :
. . · . .·
•
D
v I
•
�
.... ..... /�� t,,t,..._t, ..... :,..., l, ,t,,/,,;,,
/,
Basement
!}
... ....
... .. . .... :
= --�
<{ Cl
Q)
......
;:··_
��
Batu Raja
;;; -"
�
-
-� '""
'
�
Gumai
:J
\
- .....--r-(.
-
� �
"'
Q; Q; > c:: 5 -E "' .!: 0 co
-=:,:.:.:::l
--'
--'
"'
-�
;:0
Sub littoral
l�
v
Q;
"' 0 :::: � :.:; 0 "' :::: c. :.:; (f)
'· . . .
.
.·
<Jl
Q) 'iii
e
. "� . . ..: "--� "
-
�B "- <Jl
"'<.)
"0 > I
"
Q)
-e
Lithology
E Q) ::2:
E 0
Paleoenvt.
c:: 0
Q;
·;:;"'
UJ
Cl <{
359
/,
·. · ·
�: ,<. � ,"', : \;� � \� �
....,,_,,_,,_ ,,/, .. // <.., " / -:./
other mineral cements are extremely sparse and the sandstones are generally unconsolidated. Reservoir quality is usually excellent but the measurement of porosity and permeability from core data has inherent errors because the reservoirs are uncon solidated. However, log porosities average 29% (using a 15% cut-off) and permeabilities from production tests on initial completions range up to 30 darcies (D). A reconstruction of the palaeogeography of the area during Talang Akar times is shown in Fig. 7 (see also Beddoes, 1981). The basin margin is defined seismically by the onlap of the Talang Akar Formation on to basement (Fig. 7) and the sediment
360
R. Young, W.E. Harmony and T. Budiyento
Fig. 6. The type well, Widuri A-1.
transport directions were then defined using interval isopachs (see also Wight et a!., 1986) and individual channel isopachs. The Talang Akar clastic sediments in the Widuri area of the Asri basin were derived exclusively from a northerly direction.
RES ERVOIR
GEOMETRY AND
TRAP
A discussion o f the reservoir character, geometry and trap for each of the six main reservoir units follows. These descriptions and interpretations are based on a data set that includes a 250-m grid of very . good quality two-dimensional seismic data, a mostly full set of log data from approximately 60 wells, and a total of over 650ft (198 m) of full-hole core from seven wells. A detailed discussion of the sand-body geometries and reservoir model was the subject of a previous publication (Young et a!., 1991) and it is not the intent of the present article to duplicate this. However, reservoir geometry is so critical to the
hydrocarbon trapping that an updated and abbrevi ated version is included. 36-1 sandstone
The 36-1 sandstone is the lowermost of the Widuri reservoirs and is oil-bearing only over the structurally higher A- and D-platform areas. Few down-dip wells penetrated the entire 36-1 sandstone and therefore less information is available compared with other reservoirs. However, the reservoir was fully cored in the discovery well, Widuri-1, where it is 43 ft (13 m) thick and rests directly on top of metamorphic base ment (Fig. 8). The base of the sandstone is erosive and the top has a sharp contact with the overlying rootleted siltstone and coal. Grain size is coarse to very coarse and sedimentary structures are domi nated by trough cross-bedding in 1-2-ft (0.3-0.6 m ) sets. The sandstone does not show any overall grain size trend but at least five sedimentary units can be recognized based largely on grain size (Fig. 8). Each
Fluvial sand-body geometries
361
Java Sea
r.::l �
Sediment source areas/Sunda Shield emergent (Talang Akar FM absent) Sediment transport directions Asri Basin oil fields
Fig. 7. Local palaeogeography
during Talang Akar Formation times.
unit fines upwards slightly and they are interpreted to represent a multistorey fluvial system. Elsewhere in the field, sidewall cores indicate a more variable grain-size with a slightly finer average, and the thickness reaches a maximum of 185 ft (56 m) in Widuri E-1 (see Fig. 8). Log character is commonly blocky but occasionally, as in the case of Widuri-6 (Fig. 8), it is clear that the 36-1 is made up of several different sandstones. However, this distinction between individual sandstones cannot usually be traced with any confidence beyond adjacent wells and the thick, blocky nature of the 36-1 over most of the field is almost certainly the result of both vertical and lateral channel amalgamation. The 36-1 sandstone overlies and onlaps basement and the net sandstone isopach (Fig. 9A) defines a relatively thick, sheet sandstone, which is more extensive than the available well control in all areas except the extreme southwest of the field. The sand stone is interpreted to be a coarse grained, immature, multistorey and multilateral fluvial sheet-sandstone deposited in a high gradient, high bedload river. Reservoir quality is excellent; porosities average 28% and production test permeabilities range from 20 to 30D.
Figure 9B is a depth structure map on top of the 36-1 sandstone. The structural spill point is located approximately 5000ft ( 1525 m) north-northwest of the D platform at a depth of 3740ftSS (1140m SS) and the maximum structural closure is shaded. The oil- water contact for this reservoir is at 3680 ft SS ( 1122 m SS) and this plots inside the closing contour. The trapping mechanism for the 36-1 reservoir is, therefore, clearly structural but the structure is only partially filled. In other words, there is more struc tural closure than oil column. 35-2 sandstone
The 35-2 sandstone is separated from the 36-1 by an average of 20ft (6 m) of mudstone, siltstone and coal. The reservoir has been cored in three wells but long sections of core that are representative of the 35-2 sandstone field-wide are not available. The sandstone, however, is similar to the 36-1, but with some important differences. Grain size is more vari able and slightly finer (field-wide average is medium to coarse grained) but the sedimentary structures are still dominated by trough cross-beds in sets up to 2 ft (0.6 m) thick. Bed contacts are still abrupt and
362
R. Young, W. E. Harmony and T. Budiyento A
6000'
11 000' Widuri E-1 20
API
Widuri-7 120
20
API
5000'
10 000'
Widuri-4 120
20
120
API
A'
Widuri-1 20
API
Widuri-6 120
20
API
120
0
50
100
FT
150 Basement of weathered basement 200
GAMMA RAY
0
API
OEP TH (FT)
100
Core log 36-1 sandstone Widuri-1 Rootleted siltstone and coal
LOCATION MAP
Trough cross-bedded, pebbly, coarse to very coarse grained sandstone in at least five different sedimentary units
Metamorphic basement and weathered basement
Fig. 8. 36-1 reservoir: vertical and lateral sedimentary character.
evidence for multistorey and multilateral stacking of sand bodies is common (see Young et at., 1991); in one instance, an individual channel sandstone can be traced for up to 3000ft (915 m) along depositional
dip using log character correlation. The most sig nificant difference to the 36-1, however, is the shape of the sand body. The net sandstone isopach (Fig. lOA) defines a sandstone that is thinner and
363
Fluvial sand-body geometries
+
+
+
+
�0
+
+
+
,_� c�
+ ,co>be:, c;�
�0
�7 ,r-.:\;: {, ,
+
'
j 6
�-�-
+
+
+
+
+
� �...
0
Cl 20' 5000 FT
A
Fig. 9. 36-1 reservoir. (A) Net
sandstone isopach with the oil water contact marked. (B) Depth structure map on the top 36-1 sandstone showing that the oil water contact is 60ft above structural closure (shaded). The structural spill point is marked by a triangle.
less extensive. Net sandstone thickness from well penetrations varies from 0 to 67ft (0-20m) and a sand body is defined (or predicted from thickness trends) that is elongated roughly N-S and has an average width of approximately 11 000 ft (3354 m) with a zero-edge to the west and an inferred one to the east. The abrupt change to a westwards depo sitional strike at the north end of the sand body is defined by well penetrations to the north and west outside of Widuri field.
Cl100' 5000 FT
a
The 35-2 sandstone is interpreted to be a multi storey and multilateral fluvial system that is less extensive, slightly finer grained, and therefore probably slightly more mature than the 36-1 system. However, the sandstone was still probably deposited in a relatively high gradient, high bedload river. Reservoir quality is excellent, with log porosity averaging 29% and test permeabilities ranging from 20-30D. Figure lOB is the depth structure map on top of
364
R. Young, W.E. Harmony and T. Budiyento
+
+
+
+
+
+
+
+
+
+
+
Cl 20' 5000 FT A
Cl100' 5000 FT B
the 35-2 sandstone. The spill point has -remained in approximately the same position but is now at 3680 ft SS ( 1122 m SS) and this is coincident with the oil-water contact for this reservoir. The trapping mechanism is therefore again clearly structural, but this time the structure is full to spill. 35-1 sandstone
The Widuri E-1 core (Fig. 11) is lacking the basal
Fig. 10. 35-2 reservoir. (A) Net sandstone isopach with the oil water contact marked . (B) Depth structure map on the top 35-2 sandstone showing that the structural closure (shaded) is coincident with the oil-water contact. The structural spill point is marked by a triangle.
6 ft (1. 8 m ) of the 35-1 sandstone but adequately demonstrates the sedimentary character of the reservoir. The 35- 1 is an erosively based sandstone, which is trough cross-bedded in 1-2-ft (0. 3-0. 6 m) sets, and which averages medium grain size. Intervals of coarse grain size are common towards the base of the sandstone and fine/very fine grain size common towards the top, but the sandstone does not fine upwards systematically and several distinct units with abrupt contacts are evident (see Fig. 11). This
Fluvial sand-body geometries
365
A
A' D-8
1800'
20 API 120
�
20
D-6 API
D-5
3000' 120
L.-.-----1
20
API 120
L.-.-----1
B-3
3150' 20
4250'
API 120
A-1
4400' 20
20 API 120
.______.
B-6
1------4
API
C-3
3900' 120
�
20
API 120
�
CORE LOG 35-1 sandstone Widuri E-1 LOCATION MAP
Rootleted shale and coal
3750
3760
Predominantly medium to coarse grained,. trough cross-bedded sandstone Abundant mudstone clasts and carbonaceous debris
Fig. 11. 35-1 reservoir : vertical and lateral sedimentary character.
multistorey character is readily recognizable in all other 35-1 cores and the units are separated occasionally by coals (see Widuri A-1, Fig. 11). The 35-1 sandstone is the most variable of the Widuri reservoirs (see cross-section A-A', Fig. 11). The log character is predominantly gamma-ray increasing upwards (fining upwards) with a gra dational top, but is also commonly blocky with a sharp top. The most striking feature of the sand stone, however, is the rapid lateral thickness changes within an overall range of 0-64 ft (0-19.5 m).
Fortunately, these thickness vanatwns could be mapped and predicted seismically prior to any development drilling. The 35-1 interval is extremely well constrained between two readily correlatable markers - Coal 'A' at the top and usually a double coal at the base (see cross-section A-A', Fig. 11). Acoustic impedance contrasts associated with these coal markers are responsible for two seismic reflec tion which envelope the 35-1 sandstone and can be traced field-wide and beyond. The isochron between these two seismic markers, when linked with
366
R. Young, W.E. Harmony and T. Budiyento
well penetrations, proved to be a very effective predictor of net sandstone distribution (see Young et al., 1991). The net sandstone isopach (Fig. 12A) demon strates an anastomosing pattern and the 35-1 sandstone is interpreted to be a multistorey and multilateral fluvial system, possibly representing anastomosing rivers. Individual channels can be mapped with more confidence and over a larger area than those of the 35-2 and 36-1 (e.g. the sand body
trending southwest from the D platform is a single storey channel), and sand body widths vary from approximately 3000ft (915 m) in the west to 6500ft (1982 m) in the east. Finer grain size, thinner and narrower sand bodies, and more order to the sequence suggest a more mature fluvial system than the 35-2. Reservoir quality is again excellent, with log porosity averaging 29% and test permeabilities ranging from 10 to 20 D. The depth structure map on top of the 35-1 sand-
+
+
Cl100' 5000 FT
a
Fig. 12. 35-1 reservoir. (A) Net sandstone isopach with the oil water contact marked. (B) Depth structure map on the top 35-1 sandstone showing that the oil water contact is 20ft below structural closure (shaded).
Fluvial sand-body geometries
stone (Fig. 12B) defines a structural spill point or closing contour at 3660ft SS (1116 m SS). The oil-water contact for this reservoir, however, is the same as for the 36-1 and 35-2 at 3680 ft SS ( 1122 m SS). There is therefore 20ft (6 m) more oil column than can be accounted for by structural closure. Returning to Fig. 12A, it is clear that the structural contour of the oil-water contact does not close to the northwest but that closure is effected by the sand body pinch-out to the northwest. There is therefore a stratigraphical component to trapping and a combination stratigraphical-structural trap results. 34-2 sandstone
The 34-2 sandstone rests directly on top of Coal 'A' and the sedimentary character is demonstrated in core from the B-8 well (Fig. 13). The base is erosive on top of Coal 'A' and the thin, very coarse grained, basal lag contains coal clasts. The overlying sand stone consists of predominantly medium- to coarse grained sandstone in the basal and middle section before passing into fine- and very fine-grained sand stones in the upper part, and finally into rootleted siltstone and coal at the top. Trough cross-bedding in 1 ft (0. 3 m) sets is the dominant sedimentary struc ture. Maximum penetrated thickness for the 34-2 sandstone is 54ft (16.5 m) and in the thicker sections the log character tends to be blocky, with a gamma ray increasing (fining upwards) signature at the very top (see Fig. 13). In the thinner sections, however, the log character is mostly gamma-ray increasing (fining upwards). Immediately above the top coal is the first indication of fully marine conditions in the Talang Akar Formation. Calcareous mudstone grades up into muddy limestone containing skeletal debris of bivalves, large foraminifera, solitary corals and echinoid fragments. This limestone can be recognized throughout most of the field. The 34-2 net sandstone isopach (Fig. 14A) out lines a sinuous sand body with an average width of 4000ft ( 1220m). The shape and character suggest that the sand body is fluvial but the exact type of fluvial system is less clear. Several distinct sedi mentary units can be recognized within the 34-2 at B-8 (Fig. 13), but these are not as discrete as those of previous sandstones. Similarly, evidence from other full-hole core, dense sidewall cores in most of the other wells, and log character analysis suggests a broad systematic fining within a single-storey system (see also cross-section A- A', Fig. 13). This implies
36 7
a relatively mature fluvial system but a meandering river model (Allen, 1965) is discounted on the lack of evidence for lateral accretion and on the geometry of the reservoir which is not, so much a sheet as a ribbon of sandstone that tapers towards both sides (see Fig. 14A). Similarly, equations relating sand body thickness to channel depth and channel belt/ sand body width (see Leeder, 1973; Collinson, 1978; Lorenz et al., 1985; Fielding & Crane, 198 7), although having a wide margin of error, predict a much wider sand body if the 34-2 were a high sinuosity channel deposit. It is tempting to use a Mahakam River (East Kalimantan, island of Borneo) type model (Allen et al., 1979; Young & Atkinson, 1993) as the analogy for the 34-2 reservoir. In this case, the river channel does not meander significantly but does have a meandering thalweg. On the inside of every thalweg meander loop is a non-emergent lateral bar which is attached to the channel side and progrades down stream. Echo-sounding and dredge sampling suggest that the Mahakam River immediately above the delta has a width, sinuosity, maximum sand thick ness, grain size and bedform distribution that is remarkably similar to the 34-2 sandstone. However, irrespective of the exact mechanism of channel fill, the 34-2 reservoir is interpreted to be a single storey, fluvial sandstone. Proximity to the sea is suggested by the maturity of the fluvial system and by the overlying transgressive limestone. However, evi dence for marine conditions within the sandstone is absent. Reservoir quality is excellent, with average log porosity of 29% and test permeabilities of 10-30D. The depth structure map on top of the 34-2 sand stone is shown in Fig. 14B and the structural spill point/closing contour is at a depth of 3630ft SS (1107 m SS). The oil-water contact for the reservoir is still 3680 ft SS (1122 m SS) and there is again a discrepancy (50ft or 15m) between oil column and structural closure. It is clear on Fig. 14A, however, that the trapping mechanism is the same as for the 35-1. The structural contour of the oil-water contact does not close to the northwest but the sand body pinches out in that direction. The trapping mechan ism again combines stratigraphy and structure. 34-1 sandstone
The base of the 34-1 sandstone lies just above the limestone at the top of the 34-2 sequence and all but the top few feet of the sandstone was cored in the
368 A 20
R. Young, W.E. Harmony and T. Budiyento
B-2 API
1600' 120
20
B-3 API
1900' 120
20
B-4 API
2200' 120
5300'
Widu ri- 7
20
API
120
A'
Widuri-5
20
API
120
Top l i m estone
Base coal 'A'
DEPTH LOCATION MAP
GAMMA RAY API100 0
(FT)
Core log 34-2 sandstone Widuri B-8 Muddy l i mestone and calcareous m u dstone Bivalve and fora m i n ifera fragments Root leted si ltstone and coal
Trough cross- bedded sandstone g rading from med i u m to coarse at the base to fine-very fine at the top
4490
Si MCg ,...,... Sh�Cl SA ND I
NO RECOVERY 4500
4510
4520
Coal 'A'
Fig. 13. 34-2 reservoir: vertical and lateral sedimentary character..
B-8 well (see Fig. 15). The sandstone is erosively based and consists of mostly fine- and very fine grained, very well-sorted sandstone, which broadly fines upwards giving a predominantly gamma-ray increasing upwards log signature. Cross-bedding is sporadically present throughout but, towards the top, parallel lamination, ripple cross-lamination and ripple/wavy bedding predominate. Carbonaceous
laminations are common, and abundant mudstone clasts in the lower to middle part of the sandstone result in increased gamma-ray readings (Fig. 15). Towards the top of the sandstone (4382. 5-4391f t MJ?) a heavily bioturbated and interbedded· interlaminated unit of mudstone, siltstone and sandstone contains predominantly Thalassinoides but also Teichichnus type burrows, which are indica-
Fluvial sand-body geometries
\ ....____
)
369
+
+
'\
+
+
+
+
,
I
:
'
( )
;--s··:
(
"-.!
+
+
+
Cl 20' 5000 FT
+
Fig. 14. 34-2 reservoir : (A) Net
sandstone isopach with the oil water contact marked. (B) Depth structure map on the top 34-2 reservoir showing that the oil water contact is 50ft below structural closure (shaded).
tive of marine influence. In Widuri-1 the 34-1 was also cored and clay drapes, a common fei!ture of tidal sediments, are abundant in the upper part of the sandstone, where marine acritarchs also have been found. The net sandstone isopach (Fig. 16A) defines a sand body with a maximum penetrated thickness of 49ft ( 15 m) and an average width of 2000 ft ( 610 m). The detailed shape of the contours in Fig. 16A is not
a
E
·.-----....
n '\
,/
\
CllOO'
5000 FT
B
the result of extrapolated well data. Prior to any development drilling it was recognized that an acoustic impedance contrast between the base of the sandstone and the underlying mudstone was respon sible for a seismic peak (see Blue seismic horizon, Fig. 6, for example) which, when mapped, had a restricted areal extent. Seismic modelling suggested that the sandstone thickness could be estimated from the amplitude of the reflection through the use
370 A 20
R. Young, W.E. Harmony and T. Budiyento
D-11 API
2150' 120 20
A-12 API
1600' 120 20
Widu ri-1 API
5000' 120 20
C-6 API
DEPTH
LOCATION MAP GAMMA RAY API
(FT)
3850' 20 120
B-8 API
A' 120
Core log 34-1 sandstone Widuri B-8
Biotur bated mudstone, si ltstone and sandstone
Thalassinoides/ Teichichnus type bu rrows
Fine and very fine grained sandstone
Calcareous mu dstone with shell fragments
Fig. 15. 34-1 reservoir: vertical and lateral sedimentary character.
of a tuning curve (Young et al., 1991), and the net sandstone isopach map was constructed based on reflection amplitude. This map proved to be very accurate and has changed little after development drilling. The shape and dimensions of the sandstone, together with the log and sedimentary character, suggest a single-storey channel which was reworked by tides in the upper part. Within Widuri field, the interval that is laterally equivalent to the channel sandstone consists of mudstone and siltstone with rare, thin sandstones (see cross-section A-A',
Fig. 15). When traced southwards away from the sediment source area, however, Young & Atkinson (1993) have shown that the entire productive interval at Widuri becomes more marine and is dominated mostly by 6-24 ft thick, stacked, coarsening upwards units, which are interpreted to be distributary mouth bars and various other marginal marine bars. Dia chronism of the whole interval is clear and the overall interpretation is a deltaic succession domi nated by tidal and fluvial processes (Young & Atkinson, 1993). The 34-1 is therefore interpreted to be a distribu-
Fluvial sand-body geometries
+
+
'•? '·---)
(
371
+
��') ..
' ' Co
ON N
+
\),1
/V
r··
)
'· ·-r<'-' <;:) J ....._._ ___
+
\'
)
,/ i
I
+
! \
,...., ,. I
I I I'
,{
(/
+
+
+
,..,
,...../+�u'
I I
+
0' 20'
+
(
+
+
+
ol &J "''
+
...
+
+
<.)'
s! oJ
� +
+
.t../
+ ,
_)
Cl 20' 5000 FT A
-,__
Fig. 16. 34-1 reservoir. (A) Net
sandstone isopach with the field oil-water contact marked. (B) Depth structure map on the top 34-1 sandstone showing that the oil-water contact is 1 00ft below structural closure (shaded).
tary channel and the tidal reworking at the top is perhaps mostly a result of estuarine conditions after abandonment of the distributary. The apparent absence of delta-front bars may reflect the proximal position of the Widuri area with respect to the delta complex and/or the low preservation potential of tidal bars (oriented normal to the shoreline) as the delta prograded. The shape, dimensions and sedi mentary character of the 34-1 are very similar to
Cl100' 5000 FT
8
those of a modern Mahakam delta distributary channel (see Allen et at., 1979; Allen, 1985}. Reservoir quality is still very good with average log porosity of 29% and test permeabilities of 5-10 D. The depth structure map on top of the 34-1 sand stone (Fig. 16B) identifies the structural closing contour at 3580ftSS (1091 m SS). An oil-water contact for the 34-1 sandstone has not been penetrated in any well. However, a reservoir simu-
372
R. Young, W.E. Harmony and T. Budiyento
stratigraphical levels over a vertical interval of about: 75ft (23m) immediately above the 34-1 sandstone. None of the reservoirs has been cored, but Fig. 17 demonstrates the log character and limited lateral extent of the sandstones. Nine separate sandstones have been identified so far (Fig. 18A), although this number could increase or decrease with further drilling. The sandstones are fine to very fine grained with sharp bases and sharp or gradational tops and usually show a gamma-ray increasing upwards log signature. Penetrated net sandstone thickness does not exceed 15ft (4.5 m), with one exception, the 33-2, which is 26ft (8 m) in Widuri-6. Sand body widths average 2000 ft (610 m) but well penetrations of individual sandstones are too few to define geo-· metries accurately. However, at least some of the sandstones are thought to have shoestring geometries and occasionally, as in the case of the 33-2, reflection
lation study using the geological model described here and an oil-water contact at 3680 ft SS (1122 m SS) gave a very good match with the available production history. The oil-water contact is there fore assumed to be the same as all the underlying reservoirs. A discrepancy of 100ft (30.5 m) between the oil column and the structural closure can again be accounted for by a stratigraphical component to the trapping, where northwestwards closure is effected by pinchout of the reservoir. 33-series sandstones
There is less information on the 33-series sandstones than on any of the other reservoirs and they are the least understood and most difficult to define. The 33-series is a group of several different sandstones which, for the most part, occupy slighty different A 20
D- 1 API
3,800' 120
20
D-4 API
120
12,600'
C- 2 20
API
32 , 00' 120 20
API
0
Talang Akar formation
50 FT
150
120
Top
0
100
A'
C-1
FT
Top 33-1 sa ndst one i e �L_ �!_ .!
50
100
Top 34-1
150
san dst one inte rval
5000FT
t.----J
Fig. 17. Cross-section showing the electric-Jog character and lateral discontinuity of 33-series reservoirs.
3 73
Fluvial sand-body geometries
A
Cl100' 5000 FT
B Fig. 18. 33-series reservoirs. (A) Net sandstone isopach with the field oil-water contact marked. (B) Depth structure map
on a middle 33-series limestone marker showing that the field oil-water contact is 150ft below structural closure (shaded).
R. Young, W. E. Harmony and T. Budiyento
374
amplitude at the base of a sandstone can be used to map it. The sand bodies are laterally equivalent to finer grained sediment, which was cored in the C-2 well (Figs 1 7 & 19). The basal 32ft (9. 75m) of the core consists of carbonaceous mudstone with laminations of siltstone and very fine-grained sandstone. These laminations are millimetre-scale and sparse at the base but centimetre-scale and frequent towards the top. Both bioturbation and ripple cross-lamination increase upwards and fragments of fish bones, foraminifera and echinoids are scattered through out. The basal 27ft (8. 25 m) is a coarsening upwards unit which grades into a 5 ft ( 1.5 m) thick, fining upwards unit. The upward fining trend continues into highly bioturbated siltstone before a very sharp contact at 3899 ftMD marks the top of the unit. The overlying dark brown, carbonaceous mudstone contains scattered marine fauna. The coarsening fining couplet is thought to reflect prodelta to lower mouth-bar progradation and abandonment, whereas
DEPTH (FT)
GAMMA RAY
0
API
the overlying mudstone is clearly a shallow-marine deposit with minimal coarse clastic sediment input. However, only 16 ftMD (4. 9 mMD) above the top of the core is a 3 ft (1 m) thick coal. Using similar reasoning as for the 34-1 sandstone, the 33-series sandstones are interpreted to be single storey distributary channel deposits of a tide influenced delta. Reservoir quality is variable but log porosities average 27% and test permeabilities are of the order of 1-5 D. The depth structure map on a marker in the middle of the 33-series interval identifies a structural closing contour at 3530ft SS (1076 m SS, Fig. 18B). The oil-water contact for the 33-series sandstones is unknown and reservoir simulation studies do not help because production histories are too short. The oil-water contact is assumed to be at 3680 ft SS (1122m SS), the same as the other reservoirs, although multiple oil-water contacts are possible. However, irrespective of the position of the oil water contact, the oil column is again greater
100
3880
Widuri C-2
Fig. 19. Core log of part of the 33series interval in Widuri C-2 (see Fig. 17 for location of cored interval).
Fluvial sand-body geometries
375
Feet NW
0
1000
2000
3000
SE
Time Feet 1000 3100 1020 1040
3200
1060 3300 1080
3400
1100 1120 1140
3500
3600
1160 3700 1180 3800 1200 3900 1220 1240 4000 1260 1280 1300
4100
4200
1 320 1340 1360
4300
4400
Fig. 20. An uninterpreted seismic line from the three-dimensional volume over the Widuri area. The line is located just
south of Widuri field and shows an oblique view through one of the features that are interpreted to be incised valleys associated with the top Talang Akar horizon. See Fig. 6 for the stratigraphical location of the seismic markers.
than the structural closure and a stratigraphical component to the trapping, although less well defined, is again necessary. The interval between the top 33-1 sandstone and the top Talang Akar Formation consists of coal,
mudstone and siltstone with some thin, interbedded limestone. A low-energy, marginal marine setting is envisaged prior to the development of fully marine conditions in the overlying Batu Raja Formation.
376
R. Young, W.E. Harmony and T. Budiyento
Reservoir
33
Series
34-1
34-2
35- 1
35-2
36 -1
Average grain size
Multi storey/ m u ltilateral
fine to very fine
NO/NO
fine to very fine
medium
medium
medium to coarse
coarse
NO/NO
?NO/NO
YES/YES
YES/YES
YES/YES
®
*
Reasonable maximum thickness (ft)
715'-730'
40'
50'
60'**
65'
175'
@Average width
(ft)
72000'
2000'
4000'
6500'* *
1 1000'
Sheet
®I@ ratio
7 1 :67 to
Average porosity a n d Aerial distribution Environment of deposition (to scale) permeability range
darcies
1:50
29% 5-10
1:80
1: 108
1: 169
Sheet
�
27% 1-5
7 1 : 1 33
29% 10-30
darcies
29% 10-20
darcies
29% 20-30
darcies
28% 20-30
darcies
>CC UJ
<>: 1- w f-' ::::l Z
�
darcies
([: 1flj 0 �
"l
\�
Lithostratigraphy
g f-'� I�
cc w m
�u 0
::;; w ::;;
j:!
(3
z 0
� ::;;
UJ cc ::::>
�
::;; ::;;
cc 0 u.. cc
..: "" ..:
<.:J z
w 1Ul >Ul ...J
..: ...J
cc UJ m
..: >
::::> ...J u..
UJ cc ::::>
�
j:!
::;; UJ ::;; ..: 0 ...J UJ N
::;; :::;
__ ;
Fig. 21. Summary and comparison of sedimentary character, geometry and environment of deposition of the Widuri field
reservoirs. * Using 'reasonable' maximum thickness is an attempt to eliminate thickness anomalies; the parameter is not rigorously defined but nor is it arbitrary. * * Only the east channel system was considered ; the west channel system may be structurally controlled.
Discussion
The upper part of the productive interval at Widuri (the 34-2, 34-1 and 33-series sandstones) is inter preted as deltaic based largely on evidence outside of the Widuri area. Core data from approximately the same stratigraphical interval in the Sunda basin and the Ardjuna basin (approximately 70 miles to the southeast of Widuri in the adjacent PSC) indicate a predominance of tidal and fluvial processes at the shoreline (Young & Atkinson, 1993). In this setting, tidal reworking at the delta front fashioned delta front bars, oriented predominantly normal to the shoreline, which were thin (up to 25ft thick) compared with the channel deposits (approximately 15-40ft thick). The delta-front bars therefore had low preservation potential as the delta prograded (see e.g. Allen et al., 1979; Allen, 1985; Coleman &
Wright, 1975, their Fig. 16). The distributary channels of tidally influenced deltas are therefore not always juxtaposed against delta-front bars and, once abandoned, they become estuaries. These same characteristics (namely, channel features that erode into marine sediments and are estuarine in their upper parts), however, could be used to interpret the reservoirs (especially the 34-2 and 34-1) as incised valley fills, which formed as a response to periods of falling relative sea-level (see e.g. Van Wagoner et al., 1990). It is also reasonable to expect that relative sea-level changes caused by eustasy and/or tectonic movement had, at times, a major effect on deposition of the Talang Akar Formation over the 7.5Myr period. An example of this is at or near the top of the Talang Akar Formation, where a horizon characterized by down cutting is clearly seen on seismic records (Fig. 20).
Fluvial sand-body geometries
Reservoir
Aerial distribution ('to scale')
�
33
series
34-2
35-1
35-2
Fig. 22. Summary of Wid uri field
36- 1
Structural closing contour (ft SS)
Oil col u m n (ft)
3530
325
F i l l w.r.t. structure
Trap type
Q)
3680
q;� rf!! ��'
5'?
/f
(ft/SS )
3680 ?
�
34-1
owe
377
�1
3680
3580
3630
2 70
240
-o �
c Vl 0 0
- ·� - .c
.o c;; -o �
u rn :o · -
�u
·- ::0 LJ._ !:: "'
3680
3660
180
3680
3680
110
F i l l ed to spill
3680
3740
55
Part i a l ly fi l l ed
� 0. ::0 "' �
-
- "' Ul !::
"'
-
� 3 u 2 cil
trapping characteristics.
When mapped, this down-cutting feature has a channel form with a roughly N�S orientation, a width of approximately 1000-3000ft, and a thick ness or 'down-cut' of about 30-lOO ft. An interpret ation of recent three-dimensional seismic coverage over the northwest flank of the Asri basin has identified several more of these features, all associ ated with the same horizon. Unfortunately none of these features has yet been cored, but the event can be traced seismically all over the Asri basin and 60miles south into the western Sunda basin, where it correlates to a karst surface within limestone. The event therefore has regional extent and is interpreted to be the result of a lowstand of sea-level. There is strong seismic stratigraphical evidence that the low stand is a response to tectonic reactivation of the Asri basin. In contrast, within the productive interval at Widuri field there is no such seismic expression and no such evidence for regional lowstand surfaces of erosion (see Fig. 20). Until further evidence surfaces, therefore, the retrogressive stacking of the reservoirs at Widuri field is interpreted to be a response to essentially autocyclic switching of sediment supply within an overall transgressive systems tract. How ever, the primary aim of this paper is to demonstrate trapping configurations resulting from the interaction of sand body geometry and structure, and these are independent of the mechanism of stacking.
R E S E RV O I R
SUM M A RY
The more important aspects of the geometry and character of the Widuri reservoirs are documented in Fig. 21. The productive interval as a whole shows a gradual and systematic change from base to top. Grain size gradually decreases upwards from coarse grained in the 36-1 to fine grained - very fine grained in the 33-series reservoirs, and multistorey and multi lateral channel systems in the basal three reservoirs give way to single-channel systems in the upper three. The maximum sand body thickness gradually decreases upwards, as does the average width, and the thickness/width ratio shows a corresponding change. Porosities are similar throughout but test permeabilities show a significant decrease upwards. The change in sand body size, shape and character is the result of a maturing depositional system associated with a rise in sea-level between 25. 5 and 22Ma (see Fig. 5; Haq et a/., 1988). The immature fluvial system at the base (36-1 reservoir) is suc ceeded by progressively more mature fluvial systems before deltaic conditions prevail at the top. The change in character of the sand bodies therefore reflects the change in environment of deposition and this is supported by the available palaeontological data from the intervening shales. Rare occurrences of palynomorph taxa commonly associated with a brackish water setting have been found as low down
R. Young, W.E. Harmony and T. Budiyento
378
SE
NW
SE
Fig. 23. (A) Schematic diagram of the Widuri field trapping configurations. (B) A projection of what the Widuri trapping configuration would have been if the 34-2 reservoir had been a sheet sandstone.
as the 35-2 interval and increased frequencies in the overlying 35-1 and 34-2 intervals indicate increasing proximity to the sea.
TRAP
SUM M A R Y
The important aspects of the trap are summarized in Fig. 22. With the possible exception of the 33-series, the reservoirs share a common oil-water contact and therefore must either be in communication now or have been in communication some time after oil migration. The structural closing contour on top of the 36-1 sandstone is below the oil-water contact and the reservoir is only partially filled. The d osing contour on top of the 35-2 sandstone is coincident with the oil-water contact and is therefore full to spill. This is an extremely important point because it
is the spill point of the 35-2 reservoir that therefore defines the oil- water contact for the entire field. The lower two reservoirs are structural traps, but the upper four reservoirs are filled beyond structural closure and rely on a favourable orientation (northwestwards pinch-out) with respect to structure in order to trap oil by a combination of structure and stratigraphy. The trap model is drawn schematically in Fig. 23A but it is interesting to evaluate how the trapping would change if one of the upper four reservoirs, the 34-2 for instance, had been either a sheet sandstone or did not have a favourable orientation with respect to structure. In other words, what would have happened if the 34-2 did not have a stratigraphical component to trapping? This case is shown in Fig. 23B. The reservoirs are still in communication and so the oil would migrate upwards through the
Fluvial sand-body geometries
lower reservoirs into the 34-2, where it would be trapped with respect to a new field oil-water contact as defined by the 34-2 structural spill point (3630 ft SS or 110 7 m SS, see Fig. 22). The net effect is exaggerated in the schematic diagram of Fig. 23B, but a 50 ft (15 m) higher oil-water contact would significantly reduce the volume of oil in the field. An overall conclusion then is that the trapping at Widuri is intimately related to reservoir geometry and that the large size of the field is the result of a remarkable coincidence of events. A relative rise of sea-level helped cause a gradual change in the environment of deposition which, in turn, produced a systematic reduction in the size of sand bodies, the upper four of which each had to be oriented favour ably with respect to structure in order to trap oil stratigraphically.
APPLICATIONS
The geological model has contributed to the devel opment of Widuri field in two important ways: (i) improved drilling success and (ii) improved reservoir management. Early recognition of sand body geometries resulted in more efficient develop ment drilling, such that of the 47 development, infill and step-out wells drilled so far only two lacked economic pay, and both of these were side-tracked to successful locations. Similarly, reservoir manage ment has been optimized by having a detailed and predictive geological model, which forms the basis of reservoir simulation studies. This application will become increasingly important as the field matures, water production increases, and further infill drilling is required to optimize hydrocarbon recovery. From an exploration viewpoint, the work on Widuri field has provided an analogue model in the search for new fields in the PSC area, in which stratigraphy is an important means of trapping hydrocarbons. Widuri is by definition a large oil field but approximately 75% of its reserves have a strati graphical component to trapping. Any future dis covery of stratigraphically trapped oil could therefore potentially be large.
AC KNOWLEDGEMENTS
The authors wish to thank the management of Pertamina, Maxus and partners for permission to publish this paper. Special thanks are extended to
379
C.L. Mora and C.D. Atkinson for a critical review of the manuscript, and to J.F. Mitchell for his technical input in the early stages of the study. We also acknowledge the help and support of all the other members of the multi-disciplinary team that was responsible for developing Widuri field.
REFERENCES ALLE N ,
G.P. (1985) Delta sedimentation: modern and ancient examples in the Mahakam delta basin. Indonesia Petroleum Association, Field Trip Guide.
G.P. , LAURIER, D. & THOUVENIN , J. (1979) Etude sedimentologique du delta de Ia Mahakam. Compagnie Francaise Petrol. Notes Mem. , 15 , 156 pp. A LLEN , J.R.L. (1965) A review of the origin and character istics of recent alluvial sediments. Sedimentology , 5, 89-192. BEDDOES, L.R. (1981) Hydrocarbon plays in Tertiary basins of Southeast Asia. Energy, 6, 1141-1163 . BEN-A VRAHAM , Z. & E MERY, K.O. (1973) Structural framework of Sunda Shelf. Bull. Am. Ass. petrol. Geol. , 57, 2323-2366. B USHNELL, D . C. & TEMANSJA, A.D. (1986) A model for hydrocarbon accumulation in Sunda Basin, West Java Sea. Proceedings of the Indonesian Petroleum Associ ation, 15th Annual Convention, Vol. 1 , pp. 47-75. COLEMAN, J. & WRIG HT , L.D. (1975) Modern river deltas: variability of processes and sandbodies. In: Deltas (Ed. Broussard, M.L.), pp. 99-149. Houston Geological Society, Houston, TX. COLLINSON , J.D. (1978) Vertical sequence and sand body shape in alluvial sequences. In: Fluvial Sedimentology (Ed. Miall , A.D.), Mem. Can. Soc. petrol. Geol. , Calgary, 5, 577-588. F IELDING , C.R. & CRA N E , R.C. (1987) An application of statistical modelling to the prediction of hydrocarbon recovery factors in fluvial reservoir sequences. In: Recent Developments in Fluvial Sedimentology (Eds Ethridge , F . G. , Flores, R.M. & Harvey, M . D.), Spec. Publ. Soc. econ. Paleontol. Mineral. , Tulsa, 39, 321-327. F oLK, R.L. (1968) Petrology of Sedimentary Rocks. Hemphill's, Austin, TX, 170 pp . HAQ, B.U . , H ARDENBOL, J. & VAIL, P.R. (1988) Mesozoic and Cenozoic chronostratigraphy and eustatic cycles. In: Sea-level Changes: an Integrated Approach (Eds Wilgus , C. K. , Hastings, B.S. , Kendall, C.G.StC. , Posamentier, H . W. , Ross, C.A. & Van Wagoner , J.C.) , Spec. Pubis Soc. econ. Paleontol. Mineral. , Tulsa, 42, 71-108. H uTCHISO N , C.S. (1986) Tertiary basins of S.E. Asia their disparate tectonic origins and eustatic strati graphical similarities. Bull. geol. Soc. Malay. , 19, 109-122. LEEDER, M.R. (1973) Fluviatile fining-upward cycles and the magnitude of palaeo channels. Geol. Mag. , 110, 265-276. LORENZ, J.C. , H EINZE, D. M. , C LARK, J .A. & S EARLS, C .A. (1985) Determination of widths of meander-belt sandstone reservoirs from vertical downhole data, Mesaverde Group , Piceance Creek Basin, Colorado. ALLEN ,
380
R.
Young, W.E. Harmony and T. Budiyento
69, 71 0-721. R.M . , CAMPION , K.M. & RAHMANIAN , V . D . (1990) Siliciclastic sequence stra tigraphy in well-logs, cores and outcrops: concepts for high resolution ·correlation of time and facies. Am. Assoc. petrol. Geol. Meth. Explor. Ser. , 7, 55 pp. WIGHT, A. , SUDARMONO, & lmron, A. (1986) Stratigraphic response to structural evolution in a tensional back-arc setting and its exploratory significance: Sunda basin, west Java Sea. Proceedings of the Indonesia Petroleum Association, 15th Annual Convention, Vol. 1 , pp. 77100. Bull. Am. Ass. petrol. Geol. ,
VAN WAGONER,
J. ,
MITCHU M ,
R. & ATKINSO N , C.D. (1993) A review of Talang Akar Formation (Oligo-Miocene ) reservoirs in the off shore areas of Southeast Sumatra and Northwest Java. In: Clastic Rocks and Reservoirs of Indonesia (Eds Atkinson , C.D. , Scott, J. & Young, R.). Indonesian Petroleum Association, Core Workshop Notes, October 11, pp. 177-21 0. YOUNG, R. , HARMONY, W . E . , GUNAWAN, J. & BUDIYENTO, T. (1991) Widuri field, offshore Southeast Sumatra : sandbody geometries and the reservoir model. Proceed
YOUNG,
ings of the Indonesian
Petroleum Association,
Annual Convention, Vol.
1 , 385-417.
20th
Index
Abyssal plain,iron formations 149-51, 154 Acidic volcanic rocks,geochemistry 252 Ahwaz, Cenozoic clastic reservoirs 349-50 AI Khlata Formation 340,341 Allostratigraphy 72-3 Alluvial fan facies association, Tryfan Fach Member 172-4,184-5 petrography
178
Alluvial fans 203 Amin Formation 340 Anglesey 178 Animikie Group, iron formations 139-40, 142 Arabian Plate
335-9
Arenites lithic, geochemistry 247-8 lithic-volcanic,geochemistry 252-3 Askrigg Block 307-8,312-20 Asri basin 355-8, 377
Baby Formation,iron formations 143-6
85-94 Bideford Group 101-3, 110 Black shale facies,Bude Formation 126-7, 130 description 110-11 interpretation 120-2
Cherty iron formation Chesil Beach 91, 93
Black shales, geochemistry 248 Boss Point Formation 117 Bottom-current activity,ridge-
Computer maps,preparation Conglomerate debris, Kalvag Melange geochemical analyses 245 geochemistry 241-8 Conglomerates
trough system 296-7 Bovis Formation,Pororari Group 204,208-10 palaeocurrent pattern 212-13 Braided stream association,Tryfan Fach Member 166-9, 174-5 bounding surface with marine shelf association 177 petrography 178-9 Breccia 198-203 Bude Formation 101-5,130-1 bedtypes 108-9 facies analysis 110-26 description 110-20 interpretation 120-6 facies succession 106-10 description 126 interpretation
Burrowing, Woburn Sands 29,35-6
Basalts, geochemistry
Camaqua Basin, Brazil Cameros Basin 221 stratigraphy 248-
transitional 251-2 Basra, clastic reservoirs Cenozoic Mesozoic
349-50 344-5
Beachfaces bedforms and sedimentary structures characteristics
92-4 95
facies successions 94-6 gravelly 76 sediment transport 78-9 Beardmore Volcanic Associated Iron Formation 151-2 Beardmore-Geraldton Clastic Associated Iron Formation 146-9 Bedfordshire Strait 25 Bedforms,gravelly shorefaces
Clastic rocks, Upper Jurassic 232-3 Clasts 80-5 Coastal zone terminology 76-7
grain sizes 80-4 Guaritas sequence Ricinus 62
73
5-7,9-10
Senora de Brezales Formation 223-7 shoreface 75 Crackington Formation 104-5, 107-10 Craven Basin 308-9,312-18 Cryptobioturbation 35 Culm Basin 102-4 tectonic setting 105-7 tectonically juxtaposed successions 107-10
126-30
Back-arc basin 257 Bahrain, Mesozoic clastic reservoirs 344-5 Bars, gravelly shorefaces 85-6 alkaline 248 island arc 252 subalkaline MORB-Iike 51
139, 140
27-8,
3-4,6
Debris flows 186 breccia 203 Diorite 247 Dune migration, Guaritas sequence 10-11
220, 422
Canadian Shield, iron formations 139 Capel Curig Volcanic Formation 159, 183 Carbonate rocks,Senora de Brezales Formation 227-8 Cardium Alloformation,Alberta 48-50,73 allostratigraphy 49-50 comparison of allomembers stratigraphy and
Erosion channels 268-70 ridge-trough system storm/wave 271-4
297
surfaces 26-7 Estuaries 69-70 Estuarine environments, essential elements 41-2
65 Facies modelling
187-9
sedimentology 50-5 Cardium Formation 73, 76
Fan-deltas Feni Drift
gnivelly shorefaces 94 Carstone 38
Floodbasin facies association, Tryfan Fach Member 169-
80-4, 86-
Cenozoic clastic reservoirs 50
72, 174-7 349-
petrography 178-9 Frio Formation 267, 271-5
Channel erosion 268-70 Chert,geochemistry 247
Sedimentary Facies Analysis: A Tribute to the Research and Teaching of Harold G. Reading Edited by A. Guy Plint © 1995 The International Association of Sedimentologists ISBN: 978-0-865-42898-0 381
213, 216 297, 299
382
Index
Gabbro 247 Gardar Drift 297 Gas, Middle East 333 Gault Clay 22-3 Gharif Formation 340 Goniatite-bearing marine bands 309-11 Goniatites 111 Graphitic shear zones
149
Gravel deposition, Guaritas sequence 8-10 Gravel plunge step 90-1 Gravel wave ripples 85-6 Gravelly shorefaces 75-6 Gravels clast shape 84-5 grain size and shape 80-5 textural trends 80-4 wave-worked 90 Groundmass, olistostromal, Kalvag Melange 240-1 geochemistry 253, 256 Growth faulting 265-6 effects on sedimentation 26776 Guaritas depositional sequence, Brazil 3-5, 6 alluvial facies general features 5-11 interpretation of features 8-11 interpretation of lateral changes 11-14 lateral changes 11-14 pattern of textural dispersion 11 palaeocurrent pattern
11
Gulf of Suez, clastic reservoirs Cenozoic 350 Mesozoic 348-9 Palaeozoic 343-4 Gulf Region, Palaeozoic clastic reservoirs 342-3 Gunflint Formation, iron formation 139 Gwern Gof Tuff
159-60
Hadhramaut-Jeza rift basin 346-7 Haima Group, clastic reservoirs 340 Haradh Formation 340 Haushi Group, clastic reservoirs 340-1 Hawks Crag Breccia
199-201,
214-15 deposition 211 Hummocky cross-stratification, Bude Formation 129 Hydrocarbon trapping 378-9
Ignimbrite, geochemistry 252 Incised valley Ricinus 47-9, 66-71 Interbedded sandstone-shale facies, Bude Formation 126-7, 130 description 112-17 interpretation 122-5 Iron formations 137-9 case studies discussion objectives types
139-53 153-4 139
138
348
Jordan, Palaeozoic clastic reservoirs
241, 246-7 Middle East Clastic reservoirs
240-1
oil and gas reservoirs 333-9 palaeogeographical and palaeoclimatic synopsis
geological setting 238-40 profile through 242-3 provenance and tectonic setting 256-9 regional correlations and interpretations 259-61 Karst surfaces, Senora de Brezales Formation 228-9 Kuwait, clastic reservoirs Cenozoic 349-50 Mesozoic 344-5 Kyeburn Formation
335-9 source rocks Millstone Grit
336
cycle duration 306 cyclothem architecture and sea-level change 311-21 asymmetry 321-2 E2a(i) cyclothem architecture and sequence stratigraphy 313-17 boundaries and thickness 312 depositional model and
198
143
Laminar calcretes, Senora de Brezales Formation 228 Laminated beds, Bude Formation 112, 123 Late Jurassic unconformities 2334 Levant-Sinai region, Mesozoic clastic reservoirs 348-9 Lithic arenites, geochemistry 247-8 Lithic-volcanic arenites, geochemistry 252-6 Llewelyn Volcanic Group 15860, 183 180
tectonism 313 E2c2 cyclothem architecture and sequence stratigraphy 318-20 boundaries and thickness depositional model and tectonism 317-18
317
major cyclothems 322-3 architecture and genesis 323-4 major scale cyclicity
322-4
multi-cyclothem condensed sections 322-4 sequence model 320-2 choice 320-1 Morley Occurrence, iron formations
London- Brabant land mass
19, 20
Longshore currents, sediment transport 77-8 Lower Follifoot Grit 317-19 Lower Greensand Group, deposition 19-20 Lower Miocene, Louisiana 7, 274-6, 278
349-50
Mesozoic 344-9 Palaeozoic 339-44 geological boundaries 331-2
depositional model 257 geochemistry 241-56
volcanic source
stream association 177 petrography 179-80 Mesozoic clastic reservoirs 344-9 Metabasalts, Kalvag Melange
Cenozoic
343
Labrador Trough
Magma migration 185 Mahwis Formation 340 Marib-Jauf basin 346 Marine bands, goniatite-bearing 309-11 Marine shelf association, Tryfan Fach Member 161-6, 174 bounding surface with braided
Island arc basalts, geochemistry 252 Israel, Mesozoic clastic reservoirs
Kalvag Melange
Lower Ordovician ophiolite complexes/island arc sequences 259-60
266-
152-3 Muddy siltstone facies, Bude Formation 126-7, 129-30 description
111-13
interpretation Mudstones
122
Guaritas sequence Ricinus
8
60-2, 65, 66
Index deposition 70 distribution 63-5 Tryfan Fach Member 161-6 Mudwaves, ridge-trough system 296-7
383
pebbly sandstone and massive or graded sandstone 20310 geological setting 199 palaeocurrent pattern 212-13 tectonic implications 213-15 Pororari River 211
Namurian Basin, Central Province, 30-8 Namurian stratigraphical
Preservation potential thresholds 276 representation in growth-fault
framework 306 Neutral Zone, Mesozoic clastic reservoirs 344-5
settings 279 Pro-delta current system, ridge trough system 297-302
New Zealand 197-9 North Sea, ridge-trough system
Puysegur Formation
285-6 database and methodology 288 depositional models 296-302 external morphology 288-9 geological setting 286-8 internal reflection configurations 289-91 interpretation 291-2 mapping 292-4 palaeogeographical setting of structures 294-6
Oil, Middle East 333 Olistoliths, Kalvag Melange 240-1 sedimentary rocks geochemistry 252-3, 255 representative geochemical analyses 254 spatial distribution volcanic rock
244
geochemistry 248-52 representative geochemical analyses 249 Oman, clastic reservoirs Mesozoic 347-8 Palaeozoic 340-1 Outer shelf and slope, iron formation Outwash fans
143-6, 154
169
198
Pyroclastic material 158-9 syneruptive, aggradation 184-5
Qatar, Mesozoic clastic reservoirs 348 Quasi-planar lamination Qusayba shales 342
164
Ramp-fan system, iron formation 147-9, 154 Rhyolite, geochemistry 252 Rhyolitic braided-stream sandstones, Tryfan Fach Member 166-9 Ricinus, Alberta
47-9, 70-1
Allomember 54 comparison with other allomembers 65 cross-sections 71-2 deposition in estuary
212-13
Palaeozoic clastic reservoirs 339-44 Palestine, Mesozoic clastic reservoirs 348 Pennine Basin 308-9 Petroleum producing formations, Middle East 334-5 Planar lamination 164 Porcupine Basin 232-3 Pororari Group 199 depositional systems 211-12 facies assemblages 199-210
69-70
facies 57-62 lateral and vertical distributions 62-5 geometry 55-7 incised valley 56 sand body 56-7 incised shoreface 65-7
laminated mudstone 210 matrix- and clast-supported breccia 199-203
Sand volcano 120-1 Sand-body geometries, Widuri field 377 Sandstone-dominated facies, Bude Formation 129 description 117-20, 122-3 interpretation 124-6 Sandstones Bude Formation
112-17
graded silty 112-13, 123-5 Guaritas sequence 7-8, 10 irregularly laminated 117, 125-6 red, Senora de Brezales Formation 220-3 Ricinus 57-60, 69-70 parallel-laminated 59-60 structureless 57-9 Tryfan Fach Member conglomeratic alluvial fan 172-4 rhyolitic braided-stream 166-9 structure 179 thinly bedded floodbasin 169-72
Scandinavian Caledonides
259
Sea-level change and Millstone Grit cyclothem architecture 311-21 goniatite-bearing marine bands 309-11 Seaweed, sediment transport 78 76-9 Sedimentary rocks, olistoliths,
49
valley orientation 70 Ridge-trough system, North Sea 285-6 database and methodology 288 depositional models 296-302 external morphology 288-9 geological setting 286-8 internal reflection configurations 291-2
mapping 292-4 palaeogeographical setting of structures 294-6 Rip currents, sediment transport 78 River, trunk braided
Salinity, and marine bands 309-11 San Leonardo Formation 228-9
Sediment transport processes
interpretation 65-70 origin of valley 66-8 palaeogeography 68 structural problems 49
289-91 interpretation
347
Widuri field 360-75 Saudi Arabia, Palaeozoic clastic reservoirs 341-2
turbidite interpretation Palaeocurrents
Rutbah Formation
12-14
geochemistry 252-5 Sedimentary structures, gravelly shorefaces 85-94 Sedimentation and growth faulting 265-6, 267-76 Senora de Brezales Formation clast derivation 230-1 depositional history
230-1
depositional model 231-2 environmental synthesis 231 erosion 230-1 facies development, tectonic controls 229-30 late Jurassic unconformities, geometry 228-9 lithofacies 220-7 stratigraphy 220 Sequence stratigraphy
72, 266
234
384
Index
Shallow shelf, iron formation
139,
153-4 Shallow-marine sands 18 Shenley Limestone 38 Shorefaces bedforms and sedimentary structures 85-92 characteristics 95 facies successions 94-6 gravelly 75-6 sediment transport processes 77-8 Silica, iron formations 138-9 Skaggerak Delta 287, 297-300 Slaty iron formation 139, 140 Slumped beds, Bude Formation 116-17, 120-1, 125, 131 Slurried beds, Bude Formation 114-16, 118-19' 125 Snowdon graben 159 Solund-Stavfjord Ophiolite Complex
260-1
South Saskatchewan River 168 Staffordshire Basin 309 Stainmore Basin 307-8 Storm erosion 271-4 Submarine extrusive edifice, iron formation 152-3 Submarine lava plain, iron formation 151-2 Submarine rise, iron formation 146-9 Subsea fans 104-5, 106-7, 126,
associations
177-8
161-74
organization and distribution 174-8 relations 174-7 position of volcanic source 180 Tryfan Formation 159-61, 178 Turbidite system, Bude Formation 129-31 Turkey, Palaeozoic clastic reservoirs 343
Unayza Formation Unconformity 72 Underflows
164
Upper Howgate Edge Grit 16 Upper Jurassic clastic rocks Upper Wilcox Group
313232-3
266-70
Vertical-sequence analysis Volcanic rock olistoliths,
157-8
geochemistry 248-52 Volcanic terrains 158 bounding surface between syn and post-eruptive deposits 183-4
epiclastic interpretation 180-2 facies associations 161-74 modelling 187-9
357, 359,
376 palaeogeography 361 Tan Hill Grit 316-17 Terrace Bay Clastic Associated Iron Formation 149-51 Trace Fossils, gravelly shorefaces 91-2 Transgressiveestuarine-embayment system, depositional model 40-2 12-14 86-7 12-14
deposition 211-12 palaeocurrent pattern
212-13
Wave erosion 271-4 Wave motion and sediment transport 77 Widur field 355-6 background geology 356-8 palaeogeography 362 reservoir geometry and trap
360-77
overview 358-60 summary 377-8 stratigraphy 357, 359 trapping characteristics 378-9 Woburn Sands
377,
depositional models 20 facies 22-3 Heterolithic Sands 22-3 deposition 38-40
342
depositional systems
Talang Akar Formation
Troughs, nearshore Trunk braided river
extent of deposit facies
187
129, 130-1 Subsidence 269-70 rates 276, 279 Syria, clastic reservoirs Mesozoic 347 Palaeozoic 343
Tributary alluvial fan
Tryfan Fach Member 160, 186-9 depositional system 183-4 epiclastic interpretation 180-2 evidence for volcanism 178-80
description 28-9 interpretation 29-30 lithostratigraphical subdivision and relationships 20-4 Orange Sands 22-3 deposition 38-40 description 26-8 interpretation 28 palaeogeographic setting and depositional environments 19-20 palaeogeographical evolution 38-40 Red Sands 22-3 deposition 38, 40 description 35-6 interpretation 36-7 sedimentological characteristics 24-38 Silty Beds 22-3 deposition 40 description 34 interpretation 34-5
organization and distribution 174-8 geological framework 159-60 inter-eruptive sedimentation 187 sedimentation 185-6 terminology 158-9 Volcanic-associated iron formation 150-3, 154 Volcanism, stratigraphical record
Silver Sands 22-3 deposition 40 description 30-3 interpretation 33-4 stratigraphic framework
185-7 Volcano-tectonic uplift
Transition Series 22-3 description 37-8
183
interpretation Walther's Law 157, 187-9 Watson Formation, Pororari River 203-10
18-19
38
Yemen, Mesozoic clastic reservoirs 346-7