You are on page 1of 414

- T HE GEOLOGICAL SOCIETY

- OF AMERICA®

Special Paper 436


Formation and Applications of the Sedimentary Record
in Arc Collision Zones

Edited by

Amy E. Draut
U.S. Geological Survey
USGS Pacific Science Center
400 Natural Bridges Drive
Santa Cruz, California 95060
USA

Peter D. Clift
School of Geosciences
University of Aberdeen
Aberdeen, AB24 3UE
UK
and
DFG-Research Centre Ocean Margins
Geowissenschaften, Universität Bremen
Klagenfurter Strasse
28359 Bremen
Germany

David W. Scholl
Emeritus Senior Scientist
U.S. Geological Survey
345 Middlefield Road
Menlo Park, California 94025
USA
and
Stanford University (Emeritus)
USA

Special Paper 436


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

2008
Copyright © 2008, The Geological Society of America, Inc. (GSA). All rights reserved. GSA grants
permission to individual scientists to make unlimited photocopies of one or more items from this volume
for noncommercial purposes advancing science or education, including classroom use. For permission to
make photocopies of any item in this volume for other noncommercial, nonprofit purposes, contact the
Geological Society of America. Written permission is required from GSA for all other forms of capture
or reproduction of any item in the volume including, but not limited to, all types of electronic or digital
scanning or other digital or manual transformation of articles or any portion thereof, such as abstracts,
into computer-readable and/or transmittable form for personal or corporate use, either noncommercial
or commercial, for-profit or otherwise. Send permission requests to GSA Copyright Permissions,
3300 Penrose Place, P.O. Box 9140, Boulder, Colorado 80301-9140, USA.

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

Published by The Geological Society of America, Inc.


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

Printed in U.S.A.

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

Library of Congress Cataloging-in-Publication Data

Formation and applications of the sedimentary record in arc collision zones / edited by Amy E. Draut,
Peter D. Clift, David W. Scholl.
p. cm.
Includes bibliographical references and index.
ISBN 978-0-8137-2436-2 (pbk.)
1. Island arcs. 2. Rocks, Carbonate—Analysis. 3. Convergent margins. 4. Plate tectonics. 5. Geology,
Stratigraphic. I. Draut, Amy E. II. Clift, P. D. (Peter D.) III. Scholl, David W.
QE511.2.F67 2008
551.1′36—dc22

2007048410

Cover: Oligocene shelf sandstones and shales of the paleo-Chinese passive margin exposed on the
northern coast of Taiwan. The rocks are buried and deformed during the ongoing collision of the Luzon
Arc with the passive margin of southern China. Towards the north of Taiwan the mountains experience
gravitational collapse and extension, culminating in subsidence and formation of the Ilan Plain basin and
Okinawa Trough. Photo by Andrew Lin.

10 9 8 7 6 5 4 3 2 1
Contents

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

1. Preservation of forearc basins during island arc–continent collision: Some insights


from the Ordovician of western Ireland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Paul D. Ryan

2. Basin formation by volcanic arc loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11


Dave Waltham, Robert Hall, Helen R. Smyth, and Cynthia J. Ebinger

3. Cenozoic arc processes in Indonesia: Identification of the key influences on the


stratigraphic record in active volcanic arcs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Robert Hall and Helen R. Smyth

4. Carbonate-platform facies in volcanic-arc settings: Characteristics and controls on


deposition and stratigraphic development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Steven L. Dorobek

5. Sediment waves in the Bismarck Volcanic Arc, Papua New Guinea . . . . . . . . . . . . . . . . . . . . . . . 91


Gary Hoffmann, Eli Silver, Simon Day, Eugene Morgan, Neal Driscoll, and Daniel Orange

6. The Lichi Mélange: A collision mélange formation along early arcward backthrusts
during forearc basin closure, Taiwan arc-continent collision . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Chi-Yue Huang, Chih-Wei Chien, Bochu Yao, and Chung-Pai Chang

7. Oblique subduction in an island arc collision setting: Unique sedimentation, accretion,


and deformation processes in the Boso TTT-type triple junction area, NW Pacific . . . . . . . . . . 155
Yujiro Ogawa, Yoshihiro Takami, and Sakiko Takazawa

8. The West Crocker formation of northwest Borneo: A Paleogene accretionary prism . . . . . . . . 171
Joseph J. Lambiase, Tan Yaw Tzong, Amelia G. William, Michael D. Bidgood, Patrice Brenac,
and Andrew B. Cullen

9. Temporal changes in the composition of Miocene sandstone related to collision


between the Honshu and Izu Arcs, central Japan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
Koichi Okuzawa and Ken-ichiro Hisada

10. Cenozoic volcanic arc history of East Java, Indonesia: The stratigraphic record
of eruptions on an active continental margin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
Helen R. Smyth, Robert Hall, and Gary J. Nichols

iii
iv Contents

11. New constraints on the sedimentation and uplift history of the Andaman-Nicobar
accretionary prism, South Andaman Island . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
R. Allen, A. Carter, Y. Najman, P.C. Bandopadhyay, H.J. Chapman, M.J. Bickle, E. Garzanti,
G. Vezzoli, S. Andò, G.L. Foster, and C. Gerring

12. Post-collisional collapse in the wake of migrating arc-continent collision in the


Ilan Basin, Taiwan. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Peter D. Clift, Andrew T.S. Lin, Andrew Carter, Francis Wu, Amy E. Draut, T.-H. Lai, L.-Y. Fei,
Hans Schouten, and Louis Teng

13. The Guerrero Composite Terrane of western Mexico: Collision and subsequent
rifting in a supra-subduction zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
E. Centeno-García, M. Guerrero-Suastegui, and O. Talavera-Mendoza

14. Tectonic architecture of an arc-arc collision zone, Newfoundland Appalachians . . . . . . . . . . . 309


Alexandre Zagorevski, Cees R. van Staal, Vicki McNicoll, Neil Rogers, and Pablo Valverde-Vaquero

15. The Catalina Schist: Evidence for middle Cretaceous subduction erosion of
southwestern North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
M. Grove, G.E. Bebout, C.E. Jacobson, A.P. Barth, D.L. Kimbrough, R.L. King, Haibo Zou,
O.M. Lovera, B.J. Mahoney, and G.E. Gehrels

16. Sedimentary response to arc-continent collision, Permian, southern Mongolia . . . . . . . . . . . . 363


C.L. Johnson, J.A. Amory, D. Zinniker, M.A. Lamb, S.A. Graham, M. Affolter, and G. Badarch

17. Links among mountain building, surface erosion, and growth of an accretionary
prism in a subduction zone—An example from southwest Japan . . . . . . . . . . . . . . . . . . . . . . . . 391
Gaku Kimura, Yujin Kitamura, Asuka Yamaguchi, and Hugues Raimbourg

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
Preface
The Sedimentary Record in Arc Collision Zones

Sediment deposited in volcanic arc settings yields the best available record of active-margin evolu-
tion spanning long periods of geologic time. The tectonic, erosional, and magmatic geochemical histories
preserved in arc sedimentary records, and particularly in arc collision zones, hold the key to understanding
multiple, inter-connected geologic processes: the formation and destruction of continental crust, changes in
plate configuration and rates of plate motion, subduction-zone deformation and associated seismogenesis,
and, as is becoming increasingly apparent, links between tectonically driven rock uplift and climate.
Arc sedimentary and volcanic records from many locations have been used to decipher tectonic evolu-
tion of modern and ancient convergent margins (e.g., Dickinson et al., 1982; Scholl et al., 1983; Tatsumi et
al., 1983; Ingersoll, 1983; Harbert et al., 1986; Dewey and Ryan, 1990; Gill et al., 1994; Fackler-Adams et
al., 1997; Trop et al., 2003; Busby et al., 2005; Clift et al., 2005; Huang et al., 2006), and geochemical studies
of arc sediment and volcaniclastic deposits have done much to illuminate possible origins of continental crust
(e.g., Pearcy et al., 1990; Miller and Christensen, 1994; Draut et al., 2002). Though much has been learned
from these records and their applications thus far, many important questions remain regarding controls on for-
mation of the sedimentary record in arc environments, where tectonically generated topography is typically
complex. A greater understanding of how and where accommodation space is created, and of depositional and
reworking processes in arc settings, is essential for determining the ability of arc sediment to preserve accu-
rately the tectonic and magmatic signals for which they are commonly studied (e.g., Marsaglia and Ingersoll,
1992; Busby and Ingersoll, 1995; Draut and Clift, 2006). Better comprehension of how arc sediment forms
and evolves will facilitate new and more advanced applications of arc stratigraphy as well as, perhaps, inhibit-
ing inferences for which the sedimentary record provides insufficient or inconclusive evidence.
This compilation of papers resulted indirectly from a Geological Society of America Penrose Con-
ference held in October 2005, co-sponsored by the British Sedimentological Research Group (Geological
Society of London) and the International Association of Sedimentologists. Approximately half of the papers
in this volume were authored by participants in that meeting, which focused on lessons in tectonics, climate,
and eustasy gained from the stratigraphic record preserved in arc collision zones.
The first set of papers in this collection focuses on formation and evolution of the sedimentary record in
arc settings and arc collision zones, concentrating on modern intra-oceanic examples. Two modeling papers
examine basin formation around intra-oceanic arcs: Ryan presents a model indicating that forearc-basin
topography can evolve as a function of eclogite formation, which causes basin subsidence (forming accom-
modation space) and may explain the problematic preservation of forearc basins during arc-continent colli-
sion and orogeny. Waltham et al. model flexural subsidence due to loading of arc and forearc lithosphere by
a volcanic arc, and find that the predicted loads are sufficient to account for much or all of basin subsidence
within 100 km of active arc volcanoes, a process suggested to be particularly common in older, mature arcs.
A related study of Indonesian arc stratigraphy by Hall and Smyth identifies key influences on development of
the sedimentary record in active volcanic arcs, including complexities of basin stratigraphy such as extension,
subduction polarity reversal, collision events, and fluctuating accommodation space associated with variable
loading of the lithosphere by the arc itself. Dorobek presents a comprehensive analysis of evolving carbonate
sedimentation processes around volcanic arcs through time. He shows that although voluminous volcanic
activity is harmful to reef growth, most of the volcaniclastic output of arc volcanoes is trapped in basins.
This allows carbonate reefs to act as sensitive recorders of vertical tectonic motions in forearc regions. Hoff-
mann et al., using new high-resolution seismic images from Papua New Guinea, examine the reworking of
volcaniclastic sediment around volcanic arcs; their study demonstrates that variations in sediment thickness
caused by the formation of current-driven bedforms warrant caution when comparing bed thicknesses across
sedimentary sections sampled in cores. Huang et al. explore mélange development in the active arc-continent
collision zone of Taiwan. Ogawa et al. review the tectonic development of the unique trench-trench-trench
triple junction between the Honshu and Izu Arcs, now located south of Tokyo. They also present a physical
analogue model showing formation of synthetic and antithetic faulting and topography within this oblique
arc collision zone and evaluate stress partitioning between the basement and overlying sediment cover.

v
vi Preface

The second half of the volume presents new applications of arc sedimentary records. Lambiase et al. use
new sedimentology, micropaleontology, and structural analysis of turbidite sequences in Borneo to interpret
them as an accretionary prism and examine in detail the syn-depositional deformation structures that occur
within such a system. Okuzawa and Hisada have reconstructed the tectonic evolution of collision between
the Izu and Honshu Arcs, Japan, using provenance of pre- and syn-collisional Miocene clastic deposits.
They focus on heavy-mineral studies of clinopyroxene, garnet, and spinel to show a switch in provenance in
the trench fill during the middle Miocene driven by initial uplift and erosion of the Honshu Arc after colli-
sion with the Izu arc massif. Smyth et al. trace the complete eruptive history of a well preserved Indonesian
Eocene–Miocene volcanic arc from initiation to termination, and present evidence for contamination of vol-
canism by Archean–Cambrian continental crust underneath the arc. These authors also note that the volume
of acidic arc volcanism has previously been underestimated because weathered acidic volcaniclastic material
was misinterpreted as terrigenous erosion products. Allen et al. use sediment provenance, 40Ar-39Ar mica
dating, and fission-track data to constrain the uplift history and sources of sediment entering the accretionary
prism of the Sunda subduction zone as recorded in the Andaman Islands. This study demonstrated that the
rising Himalaya did not contribute greatly to this accretionary wedge, but that erosion switched from being
local, arc-derived to continental Myanmar around 40 Ma. Clift et al. use a variety of geological and geophysi-
cal data sets to reconstruct exhumation and subsidence rates along the Ilan Basin, Taiwan (western limit of
the Okinawa Trough), to show that subduction polarity reversal associated with a migrating, steady-state
oblique collision between the Luzon arc and the passive margin of southern China causes rapid formation
of deep marginal basins along reactivated detachment faults that are filled by detritus from the orogen. In a
comprehensive summary paper, Centeno-Garcia et al. demonstrate the utility of arc lithofacies and isotopic
signatures to reconstruct the highly complex tectonic evolution of western Mexico, including multiple stages
of terrane collision and rifting. Zagorevski et al. present new U-Pb ages constraining timing of volcanism,
sedimentation, and the age of Ordovician arc-arc collision during the closure of the Iapetus Ocean and
formation of the Appalachian-Caledonide orogen, developing a tectonic model analogous to modern arc-
arc collisions in the southwestern Pacific. Grove et al. use new detrital zircon U-Pb age data to propose a
subduction-erosion origin for portions of the Catalina Schist in the western United States, in contrast to the
long-held theory that the metamorphic complex formed entirely during nascent subduction. The final two
papers provide new insights into possible tectonic-climate coupling in arc-collisional settings: Johnson et
al. use provenance, paleoclimatic, and stratigraphic data to construct a regional synthesis of Permian arc-
continent collision in Eurasia, encompassing the timing of tectonic events and implications for climatic
changes associated with orographic effects of regional uplift. Kimura et al. investigate intriguing feedback
mechanisms among orogenesis, surface erosion via the Asian monsoon, growth of the Nankai Trough accre-
tionary prism by incorporation of sediment eroded from the orogen, and generation of large earthquakes that
promote additional sediment erosion.
It is expected that the utility and applications of arc sedimentary records will expand considerably
with upcoming research initiatives such as those of the Integrated Ocean Drilling Program (IODP) that will
address processes in modern arc settings, including the Nankai Trough, Izu-Bonin forearc, and the collision
zone between the Yakutat terrane and North America, among others. New advances in analytical techniques
will improve age controls on tectonic evolution of active margins and, it is hoped, offer greater insight into
the links and feedback processes among collision-related uplift, precipitation, sediment flux to subduction
zones, and seismic activity with its associated societal hazards.

Amy E. Draut
Peter D. Clift
David W. Scholl
Preface vii

basin of the Aleutian terrace: Relation to North Pacific tec-


REFERENCES CITED tonic events and the formation of the Aleutian subduction
complex: Geology, v. 14, p. 757–761, doi: 10.1130/0091-
Busby, C.J., and Ingersoll, R.V., eds., 1995, Tectonics of sedi- 7613(1986)14<757:MEPOAF>2.0.CO;2.
mentary basins: Blackwell Science, 592 p. Huang, C.-Y., Yuan, P.B., and Tsao, S.J., 2006, Temporal and
Busby, C.J., Bassett, K., Steiner, M.B., and Riggs, N.R., 2005, spatial records of active arc-continent collision in Taiwan:
Climatic and tectonic controls on Jurassic intra-arc basins A synthesis: Geological Society of America Bulletin,
related to northward drift of North America, in Anderson, v. 118, p. 274–288, doi: 10.1130/B25527.1.
T.H., Nourse, J.A., McKee, J.W., and Steiner, M.B., eds., Ingersoll, R.V., 1983, Petrofacies and provenance of Late Meso-
The Mojave-Sonora megashear hypothesis: Development, zoic forearc basin, northern and central California: The
assessment, and alternatives: Geological Society of Amer- American Association of Petroleum Geologists Bulletin,
ica Special Paper 393, p. 359–376. v. 67, p. 1125–1142.
Clift, P.D., Pavlis, T., Debari, S.M., Draut, A.E., Rioux, M., and Marsaglia, K.M., and Ingersoll, R.V., 1992, Compositional
Kelemen, P.B., 2005, Subduction erosion of the Juras- trends in arc-related, deep-marine sand and sandstone: A
sic Talkeetna-Bonanza arc and the Mesozoic accretion- reassessment of magmatic-arc provenance: Geological
ary tectonics of western North America: Geology, v. 33, Society of America Bulletin, v. 104, p. 1637–1649, doi:
p. 881–884. 10.1130/0016-7606(1992)104<1637:CTIARD>2.3.CO;2.
Dewey, J.F., and Ryan, P.D., 1990, The Ordovician evolution of Miller, D.J., and Christensen, N.I., 1994, Seismic signature and
the South Mayo trough, western Ireland: Tectonics, v. 9, geochemistry of an island arc: A multidisciplinary study
p. 887–901. of the Kohistan accreted terrane, northern Pakistan: Jour-
Dickinson, W.R., Ingersoll, R.V., Cowan, D.S., Helmond, nal of Geophysical Research, v. 99, p. 11623–11642, doi:
K.P., and Suczek, C.A., 1982, Provenance of Franciscan 10.1029/94JB00059.
graywackes in coastal California: Geological Society of Pearcy, L.G., Debari, S.M., and Sleep, N.H., 1990, Mass balance
America Bulletin, v. 93, p. 95–107, doi: 10.1130/0016- calculations for two sections of island arc crust and impli-
7606(1982)93<95:POFGIC>2.0.CO;2. cations for the formation of continents: Earth and Plane-
Draut, A.E., Clift, P.D., Hannigan, R.E., Layne, G., and Shi- tary Science Letters, v. 96, p. 427–442, doi: 10.1016/0012-
mizu, N., 2002, A model for continental crust genesis by 821X(90)90018-S.
arc accretion: Rare earth element evidence from the Irish Scholl, D.W., Vallier, T.L., and Stevenson, A.J., 1983, Arc,
Caledonides: Earth and Planetary Science Letters, v. 203, forearc, and trench sedimentation and tectonics; Amlia
p. 861–877, doi: 10.1016/S0012-821X(02)00931-7. corridor of the Aleutian Ridge, in Watkins, J.S., and
Draut, A.E., and Clift, P.D., 2006, Sedimentary processes in Drake, C.L., eds., Studies in continental margin geology:
modern and ancient oceanic arc settings: evidence from the American Association of Petroleum Geologists Memoir
Jurassic Talkeetna Formation of Alaska and the Mariana 34, p. 413–439.
and Tonga arcs, western Pacific: Journal of Sedimentary Tatsumi, Y., Sakuyama, M., Fukuyama, H., and Kushiro, I.,
Research, v. 76, p. 493–514, doi: 10.2110/jsr.2006.044 1983, Generation of arc basalt magmas and thermal struc-
Fackler-Adams, B.N., Busby, C.J., and Mattinson, J.M., 1997, ture of the mantle wedge in subduction zones: Journal of
Jurassic magmatism and sedimentation in the Palen Moun- Geophysical Research, v. 88, p. 5815–5825.
tains, southeastern California: Implications for regional Trop, J.M., Ridgway, K.D., and Spell, T.L., 2003, Sedimentary
tectonic controls on the Mesozoic magmatic arc: Geologi- record of transpressional tectonics and ridge subduction
cal Society of America Bulletin, v. 109, p. 1464–1484, doi: in the Tertiary Matanuska Valley–Talkeetna Mountains
10.1130/0016-7606(1997)109<1464:JMASIT>2.3.CO;2. forearc basin, southern Alaska, in Sisson, V.B., Roeske,
Gill, J.B., Hiscott, R.N., and Vidal, P., 1994, Turbidite geo- S.M., and Pavlis, T.L., eds., Geology of a transpressional
chemistry and evolution of the Izu-Bonin Arc and con- orogen during ridge-trench interaction along the North
tinents: Lithos, v. 33, p. 135–168, doi: 10.1016/0024- Pacific margin: Geological Society of America Special
4937(94)90058-2. Paper 371, p. 89–118.
Harbert, W., Scholl, D.W., Vallier, T.L., Stevenson, A.J., and
Mann, D.M., 1986, Major evolutionary phases of a forearc
The Geological Society of America
Special Paper 436
2008

Preservation of forearc basins during island arc–continent collision: Some insights from
the Ordovician of western Ireland

Paul D. Ryan*
Department of Earth and Ocean Sciences, National University of Ireland, Galway, Ireland

ABSTRACT

A new model is proposed for the problematic preservation of an Ordovician


forearc basin, which records a complete sedimentary record of arc-continent colli-
sion during the Grampian (Taconic) orogeny in the west of Ireland. The South Mayo
Trough represents an arc and forearc complex developed above a subduction zone
in which the slab dipped away from the Laurentian passive margin. The collision
of this arc with Laurentia caused the Middle Ordovician Grampian orogeny. How-
ever, the South Mayo Trough, in the hanging wall of this collision zone, remained a
site of marine sedimentation during the entire process. Early sediments show der-
ivation from an island-arc complex, an ophiolitic backstop, and polymetamorphic
trench sediments. These are conformably overlain by marine deposits derived from a
more evolved arc complex and an emerging juvenile orogen. This transition is dated
as being coeval with the Grampian metamorphism of the Laurentian footwall. The
problem remains as to why subsidence continued in a basin on the hanging wall. It is
proposed that the suppression of the expected topography is due to the nature of the
Laurentian continental margin. Geophysical and geological evidence suggests that
this was a volcanic margin during Neoproterozoic rifting. It is argued that the sub-
duction of this margin caused the formation of eclogites, which reduced its buoyancy.
Simple numerical models are presented which show that this is a viable mechanism
for the suppression of topography during early stages of arc-continent collision and
hence for the preservation of forearcs.

Keywords: forearc, preservation, Grampian Orogeny, arc-continent collision, eclogite.

INTRODUCTION (Ryan and Dewey, 1991); the early Eocene of Kamchatka (Kon-
stantinovskaia, 2000, 2001); the Neoproterozoic of the Tuva-
Island arc–continent collision is an important mechanism Mongolia Massif (Kuzmichev et al., 2001) and of the Anti-Atlas
for continental growth. However, proposed models for this pro- of Morocco (Thomas et al., 2002). This process can result in slab
cess are varied and complex. A forearc can collide with a passive break-off and a reversal of subduction polarity (McKenzie, 1969)
continental margin (Fig. 1A), initiating orogeny at that margin, such as in South Mayo (Ryan and Dewey, 1991) or Taiwan (Teng,
as suggested for the Middle Ordovician of South Mayo, Ireland 1996; Clift et al., 2003) and the creation of an active continental

*paul.ryan@nuigalway.ie

Ryan, P.D., 2008, Preservation of forearc basins during island arc–continent collision: Some insights from the Ordovician of western Ireland, in Draut, A.E., Clift,
P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special Paper 436,
p. 1–9, doi: 10.1130/2008.2436(01). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All rights reserved.

1
2 Ryan

Figure 1. Sketches showing various pro-


posed mechanisms affecting the forearc
B C region during island arc–continent col-
lision: (A) Initial precollision configu-
ration. (B) Obduction of the arc and
forearc, followed by subduction polar-
ity reversal. (C) Obduction of the arc
and forearc, followed by initiation of a
new subduction zone elsewhere in the
system, which may have either polarity.
(D) Weakness in the arc or backarc leads
D to the partial or complete subduction of
the forearc and arc without reversal of
sense of subduction with accompany-
ing ultrahigh pressure metamorphism
of the continental margin. (E) Two op-
posing magmatic arcs, or an island arc
and a magmatic arc collide. (F) Neither
forearc is obducted, and a transform
E F boundary is created.

margin (Fig. 1B). Alternatively, a new subduction zone can initi- wall of an orogeny and is likely to be removed rapidly by ero-
ate elsewhere in the plate to the rear of the arc (Fig. 1C), as has sion, especially if slab break-off occurs. Alternatively, it may be
been proposed for the late Archaean Wutaishan orogeny (Wang et subducted or dismembered by strike-slip faulting.
al., 1996). The forearc can be either thrust in its entirety over the This contribution discusses the geology of western Ireland,
footwall, the preferred model for South Mayo (Ryan and Dewey, where an Ordovician island arc is believed to have collided with
1991; Dewey and Mange, 1999; Clift et al., 2004), or become the Laurentian passive margin (Ryan and Dewey, 1991), causing
part of an imbricated stack driving subsidence in a foreland basin the Grampian orogeny (475–462 Ma) prior to the final closure of
developed upon the footwall as in Taiwan (van der Werff, 1995). the Iapetus Ocean in Late Silurian times (420–416 Ma). Remark-
Alternatively, modeling shows that if the overriding oceanic plate ably, a significant proportion of the forearc, including a possible
fails in the backarc or arc region, the arc and forearc can be wholly forearc basin (Dewey and Ryan, 1990), the South Mayo Trough, is
or partly subducted (Fig. 1D) if they are not sufficiently buoyant preserved in spite of the complex having been in the hanging wall
(Konstantinovskaia, 2000; Boutelier et al., 2003). This leads to an of the earliest orogenic phase. It will be argued that the geology
early orogenic event, with little of the arc complex preserved and of the footwall to this system, the Laurentian continental margin,
perhaps ultrahigh pressure metamorphism in the continental mar- was critical in the preservation of the South Mayo forearc.
gin. A forearc can also collide with an active margin (Fig. 1E), a
model used for the collision of the Kohistan arc with Asia during ORDOVICIAN GEOLOGY OF WESTERN IRELAND
the Late Cretaceous (Khan et al., 1997). This should produce a
climactic orogenic event following a phase of crustal growth by The main geological divisions of western Ireland are shown
addition of subduction-related magmatism. However, an alternate in Figure 2. The rocks to the north of the Fair Head–Clew Bay
model is that the Kohistan backarc collided with an active Asian Line have a Grenvillian basement (Daly, 1996, 2001; Daly and
margin (Rolland et al., 2000; O’Brien, 2001; Robertson and Col- Flowerdew, 2005), overlain by Neoproterozoic sediments of the
lins, 2002). Dalradian Supergroup, which were deposited from pre–720 Ma
In all such models the preservation potential of the forearc to post–595 Ma (Condon and Prave, 2000). Mafic dikes cut the
is low. If obducted, it lies at high structural levels in the hanging basement, postdate all Grenvillian fabrics, and are isotopically
Insights from the Ordovician of western Ireland 3

Figure 2. (A) Geology of western and northwestern Ireland. (B) Geology of South Mayo. Fm.—Formation; Gp.—Group; Cplx.—complex;
FCL—Fair Head–Clew Bay Line; SUF—Southern Upland Fault.

and geochemically matched with pre-deformational dikes in the as the westward continuation of the Highland Boundary Fault in
overlying Dalradian (Daly and Flowerdew, 2005). Both base- Scotland. The complex is highly sheared and comprises two main
ment and cover were affected by the Grampian orogeny between components: a sedimentary mélange underlying most of Clew
475 Ma and 462 Ma (Daly and Flowerdew, 2005; Flowerdew et Bay (Dewey and Ryan, 1990) and a dismembered ophiolite in
al., 2005). In the extreme south the Dalradian Supergroup con- the south (Ryan et al., 1983). Heavy mineral (Dewey and Mange,
tains mafic volcanics with crossite (Gray and Yardley, 1979). 1999) and isotopic studies (Chew, 2003) support a mature con-
This region is commonly interpreted as making up part of the tinental provenance for the sediments. The Clew Bay Complex
Laurentian passive margin to the Iapetus Ocean, which opened has been variously interpreted as part of the Dalradian (Phillips,
in latest Neoproterozoic times, the Dalradian sediments having 1973), a lateral equivalent to the Highland Border Complex of
been deposited during rifting and breakup of Rodinia (Strachan Scotland (Harper et al., 1989), a shear carpet formed beneath
and Holdsworth, 2000). advancing ophiolitic nappes (Dewey and Shackleton, 1984), or
The Clew Bay Complex lies to the south of the Fair Head– an accretionary complex above a south-verging subduction zone
Clew Bay Line and is associated with a strong magnetic lineament (Dewey and Ryan, 1990). The age of the complex is uncertain,
(Max and Riddihough, 1975; Max et al., 1983) and is regarded although a Cambrian–Ordovician age is likely on provenance
4 Ryan

grounds (Dewey and Mange, 1999), and it shares the same struc- coeval with the emplacement of the D2 gabbros in Connemara at
tural history as the Dalradian to the north with the S2 fabric in 473 Ma (Dewey and Mange, 1999). Radiometric evidence from
both dated at ca. 460 Ma (Chew et al., 2003). Provenance stud- both North Mayo and Connemara suggests that the entire event
ies also suggest a stratigraphic linkage between this complex and took <20 m.y., with initial collision taking ~10 m.y. and subduc-
the South Mayo Trough by Early to Middle Ordovician times tion flip a further 5 m.y. (Dewey and Mange, 1999; Soper et al.,
(Dewey and Ryan, 1990). 1999; Draut and Clift, 2001).
The South Mayo Trough (Fig. 2B) contains a thick sequence Evidence for the reversal of subduction polarity beneath
of Lower to Middle Ordovician sediments that are preserved in a the Laurentian margin after the Grampian orogeny is found in
large syncline, the Mweelrea-Partry Syncline. The northern limb Middle to Upper Ordovician and Silurian strata of Connemara
contains an Arenig-Llanvirn sequence that exceeds 9 km in thick- and South Mayo. The South Connemara Complex, to the south
ness in which early sediments show derivation from a primitive of the Southern Upland Fault (SUF, Fig. 2), was probably formed
island arc complex, an ophiolitic backstop, and polymetamorphic in a Late Ordovician trench of opposite polarity adjacent to the
trench sediments. These are conformably overlain by marine Laurentian margin (Ryan and Dewey, 2004). Silurian deposits
clastic sequences derived from a more evolved arc complex and in north Connemara and South Mayo (Fig. 2) contain abundant
an emerging juvenile (Grampian) orogen (Dewey and Mange, arc detritus (Menuge et al., 1995; Dewey and Mange, 1999)
1999). Associated volcanic complexes, which lie mainly in the and probably formed in a supra-subduction zone setting (Wil-
southern limb, are of primitive arc affinities near the base (Ryan liams and Harper, 1991; Menuge et al., 1995). These deposits are
et al., 1980; Clift and Ryan, 1994; Draut et al., 2004) and show an mostly shallow marine and of 2.5–3 km in thickness, suggesting
increasing proportion of Laurentian-derived melt upsection (Draut only modest stretching of a normally buoyant lithosphere without
et al., 2004). Dewey and Ryan (1990) interpreted the South Mayo topography (equivalent to a β factor of ~1.25). Critically, where
Trough as a forearc basin on the grounds that it lay between an arc the Silurian sedimentary and volcanic sequences rest unconform-
(volcanic complexes) and an accretionary prism (the Clew Bay ably upon the Ordovician rocks, the earliest slaty cleavage, if pres-
Complex) which in turn had been thrust over a passive continen- ent, cuts both sequences. Structural and stratigraphic evidence,
tal margin, Laurentia, whose leading edge had been subducted. It therefore, suggests that prior to Silurian deposition the Ordovi-
is remarkable that the South Mayo Trough, interpreted as lying cian sequence of South Mayo was only modestly deformed, of
in the obducted hanging wall, apparently provides a continuous anchimetamorphic grade, and near sea level (Dewey and Ryan,
record of sedimentation during Grampian deformation of the 1990). If the arc-continent collision model is accepted for South
Laurentian margin, interpreted as lying in the footwall. Mayo, a mechanism must, therefore, have operated to suppress
The Dalradian rocks of Connemara (Fig. 1) are in fault con- the topography of the forearc during and after this event.
tact with the South Mayo Trough (Dewey et al., 1970; Ryan and
Archer, 1978) and were probably emplaced outboard of the oce- ISOSTATIC CONSEQUENCES OF ARC-CONTINENT
anic terrane of South Mayo by Late Ordovician lateral motion COLLISION
(Hutton, 1987). The southern boundary is a SE-verging thrust,
the Mannin Thrust (Leake et al., 1983), which emplaced migma- The likely topographic response of an arc-continent colli-
titic metasediments and intrusives over evolved arc rocks dated sion is investigated in two ways. Firstly, a flexural isostatic model
at 474.6 ± 5.5 Ma (Friedrich, 1998; Draut and Clift, 2002) that is used to estimate the uplift of the hanging wall for a 10 m.y.
show chemical similarities to volcanics within the South Mayo period from 485 to 475 Ma (Dewey and Mange, 1999; Draut and
Trough. This suggests that Connemara lay structurally beneath Clift, 2002), when arc-continent collision was thought to have
an evolving arc by Llanvirn times (Draut and Clift, 2002). U/Pb taken place. Secondly, a simple model, assuming Airy isostasy,
studies of syntectonic gabbros and later granitic melts date D2 of is used to estimate the likely uplift during slab break-off, which
the Grampian orogeny at 473 Ma, and post-D3 rapid uplift that is assumed to have led to the emplacement of mafic plutons and
took place before 462 Ma (Friedrich et al., 1999a, b). quartz-diorite gneisses in Connemara between 475 and 470 Ma
The Grampian (Taconic) orogeny was a Middle Ordovician (Friedrich et al., 1999a, b). The aim is to investigate the condi-
event that affected the Laurentian margin prior to final closure tions that are required to keep the forearc basin beneath sea level
of the Iapetus Ocean, resulting from the collision of an arc with during both of these events.
the Laurentian margin, followed by a subduction polarity reversal The flexural response of the lithosphere during collision, but
(Dewey and Shackleton, 1984; Ryan and Dewey, 1991). Island prior to slab break-off, is estimated using a two-dimensional (2D)
arc–continent collision began in Arenig times (Draut and Clift, kinematic flexural model. The present across-strike width of the
2001), with hard collision being reflected in an increase of conti- rocks affected by the Grampian orogeny (the Grampian and North-
nental component in the arc melt (Draut and Clift, 2001; Draut et Western terranes of Murphy et al., 1991) is ~100–140 km. Rocks
al., 2004). Provenance studies in the South Mayo Trough suggest to the north in Donegal were metamorphosed to greenschist facies
initiation in an oceanic forearc, with later sediments recording (Pitcher and Berger, 1972), suggesting that the Grampian orogen
the stripping of an evolved arc and a juvenile orogen (Dewey and extended farther north. Strike swing and post-Grampian collapse
Mange, 1999; Clift et al., 2004). This change in provenance was (Alsop, 1991) make it difficult to estimate the original width of
Insights from the Ordovician of western Ireland 5

the orogen, but allowing for such factors, Dewey and Mange (Te) of 2.0 km (Fig. 3A) and 20.0 km (Fig. 3C) were used to
(1999) suggest 200 km as a reasonable value. It is assumed that estimate the response of flexurally weak and flexurally strong
an arc complex, including South Mayo, was emplaced over the systems, respectively. The exact position of the South Mayo
Laurentian margin along a thrust that nucleated upon the sub- forearc basin within such a system is unknown but is taken as
duction zone. A hanging wall of 210 km was moved over a flat being between 100 and 200 km south of the thrust tip.
28 km in depth (ca. 0.8 GPa) for 100 km and then up a ramp at The model with low Te (Fig. 3A) predicts some 3 km of uplift
an angle of 14.2° for a further 110 km (Fig. 3). A specific gravity in the forearc region accompanied by ~1.5 km of erosion and a
of 2200 kg m–3 was assumed for sediments in the flexural basin. flexural basin on the Laurentian margin with 0.5 km of sediment
A value of 2840 kg m–3, the mean density of diorite, was used (Fig. 3B). The model with high Te (Fig. 3C) predicts some 6 km
for both the arc material of the hanging wall and a continental of uplift in the forearc region accompanied by 3–4 km of ero-
basement containing up to 30% mafic material for the footwall. sion and a flexural basin on the Laurentian margin with 3.4 km of
The thrust was moved at 20 mm.yr –1 so that the deformation sediment (Fig. 3D). No Ordovician flexural basin is recorded in
front moved 200 km in 10 m.y. The actual rate of overthrusting Ireland, nor is there any evidence for one farther north offshore
may have been greater (Dewey and Mange, 1999). The hang- on the WIRELINES seismic reflection section (Klemperer et al.,
ing wall was moved along flow lines parallel to the ramp–flat 1991). Thus it is concluded that the overthrusting of the arc hang-
surface. Erosion was modeled using equation 5 below, and the ing wall over the continental margin footwall was likely to have
model was compensated using flexural isostasy (Watts, 2001) taken place in a low Te regime. There still remains the problem
after each increment of movement. Different elastic thicknesses that if the assumptions made in this simple mechanical model

A
B

C
D

Figure 3. (A) Flexural isostatic model in which a hanging wall of 28 km is emplaced at 20 mm.yr –1 over a continental
footwall. The heavy upper line represents the erosion surface after 10 m.y. at the end of the model run. The dashed line
represents the thrust surface. The erosion constant is 5 m.y.–1, and the elastic thickness (Te) is 2 km. (B) Subsidence-time
curve for the foreland basin produced at the toe of the thrust system in (A). (C) A similar model to that in (A) but with Te
= 20 km. (D) Subsidence-time curve for the foreland basin produced at the toe of the thrust system in (C).
6 Ryan

are reasonable (see caption for Fig. 3), the South Mayo Trough where Cz is the thickness of the crust and ρm the specific gravity
was likely to have been above sea level prior to slab break-off at of the mantle. The following values were assumed for the specific
ca. 475 Ma. gravities: ρc = 2840 kg.m–3; ρm = 3330 kg.m–3; and ρa = 3281 kg.
A simple technique is employed to estimate the possible m–3. The value for ρa places the upper surface of the lithosphere
topography associated with subsequent slab break-off. Fig- with Cz = 30 km and lz = 125 km at sea level.
ure 4A shows a typical tectonic cartoon of the process of island If topography above sea level (d) is corrected for the mean
arc–continent collision at the point of subduction reversal (based depth of the mid-ocean ridges we obtain:
upon Dewey and Mange, 1999). Airy compensation is assumed
in the light of the previous flexural analysis. Topography (h)
with respect to the summit of the oceanic ridges is given by (see d + 2.64 = h . (3)
Dewey et al., 1993):

( ρ a − ρ l ) ( lz − h ) When (2) and (3) are substituted in (1) and simplified,


h= , (1) we get:
ρc

where ρc, ρa, ρl are the specific gravities of the crust, astheno- d=
[Cz (ρm − ρc )+ lz (ρa − ρm )] − 2.64. (4)
sphere, and lithosphere, respectively, and lz is the thickness of the ρa
lithosphere. Where ρl may be estimated from:
Equation 4 can be used to estimate the topography of any

ρl =
[(Cz − h) ρc + (lz − Cz ) ρm ],
• •
(2)
geodynamic cartoon if the thicknesses of crust and mantle are
measured at regular intervals across the model (Fig. 4). Although
[lz − h] this may differ from dynamic topography, it provides a simple

Figure 4. (A) Isostatic consequences


of an arc-continent collision and sub-
duction polarity flip model taken from
Dewey and Mange (1999). The heavy
solid line in the upper topographic grid
shows the maximum topography pro-
duced by this model, assuming Airy
compensation; the dashed line shows
the likely topography after 10 Ma ero-
sion at an erosion constant of 5 m.y.–1.
(B) Isostatic consequences of both the
model in Figure 4A and the thrust mod-
els from Figure 3A, with the geology of
B South Mayo speculatively positioned
onto the hanging wall, assuming lower
crustal-phase changes. The heavy lines
in the upper topographic grid show
the likely maximum topography for
the model in Figure 3A, assuming the
lower crust below 40 km is either 100%
eclogite (solid line) or 50% eclogite
(dashed line). The lighter lines show the
maximum topography of the tectonic
model in Figure 4A, assuming that the
lower crust below 40 km is either 100%
eclogite (solid line) or 50% eclogite
(dashed line).
Insights from the Ordovician of western Ireland 7

check on the admissibility of any such cartoon. The height of a material within the basement of either Tertiary or Neoproterozoic
feature zt originally at z0, allowing for subsequent erosion during age. Mafic dikes and sills intruded in Argyll Group rocks occur
a given interval (t m.y.), can be estimated using: throughout the Dalradian sequence and are particularly preva-
lent in Donegal, composing up to 20% of the surface geology
ρc
⎛ . k .t ⎞ in some regions (Pitcher and Berger, 1972). Correlation with the
z t = z0 ⎜ 1 − e ρm ⎟, (5) Tayvallich lavas of Scotland suggest that they were emplaced at
⎜ ⎟
⎝ ⎠ ca. 595 Ma (Halliday et al., 1989). The Dalradian in northwestern
Ireland is associated with positive, moderate-frequency magnetic
where k is an erosion constant (m.y.–1). Values for k vary from anomalies (Armstrong et al., 1998), which are interpreted as hav-
1 m.y.–1 for Midwestern basins in the United States to 5 m.y.–1 for ing been caused by these mafic bodies. In Donegal Bay, Carbon-
the Moroccan Rif and 20 m.y.–1 for the New Zealand Alps (esti- iferous deposits overstep the Dalradian metasediments, and the
mated from data in Goudie, 1995). The topography associated associated broad, negative magnetic anomalies mask the pattern
with slab break-off and the remaining topography after 10 m.y. associated with the Dalradian rocks. This relationship would not
are shown in Figure 4A. Approximately 5 km of material would have occurred if the mafic rocks had been post-Carboniferous.
have been removed from the forearc region, assuming an erosion Furthermore, the WIRELINE survey did not pass close to any
constant k = 5 m.y.–1 (Fig. 4A). known Tertiary seamounts. Thus it is most likely that the base-
This analysis suggests that the South Mayo Trough, or any ment reflectivity is associated with the Neoproterozoic basalt
other similar forearc complex, would have been unlikely to sur- swarms emplaced during rifting in Argyll Group times and that
vive arc-continent collision followed by slab break-off unless the in this region the Laurentian margin was a volcanic margin.
topography was depressed by some other factor. It is unlikely The presence of voluminous basalt within the deep crust
that this was due to some dynamic process, as the South Mayo provides a mechanism for depressing topography during con-
region became a site of Silurian marine sedimentation follow- vergence (Dewey et al., 1993). If mafic rocks convert to eclo-
ing the Grampian orogeny in basins with modest β factors (see gite, their specific gravity increases by up to 20%, and other
above), suggesting that the crust was never far above sea level in continental rocks on converting to their eclogite facies equiva-
the post-Grampian interval. lents will show increases in specific gravity of between 5% and
The model suggests uplift rates in the forearc region on 10% (Dewey et al., 1993). This is likely to occur below 40 km
the order of 2 mm.yr –1 to 5 mm.yr –1 during collision and slab in a refrigerated system such as a subducting continental mar-
break-off. Similar values of 5 mm.yr –1 are recorded in Sumba gin. The effect of any such conversion to eclogite facies for the
where a forearc basin is overriding the Australian margin (van lower crust is to increase the overall specific gravity ρc' of the
der Werff, 1995); this is a region that has been suggested as an crust as follows:
analogy for the Grampian orogeny (Dewey and Mange, 1999).
However, if such rates were sustained over the duration of the ρe ∗ Ez + ρc ∗ (Cz − Ez )
Grampian orogeny, it is unlikely that the South Mayo Trough ρc' = , (6)
would have been preserved. Some feature of the Grampian col- Cz
lision zone must, therefore, have depressed the topography of
this orogen. Three possibilities exist. Firstly, South Mayo may where ρe is the density of the eclogitized crust, and Ez is the
not have been involved in the Grampian arc-continent collision, thickness of the eclogitic layer. Figure 4B shows the effect of
but the abundance of evidence reviewed above makes this seem substituting the modified values for ρc' into equation 4 on the
highly unlikely. Secondly, slab pull may have initially reduced assumption that all the crust in the collision zone beneath 40 km
the topography dynamically. However, the emplacement of mafic is eclogitized and that this region has a basaltic component of
plutons during the principal D2/D3 Grampian deformation in either 100% or 50%.
Connemara is generally believed to have been related to delami- The analysis is performed for two possible scenarios.
nation of the downgoing Iapetan slab at the time of, not after, The first is the model of Dewey and Mange (1999), shown in
peak deformation. A third possibility is that the nature of the con- Figure 4A, which assumes that the crust on the leading edge of
tinental basement in the footwall may have buffered rapid uplift. the Laurentian margin is thickened during convergence. The sec-
ond is based upon the flexural model shown in Figure 3A, which
ROLE OF THE LAURENTIAN BASEMENT assumes that the Laurentian crust is thick beneath the overriding
forearc but thins toward the ocean-continent boundary. The first
Deep reflection seismic studies (WIRELINE, Klemperer et model predicts that the forearc region will have a mean topogra-
al., 1991) show that the basement north of the Fair Head–Clew phy of a few hundred meters above sea level. The frontal region
Bay Line off the shore of western and northwestern Ireland is of the arc would be ~4 km below sea level as a result of the nega-
anomalously reflective between ~1.2 s two-way traveltime tive buoyancy attributable to the thickened eclogite-facies lower
(TWTT) and the Moho at ~10 s TWTT. Klemperer et al. (1991) crust in this region. The presence of lherzolite within the litho-
attributed this to the presence of considerable volumes of basaltic spheric mantle would also depress topography as the depth of
8 Ryan

the mantle slab increases and the aluminous phases convert from REFERENCES CITED
spinel to garnet below 90 km. However, this effect is small. If
Alsop, G.I., 1991, Gravitational collapse and extension along a mid crustal
a 10% lherzolite component containing 10% spinel is assumed, detachment: The Lough Derg Slide, northwest Ireland: Geological Maga-
the maximum depth in this model (Fig. 4B) would increase by zine, v. 128, p. 345–354.
30 m on the southern side of the forearc basin. The second model Armstrong, G.D., Brown, C., and Mars, J.E., 1998, The Irish continental
crust revealed by potential field methods: Technical Program Extended
suggests that the forearc region could be ~200 m below sea level Abstracts, Society of Exploration Geophysicists, 68th Annual Meet-
but that much of the frontal arc region would be above sea level ing, New Orleans, Louisiana, September 1998; CD-ROM and Extended
if there was 100% eclogite. The second model seems to provide Abstracts Volume.
Boutelier, D., Chemenda, A., and Burg, J.-P., 2003, Subduction versus accretion
a better fit to the geology of South Mayo, where sedimentation of intra-oceanic volcanic arcs: Insight from thermo-mechanical analogue
is believed to have continued during collision and slab break-off. experiments: Earth and Planetary Science Letters, v. 212, p. 31–45, doi:
However, given the uncertainties in any such analysis, it is diffi- 10.1016/S0012-821X(03)00239-5.
Chew, D.M., 2003, Structural and stratigraphic relationships across the con-
cult to differentiate between the two models. The important point tinuation of the Highland Boundary Fault in western Ireland: Geological
to note is that significant production of eclogite within the foot- Magazine, v. 140, p. 73–85, doi: 10.1017/S0016756802007008.
wall to the collision zone can depress the topography sufficiently Chew, D.M., Daly, J.S., Page, L.M., and Kennedy, M.J., 2003, Grampian oro-
genesis and the development of blueschist facies metamorphism in west-
to allow sedimentation to continue in the forearc region during ern Ireland: Geological Society [London] Journal, v. 160, p. 911–924.
arc-continent collision. Clift, P.D., and Ryan, P.D., 1994, Geochemical evolution of an Ordovician
island arc, South Mayo, Ireland: Geological Society [London] Journal,
v. 151, p. 329–342.
DISCUSSION AND CONCLUSIONS Clift, P.D., Schouten, H., and Draut, A.E., 2003, A general model of arc conti-
nent collision and subduction polarity reversal from Taiwan and the Irish
It is suggested that the production of significant volumes of Caledonides, in Larter, R., and Leat, P., eds., Intra-Oceanic Subduction
Systems; Tectonic and Magmatic Processes: Geological Society [Lon-
eclogite at the base of the footwall during arc-continent collision don] Special Publication 219, p. 81–98.
may depress the topography sufficiently to preserve the forearc Clift, P.D., Dewey, J.F., Draut, A.E., Chew, D.M., Mange, M., and Ryan, P.D.,
within the hanging wall. This process is likely to be important 2004, Rapid tectonic exhumation, detachment faulting and orogenic
collapse in the Caledonides of western Ireland: Tectonophysics, v. 384,
only along volcanic margins or those that contain older basic p. 91–113, doi: 10.1016/j.tecto.2004.03.009.
massifs. In this context, it is worthy of note that in the Anti-Atlas Condon, D.J., and Prave, A.R., 2000, Two from Donegal: Neoproterozoic epi-
orogen, where a forearc basin (Sarhro Group) has been preserved sodes on the northeast margin of Laurentia: Geology, v. 28, p. 951–954,
doi: 10.1130/0091-7613(2000)28<951:TFDNGE>2.0.CO;2.
during arc-continent collision, the continental basement contains Daly, J.S., 1996, Pre-Caledonian history of the Annagh Gneiss Complex, north-
dike swarms and mafic volcanics (Ifzwane Suite) related to rift- western Ireland, and correlation with Laurentia-Baltica: Irish Journal of
ing (Thomas et al., 2002). Earth Sciences, v. 15, p. 5–8.
Daly, J.S., 2001, The Precambrian, in Holland, C.H., ed., The Geology of Ire-
This analysis has the limitations of any 2D static model. land (2nd edition): Edinburgh, Scottish Academic Press, p. 7–45.
It cannot, for instance, distinguish between the effects of slab Daly, J.S., and Flowerdew, M.J., 2005, Grampian and late Grenville events
break-off along the length of a continental margin or gradual recorded by mineral geochronology near a basement cover contact in north
Mayo, Ireland: Geological Society [London] Journal, v. 162, p. 163–174,
tearing (Clift et al., 2003). However, any out-of-plane effects doi: 10.1144/0016-764903-150.
of a tearing slab should assist in depressing topography. There Dewey, J., and Mange, M., 1999, Petrography of Ordovician and Silurian sedi-
also remains the problem that a root of eclogites has not been ments in the western Ireland Caledonides: Tracers of a short-lived Ordovi-
cian continent arc collision orogeny and the evolution of the Laurentian
recorded in any deep geophysical survey of western Ireland. Appalachian Caledonian margin, in MacNiocaill, C., and Ryan, P.D., eds.,
However, the Grampian orogen underwent extensive orogenic Continental Tectonics: Geological Society [London] Special Publication
collapse in the Late Ordovician (Alsop, 1991; Clift et al., 2004) 164, p. 55–107.
Dewey, J.F., and Ryan, P.D., 1990, The Ordovician evolution of the South Mayo
during which time such a root would have been exhumed and Trough, western Ireland: Tectonics, v. 9, p. 887–901.
retrogressed to amphibolite. The buoyancy gained during the Dewey, J.F., and Shackleton, R.M., 1984, A model for the evolution of the
decompression of the eclogites could have buffered topography Grampian tract in the early Caledonides and Appalachians: Nature, v. 312,
p. 115–121, doi: 10.1038/312115a0.
and limited subsidence during this phase (Dewey et al., 1993). Dewey, J.F., McKerrow, W.S., and Moorbath, S., 1970, The relationship
Such a mechanism may account for the cessation of sedimenta- between isotopic ages, uplift and sedimentation during Ordovician times
tion within the South Mayo Trough. An alternative explanation in western Ireland: Scottish Journal of Geology, v. 6, p. 133–145.
Dewey, J.F., Ryan, P.D., and Andersen, T.B., 1993, Orogenic uplift and col-
would be to apply the model of Jull and Kelemen (2001) and lapse, crustal thickness, fabrics and metamorphic phase changes: The
suggest that the crustal eclogitic root beneath Laurentia became role of eclogites, in Prichard, H.M., et al., eds., Magmatic Processes and
detached and sank into the mantle in late post-Grampian times. Plate Tectonics: Geological Society [London] Special Publication 76,
p. 325–343.
This would have driven rapid uplift in the arc region of the model Draut, A.E., and Clift, P.D., 2001, Geochemical evolution of arc magmatism dur-
in Figure 4A and in the accretionary complex in the model of ing arc-continent collision, South Mayo, Ireland: Geology, v. 29, p. 543–
Figure 4B. Both regions were sites of Silurian sedimentation 546, doi: 10.1130/0091-7613(2001)029<0543:GEOAMD>2.0.CO;2.
Draut, A.E., and Clift, P.D., 2002, The origin and significance of the Delaney
(Fig. 2) onto a low-metamorphic-grade basement, making such Dome Formation, Connemara, Ireland: Geological Society [London]
a process unlikely, at least beneath the South Mayo Trough. Journal, v. 159, p. 95–103.
Insights from the Ordovician of western Ireland 9

Draut, A.E., Clift, P.D., Chew, D.M., Cooper, M.J., Taylor, R.N., and Hannigan, Murphy, F.C., Anderson, T.B., Daly, J.S., Gallagher, V., Graham, J.R., Harper,
R.E., 2004, Laurentian crustal recycling in the Ordovician Grampian Orog- D.A.T., Johnston, J.D., Kennan, P.S., Kennedy, M.J., Long, C.B., Morris,
eny: Nd isotopic evidence from South Mayo, western Ireland: Geological J.H., O’Keefe, W.G., Parkes, M., Ryan, P.D., Sloan, R.J., Stillman, C.J.,
Magazine, v. 141, p. 195–207, doi: 10.1017/S001675680400891X. Tietzsch-Tyler, D., Todd, S.P., and Wrafter, J.P., 1991, An appraisal of
Flowerdew, M.J., Daly, J.S., and Whitehouse, M.J., 2005, 470 Ma granitoid Caledonian suspect terranes in Ireland: Irish Journal of Earth Sciences,
magmatism associated with the Grampian Orogeny in the Slishwood v. 11, p. 11–41.
Division, NW Ireland: Geological Society [London] Journal, v. 162, O’Brien, P.J., 2001, Subduction followed by collision: Alpine and Himalayan
p. 563–575, doi: 10.1144/0016-784904-067. examples: Physics of the Earth and Planetary Interiors, v. 127, p. 277–
Friedrich, A.M., 1998, 40Ar/39Ar and U-Pb geochronological constraints on 291, doi: 10.1016/S0031-9201(01)00232-1.
the thermal and tectonic evolution of the Connemara Caledonides [Ph. Phillips, W.E.A., 1973, The pre-Silurian rocks of Clare Island, Co. Mayo, Ire-
D. thesis]: Cambridge, Department of Earth, Atmospheric and Planetary land, and the age of metamorphism of the Dalradian in Ireland: Geologi-
Sciences, Massachusetts Institute of Technology, 228 p. cal Society [London] Journal, v. 129, p. 585–606.
Friedrich, A.M., Hodges, K.V., Bowring, S.A., and Martin, M.W., 1999a, Geo- Pitcher, W.S., and Berger, A.R., 1972, The Geology of Donegal: A Study of
chronological constraints on the magmatic, metamorphic, and thermal Granite Emplacement and Unroofing: New York, John Wiley, 435 p.
evolution of the Connemara Caledonides, western Ireland: Geological Robertson, A.H.F., and Collins, A.S., 2002, Shyok Suture Zone, N Pakistan:
Society [London] Journal, v. 156, p. 1217–1230. Late Mesozoic-Tertiary evolution of a critical suture separating the oce-
Friedrich, A.M., Bowring, S.A., Martin, M.W., and Hodges, K.V., 1999b, Short- anic Ladakh Arc from the Asian continental margin: Journal of Asian
lived continental magmatic arc at Connemara, western Irish Caledonides: Earth Sciences, v. 20, p. 309–351, doi: 10.1016/S1367-9120(01)00041-4.
Implications for the age of the Grampian orogeny: Geology, v. 27, p. 27– Rolland, Y., Pecher, A., and Picard, C., 2000, Middle Cretaceous back-arc for-
30, doi: 10.1130/0091-7613(1999)027<0027:SLCMAA>2.3.CO;2. mation and arc evolution along the Asian margin: The Shyok Suture Zone
Goudie, A.S., 1995, The Changing Earth: Rates of Geomorphological Pro- in northern Ladakh (NW Himalaya): Tectonophysics, v. 325, p. 145–173,
cesses: London, Blackwell, 302 p. doi: 10.1016/S0040-1951(00)00135-9.
Gray, J.R., and Yardley, B.W.D., 1979, A Caledonian blueschist from the Irish Ryan, P.D., and Archer, J.B., 1978, The Lough Nafooey fault: A Taconic
Dalradian: Nature, v. 278, p. 736–737, doi: 10.1038/278736a0. structure in western Ireland: Geological Survey of Ireland Bulletin, v. 2,
Halliday, A.N., Graham, C.R., Aftalion, M., and Dymoke, P., 1989, The depo- p. 255–264.
sitional age of the Dalradian Supergroup: U-Pb and Sm-Nd isotopic stud- Ryan, P.D., and Dewey, J.F., 1991, A geological and tectonic cross-section of
ies of the Tayvallich Volcanics, Scotland: Geological Society [London] the Caledonides of western Ireland: Geological Society [London] Journal,
Journal, v. 146, p. 3–6. v. 148, p. 173–180.
Harper, D.A.T., Williams, D.M., and Armstrong, H.A., 1989, Stratigraphical Ryan, P.D., and Dewey, J.F., 2004, The South Connemara Group reinterpreted:
correlations adjacent to the Highland Boundary Fault in the west of Ire- A subduction-accretion complex in the Caledonides of Galway Bay, west-
land: Geological Society [London] Journal, v. 146, p. 381–384. ern Ireland: Journal of Geodynamics, v. 37, p. 513–529, doi: 10.1016/
Hutton, D.H.W., 1987, Strike slip terranes and a model for the evolution of the j.jog.2004.02.018.
British and Irish Caledonides: Geological Magazine, v. 124, p. 405–425. Ryan, P.D., Floyd, P.A., and Archer, J.B., 1980, The stratigraphy and petro-
Jull, M., and Kelemen, P.B., 2001, On the conditions for lower crustal convec- chemistry of the Lough Nafooey Group (Tremadocian), western Ireland:
tive instability: Journal of Geophysical Research, v. 106, p. 6423–6446, Geological Society [London] Journal, v. 137, p. 443–458.
doi: 10.1029/2000JB900357. Ryan, P.D., Sawal, V.K., and Rowlands, A.S., 1983, Ophiolitic mélange sepa-
Khan, M.A., Stern, R.J., Gribble, R.F., and Windley, B.F., 1997, Geochemical rates ortho- and para-tectonic Caledonides in western Ireland: Nature,
and isotopic constraints on subduction polarity, magma sources and pal- v. 302, p. 50–52, doi: 10.1038/302050a0.
aeogeography of the Kohistan Arc, northern Pakistan: Geological Society Soper, N.J., Ryan, P.D., and Dewey, J.F., 1999, Age of the Grampian Orogeny
[London] Journal, v. 154, p. 935–946. in Scotland and Ireland: Geological Society [London] Journal, v. 148,
Klemperer, S.L., Ryan, P.D., and Snyder, D.B., 1991, A deep seismic reflec- p. 173–180.
tion transect across the Irish Caledonides: Geological Society [London] Strachan, R.A., and Holdsworth, R.E., 2000, Late Neoproterozoic (<750 Ma)
Journal, v. 148, p. 149–164. to Early Ordovician passive margin sedimentation along the Laurentian
Konstantinovskaia, E.A., 2000, Geodynamics of an Early Eocene arc–continent margin of Iapetus, in Woodcock, N., and Strachan, R., eds., Geological
collision reconstructed from the Kamchatka Orogenic Belt, NE Russia: History of Britain and Ireland: London, Blackwell, p. 73–87.
Tectonophysics, v. 325, p. 87–105, doi: 10.1016/S0040-1951(00)00132-3. Teng, L.S., 1996, Extensional collapse of the northern Taiwan mountain belt:
Geology, v. 24, p. 949–952, doi: 10.1130/0091-7613(1996)024<0949:
Konstantinovskaia, E.A., 2001, Arc–continent collision and subduction rever-
ECOTNT>2.3.CO;2.
sal in the Cenozoic evolution of the Northwest Pacific: An example from
Thomas, R.J., Chevallier, L.P., Gresse, P.G., Harmer, R.E., Eglington, B.M.,
Kamchatka (NE Russia): Tectonophysics, v. 333, p. 75–94, doi: 10.1016/
Armstrong, R.A., de Beer, C.H., Martini, J.E.J., de Kock, G.S., Macey,
S0040-1951(00)00268-7.
P.H., and Ingram, B.A., 2002, Precambrian evolution of the Sirwa Win-
Kuzmichev, A.B., Bibikova, E.V., and Zhuravlev, D.Z., 2001, Neoprotero-
dow, Anti-Atlas Orogen, Morocco: Precambrian Research, v. 118, p. 1–
zoic (~800 Ma) orogeny in the Tuva-Mongolia Massif (Siberia): Island
57, doi: 10.1016/S0301-9268(02)00075-X.
arc–continent collision at the northeast Rodinia margin: Precambrian
van der Werff, W., 1995, Cenozoic evolution of the Savu Basin, Indonesia:
Research, v. 110, p. 109–126, doi: 10.1016/S0301-9268(01)00183-8.
Fore-arc basin response to arc-continent collision: Marine and Petroleum
Leake, B.E., Tanner, P.W.G., Singh, D., and Halliday, A.N., 1983, Major south-
Geology, v. 12, p. 247–262, doi: 10.1016/0264-8172(95)98378-I.
ward thrusting of the Dalradian rocks of Connemara, western Ireland:
Wang, K.Y., Li, J.L., Hao, J., Li, J., and Zhou, S., 1996, The Wutaishan oro-
Nature, v. 305, p. 210–213, doi: 10.1038/305210a0.
genic belt within the Shanxi Province, northern China: A record of late
Max, M.D., and Riddihough, R.P., 1975, Continuation of the Highland Bound- Archaean collision tectonics: Precambrian Research, v. 78, p. 95–103,
ary Fault in Ireland: Geology, v. 3, p. 206–210, doi: 10.1130/0091- doi: 10.1016/0301-9268(95)00071-2.
7613(1975)3<206:COTHBF>2.0.CO;2. Watts, A.B., 2001, Isostasy and Flexure of the Lithosphere: Cambridge, Eng-
Max, M.D., Ryan, P.D., and Inamdar, D.D., 1983, A magnetic deep structural land, Cambridge University Press, 458 p.
geology interpretation of Ireland: Tectonics, v. 2, p. 431–451. Williams, D.M., and Harper, D.A.T., 1991, End-Silurian modifications of Ordo-
McKenzie, D.P., 1969, Speculations on the consequences and causes of plate vician terranes in western Ireland: Geological Society [London] Journal,
motions: Geophysical Journal of the Royal Astronomical Society, v. 18, v. 148, p. 165–171.
p. 1–32.
Menuge, J.F., Williams, D.M., and O’Connor, P.D., 1995, Silurian turbidites
used to reconstruct a volcanic terrain and its Mesoproterozoic basement
in the Irish Caledonides: Geological Society [London] Journal, v. 152,
p. 269–278. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Basin formation by volcanic arc loading

Dave Waltham
Robert Hall
Helen R. Smyth
Cynthia J. Ebinger
SE Asia Research Group, Department of Geology, Royal Holloway University of London, Egham, Surrey, TW20 0EX, UK

ABSTRACT

This paper quantifies the flexural subsidence expected from loading by a volcanic
arc. The resulting mathematical model shows that the arc width should grow with
time and that the subsidence beneath the load can be estimated from the observed
arc width at the surface. Application of this model to the Halmahera Arc in Indonesia
shows an excellent fit to observations if a broken-plate model of flexure is assumed.
The model also gives an excellent fit to data from East Java, also in Indonesia, where it
is possible to forward model gravity anomalies. In particular, the depth, location, and
width of the depocenter-associated gravity low are accurately reproduced, although
the model does require a high density for the volcanic arc (2900 kg m–3). This may
indicate additional buried loads due, for example, to magmatic underplating. Our
main conclusion is that loads generated by the volcanic arc are sufficient to account
for much, if not all, of the subsidence in basins within ~100 km of active volcanoes at
subduction plate boundaries, if the plate is broken. The basins will be asymmetrical
and, close to the arc, will contain coarse volcaniclastic material, whereas deposits far-
ther away are likely to be volcaniclastic turbidites. The density contrast between arc
and underlying crust required to produce the Indonesian arc basins means that they
are unlikely to form in young intraoceanic arcs but may be common in older and more
mature arcs.

Keywords: volcanic arc, loading, flexural subsidence, Indonesian arcs.

INTRODUCTION mechanism responsible for basin formation. Dewey (1980) sug-


gested that these basins result from extension caused by rollback
This paper investigates whether the isostatic load of volcanic of the subduction hinge; i.e., the location of subduction moves
arcs could be responsible for a significant fraction of the subsid- progressively away from the overriding plate and so drags the
ence frequently observed in volcanic arc settings. Arc volcanism overriding plate with it. Another suggestion is that extensional
occurs in the overriding plate of a subduction system by melt- stresses are set up by secondary mantle currents created by sub-
ing of the mantle wedge owing to introduction of volatiles car- duction (Toksöz and Bird, 1977). It is also possible that basins
ried beneath it by the subducting plate. Basins commonly form result from loading by magmatic underplating, as has been sug-
both trenchward of the arc (the forearc) and behind the arc (the gested for loading of oceanic lithosphere by volcanic islands
backarc). However, there is no consensus concerning the precise (Watts et al., 1985). Another possibility (Bahlburg and Furlong,
Waltham, D., Hall, R., Smyth, H.R., and Ebinger, C.J., 2008, Basin formation by volcanic arc loading, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation
and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special Paper 436, p. 11–26, doi: 10.1130/2008.2436(02). For
permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All rights reserved.

11
12 Waltham et al.

1996; Smith et al., 2002) is that the arc volcanism itself produces (e.g., Busby and Ingersoll, 1995), although they are not floored
significant near-surface loads leading to the formation of flex- by oceanic crust and lack many features of backarc basins (Smyth
ural basins. This last suggestion is the one pursued in this paper, et al., 2007, this volume). Nor are they retro-arc foreland basins
although we emphasize that basins form in arcs by a number of formed in response to thrusting and dynamic subsidence driven
mechanisms, and we are not proposing that all basins form in this by subduction (e.g., DeCelles and Giles, 1996), as pointed out by
way. The model has particular relevance to basins that form very Moss and McCarthy (1997). The Sunda Shelf basins have many
close to the volcanic arc. The proximal parts of such basins will characteristics of rift basins (Cole and Crittenden, 1997; Hall and
be characterized by coarse volcanic debris of primary volcanic Morley, 2004). However, the basins we are concerned with are
and sedimentary origin, typically terrestrial to shallow marine, much closer to the arc itself. Seismic lines cross many of these
and may include mass-flow deposits of different types. basins, but it is rare for them to approach close to the volcanic arc
The idea that the lithosphere flexes in response to volcanic because the volcanic-rich sequences are normally considered to
loads is not new, although it has most commonly been applied be relatively unprospective as far as hydrocarbon source rocks are
to loading of oceanic lithosphere by island chains associated concerned. Also, potential reservoir rocks commonly have prob-
with hot-spot volcanism (Watts and Cochran, 1974; Watts et al., lems with loss of porosity and permeability owing to breakdown
1997). Bahlburg and Furlong (1996) applied a continuous-plate of unstable grains, and close to the arcs, seismic data are difficult
flexure model to subsidence in the Ordovician foreland of north- to acquire and interpret. We have studied a number of Indone-
western Argentina. Their model used geographically extensive sian volcanic arcs in the field and have worked on the sequences
volcanic loads and was able to produce >8 km of subsidence. In in basins close to the volcanic arcs. The ages and histories of
a similar study using the same modeling algorithm (attributed to the basins suggest that their development is related closely to the
Slingerland et al., 1994), Smith et al. (2002) modeled subsidence development of the volcanic arc. We have chosen two different
of the Abiquiu Embayment in the Rio Grande Rift, southwestern arcs in Indonesia for which the history of the basin and the arc is
USA. The Rio Grande study also employed a volcanic load that known, and where volcanic loading is a plausible explanation of
filled most of the resulting basin, but the calculations of Smith at least some of the basin subsidence. These are South Halmahera
et al. (2002) accounted for only 800 m of subsidence in that and East Java (Fig. 1), where the arcs are built on different types
area. Other authors (e.g., Karner and Watts, 1983; Nunn et al., of crust.
1987) applied similar models to continental subsidence resulting
from thrust-emplaced loads on an unbroken plate. These studies South Halmahera
showed that gravity profiles were consistent with flexural subsid-
ence, although, in most cases, additional subsurface loads were Halmahera has a long volcanic arc history extending from
required to explain the observed subsidence. the Late Jurassic, almost entirely intraoceanic (Hall et al., 1995).
In this paper we produce a semianalytical model relating the The present arc is in its final stages of activity, as subduction has
observed surface width of the loading volcanic arc to the result- nearly eliminated the Molucca Sea, and the Sangihe and Halma-
ing subsidence. We also look at the consequences of assuming a hera Arcs (Fig. 2), on the west and east sides of the Molucca Sea,
broken-plate rather than a continuous-plate model for the under- are actively colliding. The Molucca Sea plate dips east under
lying lithosphere. Our approach principally differs from that Halmahera and west under the Sangihe Arc in an inverted U-shape
employed by others in that it predicts the volcanic load from (McCaffrey et al., 1980). Seismicity shows ~200–300 km of lith-
first principles, thus providing constraints on loading that are osphere subducted beneath Halmahera, whereas the Benioff zone
independent of those obtained, for example, from gravity mod- associated with the west-dipping slab can be identified to a depth
eling. Such an approach is difficult for geometrically complex of at least 600 km beneath the Sangihe Arc.
thrust-emplaced loads (Nunn et al., 1987) but, as shown here, is The modern arc (Figs. 1 and 2) resumed activity in the Qua-
relatively straightforward for a volcanic arc load. ternary, after a brief decline in volcanic activity, and the axis of the
arc moved ~50 km west from its late Miocene to Pliocene posi-
EVIDENCE FOR ARC LOADING IN INDONESIA tion (Hall et al., 1988b, 1995). Both these arcs are built on older
volcanic arcs active during Cretaceous and early Cenozoic time.
In many volcanic arcs, thick sequences of sedimentary and The Quaternary and Neogene Halmahera Arcs have a chemical
volcaniclastic rocks in basins are close to, and on both sides of, character typical of intraoceanic arcs (Morris et al., 1983; Forde,
the volcanic arc. Much, if not all, of the material in the basins 1997). There is no evidence of continental crust beneath them
is derived from the arc itself, and much of it is very coarse. In except at the southernmost end of the volcanic arc (Morris et al.,
Indonesia (Fig. 1), many basins near modern volcanic arcs have 1983), on Bacan and Obi, where movement on strands of the
proved rich in hydrocarbons and consequently have been the tar- Sorong fault zone brought slivers of the continental crust beneath
get of exploration work. The North, Central, and South Sumatra the arc in the last few million years (Hall et al., 1995; Ali et al.,
Basins, and the offshore Northwest Java and East Java Basins are 2001). Mapping of the Halmahera Arc shows that the basement
examples. However, these basins are typically >100 km from the is ophiolitic, formed in an early Mesozoic intraoceanic arc (Hall
active volcanic arc and have been described as backarc basins et al., 1988a; Ballantyne, 1992), overlain by Cretaceous, Eocene,
Basin formation by volcanic arc loading 13

Hainan

Manila Trench
Indochina
PHILIPPINE
South SEA

Phil
China PLATE

ippi
Sea Philippines
Se man

ne
a n 10oN
law
a
da

Pat
Pa

Trench
An

ta
ni
Chumphon Con
Mergui Sulu
Son
Malay Sea
North West Sabah
Sumatra Malay Sandakan
peninsula Natuna
East Sabah
Natuna
Sarawak Brunei Celebes CAROLINE
Penyu Tarakan Sea PLATE

Sea ca
C Halmahera

c
Su ent Sarawak

Molu
ma ral Muara
tra Sunda
Borneo 0o
Su

Shelf Kutai
Kalimantan
nd

Sumatra
a

South Barito Sulawesi


Tr

Sumatra
Asem New
en

Asem Seram Guinea


ch

West
Java Banda Sea Aru
East Java
Islands
Java Lombok
Jav Bali Timor Arafura
aT 10oS
SUNDALAND ren
ch Sumba rou
gh
helf
Shelf
CORE im or T ul S
T Sah

Indian
Ocean
INDIAN-AUSTRALIAN PLATE
90oE 100oE 110oE 120oE 130oE

Figure 1. Major basins of the southern Sundaland margin and location of the Halmahera and Java Arcs.

and Oligocene arc volcanic rocks. Shallow Miocene marine car- pre-Miocene volcanic rocks or locally on shallow-water lower to
bonates unconformably overlie all the older rocks. middle Miocene limestones. The sequence fines up from fan-delta
The Neogene Halmahera Arc became active at ca. 11 Ma at conglomerates into sandstones, mudstones, and limestones. Shal-
its southern end on Obi (Hall et al., 1995). In SW Halmahera, low marine deposition continued into the Pliocene. The depth of
volcanic activity began a little later as magmatism propagated water at the time of deposition of the upper part of the sequence
north. Basins formed on each side of the arc. In South Halma- is uncertain but was no more than a few hundred meters.
hera, activity in the volcanic arc, and subsidence on each side of The Miocene–Pliocene sequence interpreted from the seis-
the arc, began in the late Miocene at ca. 8 Ma. To the west, tur- mic sections (Fig. 3) across Weda Bay varies in thickness. It
bidites and debris flows were deposited in the forearc, and to the rests unconformably on a karstified limestone surface and is up
east (Fig. 2), a marine basin developed in Weda Bay (Hall, 1987; to 2000 m thick in local depocenters but is typically ~1000 m
Hall et al., 1988b; Nichols and Hall, 1991). thick. Hence the basin was markedly asymmetric, with the great-
On the western side of the SW arm of Halmahera (Fig. 2) est thickness of sediment adjacent to the Halmahera Arc on the
there was subsidence and deposition of at least 1000 m of subma- western side of the basin. The local depocenters offshore, and
rine slope deposits in the forearc (Nichols and Hall, 1991). In the the fan deltas onshore, are thickest close to interpreted volca-
forearc, subsidence significantly exceeded the supply of mate- nic centers. There is little evidence of faulting associated with
rial. In contrast, on the backarc side, sediment supply broadly basin formation on seismic lines (Fig. 3), and there is no indica-
kept pace with subsidence. On land in the SW arm of Halmahera tion of rifting. The lower-middle Miocene limestones represent
there are between 2800 and 3800 m of sedimentary rocks in the an approximate sea-level datum and probably covered the whole
basin east of the arc (Nichols and Hall, 1991). They were depos- area; close to the arc they were removed by erosion before depo-
ited close to the arc in shallow water and rest unconformably on sition of the basin sequence. The Halmahera Basin was actively
Cotobato

Ph
Trench Talaud PHILIPPINE

ilip
0 km 200

pin
Islands SEA 4°N
N

e
Snellius

Tre
Ridge

nch
Sangihe
Morotai
Basin East
rc
CELEBES Morotai
ihe A

SEA Siau Central Rise


Molucca
Sang

Sea

t
hrus
North Sula

rc
wesi Trenc Ridge Morotai 2°N

ra A
h

ra T

ahe
ahe
Mayu

Halm
Halm
st

Halmahera
Thru

Tifore Ternate
Sulawesi North arm B
gihe

Tidore
San

MOLUCCA Waigeo 0°
Togian Islands Gorontalo SEA C Weda
Basin Gebe
Bay
Gag
Batanta
Bacan
Bird’s
East
Salawati Head
arm lt
g Fau
Obi Soron
Misool 2°S
Banggai Islands
Sula Islands
122°E 124°E 126°E 128°E 130°E A
Sangihe Halmahera
Thrust Thrust 0
W Accretionary complex E
Arc Forearc Forearc Arc ARC

km based on
seismic lines
SANGIHE HALMAHERA
MOLUCCA SEA PLATE based on
field
B 5
observations 0 km 20 C

Figure 2. (A) Location and major tectonic features of the Molucca Sea region. Small, black-filled triangles are modern volcanoes. Bathymetric
contours are at 200, 2000, 4000, and 6000 m. Large barbed lines are subduction zones, and small barbed lines are thrusts. (B) Cross section
across the Halmahera and Sangihe Arcs on section line B. Thrusts on each side of the Molucca Sea are directed outward toward the adjacent arcs,
although the subducting Molucca Sea plate dips east beneath Halmahera and west below the Sangihe Arc. (C) Inset is the restored cross section
of the Miocene–Pliocene Weda Bay Basin of SW Halmahera on section line C, flattened to the Pliocene unconformity, showing estimated thick-
ness of the section.
0
0 km 10

1
TWT secs

3
0

seabed

1 Pliocene unco
TWT secs

Miocene-Pliocene nformity
volcanic arc-derived basin fill

2 top lime
stone
top basem
ent
A
3

0
0 km 10

1
TWT secs

Halmahera
2

3
0

seabed Weda Bay
Pliocene u
1 nconformit
y
TWT secs

Miocene-Pliocene A
volcanic arc-derived basin fill

top limestone
2
B
B 128°E C
3
Figure 3. Uninterpreted seismic lines and interpreted sections across parts of the sedimentary sequences deposited at the western edge of Weda
Bay, Halmahera. Locations of lines A and B are shown on inset map C. There is little evidence of faulting on the sections and no indication of a
rift character to the basin. The Miocene–Pliocene arc-derived sequence is deposited on top of middle Miocene shallow marine limestones and is
terminated by an intra-Pliocene unconformity. The sequence thickens toward the Miocene–Pliocene volcanic arc. TWT—two-way traveltime.
16 Waltham et al.

subsiding from the late Miocene until the middle to late Pliocene, East Java
an interval of ~5 m.y. Based on field measurement of sections
and biostratigraphic dating, Nichols and Hall (1991) estimated East Java (Fig. 4) is situated on the continental margin of
that during deposition of this interval the western side of the Sundaland. There has been subduction to the south of Java, along
basin, adjacent to the arc, subsided at least 2.8 km, representing the Java Trench, since the early Cenozoic (Hall, 2002). The base-
an average subsidence rate of at least 47 cm/1000 yr. Subsidence ment of most of East Java has previously been interpreted as arc
rates for the eastern part of the basin, ~100 km from the arc, were and ophiolitic material accreted to the continental margin in the
between 17 and 34 cm/1000 yr. Late Cretaceous, but our work (Smyth et al., this volume) has
At ca. 3 Ma, sedimentary rocks of this basin were thrust west shown that there is old continental crust beneath the Southern
over the Neogene arc, and later there was thrusting of the forearc Mountains of East Java. Our work in East Java also suggests
from the west. The Quaternary Halmahera Arc is built uncon- that the present northward subduction of Indian-Australian litho-
formably upon all these rocks. The thrusting is a result of colli- sphere began in the middle Eocene. The oldest Cenozoic rocks
sion of the Sangihe and Halmahera Arcs. In Weda Bay there has resting on older basement are terrestrial conglomerates without
been recent subsidence, and its deepest parts are almost 2000 m volcanic material, but a short distance above these rocks volca-
below sea level. Part of this subsidence is probably due to thrust nic debris appears in middle Eocene sediments and increases in
loading to the west, but part may result from movements along abundance upsection (Smyth, 2005).
splays of the Sorong Fault that appear to control the form of the There is a record of two volcanic arcs (Fig. 4) in East Java
present-day depocenter (Nichols and Hall, 1991). (Smyth et al., this volume). An early Cenozoic arc formed in the

o o
110 E 112 E
o
114 E Approximate location of the Oligocene-Miocene
volcanoes of the Southern Mountains Arc
\ Deposits of the Southern Mountains Volcanic
o
7S
Arc
Modern volcanoes of the Sunda Arc

Limit of the strong negative gravity anomaly


defined by -150 µms-2 contour
Approximate line of section (after
o
8S Pertamina, 1996)

0 50 100

kilometers

Modern Arc
N S

East Java Sea: Kendeng Basin: Southern Mountains Arc:


Edge of the Sunda Shelf Thick sequence behind Eocene to Early Miocene
Eocene to Early Miocene arc volcanic arc

5 km

Basement

100 km

Figure 4. Location of the Eocene to early Miocene volcanic arc in the Southern Mountains of East Java, the Kendeng Basin, and the modern arc.
Inferred volcanic centers of the Southern Mountains Arc are shown with open triangles, and active volcanoes are shown with solid triangles. A
cross section of the Kendeng Basin along the line indicated is based on Pertamina (1996). The Kendeng Basin sequence is interpreted to thicken
toward the Southern Mountains volcanic arc.
Basin formation by volcanic arc loading 17

Southern Mountains and was active from the middle Eocene to Immediately behind, and to the north of, the Southern
the early Miocene (ca. 42–18 Ma). It erupted material ranging Mountains Arc is the deep Kendeng Basin (Fig. 4). The basin
in composition from andesite to rhyolite (Smyth, 2005). The is long (at least 400 km) and narrow (100–120 km) and trends
arc formed a chain of volcanic islands during the early Ceno- east-west, parallel to the Southern Mountains Arc. The basin
zoic that were initially similar in character to volcanoes of the is characterized by a strong negative Bouguer gravity anomaly
present-day Izu-Bonin-Mariana Arc and the Aleutian Islands, (Fig. 5), which exceeds −580 μms–2, and extends from west to
except that behind the arc was a broadly shallow marine shelf east. The basin formed during the middle Eocene (Untung and
and no deep oceanic backarc basin. The volcanoes were terres- Sato, 1978). The Kendeng Basin succession is not well exposed
trial, but some of their products were deposited close to the arc but contains much volcanic debris carried north from the arc. The
in a marine setting, and they formed separate small islands. The oldest rocks are not seen in situ but are sampled by mud vol-
Oligocene volcanic centers are well preserved and have a spac- canoes currently erupting through the basin sequence. They are
ing similar to volcanoes of the modern arc on Java (Smyth et terrestrial and shallow marine rocks similar to those deposited
al., this volume). They are predominantly andesitic and have close to the arc during the Eocene (de Genevraye and Samuel,
been described as the “Old Andesites” (van Bemmelen, 1949). 1972). The Kendeng Basin succession records a deepening of
However, the volcanoes erupted considerable volumes of more the basin with time, and during the Oligocene thick sequences
siliceous material in explosive Plinian eruptions, and this mate- of volcaniclastic turbidites and pelagic mudstones were depos-
rial was dispersed widely as ash. The volume of siliceous mate- ited, suggesting that subsidence exceeded the supply of material.
rial has been overlooked in descriptions of volcanic activity on Seismic lines across the northern parts of the depocenter show
Java. More details of the stratigraphy are given in Smyth et al. that the Kendeng Basin sequence thickens toward the Southern
(this volume). Mountains Arc and is ~3 km thick in the north (Fig. 4; Pertamina,

Figure 5. Bouguer gravity anomaly map of East Java and location of the modeled profile discussed in the text.
18 Waltham et al.

1996). Untung and Sato (1978) suggested that the deeper parts of Halmahera, local depocenters contain sequences that are thick-
the basin contain ~6 km of section. Gravity calculations suggest est close to inferred volcanic centers. Some crustal extension is
there may be as much as 10 km of sediment in its thickest parts. likely in both volcanic arcs and is probably required in order to
The increase in thickness of the sequence toward the arc and the allow magma to reach the surface. However, in Halmahera there
importance of volcanic debris suggest that most of the Kendeng is little evidence for extensional faulting in the basin sequences,
Basin fill was derived from the arc. Because the Kendeng Basin and no significant faulting is seen on seismic lines. In neither
is so poorly exposed, and there are no seismic lines crossing the basin is there evidence for a typical rift character, nor is there
entire basin, it is impossible to assess the role of faulting in the evidence for significant crustal extension. In neither region are
basin’s development. South- and north-dipping thrusts shown on there oceanic backarc basins. There is no evidence for thrusting
interpreted sections across the basin (Pertamina, 1996) could be before or during sedimentation. Thrusting, unrelated to arc devel-
reactivated normal faults. opment, occurred in both areas after the basins formed and filled.
To the north of the Kendeng Basin was a carbonate and clas- In Halmahera, arc-arc collision caused thrusting, first from the
tic shelf during the early Cenozoic. This was the edge of the Sun- backarc side and then from the forearc side, directed toward the
daland continent. During the Eocene to early Miocene there were volcanic arc. In East Java there was mainly northward-thrusting of
elongate emergent ridges, trending roughly NE-SW and oblique the Kendeng Basin at its northern edge some time after volcanic
to the arc, separated by depressions. These ridges contain ter- arc activity ceased. All these observations suggest that volca-
restrial clastic sediments and coals at their base, thought to be nic activity contributed in some way to basin formation, possibly
Eocene, overlain by shallow marine Eocene to Miocene clastic through loading by the volcanic arc itself or possibly by weaken-
sediments and platform carbonates. There are no reports of vol- ing of the plate, or by a combination of both.
canic material in the shelf sequences, although reported clay lay-
ers may be fine-grained volcanogenic air-fall tuff deposited far MATHEMATICAL MODEL
from the arc. At the shelf edge there is volcanic material, includ-
ing clays, zircons, and volcanic quartz (Smyth, 2005). The basin In this section we develop a simple analytical model for
edge at the southern edge of the shelf is partly exposed in a fold- describing how load-generated subsidence is controlled by an
thrust belt where there was 10–30 km of shortening, suggested accumulating volcanic arc. The aim is to produce the simplest
to have occurred during the Pliocene (de Genevraye and Samuel, possible model capable of testing whether the observed sub-
1972) or late Miocene. sidence is compatible with the likely size of load. The primary
The Southern Mountains Arc extends to the south coast of assumptions of our model are that subsidence is caused by a
Java, and there has been almost no hydrocarbon exploration off- volcanic arc–generated line load and that the subsidence directly
shore directly south of the coast, so little is known about this under the load is proportional to the size of the load, i.e.,
region. Marine geophysical studies show ~1 km of sedimentary
cover of unknown age on basement in small forearc basins (Kopp s = kV, (1)
et al., 2006).
The Southern Mountains Arc has been elevated and tilted where s is subsidence, k is a constant, and V is the load (Fig. 6).
since the early Miocene and now dips uniformly to the south at For flexure of a uniform, unbroken beam, k is given by (Turcotte
~30° (Smyth, 2005). On land the arc is ~40 km wide. Arc activity and Schubert, 2002, equations 3.127, 3.131, and 3.135)
ceased in the early Miocene, followed by a period of little or no
volcanic activity (Smyth et al., this volume). Volcanic arc activ- k = π/(2xbΔρ1g), (2)
ity resumed in the late Miocene ~50 km north of the Southern
Mountains Arc, and the modern volcanoes are built on the Ken- where xb is the flexural-bulge to arc-center distance, Δρ1 is the
deng Basin (Smyth et al., this volume). The products of the late density contrast between the basin fill and the asthenosphere,
Miocene to Holocene arc are more basic than the older arc and and g is the acceleration from gravity. For a broken plate the
are predominantly basaltic andesites to andesites (e.g., Soeria- equivalent formula is (Turcotte and Schubert, 2002, equations
Atmadja et al., 1994; van Bemmelen, 1949). 3.127, 3.141, and 3.144)

Causes of Basin Formation k = 3π/(4xbΔρ1g). (3)

In South Halmahera and East Java there is association The resulting basin is assumed to be filled by sediments up
between volcanic arc activity and basin formation. Both areas to a horizontal surface and also filled by the volcanic arc but with
were emergent or close to sea level at the time of basin initiation. a triangular (in cross section) subaerial load caused by the current
Subsidence began as the volcanic arc became active. Both basins volcanic edifice itself. The load therefore consists of two com-
have asymmetrical profiles, with the greatest thickness of mate- ponents, a load resulting from the density excess of the buried
rial close to the volcanic arc. Clastic input came from the vol- volcanic arc plus a load resulting from the subaerial volcanic edi-
canic arc, and both basins are dominated by volcanic debris. In fice. Hence
Basin formation by volcanic arc loading 19

Xb
W

Sediment
fill

Flexing Lithosphere

Asthenosphere

Figure 6. Schematic diagram showing concept of basin formation by volcanic arc loading. W—arc width; xb—flexural bulge to arc-center dis-
tance; s—subsidence.

edifice) is unlikely to be available, and so a theoretical model for


⎧ s
⎫ arc width as a function of subsidence is required.
V = ⎨Δρ2 ∫ wds' + ρβw2 ⎬ g , (4) Combining equations 1 and 4, differentiating with respect to
⎩ 0 ⎭ s, and using the boundary condition that w(0) = 0 produces,

where Δρ2 is the density excess of the volcanic arc over the 1 = Aw + 2Bww′, (6)
remaining basin fill, w(s) is the arc width as a function of subsid-
ence, ρ is the volcanic arc density, and β is given by where the prime indicates differentiation and

β = 0.5 tan(θ) = h/W, (5) A = kgΔρ2, (7)

where θ is the volcano slope, h is the final volcano height, and B = kgβρ, (8)
W the final arc width. Note that equation 4 implies that there
was no significant deformation of the volcanic arc during basin with k given by equation 2 or 3. The solution to ordinary differ-
subsidence. ential equation 6 is (Appendix)
Given a well-constrained cross-sectional geometry, equa-
tions 1 through 4 alone would be sufficient to test the concept w = (1/A) f(A2s/B), (9)
that basins can be formed by flexural loading resulting from a
volcanic arc. In practice the required information (e.g., detailed where f(x) is the function shown in Figure 7A. Hence, the width
geometry of the arc deposits beneath the present-day volcanic of the arc increases rapidly in the early stages of subsidence (i.e.,
A
f
0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5 6
x

B Arc Width (km)


0 5 10 15 20 25 30 35 40
0

2
Subsidence (km)

Continuous-plate (CP)
Co
6 Broken-plate (BP) nti
nu
ou
BP sediment density +10% s-p
Br

lat
o

e
ke
n-

BP arc density +10%


pl
at

8
e

BP basin width +10%


Arc

Ba
sin

BP slope +10%
de

wi
ns

dt
ity

10
+1
+1

0%
0%

12

Figure 7. (A) Loading function derived in Appendix. Arc width increases with subsidence along a suitably scaled version of this curve. (B) Sen-
sitivity analysis. The highest dashed line shows the result for the continuous-plate model, equation 9, and the average values from Table 1. The
solid line shows the result for the broken-plate model, equation 9, and the average values from Table 1. The other curves show the result of
increasing each parameter by 10% using the broken-plate model.
Basin formation by volcanic arc loading 21

when A2s/B is small) and asymptotically approaches a maximum Although the backarc basin has been thrust westward, the very
width of 1/A as subsidence becomes large. Figure 7B illustrates the coarse character of the Neogene fan-delta deposits suggests they
effect of arc width on basin subsidence predicted from equation 9. were deposited within a few kilometers of the arc. The seismic
The “standard” simulation uses the average values of the param- lines show that the basin deposits thin eastward, away from the
eters given in Table 1 and assumes a broken-plate model (i.e., it arc, and suggest that the basin width was between 60 and 80 km.
uses equation 3 rather than equation 2). This simulation shows The estimated thickness of the sequence close to the Neogene arc
that highly significant subsidence can be obtained from reasonable is between 2800 and 3800 m (Nichols and Hall, 1991) and up to
parameter values given sensible widths for a volcanic arc. 2 km offshore farther from the arc.
Figure 7B also shows the sensitivity to parameter uncertain- In addition to this information, the mathematical model
ties. Each of the curves is obtained by increasing a single parame- requires densities for the basin fill, mantle, and volcanic arc.
ter 10% above its standard value. The result of using a continuous Based upon Hamilton (1979), we assume a basin-fill density of
rather than a broken plate is also shown. Figure 7B demonstrates 2300 ± 100 kg/m3, a mantle density of 3400 kg/m3, and a vol-
that the most important factor is whether or not the plate can be canic arc density of 2700 kg/m3. These densities, together with
considered broken. Other factors have relatively minor influ- the observed basin width of 70 ± 10 km, subsidence of 3300
ences, although uncertainties in the arc density are, perhaps, the ± 500 m, and β = 0.07 ± 0.02 (i.e., volcano slopes of 8 ± 2°), give
most significant factor. However, it should be noted that a 10% a predicted arc width of 23 ± 5 km for a broken-plate model, in
uncertainty in arc density is unlikely, whereas a 10% uncertainty good agreement with the observed volcano diameters. The con-
in volcanic slope or in basin width is realistic. tinuous plate model predicts an arc width of 28 ± 7 km. Hence,
In the following section, observations on the geometry and the observed subsidence and arc width are compatible with basin
gravity profiles from two Indonesian arcs are used to test the formation by arc loading and suggest that a broken-plate model
model developed in this section. fits the data better than a continuous-plate model of lithospheric
flexure. The parameters used, and the results obtained, for the
APPLICATION TO INDONESIAN ARCS Halmahera example are summarized in Table 1.
In our second test case we do not, unfortunately, have good
The modern Halmahera volcanoes have elevations up to constraints upon basin subsidence, but, on the other hand, we
1730 m and diameters at sea level of their base of between 10 do have a gravity profile that can be directly compared with a
and 25 km, typically ~20 km for those on land. Most of the active simulated profile based upon our mathematical model of arc-
volcanoes with smaller diameters are offshore volcanic islands loaded flexure.
(Makian, Tidore, and Ternate) rising from sea-floor depths In East Java the volcanoes are much larger than those on
of several hundred meters, and therefore their diameter at the Halmahera. Several of the present-day volcanoes have elevations
base is not known. A diameter of 20 km for all the volcanoes >3000 m, and all except two of East Java’s modern volcanoes are
at their base is reasonable. The volcanoes of the Quaternary arc between 1600 and 3300 m high. The volcanoes typically have
have been active for no more than 2 m.y. and probably less. The diameters of 50–55 km at their base. The modern arc began activ-
dimensions of the Neogene volcanic centers are not known pre- ity ~10–12 m.y. ago. As in Halmahera, the size of the volcanic
cisely, partly because of the nature of the exposure in rainforest centers of the Southern Mountains Arc is more difficult to assess.
terrain, and partly because they are overthrust by backarc and They have a similar distribution and spacing to the modern vol-
forearc basin rocks. The distribution of their products identified canoes. Several of the centers can be mapped up to 40 km north
during mapping of the islands at 1:250,000 scale suggests that of the coast, and the present steep coastline is not the southern
they were larger than the Quaternary volcanoes, consistent with limit of the volcanoes. Thus, an arc width of 50 km, similar to the
their longer period of activity from 8 to 3 Ma. A width of the arc diameter of the modern volcanoes, is reasonable and consistent
of 20 km is therefore a reasonable value. Like the Neogene arc, with an interval of arc activity of ~23 m.y. The basin width can be
the exact width of the Halmahera backarc basin is uncertain. Its determined with greater certainty; it is between 100 and 120 km
eastern edge is in Weda Bay at water depths of 2 km, and detailed from the arc margin to the shelf edge, and there is estimated to
seismic lines cover only the western side of the offshore basin. have been 10–30 km of shortening on thrusts at the shelf edge.

TABLE 1. PARAMETERS USED IN THE FLEXURAL MODEL FOR HALMAHERA


Parameter Minimum Maximum Comments
3
Basin fill density (kg/m ) 2400 2200 Minimum density gives maximum arc width
3
Volcanic arc density (kg/m ) 2700 2700 Andesite
Mantle density (kg/m3) 3400 3400 Standard
Basin width (km) 60 80 Gives estimated elastic thickness of 4.5–7.1 km
Subsidence (m) 3250 3350 See text
Slope (°) 10 6 Minimum slope gives maximum arc width

Arc width (km) 10 25 O bs e r v e d


Arc width (km) 17 28 Predicted broken-plate model
Arc width (km) 21 34 Predicted continuous-plate model
22 Waltham et al.

Independent estimates of basin thickness are unavailable for this contrast between the mantle lithosphere and the asthenosphere
area, but gravity data (see Fig. 3) can be used to test the model. was ignored, and a single density was assumed for the entire
The Bouguer anomaly data along an East Java profile is mantle. A forward model of gravity over an arc-loaded, flexing,
shown in Figure 8. Figure 9 shows the same data after a regional broken elastic lithosphere is also shown in Figure 9. Flexure and
trend was removed owing to the gravity signature of a subduct- gravity model parameters are given in Table 2. Note that β =
ing lithospheric plate deep beneath Java. The gravity modeling 0.05 implies a modern volcano height of 2500 m and a slope
was performed by an add-on to our main arc-load modeling pro- of 6°. Using the parameters listed in Table 2, the mathematical
gram, which simply summed contributions from a large number model predicts a subsidence of 6.9 km beneath the arc.
of small, rectangular, constant-density slabs using well-known The key features of Figure 9 are the gravity low owing to
expressions (e.g., see Telford et al., 1990). The small density the basin depocenter, but also the gravity high at the south end

A 1400 S N
1200

1000

800
Bouguer Anomaly (gu)

600

400

200

-200

-400

-600

-800
B S N
900

800

700

600
Topography (m)

500

400

300

200

100

0
50 100 150 200 250 300
Distance (km)
Figure 8. Bouguer anomaly (A) and topographic profile (B) for East Java for the line shown in Figure 5.
Basin formation by volcanic arc loading 23

600
S N
400

200
Bouguer Anomaly (gu)

-200

-400

-600

-800

-1000

-1200
50 100 150 200 250 300 350

Distance (km)
Figure 9. Bouguer anomaly from Figure 8 after removal of regional trend. The solid line is the predicted gravity anomaly based on a filled basin
produced by subsidence caused by volcanic arc loading of a broken elastic plate.

TABLE 2. PARAMETERS USED IN THE GRAVITY AND is a simple relationship between arc width at the surface and
FLEXURAL MODELING OF THE EAST JAVA BASIN the amount of subsidence. This model is strongly supported by
Parameters Units analysis of the Halmahera Arc, whose width agrees closely with
3
Basin fill density (kg/m ) 2200 that predicted by our model from the observed subsidence. Our
3
Volcanic arc density (kg/m ) 2900
Crustal density (kg/m )3
2700 model is further supported by the strong similarity between the
3
Crustal thickness (kg/m ) 40 observed Bouguer anomaly on East Java to that predicted by a
Mantle density (kg/m3) 3400 model of basin subsidence above a broken plate. The gravity pro-
Basin width (km) 150
Arc width (km) 50 files agree closely in both the depth and width of the anomaly
Slope (°) 6 produced by the basin.
Subsidence pred icted (m) 6900 For Halmahera, the subsidence observed can be accounted
for entirely by arc loading if a broken-plate model is assumed.
of the profile between 50 and 150 km, which is due to the posi- This is consistent with weakening of the arc crust by magmatic
tive effect of the volcanic arc itself. The remaining discrepancies heating. For East Java, the modeling suggests that an additional
between modeled and observed gravity profiles occur at distinct load is required to produce the observed subsidence. One obvi-
topographic highs and would be removed by suitable topographic ous contributor could be crustal extension, and we cannot rule
corrections. The basin-fill density is at the low end of figures given out extensional faulting, as the Kendeng Basin is poorly exposed
by Hamilton (1979). The crustal thickness of 40 km is possibly and there are no seismic lines crossing the basin. However, the
too high, although Hamilton (1979) shows a speculative thick- large volumes of highly acidic volcanic products that erupted in
ness beneath the Java magmatic arc of >30 km, and the value is East Java in the Southern Mountains Arc suggest significant dif-
well within the plausible range for a mature arc built on conti- ferentiation of arc magmas before eruption, which should have
nental crust (e.g., Kay and Mpodozis, 2001). However, the arc produced large volumes of basic cumulates (S. Sparks, 2005, per-
density has been set rather high, and this may reflect the need for sonal commun.). The relatively high density used in the model
further loads to produce the observed gravity high near the arc. would be consistent with large volumes of dense cumulates at
deep crustal levels. Magmatic underplating by such cumulates
DISCUSSION would have provided an additional load as in volcanic islands
(Watts et al., 1985).
The key prediction of the mathematical model of basin for- Probably the biggest uncertainty in the modeling and inter-
mation by volcanic arc loading, as presented here, is that there pretation is the density contrast between crust and arc. Little is
24 Waltham et al.

known about the crustal structure and thickness of any Indone- volcanoes and the underlying crust, and therefore the volcanoes
sian arc. The model assumes a normal continental crust for the do not act as a load. In continental margin arcs such as Java, and
arc basement with a load similar in composition to andesites. long-lived intraoceanic arcs such as Halmahera, there is a larger
This is reasonable in East Java, where the basement is continental density contrast between the deeper crust and the eruptive prod-
crust and preserved volcanic centers are predominantly andes- ucts of the arc. In these cases the volcanoes act as a significant
ites. It is also likely, based on observations of the character of load and produce flexural basins close to the arc. The absence of
small areas of exposed basement and oil-company drilling of the deep basins in arcs in some parts of Indonesia and elsewhere in
basement, that accreted arc and ophiolitic material includes ser- the world may be due to the small density contrast between vol-
pentinites, which would lower the average crustal density. For canoes and the underlying crust in those regions.
Halmahera, Milsom et al. (1996) concluded that the crust was at
least 20 km thick and had a bulk density approaching that of con- APPENDIX
tinental rather than oceanic crust. Local gravity highs are associ-
ated with Paleogene arc volcanic rocks rather than the ophiolites, Equation 6 is
and nowhere do the ophiolites have the high gravity fields asso-
ciated with classic ophiolites. Although much of Halmahera is 1 = Aw + 2Bww′. (A1)
underlain by ophiolitic rocks, significant serpentinization of the
ultramafic parts could account for the low crustal density. The Changing the variable to
products of the modern arc (Morris et al., 1983) and the Neogene
arc (Forde, 1997; Macpherson et al., 2003) are almost entirely x = A2s/B (A2)
basaltic andesites and andesites.
and scaling using
CONCLUSIONS
f = Aw (A3)
The role of volcanoes in the development of basins in arc
regions has generally been ignored, overlooked, or not consid- then gives
ered important. This is possibly because most studies have been
concerned with larger true backarc basins and forearc basins both 1 = f + 2 f f ′, (A4)
of which are formed much farther from the arc. However, in sev-
eral Indonesian arcs there are thick sequences of sedimentary where f ′ now indicates differentiation with respect to x. Equation
and volcaniclastic rocks in deep basins very close to the arc. The A4 is a nonlinear ordinary differential equation (ode), which does
mathematical modeling shows that volcanic loading can make not have an analytical solution in terms of elementary functions.
a contribution to basin subsidence in these settings, and may However, at small values of f the second term on the right side
be the primary cause. Volcanic loading can account for basins dominates, so that 1 ~2 f f ′ with a solution of f = x0.5. At large val-
close to the arc but is not relevant to basins >~100 km from the ues of f the first term dominates, and the solution asymptotically
arc. The basins produced by arc loading are asymmetrical, and approaches f = 1. The function f can be estimated using finite
basin sequences are thickest close to the arc. These are likely to differencing with
be dominated by very coarse, terrestrial and shallow marine, pri-
mary volcanic and volcaniclastic rocks that may include a variety f ~(fi + fi+1)/2 (A5)
of mass-flow deposits. Locally and intermittently the supply of
material from the arc may exceed subsidence, and there may be and
rapid vertical and lateral variations in grain size and changes from
shallow marine to terrestrial deposits. Farther from the arc, sub- f ′ ~(fi+1 – fi)/Δx, (A6)
sidence typically exceeds supply, and the more proximal deposits
pass laterally into deeper water sedimentary deposits such as tur- which, after substitution into (A4), yields
bidites. The distribution and character of the material, particularly
close to the arc, has some similarities to rocks described by Draut 2
− Δx ⎛ Δx ⎞ Δx
and Clift (2006) from the Jurassic Talkeetna Formation of Alaska. f i+1 = + ⎜ ⎟ + fi 2 − f i + Δx , (A7)
However, in the Alaskan Jurassic arc and in the modern Mari- 4 ⎝ 4 ⎠ 2
ana and Tonga Arcs, accommodation space is not likely to have
been created by volcanic loading but had already existed. This is which, together with the boundary condition that f0 = 0, allows
because in young intraoceanic settings the volcanoes rise from the function in Figure 2 to be calculated. The resultant function
great depths above the ocean floor. In young intraoceanic arcs has the required properties that f ~x0.5 for small x and f ~1 for
there is also likely to be little density difference between the arc large x.
Basin formation by volcanic arc loading 25

ACKNOWLEDGMENTS Hall, R., Ali, J.R., Anderson, C.D., and Baker, S.J., 1995, Origin and motion
history of the Philippine Sea Plate: Tectonophysics, v. 251, p. 229–250,
doi:10.1016/0040-1951(95)00038-0.
Our work on Indonesian arcs has been supported at different Hamilton, W., 1979, Tectonics of the Indonesian region: U.S. Geological Sur-
times by the University of London Central Research Fund, vey Professional Paper 1078, 345 p.
NERC, and the Royal Society, but mainly by a consortium of Karner, G.D., and Watts, A.B., 1983, Gravity anomalies and flexure of the
lithosphere at mountain ranges: Journal of Geophysical Research, v. 88,
oil companies whose membership has changed with time. We p. 10,449–10,477.
thank colleagues in Indonesia at the Geological Research and Kay, S.M., and Mpodozis, C., 2001, Central Andean ore deposits linked
Development Center Bandung (now the Geological Survey of to evolving shallow subduction systems and thickening crust: GSA
Today, v. 11, no. 3, p. 4–9, doi: 10.1130/1052-5173(2001)011<0004:
Indonesia), LIPI, and Institut Teknologi Bandung for help with CAODLT>2.0.CO;2.
field work. We thank Chris Elders and Andi Salahuddin for dis- Kopp, H., Flueh, E.R., Petersen, C.J., Weinrebe, W., Wittwer, A., and Meramex
cussion of Weda Bay seismic interpretation. Scientists, 2006, The Java margin revisited: Evidence for subduction ero-
sion off Java: Earth and Planetary Science Letters, v. 242, p. 130–142,
doi: 10.1016/j.epsl.2005.11.036.
Macpherson, C.G., Forde, E.J., Hall, R., and Thirlwall, M.F., 2003, Geochemi-
REFERENCES CITED cal evolution of magmatism in an arc-arc collision: The Halmahera and
Sangihe arcs, eastern Indonesia, in Larter, R.D., and Leat, P.T., eds., Intra-
oceanic Subduction Systems: Tectonic and Magmatic Processes: Geolog-
Ali, J.R., Hall, R., and Baker, S.J., 2001, Palaeomagnetic data from a Meso- ical Society [London] Special Publication 219, p. 207–220.
zoic Philippine Sea Plate ophiolite on Obi Island, Eastern Indonesia: McCaffrey, R., Silver, E.A., and Raitt, R.W., 1980, Crustal structure of the
Journal of Asian Earth Sciences, v. 19, p. 535–546, doi: 10.1016/S1367- Molucca Sea collision zone, Indonesia, in Hayes, D.E., ed., The Tectonic
9120(00)00053-5. and Geologic Evolution of Southeast Asian Seas and Islands, Part 1:
Bahlburg, H., and Furlong, K.P., 1996, Lithospheric modeling of the Ordovi- American Geophysical Union Geophysical Monograph 23, p. 161–178.
cian foreland basin in the Puna of northwestern Argentina: Tectonophys- Milsom, J., Hall, R., and Padmawidjaja, T., 1996, Gravity fields in eastern
ics, v. 259, p. 245–258, doi: 10.1016/0040-1951(95)00129-8. Halmahera and the Bonin Arc: Implications for ophiolite origin and
Ballantyne, P.D., 1992, Petrology and geochemistry of the plutonic rocks of emplacement: Tectonics, v. 15, p. 84–93, doi: 10.1029/95TC02353.
the Halmahera ophiolite, eastern Indonesia, an analogue of modern oce- Morris, J.D., Jezek, P.A., Hart, S.R., and Gill, J.B., 1983, The Halmahera
anic forearcs, in Parson, L.M., et al., eds., Ophiolites and Their Modern island arc, Molucca Sea collision zone, Indonesia: A geochemical survey,
Oceanic Analogues: Geological Society [London] Special Publication 60, in Hayes, D.E., ed., The Tectonic and Geologic Evolution of Southeast
p. 179–202. Asian Seas and Islands, Part 2: American Geophysical Union Geophysi-
Busby, C.J., and Ingersoll, R.V., 1995, Tectonics of Sedimentary Basins: Cam- cal Monograph 27, p. 373–387.
bridge, Massachusetts, Blackwell, 579 p. Moss, S.J., and McCarthy, A.J., 1997, Discussion—Foreland basin systems:
Cole, J.M., and Crittenden, S., 1997, Early Tertiary basin formation and the Basin Research, v. 9, p. 171–176.
development of lacustrine and quasi-lacustrine/marine source rocks on Nichols, G.J., and Hall, R., 1991, Basin formation and Neogene sedimentation
the Sunda Shelf of SE Asia, in Fraser, A.J., et al., eds., Petroleum Geol- in a backarc setting, Halmahera, eastern Indonesia: Marine and Petroleum
ogy of SE Asia: Geological Society [London] Special Publication 126, Geology, v. 8, p. 50–61, doi: 10.1016/0264-8172(91)90044-2.
p. 147–183. Nunn, J.A., Czerniak, M., and Pilger, R.H., Jr., 1987, Constraints on the struc-
DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems: Basin Research, ture of Brooks Range and Colville Basin, Northern Alaska, from flexure
v. 8, p. 105–123, doi: 10.1046/j.1365-2117.1996.01491.x. and gravity analysis: Tectonics, v. 6, p. 603–617.
de Genevraye, P., and Samuel, L., 1972, The geology of Kendeng Zone (East Pertamina, B.P.P.K.A., 1996, Petroleum geology of Indonesian basins, prin-
Java): Proceedings of Indonesian Petroleum Association 1st Annual Con- ciples, methods and application. Volume IV: East Java Basins, 107 p.
vention, Jakarta, p. 17–30. Slingerland, R., Harbaugh, J.W., and Furlong, K., 1994, Simulating Clastic
Dewey, J.F., 1980, Episodicity, sequence and style at convergent plate margins, Sedimentary Basins: New York, Prentice Hall, 220 p.
in Strangeway, D.W., ed., The Continental Crust and Its Mineral Deposits: Smith, G.A., Moore, J.D., and McIntosh, W.C., 2002, Assessing roles of vol-
Geological Association of Canada Special Paper 20, p. 553–573. canism and basin subsidence in causing Oligocene–Lower Miocene sedi-
Draut, A.E., and Clift, P.D., 2006, Sedimentary processes in modern and ancient mentation in the northern Rio Grande Rift, New Mexico, USA: Journal of
oceanic arc settings: Evidence from the Jurassic Talkeetna Formation of Sedimentary Research, v. 72, p. 836–848.
Alaska and the Mariana and Tonga Arcs, Western Pacific: Journal of Sedi- Smyth, H.R., 2005, Eocene to Miocene basin history and volcanic history in
mentary Research, v. 76, p. 493–514, doi: 10.2110/jsr.2006.044. East Java, Indonesia [Ph.D. thesis]: University of London, 476 p.
Forde, E.J., 1997, The geochemistry of the Neogene Halmahera Arc, eastern Smyth, H.R., Hall, R., and Nichols, G.J., 2008, Early Cenozoic volcanic arc
Indonesia [Ph.D. thesis]: University of London, 268 p. history of East Java, Indonesia: The stratigraphic record of eruptions on a
Hall, R., 1987, Plate boundary evolution in the Halmahera region, Indonesia: continental margin in a tropical setting, in Draut, A.E., et al., eds., Lessons
Tectonophysics, v. 144, p. 337–352, doi: 10.1016/0040-1951(87)90301-5. from the Stratigraphic Record in Arc Collision Zones: Geological Society
Hall, R., 2002, Cenozoic geological and plate tectonic evolution of SE Asia and of America Special Publication 436 (this volume).
the SW Pacific: Computer-based reconstructions, model and animations: Soeria-Atmadja, R., Maury, R.C., Bellon, H., Pringgoprawiro, H., Polvé, M., and
Journal of Asian Earth Sciences, v. 20, p. 353–434, doi: 10.1016/S1367- Priadi, B., 1994, Tertiary magmatic belts in Java: Journal of Southeast Asian
9120(01)00069-4. Earth Sciences, v. 9, p. 13–17, doi: 10.1016/0743-9547(94)90062-0.
Hall, R., and Morley, C.K., 2004, Sundaland Basins, in Clift, P., et al., eds., Con- Telford, W.M., Geldart, L.P., and Sheriff, R.E., 1990, Applied Geophysics:
tinent-Ocean Interactions within the East Asian Marginal Seas: American Cambridge, UK, Cambridge University Press, 770 p.
Geophysical Union Geophysical Monograph 149, p. 55–85. Toksöz, M.N., and Bird, P., 1977, Formation and evolution of marginal basins
Hall, R., Audley-Charles, M.G., Banner, F.T., Hidayat, S., and Tobing, S.L., and continental plateaus, in Talwani, M., and Pitman, W.C., eds., Island
1988a, Basement rocks of the Halmahera region, Eastern Indonesia: A Arcs, Deep Sea Trenches and Back Arc Basins: American Geophysical
Late Cretaceous–Early Tertiary arc and fore-arc: Geological Society Union Maurice Ewing Series 1, p. 379–393.
[London] Journal, v. 145, p. 65–84. Turcotte, D.L., and Schubert, G., 2002, Geodynamics (2nd edition): Cambridge,
Hall, R., Audley-Charles, M.G., Banner, F.T., Hidayat, S., and Tobing, S.L., UK, Cambridge University Press, 456 p.
1988b, Late Paleogene–Quaternary geology of Halmahera, Eastern Indo- Untung, M., and Sato, Y., 1978, Gravity and geological studies in Java, Indo-
nesia: Initiation of a volcanic island arc: Geological Society [London] nesia: Geological Survey of Indonesia and Geological Survey of Japan,
Journal, v. 145, p. 577–590. Special Publication 6, 207 p.
26 Waltham et al.

van Bemmelen, R.W., 1949, The Geology of Indonesia: Nijhoff, The Hague, Watts, A.B., Peirce, C., Collier, J., Dalwood, R., Canales, J.P., and Henstock,
Government Printing Office, 732 p. T.J., 1997, A seismic study of lithospheric flexure in the vicinity of Tener-
Watts, A.B., and Cochran, J.R., 1974, Gravity anomalies and flexure of the ife, Canary Islands: Earth and Planetary Science Letters, v. 146, p. 431–
lithosphere along the Hawaiian-Emperor seamount chain: Royal Astro- 447, doi: 10.1016/S0012-821X(96)00249-X.
nomical Society Geophysical Journal, v. 38, p. 119–141.
Watts, A.B., ten Brink, U.S., Buhl, P., and Brocher, T., 1985, A multichannel
seismic study of lithospheric flexure across the Hawaiian-Emperor sea-
mount chain: Nature, v. 315, p. 105–111, doi:10.1038/315105a0. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Cenozoic arc processes in Indonesia: Identification of the key influences on the


stratigraphic record in active volcanic arcs

Robert Hall
Helen R. Smyth
SE Asia Research Group, Department of Geology, Royal Holloway University of London, Egham, Surrey, TW20 0EX, UK

ABSTRACT

The Indonesian region includes several volcanic island arcs that are highly active
at the present day, and also contains a record of Cenozoic volcanic activity owing to
subduction of oceanic lithosphere at the margins of SE Asia. As a result of long-term
subduction, there is a high regional heat flow, and a weak crust and lithosphere, as
identified in other subduction zone backarcs. The stratigraphic record in the Indo-
nesian region reflects a complex tectonic history, including collisions, changing plate
boundaries, subduction polarity reversals, elimination of volcanic arcs, and exten-
sion. The arcs have not behaved as often portrayed in many arc models. They mark
subduction but were not continuously active, and it is possible to have subduction
without magmatism. Subduction hinge retreat was accompanied by significant arc
volcanism, whereas periods of hinge advance were marked by reduction or cessation
of volcanic activity. Growth of the region occurred in an episodic way, by the addition
of ophiolites and continental slivers, and as a result of arc magmatism. In Indonesia,
relatively small amounts of material were accreted from the downgoing plate during
subduction, but there is also little evidence for subduction erosion. During collision
the arc region may fail, resulting in thrusting, and the weakest point is the position of
the active volcanic arc itself. Volcanic arcs shift position suddenly, and arcs can disap-
pear during collision by overthrusting. Arcs are geologically ephemeral features and
may have very short histories in comparison with most well-known older orogenic
belts. The stratigraphic record of the basins within arc regions is complex. Because of
a weak lithosphere the character of sedimentary basins may be unusual, and basins
are commonly very deep and subside rapidly. There is a high sediment flux. The vol-
canic arc itself influences the stratigraphic record and basin development. The load
imposed by the volcanic arc causes flexure and provides accommodation space. The
volcanic arc thus can form the basin and supply most of its sediment. Tropical pro-
cesses influence the mineralogy and apparent maturity of the sediment, especially
volcanogenic material. A complex stratigraphy will result from the waxing and wan-
ing of volcanic activity.

Keywords: Cenozoic, Sunda Arc, Banda Arc, Sangihe Arc, Halmahera Arc, Indone-
sia, stratigraphic record.

Hall, R., and Smyth, H.R., 2008, Cenozoic arc processes in Indonesia: Identification of the key influences on the stratigraphic record in active volcanic arcs, in
Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America
Special Paper 436, p. 27–54, doi: 10.1130/2008.2436(03). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All
rights reserved.

27
28 Hall and Smyth

INTRODUCTION background on the region, and a summary of the history of several


of the most important Cenozoic arcs in Indonesia. We then dis-
The size of Indonesia, the abundance of volcanic activity, cuss particular features of the record of arc activity and attempt to
and its long volcanic history make it an important area for under- identify those that may be of general interest in understanding arc
standing arc processes and influences of volcanic arcs on the processes and those that may be useful in interpreting the record
stratigraphic record. Hamilton (1988) commented that the arcs of activity in older arcs elsewhere in the world. Our discussion
of Indonesia–southern Philippines–western Melanesia are in “the of the stratigraphic record in Indonesian arcs is based mainly on
region of greatest modern variety and complexity,” but the region studies carried out by the SE Asia Research Group in several dif-
remains understudied and often overlooked. The long history of ferent arcs over many years, especially in Java and Halmahera,
subduction beneath Indonesia has influenced arc development. but also in arcs of Sumatra, Sulawesi, and Borneo. These studies
Subduction has produced a thin and warm lithosphere beneath have included extensive field work and related work documented
the upper plate, which is unlike that of older stable continents and in theses, but some of the results are unpublished and in places
their margins. These features have influenced the development of we may appear to make statements unsupported by easily acces-
the Indonesian arcs, which differ in many ways from textbook sible literature; we hope this will be forgiven, and we felt it was
examples of arcs described from other parts of the world. They justified in our attempt to give an account of the region in a paper
were formed at the margins of a large, weak subduction backarc of reasonable length.
region and consequently seem to have been unusually responsive
to plate boundary forces. Because of their youth it is not pos- Indonesia and Large Explosive Volcanic Events
sible to study the deep levels of the arcs, but the arc history that
is revealed by their stratigraphy offers insights into the develop- Indonesia is a large country that includes >18,000 islands
ment of older orogens and older volcanic arcs. and stretches >5000 km from west to east (Fig. 1). Wallace (1869)
In this paper we identify some important features of Indo- tried to draw attention to its size in his book The Malay Archi-
nesian volcanic arcs, particularly those that influence the strati- pelago by comparing the island of Borneo to the British Isles,
graphic record of arc activity. Because the geology of Indonesia is and subsequently Umbgrove (1938) and van Bemmelen (1949)
not familiar to many readers, we begin our discussion with some compared the region to areas more familiar to North American

Andaman
10°N Sea South China Sea
Philippines

5°N Malay Sunda


Peninsula Shelf Sangihe
Celebes
Sea
Borneo Bird’s
Nias Halmahera Head

ts

Molucca
ai

Sea
Str
sar

Sulawesi
kas

Indian
Ocean Sumatra
Ma

5°S Java Sea


Banda
Flores Sea
Sea
Java

Timor
0 500 1,000
10°S Northwest Shelf
kilometers Australia
100°E 110°E 120°E 130°E

Figure 1. Indonesia shown on a digital elevation model (DEM), based on Shuttle Radar Topography Mission (SRTM) data merged with the
Sandwell and Smith (1997) bathymetry.
Cenozoic arc processes in Indonesia 29

and European readers. Figure 2 is a modern attempt to show the island of Sumatra, was the largest in the last 2 m.y., and its effect
size of the region by comparing Indonesia to the conterminous on climate must have been even more devastating. It has been
United States. According to the Smithsonian Global Volcanism speculated that the eruption led to accelerated glaciation and a
Program (Smithsonian, 2006), Indonesia includes 150 Holocene bottleneck in human evolution (Ambrose, 2003; Rampino and
volcanoes; 95 of these could be considered active because they Self, 1992, 1993a, b).
have erupted since A.D. 1500. Indonesia has a long record of For the last 36 m.y., 42 large explosive events are recorded
volcanic activity, recording subduction related to northward in a recent global compilation (Mason et al., 2004); 33 of
movement of the Indian-Australian plate and westward motion these large explosive events are recorded from North America,
of the Pacific plates. Cenozoic subduction of Indian-Australian whereas Toba alone is situated in a humid tropical climate. The
oceanic lithosphere beneath most of Indonesia began after Aus- record has been interpreted to indicate that there were two pulses
tralia separated from Antarctica and moved rapidly northward in Cenozoic global volcanism: one between 36 and 25 Ma, and
from ~45 m.y. ago (Hall, 2002). the other since 13.5 Ma. However, Indonesia remains a rela-
With a long history of volcanic activity, and a well-known tively remote area, which is difficult to explore because of the
explosive character in historical times, it is surprising that the absence of infrastructure, difficult terrain, and rain forest cover.
only exceptionally large eruptions known from Indonesia in the We suggest that the absence of exceptionally large eruptions
Cenozoic are those of Toba. Of the Holocene volcanoes, 32 have in the stratigraphic record from Indonesia has more to do with
records of very large eruptions with a volcanic explosive index preservation and sampling than with volcanic activity. As we
(VEI) >4; 19 of these have erupted in the past 200 yr, includ- report below, there is indeed evidence of large explosive events
ing Tambora in 1815 (VEI = 7) and Krakatau in 1883 (VEI = in Indonesia during the Cenozoic, and no doubt more of these
6). There are only 4 volcanoes in the Smithsonian global data remain to be discovered. The lack of information on large erup-
set of large-volume Holocene explosive eruptions with a VEI of tions in Indonesia is not dissimilar to the general lack of infor-
7; the other 3 erupted in prehistorical times and pre-date 4000 mation on Indonesia’s volcanic history. This partly reflects the
B.C. Tambora, on the island of Sumbawa, is known for its impact absence of modern detailed studies, but ideas about arcs may
on global climate, and its 1815 eruption resulted in the North- be unduly influenced by results from more accessible or bet-
ern Hemisphere’s “year without a summer” in 1816, when crops ter exposed regions, from ocean drilling, and from geochemi-
failed, causing famine and population movements (Harington, cal studies. However, as with large eruptions the stratigraphic
1992; Zeilinga de Boer and Sanders, 2002). The 74 ka eruption record in Indonesia has some valuable insights to reveal; this
of Toba (Chesner and Rose, 1991; Chesner et al., 1991), on the paper identifies some of them.

5000 km

Figure 2. Indonesia and Malaysia compared to the USA. The box is 60° from east to west and 25° from north to south.
30 Hall and Smyth

ARC ACTIVITY IN INDONESIA active since the Eocene (e.g., Garwin et al., 2005; Hall, 2002;
Hamilton, 1977) and is the product of subduction of the Indian-
The distribution of modern-day volcanoes in Indonesia Australian plate beneath the Sundaland margin. Sundaland
shows clearly that most volcanic activity is related to subduction (Fig. 4) is the continental core of SE Asia, comprising Indochina,
(Fig. 3). This is also emphasized by the distribution of seismic- the Thai-Malay Peninsula, Sumatra, Java, Borneo, and the shal-
ity (Cardwell and Isacks, 1978; Cardwell et al., 1980; Engdahl low marine shelf (Sunda Shelf) between them, and formed by
et al., 1998; England et al., 2004). Global positioning system amalgamation of continental blocks during the Triassic. There
(GPS) observations (Bock et al., 2003; Kreemer et al., 2000; is a Proterozoic basement (Liew and McCulloch, 1985; Liew
Rangin et al., 1999), seismicity, and volcanic activity indicate and Page, 1985) intruded by Permian–Triassic granites formed
rapid plate movements and considerable tectonic complexity. by subduction and postcollisional thickening of the continental
Present-day Indonesian volcanic activity can be regarded as the crust (Hutchison, 1989, 1996). In Borneo the oldest rocks known
products of four separate volcanic arcs (Fig. 3): the Sunda Arc, are Paleozoic metamorphic rocks intruded by Mesozoic granites.
stretching from Sumatra through Java to the east, the horseshoe- From Sumatra to Borneo the continental core is surrounded by
shaped Banda Arc of eastern Indonesia, the Halmahera Arc Mesozoic ophiolitic and arc igneous rocks, and possible frag-
of the North Moluccas, and the Sangihe Arc, extending from ments of continental material, accreted during the Cretaceous.
Sulawesi into the southern Philippines. Here we give first a brief In Sumatra and West Java the Sunda Arc is built in part on the
summary of the Cenozoic history of each of these arcs and the continental margin of Sundaland. In East Java it is constructed
character of the crust on which they are constructed to provide on arc and ophiolitic rocks accreted at the Mesozoic active mar-
the background to our commentary on Java, east Indonesia, and gin and in part on a continental fragment that was added by a
the Moluccas. More detailed discussion of each of these arcs, Cretaceous collision.
literature references, and animated reconstructions of the tec- In Sumatra a record of subduction-related magmatism
tonic development are in Hall (2002) and Garwin et al. (2005). extends back into the early Mesozoic or late Paleozoic (McCourt
et al., 1996), but in Java the oldest record of subduction is Cre-
Java: The Sunda Arc taceous. The Sundaland Cretaceous active margin is interpreted
to have run the length of Sumatra into West Java and then turned
Indonesia is situated on the southern margin of the Eurasian northeast into SE Borneo (Hamilton, 1979). In Central Java and
plate at the edge of Sundaland (van Bemmelen, 1949; Hamilton, SE Borneo there is evidence for Cretaceous high pressure–low
1979; Hutchison, 1989; Katili, 1975). The Sunda Arc has been temperature subduction-related metamorphism (Parkinson et

5°N

Sangihe Talaud
Arc
Morotai
Mayu Halmahera
Borneo Arc

Sunda
Kalimantan
Su

Shelf
m
at

Sulawesi
ra

Obi
ait
Su

r Str
nd

5°S Seram
a

Buru
assa
Ar

Java Sea
c

Mak

Banda Arc
Java Wetar
Bali Flores
Sunda
10°S
Arc
Lombok
Timor

95°E 100°E 105°E 110°E 115°E 120°E 125°E 130°E

Figure 3. Present-day volcanoes in Indonesia (Smithsonian, 2006) and the principal volcanic arcs discussed in the text: Sunda Arc, Banda Arc,
Halmahera Arc, and Sangihe Arc. Indonesia is shaded in gray, and the area of Malaysia is unfilled.
Cenozoic arc processes in Indonesia 31

Manila Trench
EURASIAN PHILIPPINE
PLATE SEA PACIFIC
PLATE OCEAN

Philip
South
China Philippines

pine
Andaman Sea
10°N
PACIFIC

Tren
Sea

c
PLATE

h
MESOZOIC Borneo Sulu
SUNDALAND Sea
CORE Suture


Molucca


Sunda
Shelf
Celebes
Sea
Suture CAROLINE
PLATE
Molucca
Sea
Sorong
SUNDALAND Suture 0°

t
trai
Sun

S
sar
da

Se
ram
kas
Bismarck
Tre

Sea
Ma
nc

Tro
Sulawesi
h

Banda

u
Java Sea Banda

gh
Suture Sea
Suture New Guinea
ADDED BY
EARLY CENOZOIC
Arafura 10°S
g h
Java Trou Shelf
INDIAN Trench
Tim
or
OCEAN Sahul
ADDED BY Shelf
EARLY MIOCENE
Coral

0 km 1000
INDIAN-AUSTRALIAN Sea

PLATE AUSTRALIA
90°E 100°E 110°E 120°E 130°E 140°E 20°S

Figure 4. Wide plate boundary zone of Indonesia. The light shaded area is the zone of collision between the Eurasian, Indian-Australian, and
Pacific–Philippine Sea plates, which can be considered as a wide suture zone, within which there are smaller sutures shown in darker shading.
Lines show incremental growth of the Indonesian region at different stages by addition of fragments of continental, arc, and ophiolitic rocks.
Modified from Hall and Wilson (2000).

al., 1998). In Indonesia there is almost no Paleocene record, and plate. In the Java-Sulawesi sector of the Sunda Arc, volca-
it appears that most of Sundaland was emergent. All published nism greatly diminished during the early and middle Miocene,
plate tectonic reconstructions show subduction beneath Java although northward subduction of Indian-Australian lithosphere
before 45 Ma, but there is no record of volcanic activity in Java continued (Fig. 5). Following Australian continental collision
in the early Cenozoic (Smyth et al., this volume). We suggest in east Indonesia, the Sunda subduction hinge advanced north-
there was a cessation of subduction beneath Java in the Late ward as a result of counterclockwise rotation of the Borneo-Java
Cretaceous following the Cretaceous collision of a continental part of the Sundaland margin, and this is interpreted to have led
fragment at the Sundaland margin (Smyth et al., this volume). to termination of magmatism, despite continued subduction,
Our recent and current work in Java suggests that subduction because of the absence of fresh material replenishing the mantle
resumed in the middle Eocene, forming a volcanic arc in the wedge (Macpherson and Hall, 1999, 2002). At the end of the
Southern Mountains that ran the length of Java. At this time, middle Miocene at ca. 10 Ma, volcanic activity resumed in the
ca. 45 Ma (Fig. 5), the rate of Australia-Antarctica separation Java-Sulawesi sector of the Sunda Arc, after the termination
increased, and Australia began to move northward relatively rap- of Borneo-Java rotation. Magmatism began again in Java with
idly. Since then there has been continuous subduction of Indian increased vigor, but in a position north of the Paleogene arc,
Ocean lithosphere beneath the Sunda Arc. During the Eocene forming the modern arc (Fig. 5). During most of the Cenozoic
and Oligocene the position of the volcanic arc remained broadly since the Eocene, Java was a chain of volcanic islands, and only
fixed, and from Sumatra to Sulawesi abundant volcanic activ- in the last 5 m.y. has the island emerged as an extensive area
ity accompanied northward subduction of the Indian-Australian of land.
90°E 120°E 30°N 90°E 120°E 30°N
45 Ma 15 Ma
Shikoku PACIFIC
Basin PLATE

PACIFIC
PLATE
Luzon
Parece
Proto-South Vela
China Sea Izu
Palawan Basin

Bo
Sulu

nin
Sea

Ma

Su
ma
rian
Su

tra
as
Celebes

ma
Sea West 0° 0°

tr
Borneo

a
Borneo Philippine
c
Ar
Basin

Su
gihe

nd
San Molucca

a
Ja
va Sulawesi
Eas Sea
Ar
c West t P Java
Sulawesi hil Halmahera
ip pin
es

Ha INDIAN-AUSTRALIAN Banda
lm PLATE Embayment New Solomon
INDIAN-AUSTRALIAN ah
er Halmahera Guinea Sea
PLATE a
Arc

20°S 20°S

90°E 120°E 30°N


25 Ma 5 Ma
PACIFIC
PLATE
South
China
Sea
Mariana
Trough

ea th

o
aS u
h in -So
Pr
C ot
Parece Celebes
Vela Sea
c

Basin Ayu
Sea ca


Molu

Trough

Celebes
Sea Bismarck
Caroline Sea
Sea Banda
Sea New
Guinea
Bird’s
Head
Sula INDIAN-AUSTRALIAN
INDIAN-AUSTRALIAN Spur
PLATE PLATE
New
Guinea 20°S
Cenozoic arc processes in Indonesia 33

Figure 5. Reconstructions of the Indonesian region, based on Hall (2002). 45 Ma: The Pacific plate was moving broadly NW and subducting be-
neath the east Asian margin and the eastern margin of the Philippine Sea plate, forming the Izu-Bonin-Marianas Arc. Spreading centers linked the
West Philippine Sea Basin and the Celebes Sea. Australia was moving NNE on the Indian-Australian plate, and northward subduction produced
the Sunda Arc and the East Philippines–Halmahera Arc. The proto–South China Sea was subducting southward beneath the north Borneo–Luzon
margin. The parts of the Pacific and Indian-Australian plates shown in blue without anomalies have been subducted since 45 Ma. 25 Ma: The
East Philippines–Halmahera–South Caroline Arc collided with the Australian margin in New Guinea. Collision between Australia and SE Asia
began with initial contact of the Sula Spur, the continental promontory east of the Bird’s Head of New Guinea, and East Sulawesi. These events
caused major reorganization of plate boundaries. Active spreading in the South China Sea was caused by southward subduction of the proto–South
China Sea. 15 Ma: Subduction continued beneath the Sunda Arc, but arc volcanic activity ceased in Java between ca. 20 and 10 Ma as the hinge
advanced, owing to collision of the Australian margin in Sulawesi. Arc terranes in New Guinea moved westward in a wide left-lateral strike-
slip zone. Locking of splays at the western end of the fault zone induced eastward subduction of the Molucca Sea; westward subduction of the
Molucca Sea beneath the Sangihe Arc had begun at ca. 25 Ma. 5 Ma: Volcanic activity resumed in Java at ca. 10 Ma and has continued until the
present. Hinge rollback into the Banda Embayment south of the Bird’s Head produced the arcuate subducted slab beneath the Banda Arc and led
to backarc spreading in the Banda Sea. The Molucca Sea was in the process of elimination by subduction at its west and east sides beneath the
Sangihe and Halmahera Arcs.

Eastern Indonesia: The Banda Arc longer be accommodated by orogenic contraction. At this time the
oldest oceanic lithosphere in the Indian Ocean, of Late Jurassic
East of Java and Borneo, Indonesia can be considered to age, arrived at the eastern end of the Java Trench. This area of
be a wide and complex suture zone (Fig. 4), and even today old crust north of the NW Shelf of Australia formed an embay-
it can be tectonically described only in terms of several small ment in the Australian margin, the proto–Banda Sea. Because of
plates and multiple subduction zones (Hall, 2002; Hall and Wil- its age and thickness, the Jurassic ocean lithosphere fell away rap-
son, 2000; Hamilton, 1979). Eastern Indonesia is the product of idly, causing the Banda subduction hinge to roll back rapidly to
a long period of extension, subduction, and collision and has the south and east, inducing massive extension in the overlying
a complex basement; there is some continental crust but also plate. In western Sulawesi this first induced extensional magma-
much more arc and ophiolitic basement in comparison with tism, which began at ca. 11 Ma (Polvé et al. 1997). As the hinge
western Indonesia. The Banda Arc is the horseshoe-shaped arc, rolled back into the Banda Embayment it led to formation of the
which today extends east from Flores to Buru (Fig. 3), pass- Banda Arc and the opening of the North Banda Sea, as described
ing through Timor and Seram and includes both an outer non- by Hamilton (1988). Further rollback caused the opening of the
volcanic arc and an inner volcanic arc. Although there was a Flores Sea and later the South Banda Sea (Fig. 5). South of the
volcanic arc in eastern Indonesia during the Paleogene, and Bird’s Head microcontinent the rollback of the subduction hinge
parts of this and possibly older arcs may be found in the highest resulted in collision of the Banda volcanic arc at ca. 3 Ma with
nappes of Timor and other Banda islands, we do not term them the Australian margin in the region of Timor and the cessation of
the Banda Arc. The older volcanic arc rocks formed part of the volcanic activity in this segment of the arc. After collision, conver-
Paleogene Sunda Arc. The Banda volcanic arc is young and has gence ceased south of the volcanic arc, and new plate boundaries
been active only for ~10 m.y. (Abbott and Chamalaun, 1981; developed north of the arc between Flores and Wetar and to the
Honthaas et al., 1999), and the horseshoe-shaped Banda Arc is a north of the South Banda Sea. Within the Bird’s Head microconti-
phenomenon of the same time interval. It formed by subduction nent there has been significant shortening and probable intraconti-
of an embayment within the northward-moving Australian plate nental subduction within the last 3 m.y. at the Seram Trough.
(Charlton, 2000; Hall, 1996, 2002; Hamilton, 1979). The arc
developed within the collision zone after the Australian margin Moluccas: The Halmahera and Sangihe Arcs
collided with the former active margin of the Sunda Arc.
From the Eocene the Sunda Arc can be traced east through The Halmahera and Sangihe Arcs both have long volcanic
Sulawesi into the East Philippines and Halmahera Arcs in the histories. The modern Halmahera Arc is constructed on older
Pacific. During the Paleogene (Fig. 5) subduction was northward arcs, of which the oldest known is an intraoceanic arc formed
beneath Sulawesi and the East Philippines until the first collision in the Pacific in the Mesozoic (Hall et al., 1988a, 1995) pre-
of Australian crust, the Sula Spur, with SE Asia in the early Mio- sumably built on older oceanic crust. The modern Sangihe Arc
cene ~25 m.y. ago (Fig. 5). The collision caused the long subduc- is constructed on arcs formed at the Pacific margin in the early
tion system to separate into two parts. West of Sulawesi, northward Cenozoic (Hall, 2002); again the deepest parts are built on older
subduction continued in the Sunda Arc, but to the east subduction oceanic crust (Evans et al., 1983). Both of the currently active
ceased, and the Australia–Philippine Sea plate boundary became arcs formed during the Neogene. They are unusual in that they
a strike-slip system. Between the two, the Australia–SE Asia col- are the only arcs in the world that are currently colliding.
lision formed a mountain belt in East Sulawesi. By the late middle The pre-Eocene history of these arcs is not well known. At
Miocene, at ca. 12 Ma, convergence in East Sulawesi could no 45 Ma (Fig. 5) the Halmahera Arc was far out in the western
34 Hall and Smyth

Pacific but was situated on the southern margin of the Philip- Molucca Sea has been eliminated by subduction at both its east-
pine Sea plate beneath which there was northward subduction ern and western sides (Fig. 5). The Sangihe–North Sulawesi Arc
of Indian-Australian lithosphere. Between 45 and 25 Ma the is now thrusting over the Halmahera Arc in the Molucca Sea,
Philippines-Halmahera Arc remained at approximately the same as discussed below. The central Molucca Sea mélange wedge
latitude above a north-dipping subduction zone north of Austra- and ophiolites represent the forearc basin and basement of the
lia. At ca. 25 Ma (Fig. 5) there was the most important Cenozoic Sangihe Arc, which will soon have completely overridden the
change in plate boundaries in the region. Arc-continent collision Halmahera Arc (Hall, 2000).
between the East Philippines–Halmahera Arc and the New Guinea
margin terminated northward subduction of oceanic lithosphere RESULTS OF LONG-TERM SUBDUCTION
north of Australia, and a major left-lateral strike-slip boundary
developed in northern New Guinea. Arc terranes were translated There is a great contrast between present tectonic activity,
westward within this strike-slip system. The arc terranes at the manifested by seismicity and volcanism at the margins, and the
southern edge of the Philippine Sea plate moved in a clockwise apparent interior stability of Sundaland, which has led some
direction along the northern New Guinea margin within the left- authors to describe the core as a shield or craton (e.g., Ben-
lateral Sorong strike-slip zone. At the western end of the left- Avraham and Emery, 1973; Gobbett and Hutchison, 1973; Tjia,
lateral fault system there was subduction beneath the Sangihe Arc, 1996). It is true that most of the areas of extreme relief, high
and shortening and uplift in East Sulawesi. The Philippine Sea elevations, and actively rising mountains are in eastern Indonesia,
plate moved northward as it rotated clockwise, accompanied by but the core is certainly not a craton. Most of the Sunda Shelf is
complex strike-slip faulting and minor subduction at its western flat and close to sea level (Fig. 1), and even the large island of
edge in the Sangihe Arc and Philippines (Fig. 5). The Molucca Borneo has generally low elevations with the exception of the
Sea double subduction system was initiated at ca. 15 Ma (Fig. 5). isolated 4 km peak of Mount Kinabalu in the north. However, the
Subduction on the west side of the Molucca Sea in the Sangihe picture of Indonesia as an active volcanic margin surrounding a
Arc had started soon after the 25 Ma plate reorganization, but stable continental region is misleading. There has been significant
the oldest Neogene volcanic rocks in the Halmahera Arc have deformation within Sundaland during the Cenozoic. Heat flow,
ages of ca. 11 Ma (Baker and Malaihollo, 1996). Initiation of seismic tomography, and geological observations indicate that
east-directed Halmahera subduction probably resulted from the the Sundaland continental core north of Indonesia is unusual.
locking of one of the strands of the left-lateral Sorong fault zone Sundaland has high surface-heat-flow values (Fig. 6), typi-
at the southern edge of the Molucca Sea. In eastern Indonesia the cally more than 80 mW/m2 (Artemieva and Mooney, 2001; Hall

Figure 6. Contoured heat-flow map for SE Asia, based on the database of Pollack et al. (1990, 1993) and oil company
compilations (Kenyon and Beddoes, 1977; Rutherford and Qureshi, 1981).
Cenozoic arc processes in Indonesia 35

and Morley, 2004). At the margins of Sundaland, high heat flow is notably at the Indonesian arcs. These features have influenced arc
related to subduction-related processes and magma rise, but in the development in some ways not identified in many arc models.
interior this is not a result of arc magmatism as sometimes sug-
gested (e.g., Nagao and Uyeda, 1995). High heat flow values are ISLAND ARC MODELS
recorded in the interior Sundaland basins from the Gulf of Thai-
land to west Borneo >800 km from the active volcanoes of the Many models of island arc development were formulated in
Sunda Arc and similar distances from oceanic crust of the South the 1960s and 1970s during the rapid period of development of the
China Sea. The hot interior of Sundaland appears to be the conse- plate tectonic theory (e.g., Dewey, 1980; Dewey and Bird, 1970;
quence of high upper-crustal heat flow from radiogenic granites Dickinson, 1973, 1974, 1977; Dickinson and Seely, 1979; Karig,
and their erosional products, the insulation effects of thick sedi- 1971), and they are still commonly reproduced in textbooks.
ments, and a high mantle heat flow (Hall and Morley, 2004). Much of the subsequent work on island arcs in the past 30 yr has
P and S wave seismic-tomography models (Fig. 7) show focused on geochemistry and magmatic processes, as it seems
that Sundaland is an area of low velocities in the lithosphere to be widely thought that our knowledge of tectonic processes
and underlying mantle (e.g., Bijwaard et al., 1998; Lebedev and in volcanic arcs is complete. A common arc model (Fig. 8) sug-
Nolet, 2003; Ritsema and van Heijst, 2000; Widiyantoro and gests that after the initiation of subduction (a problem in itself),
van der Hilst, 1997) in contrast to Indian and Australian conti- a volcanic arc develops, typically between 70 and 100 km from
nental lithosphere to the NW and SE, which are colder, thicker, the subduction trench (Dickinson, 1973, 1977; Karig and Shar-
and stronger. Low mantle velocities are commonly interpreted man, 1975); this width increases with the age since arc initiation.
in terms of elevated temperature, and this is consistent with An accretionary prism formed by progressive transfer of material
regional high heat flow, but they may also partly reflect elevated from the downgoing slab to the upper plate develops near the
mantle volatile contents, partial melting, or seismic anisotropy trench (Scholl et al., 1986). As time proceeds the accretionary
(Lebedev and Nolet, 2003). The high heat flows seen region- prism becomes larger and wider (Dickinson, 1973, 1977), under-
ally across Sundaland, and the generally low mantle velocities thrusting leads to elevation of the arcward part of the prism which
observed in tomographic models, suggest mantle heat-flow val- may form a forearc high, and the arc-trench gap progressively
ues on the order of 40 mW/m2 (Hall and Morley, 2004), which widens. At the same time, debris from the volcanic arc is depos-
are at the high end of the range estimated globally (Artemieva ited in the forearc in basins in front of the arc, and the size and
and Mooney, 1999). Thus they suggest that the region is under- width of these basins increase with time. Continued arc magma-
lain by a thin and weak lithosphere. tism is interpreted to lead to thickening of the crust beneath the
Similar lithosphere has been identified in other subduction arc. The backarc region is usually not shown in the tectonic mod-
zone backarcs (Hyndman et al., 2005). Subduction has produced els; in some arcs backarc basins are formed, floored by oceanic
a different mantle beneath the upper plate, associated with a lith- crust, whereas in others no oceanic backarc basins are formed.
osphere that is not like that of older stable continents and their The reasons for the formation of backarc basins, which are found
margins. The high heat flow and thin, weak lithosphere are the mainly in the western Pacific, are still not clear. However, these
consequence of long-term subduction at the Sundaland margins, basins are some distance from the arc, typically >100 km. Little

Figure 7. Depth slices through S20RTS S-wave tomographic model (Ritsema and van Heijst, 2000) for SE Asia. High
shear velocities are represented by blue, and low shear velocities by red, with an intensity that is proportional to the am-
plitude of the shear velocity perturbations. The range in shear velocity variation is given below each map.
36 Hall and Smyth

Volcanic arc Forearc basin


A Trench
Oceanic crust B Structural in
folding (kneading)

crystalline

Offscraped oceanic
deposits

SEDIMENT
sedimentary SUBDUCTION
C
Subduction complex
ocean crust
crystalline

Lithosphere
small
Forearc basin accretionary
body
D
Asthenosphere

crystalline
Forearc basin
SUBDUCTION
EROSION

SEDIMENT
SUBDUCTION
E

30 km
CRATONIC
MASSIF

SUBDUCTION 50 km
100 km EROSION

Figure 8. Some models for arc development. (A) The arc-trench gap increases in width as progressive accretion occurs at the trench (modified
from Dickinson, 1977). As discussed in the text, this type of model is based in part on observations from Indonesia, but new mapping suggests
a less important role for subduction accretion. (B, C, D, E) The range of tectonic processes at subduction margins, from accretion to erosion,
modified from Scholl et al. (1980).

attention has been given to the area directly behind the volca- knowledge acquired by hydrocarbon exploration in Indonesia to
nic arc, and most arc-derived debris is usually shown to move interpret arc history and tectonic processes.
into the forearc rather than into the area behind the arc. In recent The Indonesian arcs differ from these models in a number
years, in many accounts of arcs in other regions, and in modeling of ways. Arc magmatism may be discontinuous despite continu-
of arcs, the arc model has become more sophisticated. Partly as ous subduction. In many Indonesian arcs there is little evidence
a result of discoveries made by the Ocean Drilling Program, fea- of major addition of new material by accretion at the trench, but
tures such as subduction erosion (e.g., Kay et al., 2005; Ranero equally there is little evidence for subduction erosion. There are
and von Huene, 2000; von Huene and Scholl, 1991) and chang- sudden movements of the position of the volcanic arc, and there
ing slab dips (e.g., Funiciello et al., 2003; Kay et al., 1987; Van- may be evidence of significant thrusting within the arc itself,
nucchi et al., 2004) have been interpreted as important influences not always with obvious cause; collisions, subduction accretion,
on arc development (Fig. 8). However, there has been no ocean subduction erosion, or changing angles of subduction do not
drilling of the Indonesian arcs, and few offshore investigations offer solutions. In general, in the Indonesian arcs there is evi-
of the forearcs; we are still dependent on field studies and the dence of a staccato development, possibly related to processes
Cenozoic arc processes in Indonesia 37

of arc-continent collision but in some cases with no clear reason. cessation of arc activity. Hamilton (1988, 1995) pointed out that
Observations of seismicity on the basis of well-located hypo- subduction zones are often portrayed as static systems with the
centers (Engdahl et al., 1998; England et al., 2004) show that subducting slab rolling over a stationary hinge and sliding down
along the modern arc there are sectors that are aseismic or signif- a slot fixed in the mantle, but this is not a good model. He empha-
icantly less seismically active in comparison with other sectors. sized the necessity of viewing subduction in a dynamic way and
This may indicate differences in coupling of the downgoing and argued that it is most commonly driven by gravity acting on the
overriding plate at the subduction zone, and variations in cou- subducting slab, and most subduction zones are characterized by
pling and decoupling in time as well as space could contribute retreat of the hinge with time, or rollback. Extension of the over-
to a discontinuous tectonic history. This could also be related to riding plate is known to be common in arc settings (e.g., Chase,
dynamics of flow in the mantle wedge (S. Lamb, 2006, personal 1978; Dewey, 1980; Elsassar, 1971; Hamilton, 1988, 1995) and
commun.; Oncken et al., 2006). may be the normal condition for a volcanic arc. For the period
Large-scale dynamic processes have undoubtedly influenced since 25 Ma, tectonic models (Hall, 1996, 2002) suggest that
the stratigraphic record of the Indonesian arcs. However, smaller hinge retreat (Fig. 9) in most subduction zones in SE Asia and
scale effects are also observed. The growth of arc volcanoes has the western Pacific was accompanied by significant arc volca-
in some parts of the region contributed to basin development by nism (Macpherson and Hall, 1999, 2002) and in some zones by
flexural loading. Furthermore the products of the Indonesian vol- marginal basin formation.
canic arcs have been modified by processes related to eruption and Extension in arcs may be essential at regional and local
weathering in a tropical setting. Both these features may have rel- scales to induce melting, to provide pathways for magmas, and
evance to the interpretation of ancient arcs and their stratigraphic to vary the rate of supply of volatile components to the mantle
record, as discussed below. We begin with the larger scale effects wedge. In particular, hinge retreat allows replenishment of the
and move on to more subtle features of arc stratigraphy. mantle wedge by an inflow of hot, undepleted mantle (Andrews
and Sleep, 1974; Furukawa, 1993). This can also cause ablation
Magmatism and Subduction of the base of the overriding plate (Furukawa, 1993; Iwamori,
1997; Rowland and Davies, 1999), and the mantle that upwells
The association between subduction and magmatism is well may melt by decompression. These processes can maintain
known. Early plate tectonic models (e.g., Oxburgh and Turcotte, steady-state arc magmatism at a relatively constant distance from
1970) attributed magmatism to frictional heating at the Benioff the trench over prolonged periods.
zone, melting of the subducting slab, and compressional shorten- Where subduction is characterized by a fixed hinge, or
ing of the overriding plate in the arc-trench gap. Melts are now by hinge advance, magmatism may decline or cease, although
considered to originate primarily in the mantle wedge above subduction continues. A fixed hinge may result from coupling
the subduction zone and are thought to result from the input of of the subducting and overriding plate or from collision, caus-
volatiles that lower the mantle solidus. In Indonesia, almost all ing the two plates to move in the same direction as the slab is
Cenozoic magmatism seems to have been influenced by subduc- being subducted. In these cases, magmatism may cease, although
tion, although it has not always been the direct result of melting subduction continues, simply because the mantle wedge is not
of the mantle wedge above the subducting slab. In some cases being replenished, becomes depleted, and can no longer melt. For
a subduction signature is inherited from previous subduction the period since 25 Ma (Fig. 9), hinge advance in SE Asia and
events. The geochemistry of igneous rocks (e.g., Pearce and the western Pacific was accompanied by reduction in magma-
Peate, 1995) is often used to infer the tectonic setting in ancient tism (Macpherson and Hall, 1999, 2002). The Sunda Arc pro-
arcs, but if a subduction signature is inherited, volcanic arcs may vides a good example. From Java eastward the hinge advanced
be mistakenly identified in ancient orogenic belts as subduction from ca. 20 to 10 Ma, and magmatic activity declined. In East
related when they are not. Macpherson and Hall (1999, 2002) Java, Paleogene volcanic activity culminated with a short-lived
highlighted areas in Indonesia where there has been arc-like explosive phase (Smyth et al., this volume) and then declined sig-
magmatism but subduction is not active, such as south Sulawesi nificantly or terminated; magmatism resumed at ca. 12–10 Ma.
(Polvé et al., 1997) or northern Java, where volcanoes are very far There is no indication that subduction ceased or slowed during
from the active arc and have an ultra-potassic character (Edwards the period of reduced magmatism, and the Indian-Australian
et al., 1994; Macpherson, 1994). In these places the character of plate continued to move northward relative to Eurasia at a similar
magmatism reflects older subduction events that have enriched rate to that during the Paleogene. The reduction in magmatism
the mantle. can be understood as the result of hinge advance driven by the
However, the association between magmatism and subduc- collision of Australia in eastern Indonesia (Hall, 1996, 2002) and
tion is commonly not straightforward. Some arc models portray manifested by counterclockwise rotation of the Borneo region
compressional arcs with active magmatism, and contraction is (Fuller et al., 1999). When rotation of Borneo ceased, the hinge
often thought to be typical of active arcs, probably because the retreated again and magmatism resumed, especially in the Banda
subduction model causes us to think of plate convergence. In Arc, where dramatic retreat of the subduction hinge required a
the Indonesian arcs, contraction is commonly associated with massive influx of fertile mantle beneath the overriding plate.
25-5 Ma
Japan
25-5 Ma

u
u ky Izu-Bonin
Ry 25-5 Ma

Hinge
Advance

Phil
ippi
Hinge
nes
Mariana Retreat
25-15 Ma
Sangihe Mariana
Sunda 15-5 Ma
25-5 Ma

Halmahera
15-5 Ma Solomons
New
Sunda-Banda Hebrides
20-10 Ma 10-5 Ma

A: Andaman Sea
5-0 Ma BS: Banda Sea
Bi: Bismarck Sea
M: Mariana Trough
Wd: Woodlark Basin
Japan

u
u ky Izu
Ry Bonin

Hinge
Phil

Manila Advance
ippi

Trench M
ne T

Hinge
A
Mariana Retreat
renc
Sa
ng

h
ihe

Halmahera
North
Sulawesi
Bi New Britain
Solomons
Sunda
BS
da Wd
Ban

Figure 9. Movements of subduction hinges in the Neogene, modified from Macpherson and Hall (2002). Regional plate reorganizations occurred
at ca. 25 Ma and 5 Ma. Major coastal outlines are shown for reference. For subduction zones shown without shading, there was no significant
movement of the hinge. Bold letters indicate areas of young marginal basin formation.
Cenozoic arc processes in Indonesia 39

Subduction Accretion and Subduction Erosion Near the trench there is some evidence for a young accretion-
ary complex constructed against an older backstop (Fig. 10),
A feature of many early arc models is an increase in the width interpreted to be an older accretionary complex (Kopp et al.,
of the arc-trench gap with increasing age, commonly attributed to 2002, 2006; Schlüter et al., 2002), but this is only ~50 km wide.
accretion, and some of these models were based on observations Schlüter et al. (2002) suggest there was a progressive growth of
from Indonesia (e.g., Dickinson, 1973; Moore and Karig, 1980). the arc-trench gap south of Sumatra since the Paleocene, but the
In contrast, in many arcs the role of subduction erosion in the age control on this development is poor. In contrast, Kopp et al.
development of the arc has become widely recognized as impor- (2006) suggest there has been subduction erosion at the Sunda
tant (e.g., Scholl et al., 1980; von Huene et al., 2004). In Indo- margin south of Java, and if this is correct, the arc-trench gap
nesia these contrasting processes are difficult to assess, because must have been wider in the past. The seismic lines on which
the age and nature of the crust between the arc and the trench in these interpretations are based do not show a progressive growth
Indonesian arcs is unexplored or poorly known. There is a great of the accretionary zone but rather a young accretionary com-
need for detailed exploration of these parts of all arcs. However, plex juxtaposed against an older one (Fig. 10). This relationship
observations on land suggest that neither subduction accretion could equally be interpreted in terms of an abrupt change in the
nor erosion has had a significant impact on arc history. subduction history, and we note that on land there is evidence for
such a change; the volcanic arc ceased activity at ca. 20 Ma and
Change in Width of Arc-Trench Gap with Age resumed activity ~10 m.y. later, 50 km north of its previous posi-
There is good evidence of Early Cretaceous subduction and tion. Why this happened is not clear. At present the available data
widespread evidence from Sumatra to SE Borneo of middle to are simply not adequate to distinguish different models.
Late Cretaceous collisions accompanied by ophiolite emplace- However, investigations of the Sumatra forearc in the Nias
ment (Miyazaki et al., 1998; Parkinson et al., 1998; Wakita, 2000). region do provide the basis for testing the interpretation of arc
As noted above, there is almost no Paleocene stratigraphic record growth by continuous accretion. Nias, at the north end of the
in Indonesia, and the whole region seems to have been emergent. Sumatra forearc (Fig. 1), is commonly cited as the classic exam-
In many parts of Indonesia, poorly dated, typically terrestrial ple of a forearc high elevated by accretion, following the work of
rocks interpreted to be Paleocene or Eocene rest unconformably Moore and Karig (1980) on the basis of their mapping of part of
on older basement rocks. In East Java the oldest sedimentary the island. Later mapping of the whole island and other islands of
rocks that rest unconformably on ophiolitic basement are middle the forearc does not support this model (Samuel, 1994; Samuel
Eocene or older and lack volcanic debris (Smyth, 2005; Smyth et al., 1995, 1997; Samuel and Harbury, 1996). There is little evi-
et al., this volume). There is little evidence for latest Cretaceous dence on Nias for the progressive addition of material. The ophi-
to early Eocene volcanic activity in most of the Sundaland mar- olitic basement includes Cretaceous rocks and locally is overlain
gin between Sumatra and Sulawesi. The Cenozoic stratigraphic unconformably by a thin Eocene marine cover. Most of the island
record, commonly volcanogenic, begins in the middle Eocene, consists of a thick sequence of Oligocene to lower Miocene deep
and in Java the current phase of subduction along the arc began marine clastic sedimentary rocks and a thick sequence of lower to
after 45 Ma, when Australia began to move rapidly northward. upper Miocene shallow marine clastic sedimentary rocks. There
Although it is commonly assumed that subduction at the Sunda are a few limestones and some tuff layers in the sequence. The
Trench in the Late Cretaceous (e.g., Metcalfe, 1996; Hall, 2002) island became emergent in the Pliocene. The ophiolitic base-
continued into the Cenozoic, we question this assumption. Ham- ment was in place by the Eocene, and the deep marine sedimen-
ilton (1988) stated that the modern subduction system in Java was tary rocks above the basement were deposited on it. However,
inaugurated no earlier than late Oligocene time. We, too, consider the sedimentary rocks are not material scraped off the downgo-
that the present subduction began only in the early Cenozoic, ing plate but represent the fill of a forearc basin, later inverted
although we now know of evidence for subduction beneath Java (Fig. 11), in which most material was carried from the direction
before the late Oligocene (Smyth, 2005; Smyth et al., this vol- of Sumatra. The Oligocene to early Miocene development of the
ume). However, there is no evidence in Java to support the sugges- outer forearc was characterized by extension, followed by early
tion of late Paleogene southward subduction or an arc collision in Miocene inversion. Mélanges are not subduction accretion com-
or with Java in the Paleogene (Hamilton, 1988). plexes but are diapiric in origin and formed by mud volcanism
In the Sunda Arc between Sumatra and Java the arc-trench that continues today. The forearc islands have a complex history
gap is rather large, between 300 and 350 km (Fig. 1), and subduc- of local uplift and subsidence to which addition of material at the
tion began in the early Cenozoic. If the arc-trench gap increased trench has contributed, but the Sumatra forearc does not fit very
progressively with age, there should be evidence of movement well to the model of continuous widening and accretion inter-
of either the volcanic arc or the trench during the Cenozoic. In preted from the early land-based studies.
East Java the volcanic arc remained in the same position from the
Eocene to the early Miocene (Smyth, 2005; Smyth et al., this vol- Mélanges
ume). It then ceased activity, and a new arc formed ~10 m.y. later, Mélanges were formed at numerous stages in the develop-
50 km to the north, where it has remained since the late Miocene. ment of arcs in Indonesia (Hall and Wilson, 2000; Hamilton,
40 Hall and Smyth

SW NE
0 JAVA
A SEA Outer Arc High Mentawai
2 Forearc Basin
INDIAN B
OCEAN C
4
TWT (s)

10°S Java Mentawai


Trenc Wrench
100°E h 110°E Accretionary Wedge I
6 Fault
Accretionary Wedge II
Sunda Trench
8 Continental crust

10 A Oceanic crust of Indian Ocean Plate 50 km

SW ACCRETIONARY DOMAIN NE
0 FOSSIL Forearc Strata
ACTIVE
Trench infill
FRONTAL PRISM
-10 OUTER HIGH OCEANIC-TYPE SHELF
CRUST
Depth (km)

Oceanic mantle Sub Shallow Mantle


duct
ing S
-30 lab

B 50 km
-40
S OUTER FOREARC HIGH FOREARC BASIN N
0
Basin Sediments:
1 Slope Deposits Hemipelagic / Pelagic sediment Normal
and Volcanic Ashes
Faulting
2
Trench Vertical Movement Post-sedimentary
TWT (s)

and Doming buckling


3

6
C 40 km

Figure 10. Cross sections across the Sunda forearc, based on marine geophysical investigations. (A) At the south end of Sumatra the accretionary
wedge is interpreted to be ~300 km wide and constructed against continental crust beneath the forearc basin (Schlüter et al., 2002). (B) Interpret-
ed tectonic units of the Sunda margin south of West Java (Kopp et al., 2002), where the accretionary domain is interpreted to have an active part
and an older part, juxtaposed against oceanic-type crust beneath the forearc basin. (C) Section south of Central Java (Kopp et al., 2006) shows
an uplifted forearc high, covered by thin sediments interpreted as evidence of subduction erosion. The age of the unconformity is unknown, but
the thin sediments above the unconformity suggest a young age. A similar unconformity on land is late Miocene to Pliocene, which developed
after northward thrusting within Java.

1979). Marine geophysical studies suggest that mélanges are and seismically incoherent volume of sediment in the central
formed at Indonesian subduction zones (Kopp et al., 2002; Molucca Sea. Mélanges have also formed far from subduction
Schlüter et al., 2002), but they are also formed at later stages dur- zones without collision (Samuel et al., 1997). In Timor, mélange
ing collision (Silver and Moore, 1978). In the central Molucca formation is attributed by some authors mainly to thrusting
Sea (Fig. 12) mélanges were formed at two stages. Mélanges related to collisional processes (Harris et al., 1998) and by oth-
reported from Talaud (Moore et al., 1981) and present on Mayu ers mainly to diapiric processes expressed as modern eruption of
were not formed during the present arc-arc collision but are older mud volcanoes (Barber et al., 1986). Active mud volcanism in
rocks forming part of the pre-Neogene basement of the Sangihe East Java (Smyth, 2005), behind both the Paleogene and mod-
forearc. Presumed mélanges of the present collision complex are ern arcs, erupts rock samples from deeper parts of the basin
all submarine and constitute part of the bathymetrically shallow directly north of the modern arc and brings to the surface blocks
Cenozoic arc processes in Indonesia 41

Lahewa Sub-basin Mujoi Sub-basin Gomo Sub-basin Mola Basement High

MID-MIOCENE– SEDIMENT & OLIGOCENE–


MELANGE LATE PLIOCENE–
EARLY PLIOCENE EARLY MIOCENE
COMPLEX PLEISTOCENE
Coast SEDIMENT Coast

MID-MIOCENE–
EARLY PLIOCENE

BASEMENT
SW NE

0 V=H 10 Km

SW NE
Accretionary Forearc Sumatran fault zone,
Indian Ocean Trench wedge Nias magmatic arc & backarc
basin

Continental crust

0 V=H 60 Km
Moho

+
Overriding
continental
lithosphere

Figure 11. Cross sections across the Sumatra forearc, modified from Samuel and Harbury (1996). Mapping of Nias by Samuel (1994) shows
that the area between Sumatra and Nias was in extension for much of the Cenozoic and that most of the island’s sediments were derived from
Sumatra. The island has not emerged as a forearc high as the result of progressive subduction accretion. V=H—vertical and horizontal scales
are equal.

of the oldest parts of the basin sequence entrained in overpres- of the two arcs (Fig. 13). However, on Talaud a post–middle
sured muds. This is clearly not related to subduction because the Miocene sedimentary sequence rests unconformably on ophio-
Benioff zone is >100 km below the site of mud volcanism. The lites that are middle Eocene or older (Moore et al., 1981). The
mélanges at the surface, and the presumed equivalents beneath middle Miocene to Pleistocene rocks are tuffaceous sandstones,
the surface, are the result of overpressures related to rapid depo- siltstones, and shales with intercalations of limestone, marl, and
sition of thick sedimentary sequences, hydrocarbon generation, conglomerate in which the sediments are dominated by volcanic
and young thrusting very far from the subduction zone. debris. They were deposited in deep water by turbidity currents.
On Talaud, Neogene strata are in many places little deformed,
Mélanges and Ophiolites and they are moderately to strongly deformed locally and typi-
The emplacement of ophiolites is often linked to the pro- cally in narrow zones adjacent to mélange rocks or ophiolites.
cesses of mélange formation and accretion. Between the Sangihe On Mayu, coastal exposures of highly indurated mélange yield
and Halmahera Arcs (Fig. 12) is a collision complex in which Eocene ages (Baker, 1997). All of the mélange is probably older
there are ophiolites and mélanges. The Talaud Ridge at the center than middle Miocene, and none appears to be related to the pres-
of the Molucca Sea was interpreted on the basis of marine geo- ent collision.
physics as a mélange wedge (McCaffrey et al., 1980; Silver and The similarity in character and structure of the ophiolitic
Moore, 1978) including slices of the Molucca Sea lithosphere rocks to those of the basement complex on Halmahera (Hall
(Fig. 13). These rocks are exposed on Talaud and on two tiny et al., 1988a) suggest that much of the mélange formed dur-
islands along the ridge, Mayu and Tifore. The ophiolites were ing Eocene or older deformation events. The interpretation of
interpreted by McCaffrey (1983, 1991) as part of the subducted the ophiolites as part of the subducted Molucca Sea lithosphere
Molucca Sea plate, which has been emplaced by the collision requires some complex arguments (McCaffrey, 1991) to account
42 Hall and Smyth

Philippine
Fault Philippine
Trench
EURASIAN PLATE

Sulu Sea
PHILIPPINE SEA
Mindanao PLATE
km
00
18
Celebes Sea Cotobato
Trench
Sangihe Arc
Halmahera Arc

Halmahera

MOLUCCA SEA
PLATE

SORONG FA
Sulawesi ULT ZONE

New Guinea
Banda Sea
AUSTRALIAN PLATE

1800 km

Figure 12. Molucca Sea region, showing the converging Halmahera and Sangihe Arcs, displayed in a three-dimensional (3D) diagram that
represents, in simplified form, the geometry of the converging plates in east Indonesia. Modified from Hall et al. (1995).

for their elevation instead of their disappearance by subduction. gap is just as wide. The role of the upper plate in providing mate-
In contrast, if the ophiolites are interpreted as part of the Eocene rial is well illustrated at the eastern edge of the Philippine Sea
or older basement (Fig. 13), as they are in Halmahera, and they plate in the Nankai margin, where classic fold and thrust struc-
formed the basement to the forearc, the gravity model is much tures are formed in the thick “accretionary wedge” (Mikada et
easier to understand. Where the Halmahera forearc and arc have al., 2005; Taira et al., 1991). The sediment in the upper part of the
been significantly overthrust the Sangihe forearc has been jacked prism is dominated by clastic material carried along the trough
up. The wide Molucca Sea collisional complex is composed from the Japan–Izu-Bonin Arc collision zone, and relatively
of the accretionary wedges of both arcs. The basement of the small amounts of material are contributed from hemipelagic
Sangihe forearc is exposed where it thrusts over this wedge. The sediments on the subducting slab (Karig and Ingle et al., 1975;
ophiolitic rocks of the central Molucca Sea are not part of the Klein and Kobayashi et al., 1980; Taira et al., 1991). In Indonesia
Molucca Sea plate but are the basement of the Sangihe forearc the role of the structurally upper plate is well illustrated by the
(Hall, 2000). The ophiolites were the basement before the mid- eastern Makassar Strait, where there has been no subduction, yet
dle Miocene. accretionary-style structures have developed in sand-dominated
sedimentary wedges that have built out from Sulawesi (Puspita et
Accretion or Erosion? al., 2005). These structures closely resemble the structures devel-
In many Indonesian arcs, there seems to have been relatively oped in subduction accretionary complexes, and major eastward
little growth of the forearc region by the addition of material subduction during the Cenozoic has been suggested by some
from the downgoing plate. Many so-called accretionary com- authors (e.g., Charlton, 2000; Guntoro, 1999). However, most
plexes are in fact dominated by material derived from the arc. At authors have interpreted the West Sulawesi margin to be either
the northwestern end of the Sunda Arc there is undoubtedly sedi- a passive margin with oceanic crust to the west (e.g., Bergman
ment at the distal end of the Bengal Fan (Curray, 1994) that is et al., 1996; Cloke et al., 1999; Hall, 1996; Hamilton, 1979) or a
being added to the overriding plate, although most of this is prob- rifted margin with the thinned continental crust to the west (e.g.,
ably mud-dominated sediment (Stow et al., 1989). Farther south, Calvert, 2000; Situmorang, 1982). Recent offshore seismic data
south of south Sumatra and Java, the amount of sedimentary show that there has been no subduction (Fraser et al., 2003; Pus-
cover on the subducting plate is much thinner, yet the arc-trench pita et al., 2005) and suggest that the Makassar Strait is floored
Accretionary complex
Accreted
SANGIHE TALAUD sediments and crust SNELLIUS RIDGE
of Halmahera forearc MINDANAO
Forearc basin Backarc basin
0 v
v
v
v
v
v
v v v v v v
v v v v v v v v v v
10 v v v v v v v v
v Arcv crust v
v v Forearc crust v v v v Backarc crust
v v v v v
v v
v v v
v Arc
v
crust v
v
v
v
20 v
v
v
v
v
v
v
v
v v
v
v v
v v v v v v
Talaud
30
EURASIAN PLATE rc c
ru st
PHILIPPINE SEA PLATE A
Kilometers

40 Sangihe Arc rea Halmahera Arc

1
Fo
50 Snellius
Sangihe Ridge
60 Sangihe mantle Halmahera mantle

70 B
80
90
MOLUCCA SEA PLATE A Mayu
Morotai

A
ES
C

ER
100

AH
LA

LM
SU

HA
s

NW SANGIHE
Accretionary complex
MOROTAI
SE
Forearc basin Backarc basin
Forearc basin
0 v v v v v v
v v v v v
v v v v v v v v v
v v v
10 v v v v v Forearc crust Forearc crust v v
Arc crust v v v
v v v v v v
Backarc crust
v Arc crust v v v v v
v
v
v
v
v
v
v v v
20 v v v v v v v v v v v v v
v v v v v v v
v v
30
Kilometers

40
EURASIAN PLATE PHILIPPINE SEA PLATE
Sangihe Arc Halmahera Arc
50
60
Sangihe mantle Halmahera mantle
70 MOLUCCA SEA PLATE
80
90
B
100
100
50
Calculated MGAL
0
Observed -50
-100
Talaud-Mayu -150
0 50 100 Ridge 150 200 250

-0.59 -0.59 -1.83 km


-0.43 -0.43 10
0.40

20

0.44 C 30

WNW Sangihe Arc Halmahera Arc ESE


100 0 km 100 200 300
0
10
20
2.86
30
Kilometers

40
50 3.40 g/cm3
60
70
+0.20 +0.20
80 3.34 +0.05 +0.05
90 Mantle density D
100

Figure 13. Cross sections across the Molucca Sea (Hall, 2000) drawn at the same vertical and horizontal scales (A, B) in comparison with those
based on gravity modeling (C, D) by McCaffrey et al. (1980). In section A at the latitude of Talaud, the entire arc and forearc of the Halmahera
Arc has been overridden by the Sangihe forearc. Ophiolites interpreted as part of the Sangihe forearc basement are exposed in the Talaud Islands.
Farther south (section B), only part of the forearc has been overridden, but the Halmahera Arc in Morotai was overridden by its own backarc in
an earlier thrusting episode. In contrast, McCaffrey et al. (1980) interpret the ophiolites as part of the subducted Molucca Sea plate.
44 Hall and Smyth

by extended continental crust (Nur’aini et al., 2005; Puspita et the whole of the forearc and volcanic arc have disappeared. This
al., 2005). The structures are foreland-type fold and thrust belts, process is not complete, and it is likely that much, if not all, of
and developed as westward thrusting has progressed since the the Halmahera Arc will disappear beneath the Sangihe Arc. This
Pliocene, all the material having been derived from Sulawesi. process is certainly not subduction erosion, as intended by most
There certainly has been addition of material to the continen- authors (Scholl et al., 1980; von Huene and Scholl, 1991; von
tal margin of Sundaland during the Cenozoic (Fig. 4). However, Huene et al., 2004). However, neither is it subduction accretion,
the additions occurred during relatively short periods of time, at as usually envisaged. The arc-trench gap of the Sangihe Arc has
widely spaced intervals, and were related to the collision of con- increased since the two arcs came into collision, but the arc-
tinental fragments at the Sundaland margin. The material added trench gap of the Halmahera Arc has diminished.
included arc and ophiolitic rocks, which may in some cases rep-
resent backarc basins within the continental margin or the rem- Arc Movements and Changing Slab Dips
nants of material subducted during the arc-continent collision. In
Indonesia, arc growth seems to have occurred discontinuously Observations in Indonesian arcs raise the question of whether
rather than by continuous steady-state addition of material at the there is such a thing as a compressional arc. Arcs may certainly
trench that led to gradual widening of the arc-trench gap. be compressed, but when contraction occurs volcanic arc activity
ceases. Magmatism may resume when compression ceases, but
Disappearance of Arcs the stratigraphic record suggests it does not persist during con-
Westward subduction of the Molucca Sea beneath the traction. In the Banda Arc, magmatic activity terminated in the
Sangihe Arc probably began in the early Miocene. Eastward sub- Wetar sector of the volcanic arc at ca. 4 Ma, when arc-continent
duction of the Molucca Sea plate beneath Halmahera began in collision began in Timor, and was never resumed during stacking
the middle Miocene. The double subduction zone was initiated of nappes (Audley-Charles, 2004). In the Halmahera Arc, vol-
at this time, forming a new plate, the Molucca Sea plate, separate canic activity ceased as arc-arc collision began but resumed in a
from the Philippine Sea plate. The oldest volcanic rocks dated new location shortly afterward. The repeated failures in different
from the Halmahera Arc are 11 Ma in Obi at its southern end, and places in the entire arc system in Halmahera suggest that inter-
they become younger to the north (Baker and Malaihollo, 1996). mittent thrusting was followed by periods of relaxation in which
The Molucca Sea was eliminated from south to north, and the arc magmatism resumed.
two forearcs began to collide. The result of the collision was that Such abrupt movements in the position of the volcanic arc
the Halmahera Arc system (meaning the entire region between are not uncommon and may follow periods when an arc is com-
the trench and backarc region) failed repeatedly, with thrusting in pressed, but they may be due to other tectonic causes. Sudden
different directions at different stages in the collision. First, the shifts in position of the volcanic arc could be related to shorten-
backarc was thrust over the volcanic arc, and later the forearc was ing of the entire arc system (forearc, arc, and backarc), which
thrust toward the volcanic arc. In south Halmahera the backarc may involve thrusting of the backarc or the forearc toward the
region was thrust onto the forearc, in places entirely eliminating arc, the magmatic arc itself being the weakest point. Field obser-
the Neogene arc. At the southern end of the Halmahera Arc on vations in Halmahera show that the site of thrusting was close
Obi the arc was thrust onto the forearc (Ali and Hall, 1995; Ali et to the former active arc. Experimental and numerical modeling
al., 2001). After west-vergent thrusting, volcanism in the Halma- studies also show that the active arc is the weakest point in the
hera Arc resumed between Bacan and north Halmahera. At the whole arc system (Shemenda, 1994; Tang and Chemenda, 2000;
south end of the arc on Obi, and at the north end from Morotai Tang et al., 2002). These studies show that failure occurs directly
northward, volcanism ceased. In the northern Molucca Sea the beneath the arc. It can take place on faults dipping toward the
Sangihe forearc was then thrust east onto the Halmahera forearc trench or arc, and this depends on the thickness of subducted
and arc. In the region between Morotai and the Snellius Ridge, crust and flexural rigidity of the upper plate.
parts of the Neogene Halmahera Arc and forearc have now dis- Since the middle Pliocene (Fig. 14) the Halmahera Arc sys-
appeared. Farther south, this east-vergent thrusting carried the tem failed more than once close to the site of the active volcanic
Halmahera forearc onto the flanks of the active Halmahera Arc, arc, presumably reflecting its weakness owing to mineralogy and
and pre-Neogene rocks of the Halmahera forearc basement are magmatism. Active volcanism has most recently resumed in a
now exposed in islands of the Bacan group and off the coast of restricted area at the center of the arc chain, over a distance of
northwest Halmahera. ~200 km, in comparison with the 700 km length of the early Plio-
Cross sections drawn across the present-day collision zone cene arc. The arc-arc collision explains why the arc volcanism
from south to north can also be considered to display the sequence may have ceased and then resumed as stresses changed, but this
of events in time, and a series of sections illustrating the earlier does not really account for the change in position of the arc. The
stages in the collision can be inferred from the geology of the Quaternary volcanoes are situated ~50 km west of the Pliocene
Halmahera Arc (Fig. 14). An interesting consequence of the col- centers (Fig. 15). Hall et al. (1988b) speculated that the shift
lision is that the Halmahera Arc is being progressively overrid- may be related to a steepening dip of the subducted slab beneath
den by the Sangihe Arc, and in the northern Molucca Sea almost Halmahera. A highly complex geometry of subducted slabs lies
Talaud ridge

0 Ma

Philippine Sea plate

Halmahera Arc and forearc overridden


by Sangihe forearc
Molucca Sea plate sinks deeper
Emergence of ophiolitic basement of
Sangihe forearc

0 Ma

Halmahera forearc overridden by Sangihe forearc


Molucca Sea plate sinks
Local emergence of islands in ‘collision complex’

0 Ma

Halmahera Arc fails again


Overthrusting of forearc region: detachments
deep within basement
Minor backthrusting at back of Sangihe forearc

Sangihe Arc Halmahera Arc


2 Ma

Molucca Sea plate

Volcanic activity ceases in Halmahera arc


Arc completely overridden by backarc region

Figure 14. Cross sections across the Molucca Sea drawn at same vertical and horizontal scales to illustrate the sequence of convergence of the
Halmahera and Sangihe Arcs since 2 Ma. The lowermost section is inferred from geological mapping on land. The upper three sections are drawn
at different latitudes across the Molucca Sea from south (bottom) to north (top). They can be considered to represent the sequence of events in
the last ~2 m.y. that have led to the structure at the latitude of Talaud, where convergence is most advanced. Collision has resulted in the almost
complete elimination of the Halmahera Arc and forearc at the latitude of Talaud.
A 127oE 129oE

I
TA
RO
O
M
o
2N

Modern volcanoes of the


Halmahera Arc
Pleistocene to modern
Molucca Sea Halmahera Arc

Miocene to Pliocene Arc

HALMAHERA

0o Weda Bay

0 50 100
BACAN
Km

B o
110 E
o
112 E
Modern volcanoes of the Sunda Arc

Oligo-Miocene volcanic centers of


the Southern Mountains Arc

7oS

EAST JAVA

SUND
A ARC

SOUT
MOUN HERN
TAIN
ARC S
8oS

0 50 100

Km

Figure 15. Abrupt movements of the volcanic arcs in (A) Halmahera (ca. 2 Ma) and (B) Java (ca. 10 Ma) took place, as discussed in the text.
Cenozoic arc processes in Indonesia 47

beneath the Halmahera and Sangihe Arcs and the southern Philip- considered backarc basins. They are found behind the arc but
pines (Fig. 12). The west-dipping Philippine slab has now arrived are much closer to it than typical backarc basins. They are not
at depths where it would hit the subducted Halmahera slab and underlain by newly formed oceanic crust, nor are they charac-
increase its dip, thus shifting the arc to the west if melting occurs terized by obvious extension. They are filled with volcanic arc
at a constant depth. Macpherson et al. (2003) showed that dif- material, mainly reworked as sediments. The basins formed in
ferences in the geochemistry of Quaternary lavas in comparison a setting that may have been extensional (as noted above, pos-
with Neogene lavas of the Halmahera Arc are consistent with an sibly the typical condition of a volcanic arc) or neutral, using
increased sediment flux interpreted to indicate an increase in dip the terminology of Dickinson (1995) and Busby and Ingersoll
of the subducted slab. (1995), but certainly not compressional. They are not retro-arc
In East Java, activity in the Southern Mountains Arc termi- basins, and although there may be evidence of thrusting, this
nated in the early Miocene after a period of >20 m.y. at the same occurred after the basin formed.
location. About 10 m.y. later a new arc formed, 50 km north of East Java and Halmahera provide two examples. In Java the
the older arc, that has remained at the same position since the deep Kendeng Basin behind the arc is filled with a thick sequence
late Miocene (Fig. 15). Why the volcanic arc moved north is not of volcanic and sedimentary rocks mostly derived from the arc
known. There is no evidence for a collision at the Java Trench. itself. The basin formed in the middle Eocene, when the volca-
There was Neogene contraction in Java, although it is not well nic arc began its activity. Modeling by Smyth (2005) suggested
dated. It is possible that the slab dip remained at the same angle, that at least part of the accommodation space was created by the
the depth to the Benioff zone remained constant, and therefore load of the volcanic arc. The basin in Weda Bay (Nichols and
contraction of the arc meant that when volcanism resumed the Hall, 1991) behind the Halmahera Arc has been investigated as
new arc formed at the same distance from the trench. Several part of oil exploration, and seismic lines close to the arc show no
possible explanations are suggested by Smyth et al. (this vol- indication of fault control on subsidence (see Waltham et al., this
ume), but all are speculative. volume). Again, a likely important contribution to subsidence is
The abrupt change in position of the volcanic arc is a feature flexural loading by the volcanic arc. This suggestion is explored
of several Indonesian arcs, and the causes are not known. How- in more detail by Waltham et al. (this volume) in a mathematical
ever, we do not know of examples of gradual movement of the model. They conclude that volcanic loading can make a contri-
arc with increasing age of the magmatic arc, as might be expected bution to basin subsidence in arc settings within ~200 km of the
from many arc models. This could mean that slab dip gradually arc, which may be the primary cause. In continental margin arcs
declines as the arc-trench gap increases, so that the arc remains in such as Java, and long-lived intraoceanic arcs such as Halmahera,
the same position. Obviously this idea is difficult to test, as there there is a significant density contrast between the deeper crust
is no way of measuring slab dip in the past. The dip of slabs in and the eruptive arc products. In these cases the volcanoes form
most Indonesian arcs is typically steep where the slab reaches a load that can produce or contribute to the formation of flexural
>~200 km in depth. The depth to the Benioff zone beneath active basins close to the arc. Elsewhere the absence of deep basins in
arcs is also much more variable than previously thought (England arcs may be due to the small density contrast between volcanoes
et al., 2004). Perhaps magmatic activity essentially pins the arc in and the underlying crust.
one position until it ceases, and then there are a number of factors
which could control the location of the resumption of volcanic Sediment Character in Indonesian Arcs
activity, including thickness or strength of the crust, preexisting
weaknesses in the crust, or changing dip of the subducting slab. All the Indonesian arcs have been in an equatorial position
At present, all are conjectures and illustrate how incomplete our throughout the Cenozoic (Hall, 2002), and as a result the arc
knowledge of arcs remains. products have formed in a tropical climate. Tropical processes
have several effects on rocks and grains within them. Deep tropi-
Volcanic Arc Loading cal weathering of well-jointed rocks can lead to a high degree of
in situ rounding of material that eventually falls out of the out-
Volcanoes exert a load. Since most basic to intermediate crop (Fig. 16). The onion-skin type of weathering of outcrops
arc volcanoes are denser than average continental crust, long- is common in Indonesia, and rounded clasts are formed at all
lived volcanism should contribute to subsidence. Several of scales from huge boulders to pebbles. The rounded clasts may
the Indonesian arcs have deep filled basins close to, and both be incorporated in volcaniclastic and sedimentary rocks without
in front of and behind, the arc. Few of these basins have been any transport whatsoever. Weathering of material also leads to
explored seismically, and their deep structure is unknown. How- rapid destruction of unstable rock fragments and minerals. This
ever, there are indications that arc loading may have contributed is particularly important for volcanic-derived material that con-
to their formation, because the basin sequences thicken toward tains many unstable minerals and rock fragments such as mafic
the arc, and the timing of basin development is closely linked to minerals, feldspars, and clays. The interpretation of transport his-
activity in the arc. As discussed by Nichols and Hall (1991) and tory and maturity based on grain shape and light mineral modes
Smyth et al. (this volume), the basins are not what are normally can be very misleading in tropical settings. Two recent studies of
48 Hall and Smyth

Figure 16. Tropical weathering of rocks, resulting in rounding of clasts before they are even removed from the outcrop. Rounded clasts are
produced at all scales from small pebbles to boulders. Both photographs show rocks exposed by quarrying of the Pendul Diorite, Jiwo Hills,
East Java.

sedimentary rocks deposited close to active margins in the Indo- reflects weathering rather than provenance. Apatite is stable dur-
nesian region illustrate the effects of tropical weathering (van ing burial but is susceptible to acidic weathering, which is com-
Hattum, 2005; Smyth, 2005). In both cases the tectonic setting mon in humid tropical settings (Morton, 1984). Plots of chemical
is known, but if these examples had come from much older oro- indices of weathering and alteration (Fig. 17) show that very few
genic belts the interpretation could well be very different. of the sandstones are similar in composition to the fresh source
There was subduction of the proto–South China Sea beneath rock (van Hattum, 2005). The composition has changed between
north Borneo (Hall, 2002; Hazebroek and Tan, 1993; Hutchi- erosion and deposition. The tropical setting of the sandstones
son, 1996; Tongkul, 1991) between the Eocene and early Mio- needs to be considered before provenance can be interpreted.
cene. A large sedimentary fan, the Crocker Fan, formed at this In East Java, Eocene to lower Miocene quartz-rich sand-
active margin. All sandstones are quartz-rich, and their composi- stones plot on commonly used ternary diagrams as recycled oro-
tions plotted on conventional ternary diagrams suggest they are genic or cratonic interior–derived sediments (Fig. 18), but this
mature recycled orogenic products, but there are some anomalies interpretation is even more misleading than that for the Borneo
(van Hattum et al., 2003; van Hattum, 2005). The oldest sand- sandstones. Examination of quartz grains shows that many have
stones show the greatest compositional maturity, and although a volcanic origin, and a volcanic provenance is supported by the
younger sandstones plot as recycled orogenic sediments they abundance of fresh euhedral zircons. The sandstones have previ-
are less mature, which is inconsistent with their derivation by ously been interpreted as eroded from a continental Sundaland
recycling of older sedimentary rocks as suggested by Hutchison source because they are rich in quartz and well sorted, but careful
et al. (2000) and William et al. (2003). The sandstone textures examination reveals abundant evidence of their volcanic origin
are in marked contrast to their apparent compositional maturity on the basis of textures, light mineral constituents, quartz charac-
(van Hattum, 2005; van Hattum et al., 2006). They are immature, ter, clay mineralogy, and zircon character and ages (Smyth, 2005;
typical first-cycle sandstones; grains are angular to subangular, Smyth et al., this volume). These quartz-rich rocks are not the
and there are few clasts that suggest recycling. The sandstones product of recycling of material derived from continental source
are poorly sorted and have a muddy matrix and very low poros- regions but are the result of explosive volcanic events such as the
ity, and they contain abundant euhedral and subhedral zircon eruption of crystal-rich magma, followed by air-fall sorting and
grains typical of first-cycle sandstones. Subrounded and (rare) subsequent epiclastic reworking (Cas and Wright, 1987; Walker,
rounded grains are less abundant. Tourmaline also occurs pre- 1972). In tropical settings, intense weathering rapidly causes the
dominantly as unabraded grains. The shapes and lack of abrasion breakdown of labile minerals, mineral aggregates, and lithic frag-
of zircons and tourmalines indicate that long-distance transport is ments. The redeposited sediment will be rich in resistant minerals
unlikely. Zircon and tourmaline typically make up >70% of the such as quartz and heavy minerals like zircon, and it has a higher
heavy mineral assemblages in the Crocker Fan sandstones. Their percentage of quartz per unit volume than the source material.
abundance indicates erosion from acid plutonic rocks, and their The high rates of weathering observed in tropical settings
shapes suggest a nearby source area. Apatite would normally also can also have a significant influence on the preservation of
be abundant in material derived from acid plutonic rocks, but it unconsolidated volcanic deposits. The potential for preserving
is commonly absent, and those grains present are typically pitted volcaniclastic deposits on steep terrestrial slopes is low unless
or partially dissolved. This suggests that the abundance of apatite they are rapidly overlain by lavas or buried under younger arc
Cenozoic arc processes in Indonesia 49

100 100
Strongly weathered
90 95 Inc
rea
sin
CIW (100)[Al2O3/(Al2O3+CaO+Na2O)]

80 gc
90 he
mi
ca
70
lm
85 atu
rity
60 80
HUMID

SiO2%
50 75

40 70

30 65
Fresh (source) rock
ARID
20 60

10 55
Strongly altered
0 50
0 10 20 30 40 50 60 70 80 90 100 0 5 10 15 20 25
CIA (100)[Al2O3/(Al2O3+CaO+Na2O+K2O)] Al2O3+K2O+Na2O

Figure 17. (A) Chemical index of weathering (CIW; Harnois, 1988) plotted against chemical index of alteration (CIA; Nesbitt and Young, 1984).
(B) Chemical maturity, expressed as a function of percentage of SiO2 and total percentage of Al2O3 + K2O + Na2O. After Suttner and Dutta (1986)
for Sabah sandstones from van Hattum (2005).

deposits. The breakdown and recycling of soft, nonwelded ashes inaccessible areas. An extremely large eruption, possibly on the
can be very rapid in terrestrial settings with intense precipita- scale of Toba, terminated the Eocene to early Miocene phase of
tion. Lahars are common hazards on many of Java’s volcanoes arc activity in the Southern Mountains of East Java (Smyth, 2005)
such as Merapi (Lavigne et al., 2000). Preservation is more likely at ca. 20 Ma, and other such events will surely be recognized in
if the volcanic material is deposited in a marine setting below the future. The tropical setting means that some features may be
the storm-wave base, but even this material may be apparently typical only of tropical arcs—for example, the rapidly enhanced
more mature than would be expected in a nontropical environ- maturity of sandstones and enrichment in quartz in volcano-
ment because of eruptive sorting of material in the ash cloud. genic debris deposited in arc basins. Nonetheless, in ancient arcs
The East Java sandstones are well sorted, quartz-rich, and clearly it is still necessary to consider the likely climatic setting before
deposited in shallow marine settings, but they were derived from interpreting provenance, both in terms of transport distances and
volcanic material produced by explosive eruptions that occurred source regions.
a short period before their deposition. We know of other quartz- Other features, such as the importance of extension in arcs,
rich sandstones in Indonesia that we and others have previously the influence of volcanic arc loading on sedimentary basin devel-
interpreted as having a continental provenance, and our experi- opment, the critical role of hinge movement in magmatism, and
ence of sandstones in Borneo and East Java now causes us to the weakness of the arc, have more general applicability. Certain
consider more carefully if some of these interpretations may have features of Indonesian arcs could also be of general relevance but
been wrong. Quartz character and heavy mineral studies (Smyth, could simply reflect particular differences between tectonic pro-
2005; Smyth et al., this volume) are of particular importance for cesses in SE Asia and those of other arc regions—for example,
reassessment. In older orogenic belts, similar misinterpretations the absence of a relationship between arc-trench gaps and dura-
could easily be made, especially if the ancient climate is not con- tion of magmatic arc activity, emplacement of ophiolites, and for-
sidered or known. mation of mélanges. The long history of subduction in the region
has resulted in a thin, warm, and weak lithosphere, and although
CONCLUSIONS: LESSONS FOR ANCIENT ARCS the region is not typical of continental crust it does share many
features with other subduction zone backarcs (Hyndman et al.,
The Indonesian arcs differ in many ways from those com- 2005). The behavior of Indonesian arcs may be more representa-
monly portrayed in many arc models. Are the Indonesian arcs typi- tive of arcs in the upper plate of long-lived subduction margins
cal, or are they one member of a spectrum of arcs? The absence of in which the lithosphere becomes unusually responsive to small
well-documented major eruptions in the Cenozoic stratigraphic changes in plate boundary forces.
record of Indonesian arcs is likely to reflect their tropical setting Indonesian arcs appear to have had relatively short lives
and the lack of detailed studies in difficult terrain and relatively in comparison with many ancient arcs, but this is probably a
50 Hall and Smyth

A Q B
CRATON Continental Block
INTERIOR
Recycled Orogen
TRANSITIONAL Magmatic Arc
CONTINENTAL
Q - Total free quartz
F - Feldspar
L - Total lithic
RECYCLED fragments
OROGEN
T

DISSECTED ARC
LIF
UP
NT
ME

TRANSITIONAL
SE

ARC UNDISSECTED
BA

ARC
1mm
F L

C D

200 µm 200 µm

Figure 18. (A) Quartz sandstones from East Java plotted on a QFL ternary diagram, which could be interpreted as indicating a reworked conti-
nental orogenic provenance. The sandstones are actually first-cycle products of rapid reworking of volcanogenic material. Examples of volcanic
quartz grains from the Jaten Formation, Pacitan, East Java. (B) Bipyramidal quartz grain. (C) Scanning electron microscope (SEM) image of a
quartz shard. (D) SEM image of rounded melt embayments in quartz.

reflection of their young age and the greater ease with which time and may ultimately leave little trace in the stratigraphic
events of different arcs can be resolved in the relatively recent record. Even in the young arcs of Indonesia, dating of events is
past. Pre-Cenozoic arcs with lives of tens of millions of years not good, and more and better dating is required to understand
may include several phases of arc activity, and several different and interpret arc processes. We also need to know more about
arcs. The complexity of tectonic evolution in the region shows the deep structure of arcs. For example, very little is known
that a great deal can happen in a 45 m.y. time span, and much about the crust between the arc and trench in any Indonesian
of this complexity could be missed in the study of ancient arcs. arc, and almost nothing is known of the thickness and character
This could be an explanation of the apparently compressional of the crust beneath the volcanic arc. Most arcs in the western
arc; short-lived contractional events in older arcs may appear Pacific are similarly poorly known. At the very least, the record
to be contemporaneous with arc magmatism when in fact they of the Indonesian arcs shows that some features of arc models
punctuate arc activity. The history of the Neogene Halmahera need to be questioned and that our understanding of tectonic
Arc shows that an arc can be formed and destroyed in a short processes in volcanic arcs needs to be improved.
Cenozoic arc processes in Indonesia 51

ACKNOWLEDGMENTS Bijwaard, H., Spakman, W., and Engdahl, E.R., 1998, Closing the gap between
regional and global travel time tomography: Journal of Geophysical
Research, v. 103, p. 30,055–30,078, doi: 10.1029/98JB02467.
Our knowledge of Indonesian arcs has been acquired through Bock, Y., Prawirodirdjo, L., Genrich, J.F., Stevens, C.W., McCaffrey, R., Sub-
many projects carried out by the SE Asia Research Group at arya, C., Puntodewo, S.S.O., and Calais, E., 2003, Crustal motion in Indo-
London University and Royal Holloway by many different nesia from Global Positioning System measurements: Journal of Geo-
physical Research, v. 108, doi:10.1029/2001JB000324.
people. We thank all of them. Our work has been supported at Busby, C.J., and Ingersoll, R.V., eds., 1995, Tectonics of Sedimentary Basins:
different times by the University of London Central Research Cambridge, Massachusetts, Blackwell Science, 579 p.
Fund, the Natural Environment Research Council (NERC), and Calvert, S.J., 2000, The Cenozoic geology of the Lariang and Karama regions,
Western Sulawesi, Indonesia [Ph.D. thesis]: University of London, 353 p.
the Royal Society, but mainly by a consortium of oil compa- Cardwell, R.K., and Isacks, B.L., 1978, Geometry of the subducted lithosphere
nies whose membership has changed with time. We have been beneath the Banda Sea in eastern Indonesia from seismicity and fault
fortunate with help and support from colleagues in Indonesia plane solutions: Journal of Geophysical Research, v. 83, p. 2825–2838.
Cardwell, R.K., Isacks, B.L., and Karig, D.E., 1980, The spatial distribution
at the Geological Research and Development Centre Bandung of earthquakes, focal mechanism solutions and subducted lithosphere in
(now the Geological Survey of Indonesia), Lemigas, Indone- the Philippines and northeast Indonesian islands, in Hayes, D.E., ed., The
sian Institute of Sciences (LIPI), and Institut Teknologi Band- Tectonic and Geologic Evolution of Southeast Asian Seas and Islands:
American Geophysical Union, Geophysical Monograph 23, p. 1–35.
ung. We thank Tim Charlton, Ron Harris, and Dave Scholl for Cas, R.A.F., and Wright, J.V., 1987, Volcanic Successions: London, Allen and
helpful comments on an earlier version of the paper, and Wim Unwin, 528 p.
Spakman for help in producing Figure 7. Charlton, T.R., 2000, Tertiary evolution of the Eastern Indonesia collision com-
plex: Journal of Asian Earth Sciences, v. 18, p. 603–631, doi: 10.1016/
S1367-9120(99)00049-8.
REFERENCES CITED Chase, C.G., 1978, Extension behind island arcs and motions relative to hot-
spots: Journal of Geophysical Research, v. 83, p. 5385–5387.
Abbott, M.J., and Chamalaun, F.H., 1981, Geochronology of some Banda Arc Chesner, C.A., and Rose, W.I., 1991, Stratigraphy of the Toba tuffs and the
volcanics, in Barber, A.J., and Wiryosujono, S., eds., The Geology and evolution of the Toba caldera complex, Sumatra, Indonesia: Bulletin of
Tectonics of Eastern Indonesia: Bandung, Indonesia, Geological Research Volcanology, v. 53, p. 343–356, doi: 10.1007/BF00280226.
and Development Centre Special Publication 2, p. 253–268. Chesner, C.A., Rose, W.I., Deino, A., Drake, R., and Westgate, J.A., 1991, Erup-
Ali, J.R., and Hall, R., 1995, Evolution of the boundary between the Philippine tive history of earth’s largest Quaternary caldera (Toba, Indonesia) clarified:
Sea Plate and Australia: Palaeomagnetic evidence from eastern Indonesia: Geology, v. 19, p. 200–203, doi: 10.1130/0091-7613(1991)019<0200:
Tectonophysics, v. 251, p. 251–275, doi: 10.1016/0040-1951(95)00029-1. EHOESL>2.3.CO;2.
Ali, J.R., Hall, R., and Baker, S.J., 2001, Palaeomagnetic data from a Meso- Cloke, I.R., Milsom, J., and Blundell, D.J.B., 1999, Implications of gravity data
zoic Philippine Sea Plate ophiolite on Obi Island, Eastern Indonesia: from East Kalimantan and the Makassar Straits: A solution to the origin
Journal of Asian Earth Sciences, v. 19, p. 535–546, doi: 10.1016/S1367- of the Makassar Straits?: Journal of Asian Earth Sciences, v. 17, p. 61–78,
9120(00)00053-5. doi: 10.1016/S0743-9547(98)00056-7.
Ambrose, S.H., 2003, Did the super-eruption of Toba cause a human population Curray, J.R., 1994, Sediment volume and mass beneath the Bay of Bengal: Earth
bottleneck? Reply to Gathorne-Hardy and Harcourt-Smith: Journal of and Planetary Science Letters, v. 125, p. 371–383, doi: 10.1016/0012-
Human Evolution, v. 45, p. 231–237, doi: 10.1016/j.jhevol.2003.08.001. 821X(94)90227-5.
Andrews, D.J., and Sleep, N.H., 1974, Numerical modelling of tectonic flow Dewey, J.F., 1980, Episodicity, sequence and style at convergent plate boundar-
behind island arcs: Geophysical Journal of the Royal Astronomical Soci- ies: Geological Association of Canada Special Paper 20, p. 553–573.
ety, v. 38, p. 237–251. Dewey, J.F., and Bird, J.M., 1970, Mountain belts and the new global tectonics:
Artemieva, I.M., and Mooney, W.D., 1999, Mantle heat flow in stable conti- Journal of Geophysical Research, v. 75, p. 2625–2647.
nental regions: A global study: Eos (Transactions, American Geophysical Dickinson, W.R., 1973, Widths of modern arc–trench gaps proportional to past
Union), v. 80, p. F967. duration of igneous activity in associated magmatic arcs: Journal of Geo-
Artemieva, I.M., and Mooney, W.D., 2001, Thermal thickness and evolution physical Research, v. 78, p. 3395–3417.
of Precambrian lithosphere: A global study: Journal of Geophysical Dickinson, W.R., 1974, Plate tectonics and sedimentation, in Dickinson, W.R.,
Research, v. 106, p. 16,387–16,414, doi: 10.1029/2000JB900439. ed., Tectonics and Sedimentation: Society of Economic Paleontologists
Audley-Charles, M.G., 2004, Ocean trench blocked and obliterated by Banda and Mineralogists Special Publication 22, p. 1–27.
forearc collision with Australian proximal continental slope: Tectono- Dickinson, W.R., 1977, Tectono-stratigraphic evolution of subduction con-
physics, v. 389, p. 65–79, doi: 10.1016/j.tecto.2004.07.048. trolled sedimentary assemblages, in Talwani, M., and Pitman, W.C., eds.,
Baker, S., and Malaihollo, J., 1996, Dating of Neogene igneous rocks in the Island Arcs, Deep Sea Trenches, and Back-Arc Basins: Maurice Ewing
Halmahera region: Arc initiation and development, in Hall, R., and Series 1: Washington, American Geophysical Union 1, p. 33–40.
Blundell, D.J., eds., Tectonic Evolution of SE Asia: Geological Society Dickinson, W.R., 1995, Forearc basins, in Busby, C.J., and Ingersoll, R.V., eds.,
[London] Special Publication 106, p. 499–509. Tectonics of Sedimentary Basins: Cambridge, Massachusetts, Blackwell
Baker, S.J., 1997, Isotopic dating and island arc development in the Halmahera Science, p. 221–262.
Region, Eastern Indonesia [Ph.D. thesis]: University of London, 331 p. Dickinson, W.R., and Seely, D.R., 1979, Structure and stratigraphy of forearc
Barber, A.J., Tjokrosapoetro, S., and Charlton, T.R., 1986, Mud volcanoes, regions: American Association of Petroleum Geologists Bulletin, v. 63,
shale diapirs, wrench faults, and melanges in accretionary complexes, p. 2–31.
Eastern Indonesia: American Association of Petroleum Geologists Bul- Edwards, C.M.H., Menzies, M.A., Thirlwall, M.F., Morris, J.D., Leeman, W.P.,
letin, v. 70, p. 1729–1741. and Harmon, R.S., 1994, The transition to potassic alkaline volcanism in
Ben-Avraham, Z., and Emery, K.O., 1973, Structural framework of Sunda island arcs: The Ringgit-Besar Complex, east Java, Indonesia: Journal of
Shelf: American Association of Petroleum Geologists Bulletin, v. 57, Petrology, v. 35, p. 1557–1595.
p. 2323–2366. Elsassar, W.M., 1971, Sea-floor spreading as thermal convection: Journal of
Bergman, S.C., Coffield, D.Q., Talbot, J.P., and Garrard, R.A., 1996, Tertiary Geophysical Research, v. 76, p. 1101–1112.
tectonic and magmatic evolution of western Sulawesi and the Makassar Engdahl, E.R., van der Hilst, R., and Buland, R., 1998, Global teleseismic
Strait, Indonesia: Evidence for a Miocene continent–continent collision, earthquake relocation with improved travel times and procedures for
in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE Asia: Geo- depth determination: Bulletin of the Seismological Society of America,
logical Society [London] Special Publication 106, p. 391–429. v. 88, p. 722–743.
52 Hall and Smyth

England, P., Engdahl, R., and Thatcher, W., 2004, Systematic variation in the Harington, C.R., ed., 1992, The Year without a Summer: World Climate in
depths of slabs beneath arc volcanoes: Geophysical Journal International, 1816: Ottawa, Canadian Museum of Nature, 576 p.
v. 156, p. 377–408, doi: 10.1111/j.1365-246X.2003.02132.x. Harnois, L., 1988, The CIW index: A new chemical index of weathering: Sedi-
Evans, C.A., Hawkins, J.W., and Moore, G.F., 1983, Petrology and geochem- mentary Geology, v. 55, p. 319–322, doi: 10.1016/0037-0738(88)90137-6.
istry of ophiolitic and associated volcanic rocks on the Talaud Islands, Harris, R.A., Sawyer, R.K., and Audley-Charles, M.G., 1998, Collisional
Molucca Sea Collision Zone, Northeast Indonesia, in Hilde, T.W.C., and melange development: Geologic associations of active melange-forming
Uyeda, S., eds., Geodynamics of the Western Pacific–Indonesian Region: processes with exhumed melange facies in the western Banda orogen,
American Geophysical Union and Geological Society of America, Geo- Indonesia: Tectonics, v. 17, p. 458–479, doi: 10.1029/97TC03083.
dynamic Ser. 11, p. 159–172. Hazebroek, H.P., and Tan, D.N.K., 1993, Tertiary tectonic evolution of the NW
Fraser, T.H., Jackson, B.A., Barber, P.M., Baillie, P.W., and Myers, K., 2003, Sabah continental margin: Geological Society of Malaysia Bulletin, v. 33,
The West Sulawesi Fold Belt and other new plays within the North Makas- p. 195–210.
sar Straits—A prospectivity review: Proceedings of Indonesian Petroleum Honthaas, C., Maury, R.C., Priadi, B., Bellon, H., and Cotten, J., 1999, The
Association 29th Annual Convention, Jakarta, p. 429–450. Plio–Quaternary Ambon arc, eastern Indonesia: Tectonophysics, v. 301,
Fuller, M., Ali, J.R., Moss, S.J., Frost, G.M., Richter, B., and Mahfi, A., 1999, p. 261–281, doi: 10.1016/S0040-1951(98)00227-3.
Paleomagnetism of Borneo: Journal of Asian Earth Sciences, v. 17, p. 3– Hutchison, C.S., 1989, Geological Evolution of South-East Asia, Oxford
24, doi: 10.1016/S0743-9547(98)00057-9. Monographs on Geology and Geophysics: Oxford, UK, Clarendon Press,
Funiciello, F., Faccenna, C., Giardini, D., and Regenauer-Lieb, K., 2003, 376 p.
Dynamics of retreating slabs: 2. Insights from three-dimensional labora- Hutchison, C.S., 1996, The ‘Rajang Accretionary Prism’ and ‘Lupar Line’
tory experiments: Journal of Geophysical Research, v. 108, p. 2207, doi: problem of Borneo, in Hall, R., and Blundell, D.J., eds., Tectonic Evolu-
10.1029/2001JB000896. tion of SE Asia: Geological Society [London] Special Publication 106,
Furukawa, Y., 1993, Depth of the decoupling plate interface and thermal p. 247–261.
structure under arcs: Journal of Geophysical Research, v. 98, p. 20,005– Hutchison, C.S., Bergman, S.C., Swauger, D.A., and Graves, J.E., 2000, A
20,013. Miocene collisional belt in north Borneo: Uplift mechanism and isostatic
Garwin, S., Hall, R., and Watanabe, Y., 2005, Tectonic setting, geology and gold adjustment quantified by thermochronology: Geological Society [Lon-
and copper mineralization in Cenozoic magmatic arcs of Southeast Asia don] Journal, v. 157, p. 783–793.
and the West Pacific: Economic Geology and Bulletin of the Society of Hyndman, R.D., Currie, C.A., and Mazzotti, S., 2005, Subduction zone back-
Economic Geologists, v. 100, p. 891–930. arcs, mobile belts, and orogenic heat: GSA Today, v. 15, no. 2, p. 4–9.
Gobbett, D.J., and Hutchison, C.S., eds., 1973, Geology of the Malay Pen- Iwamori, H., 1997, Heat sources and melting in subduction zones: Journal of Geo-
insula (West Malaysia and Singapore): New York, Wiley-Interscience, physical Research, v. 102, p. 14,803–14,820, doi: 10.1029/97JB01036.
438 p. Karig, D.E., 1971, Origin and development of marginal basins in the Western
Guntoro, A., 1999, The formation of the Makassar Strait and the separation Pacific: Journal of Geophysical Research, v. 76, p. 2542–2561.
between SE Kalimantan and SW Sulawesi: Journal of Asian Earth Sci- Karig, D.E., and Ingle, J.C., Jr., et al., eds., 1975, Initial Reports of the Deep Sea
ences, v. 17, p. 79–98, doi: 10.1016/S0743-9547(98)00037-3. Drilling Project, Leg 31: Washington, U.S. Government Printing Office,
Hall, R., 1996, Reconstructing Cenozoic SE Asia, in Hall, R., and Blundell, 927 p., doi: 10.2973/dsdp.proc.31.1975.
D.J., eds., Tectonic Evolution of SE Asia: Geological Society [London] Karig, D.E., and Sharman, G.F., III, 1975, Subduction and accretion in
Special Publication 106, p. 153–184. trenches: Geological Society of America Bulletin, v. 86, p. 377–389, doi:
Hall, R., 2000, Neogene history of collision in the Halmahera region, Indone- 10.1130/0016-7606(1975)86<377:SAAIT>2.0.CO;2.
sia, in Proceedings of the Indonesian Petroleum Association 27th Annual Katili, J.A., 1975, Volcanism and plate tectonics in the Indonesian island arcs:
Convention, Jakarta, p. 487–493. Tectonophysics, v. 26, p. 165–188, doi: 10.1016/0040-1951(75)90088-8.
Hall, R., 2002, Cenozoic geological and plate tectonic evolution of SE Asia and Kay, S.M., Godoy, E., and Kurtz, A., 2005, Episodic arc migration, crustal
the SW Pacific: Computer-based reconstructions, model and animations: thickening, subduction erosion, and magmatism in the south-central
Journal of Asian Earth Sciences, v. 20, p. 353–434, doi: 10.1016/S1367- Andes: Geological Society of America Bulletin, v. 117, p. 67–88, doi:
9120(01)00069-4. 10.1130/B25431.1.
Hall, R., and Morley, C.K., 2004, Sundaland Basins, in Clift, P., et al., eds., Con- Kay, S.M., Maksaev, V., Moscoso, R., Mpodozis, C., and Nasi, C., 1987, Prob-
tinent–Ocean Interactions within the East Asian Marginal Seas: American ing the evolving Andean lithosphere: Mid–late Tertiary magmatism in
Geophysical Union Geophysical Monograph 149, p. 55–85. Chile (29°–30°30′S) over the modern zone of subhorizontal subduction:
Hall, R., and Wilson, M.E.J., 2000, Neogene sutures in eastern Indonesia: Journal of Geophysical Research, v. 92, p. 6173–6189.
Journal of Asian Earth Sciences, v. 18, p. 781–808, doi: 10.1016/S1367- Kenyon, C.S., and Beddoes, L.R., eds., 1977, Geothermal Gradient Map of
9120(00)00040-7. Southeast Asia: Jakarta, Indonesia, South East Asia Petroleum Explora-
Hall, R., Audley-Charles, M.G., Banner, F.T., Hidayat, S., and Tobing, S.L., tion Society and Indonesian Petroleum Association, 50 p.
1988a, Basement rocks of the Halmahera region, eastern Indonesia: A Klein, G. deV., and Kobayashi, K., et al., editors, 1980, Initial Reports of the
Late Cretaceous–Early Tertiary arc and fore-arc: Geological Society Deep Sea Drilling Project, Leg 58: Washington, U.S. Government Print-
[London] Journal, v. 145, p. 65–84. ing Office, 1022 p., doi: 10.2973/dsdp.proc.58.1980.
Hall, R., Audley-Charles, M.G., Banner, F.T., Hidayat, S., and Tobing, S.L., Kopp, H., Klaeschen, D., Flueh, E.R., Bialas, J., and Reichert, C., 2002, Crustal
1988b, Late Paleogene–Quaternary geology of Halmahera, eastern Indo- structure of the Java margin from seismic wide-angle and multichannel
nesia: Initiation of a volcanic island arc: Geological Society [London] reflection data: Journal of Geophysical Research, v. 107, doi:10.1029/
Journal, v. 145, p. 577–590. 2000JB000095.
Hall, R., Ali, J.R., Anderson, C.D., and Baker, S.J., 1995, Origin and motion Kopp, H., Flueh, E.R., Peterson, C.J., Weinrebe, W., Wittwer, A., and Meramax
history of the Philippine Sea Plate: Tectonophysics, v. 251, p. 229–250, Scientists, 2006, The Java margin revisited: Evidence for subduction ero-
doi: 10.1016/0040-1951(95)00038-0. sion off Java: Earth and Planetary Science Letters, v. 242, p. 130–142, doi:
Hamilton, W., 1977, Subduction in the Indonesian region, in Talwani, M., 10.1016/j.epsl.2005.11.036.
and Pitman, W.C., eds., Island Arcs, Deep Sea Trenches, and Back-Arc Kreemer, C., Holt, W.E., Goes, S., and Govers, R., 2000, Active deforma-
Basins: Maurice Ewing Series 1: Washington, American Geophysical tion in eastern Indonesia and the Philippines from GPS and seismic-
Union 1, p. 15–31. ity data: Journal of Geophysical Research, v. 105, p. 663–680, doi:
Hamilton, W., 1979, Tectonics of the Indonesian region: U.S. Geological Sur- 10.1029/1999JB900356.
vey Professional Paper 1078, 345 p. Lavigne, F., Thouret, J.C., Voight, B., Suwa, H., and Sumaryono, A., 2000,
Hamilton, W.B., 1988, Plate tectonics and island arcs: Geological Soci- Lahars at Merapi volcano, Central Java: An overview: Journal of Vol-
ety of America Bulletin, v. 100, p. 1503–1527, doi: 10.1130/0016- canology and Geothermal Research, v. 100, p. 423–456, doi: 10.1016/
7606(1988)100<1503:PTAIA>2.3.CO;2. S0377-0273(00)00150-5.
Hamilton, W.B., 1995, Subduction and magmatism, in Smellie, J.L., ed., Volca- Lebedev, S., and Nolet, G., 2003, Upper mantle beneath Southeast Asia
nism Associated with Subduction at Consuming Plate Margins: Geologi- from S velocity tomography: Journal of Geophysical Research, v. 108,
cal Society [London] Special Publication 81, p. 3–28. doi:10.1029/2000JB000073.
Cenozoic arc processes in Indonesia 53

Liew, T.C., and McCulloch, M.T., 1985, Genesis of granitoid batholiths of Pen- Indonesian Petroleum Association 30th Annual Convention, Jakarta,
insular Malaysia and implications for models of crustal evolution: Evi- p. 483–497.
dence from a Nd–Sr isotopic zircon study: Geochimica et Cosmochimica Oncken, O., Chong, G., Franz, G., Giese, P., Götze, H.-J., Ramos, V.A., Strecker,
Acta, v. 49, p. 587–600, doi: 10.1016/0016-7037(85)90050-X. M.R., and Wigger, P., eds., 2006, The Andes—Active Subduction Orog-
Liew, T.C., and Page, R.W., 1985, U–Pb zircon dating of granitoid plutons from eny: Berlin, Springer, 570 p.
the West Coast of Peninsular Malaysia: Geological Society [London] Oxburgh, E.R., and Turcotte, D.L., 1970, Thermal structure of island arcs:
Journal, v. 142, p. 515–526. Geological Society of America Bulletin, v. 81, p. 1665–1688, doi:
Macpherson, C.G., 1994, New analytical approaches to the oxygen and carbon 10.1130/0016-7606(1970)81[1665:TSOIA]2.0.CO;2.
stable isotope geochemistry of some subduction-related lavas [Ph.D. the- Parkinson, C.D., Miyazaki, K., Wakita, K., Barber, A.J., and Carswell, D.A.,
sis]: University of London, 206 p. 1998, An overview and tectonic synthesis of the pre-Tertiary very-high-
Macpherson, C.G., and Hall, R., 1999, Tectonic controls of geochemical evolu- pressure metamorphic and associated rocks of Java, Sulawesi and Kali-
tion in arc magmatism of SE Asia, in Proceedings of 4th PACRIM Con- mantan, Indonesia: The Island Arc, v. 7, p. 184–200, doi: 10.1046/j.1440-
gress: Bali, Indonesia, Australian Institute of Mining and Metallurgy, 1738.1998.00184.x.
p. 359–368. Pearce, J.A., and Peate, D.W., 1995, Tectonic implications of the composition
Macpherson, C.G., and Hall, R., 2002, Timing and tectonic controls in the of volcanic arc magmas: Annual Review of Earth and Planetary Sciences,
evolving orogen of SE Asia and the western Pacific and some implications v. 23, p. 251–285, doi: 10.1146/annurev.ea.23.050195.001343.
for ore generation, in Blundell, D.J., et al., eds., The Timing and Location Pollack, H.N., Hurter, S., and Johnson, J.R., 1990, The new global heat flow
of Major Ore Deposits in an Evolving Orogen: Geological Society [Lon- data compilation: Eos (Transactions, American Geophysical Union),
don] Special Publication 204, p. 49–67. v. 71, p. 1604.
Macpherson, C.G., Forde, E.J., Hall, R., and Thirlwall, M.F., 2003, Geochemi- Pollack, H.N., Hurter, S.J., and Johnson, J.R., 1993, Heat flow from the Earth’s
cal evolution of magmatism in an arc–arc collision: The Halmahera and interior: Analysis of the global data set: Reviews of Geophysics, v. 31,
Sangihe arcs, eastern Indonesia, in Larter, R.D., and Leat, P.T., eds., Intra- p. 267–280, doi: 10.1029/93RG01249.
Oceanic Subduction Systems: Tectonic and Magmatic Processes: Geo- Polvé, M., Maury, R.C., Bellon, H., Rangin, C., Priadi, B., Yuwono, S., Joron,
logical Society [London] Special Publication 219, p. 207–220. J.L., and Soeria-Atmadja, R., 1997, Magmatic evolution of Sulawesi
Mason, B.G., Pyle, D.M., and Oppenheimer, C., 2004, The size and frequency (Indonesia): Constraints on the Cenozoic geodynamic history of the Sun-
of the largest explosive eruptions on Earth: Bulletin of Volcanology, v. 66, daland active margin: Tectonophysics, v. 272, p. 69–92, doi: 10.1016/
p. 735–748, doi: 10.1007/s00445-004-0355-9. S0040-1951(96)00276-4.
McCaffrey, R., 1983, Seismic wave propagation beneath the Molucca Sea Puspita, S.D., Hall, R., and Elders, C.F., 2005, Structural styles of the offshore
arc–arc collision zone, Indonesia: Tectonophysics, v. 96, p. 45–57, doi: West Sulawesi fold belt, North Makassar Straits, Indonesia: Proceedings
10.1016/0040-1951(83)90243-3. of Indonesian Petroleum Association 30th Annual Convention, Jakarta,
McCaffrey, R., 1991, Earthquakes and ophiolite emplacement in the Molucca p. 519–542.
Sea collision zone, Indonesia: Tectonics, v. 10, p. 433–454. Rampino, M.R., and Self, S., 1992, Volcanic winter and accelerated glacia-
McCaffrey, R., Silver, E.A., and Raitt, R.W., 1980, Crustal structure of the tion following the Toba super-eruption: Nature, v. 359, p. 50–52, doi:
Molucca Sea collision zone, Indonesia, in Hayes, D.E., ed., The Tectonic 10.1038/359050a0.
and Geologic Evolution of Southeast Asian Seas and Islands: American Rampino, M.R., and Self, S., 1993a, Bottleneck in human evolution and the
Geophysical Union, Geophysical Monograph Ser. 23, p. 161–178. Toba eruption: Science, v. 262, p. 1955, doi: 10.1126/science.8266085.
McCourt, W.J., Crow, M.J., Cobbing, E.J., and Amin, T.C., 1996, Mesozoic and Rampino, M.R., and Self, S., 1993b, Climate–volcanism feedback and the Toba
Cenozoic plutonic evolution of SE Asia: Evidence from Sumatra, Indone- eruption of ca. 74,000 years ago: Quaternary Research, v. 40, p. 269–280,
sia, in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE Asia: doi: 10.1006/qres.1993.1081.
Geological Society [London] Special Publication 106, p. 321–335. Ranero, C.R., and von Huene, R., 2000, Subduction erosion along the Mid-
Metcalfe, I., 1996, Pre-Cretaceous evolution of SE Asian Terranes, in Hall, R., dle America convergent margin: Nature, v. 404, p. 748–752, doi:
and Blundell, D.J., eds., Tectonic Evolution of SE Asia: Geological Soci- 10.1038/35008046.
ety [London] Special Publication 106, p. 97–122. Rangin, C., Spakman, W., Pubellier, M., and Bijwaard, H., 1999, Tomographic
Mikada, H., Moore, G.F., Taira, A., Becker, K., Moore, J.C., and Klaus, A., and geological constraints on subduction along the eastern Sundaland
eds., 2005, Proceedings of Ocean Drilling Program, Scientific Results, continental margin (South-East Asia): Bulletin de la Société Géologique
190/196 [Online]: http://www-odp.tamu.edu/publications/190196SR/ de France, v. 170, p. 775–788.
190196sr.htm (November 2006). Ritsema, J., and van Heijst, H.-J., 2000, Seismic imaging of structural heteroge-
Miyazaki, K., Sopaheluwakan, J., Zulkarnain, I., and Wakita, K., 1998, A jadeite- neity in the Earth’s mantle: Evidence for large-scale mantle flow: Science
quartz-glaucophane rock from Karangsambung, central Java, Indonesia: Progress, v. 83, p. 243–259.
The Island Arc, v. 7, p. 223–230, doi: 10.1046/j.1440-1738.1998.00164.x. Rowland, A., and Davies, J.H., 1999, Buoyancy rather than rheology con-
Moore, G.F., and Karig, D.E., 1980, Structural geology of Nias island, Indo- trols the thickness of the overriding mechanical lithosphere at subduc-
nesia: Implications for subduction zone tectonics: American Journal of tion zones: Geophysical Research Letters, v. 26, p. 3037–3040, doi:
Science, v. 280, p. 193–223. 10.1029/1999GL005347.
Moore, G.F., Kadarisman, D.K., Evans, C.A., and Hawkins, J.W., 1981, Geol- Rutherford, K.J., and Qureshi, M.K., eds., 1981, Geothermal gradient map
ogy of the Talaud Islands, Molucca Sea Collision Zone, northeast Indone- of Southeast Asia (2nd edition): Jakarta, Indonesia, South East Asia
sia: Journal of Structural Geology, v. 3, p. 467–475, doi: 10.1016/0191- Petroleum Exploration Society and Indonesian Petroleum Association,
8141(81)90046-8. 51 p.
Morton, A.C., 1984, Stability of detrital heavy minerals in Tertiary sandstones Samuel, M.A., 1994, The structural and stratigraphic evolution of islands at the
from the North Sea basin: Clay Minerals, v. 19, p. 287–308, doi: 10.1180/ active margin of the Sumatran Forearc, Indonesia [Ph.D. thesis]: Univer-
claymin.1984.019.3.04. sity of London, 345 p.
Nagao, T., and Uyeda, S., 1995, Heat flow distribution in southeast Asia with Samuel, M.A., and Harbury, N.A., 1996, The Mentawai fault zone and defor-
consideration of volcanic heat: Tectonophysics, v. 251, p. 153–159, doi: mation of the Sumatran forearc in the Nias area, in Hall, R., and Blundell,
10.1016/0040-1951(95)00084-4. D.J., eds., Tectonic Evolution of SE Asia: Geological Society [London]
Nesbitt, H.W., and Young, G.M., 1984, Prediction of some weathering trends Special Publication 106, p. 337–351.
of plutonic and volcanic rocks based on thermodynamic and kinetic con- Samuel, M.A., Harbury, N.A., Jones, M.E., and Matthews, S.J., 1995, Inver-
siderations: Geochimica et Cosmochimica Acta, v. 48, p. 1523–1534, doi: sion of an outer-arc ridge: The Sumatran Forearc, Indonesia, in Buchanan,
10.1016/0016-7037(84)90408-3. J.G., and Buchanan, P.G., eds., Basin Inversion: Geological Society [Lon-
Nichols, G.J., and Hall, R., 1991, Basin formation and Neogene sedimentation don] Special Publication 88, p. 473–492.
in a backarc setting, Halmahera, eastern Indonesia: Marine and Petroleum Samuel, M.A., Harbury, N.A., Bakri, A., Banner, F.T., and Hartono, L., 1997,
Geology, v. 8, p. 50–61, doi: 10.1016/0264-8172(91)90044-2. A new stratigraphy for the islands of the Sumatran Forearc, Indonesia:
Nur’Aini., S., Hall, R., and Elders, C.F., 2005, Basement architecture and Journal of Asian Earth Sciences, v. 15, p. 339–380, doi: 10.1016/S0743-
sedimentary fills of the North Makassar Straits basin: Proceedings of 9547(97)87720-3.
54 Hall and Smyth

Sandwell, D.T., and Smith, W.H.F., 1997, Marine gravity anomaly from Geosat Tjia, H.D., 1996, Sea-level changes in the tectonically stable Malay–Thai pen-
and ERS 1 satellite altimetry: Journal of Geophysical Research, v. 102, insula: Quaternary International, v. 31, p. 95–101, doi: 10.1016/1040-
p. 10,039–10,054, doi: 10.1029/96JB03223. 6182(95)00025-E.
Schlüter, H.U., Gaedicke, C., Roeser, H.A., Schreckenberger, B., Meyer, H., Tongkul, F., 1991, Tectonic evolution of Sabah, Malaysia: Journal of South-
Reichert, C., Djajadihardja, Y., and Prexl, A., 2002, Tectonic features of east Asian Earth Sciences, v. 6, p. 395–406, doi: 10.1016/0743-
the southern Sumatra–western Java forearc of Indonesia: Tectonics, v. 21, 9547(91)90084-B.
p. 1047, doi: 10.1029/2001TC901048, doi: 10.1029/2001TC901048. Umbgrove, J.H.F., 1938, Geological history of the East Indies: American Asso-
Scholl, D., von Huene, R., Vallier, T.L., and Howell, D.G., 1980, Sedimentary ciation of Petroleum Geologists Bulletin, v. 22, p. 1–70.
masses and concepts about tectonic processes at underthrust ocean mar- van Bemmelen, R.W., 1949, The Geology of Indonesia: Nijhoff, The Hague,
gins: Geology, v. 8, p. 564–568, doi: 10.1130/0091-7613(1980)8<564: Government Printing Office, 732 p.
SMACAT>2.0.CO;2. van Hattum, M.W.A., 2005, Provenance of Cenozoic sedimentary rocks of
Scholl, D., Vallier, T.L., and Stevenson, A.J., 1986, Terrane accretion, produc- northern Borneo [Ph.D. thesis]: University of London, 457 p.
tion and continental growth: A perspective based on the origin and tec- van Hattum, M.W.A., Hall, R., and Nichols, G.J., 2003, Provenance of north-
tonic fate of the Aleutian–Bering Sea region: Geology, v. 14, p. 43–47, ern Borneo sediments: Proceedings of Indonesian Petroleum Association
doi: 10.1130/0091-7613(1986)14<43:TAPACG>2.0.CO;2. 29th Annual Convention, Jakarta, p. 305–319.
Shemenda, A.I., 1994, Subduction: Insights from Physical Modelling: Norwell, van Hattum, M.W.A., Hall, R., Pickard, A.L., and Nichols, G.J., 2006, SE Asian
Massachusetts, Kluwer Academic Publishers, 215 p. sediments not from Asia: Provenance and geochronology of North Bor-
Silver, E.A., and Moore, J.C., 1978, The Molucca Sea collision zone, Indone- neo sandstones: Geology, v. 34, p. 589–592, doi: 10.1130/G21939.1.
sia: Journal of Geophysical Research, v. 83, p. 1681–1691. Vannucchi, P., Galeotti, S., Clift, P.D., Ranero, C.R., and von Huene, R., 2004,
Situmorang, B., 1982, The formation and evolution of the Makassar basin, Long-term subduction-erosion along the Guatemalan margin of the Mid-
Indonesia [Ph.D. thesis]: University of London, 313 p. dle America trench: Geology, v. 32, p. 617–620, doi: 10.1130/G20422.1.
Smithsonian, 2006, Global volcanism program: http://www.volcano.si.edu/ von Huene, R., and Scholl, D., 1991, Observations at convergent margins con-
(March 2006). cerning sediment subduction, subduction erosion, and even the growth of
Smyth, H.R., 2005, Eocene to Miocene basin history and volcanic history in the continental crust: Reviews of Geophysics, v. 29, p. 279–316.
East Java, Indonesia [Ph.D. thesis]: University of London, p. 476. von Huene, R., Ranero, C.S., and Vannucchi, P., 2004, Generic model of sub-
Smyth, H.R., Hall, R., and Nichols, G.J., 2008, this volume, Cenozoic volcanic duction erosion: Geology, v. 32, p. 913–916, doi: 10.1130/G20563.1.
arc history of East Java, Indonesia: The stratigraphic record of eruptions Wakita, K., 2000, Cretaceous accretionary–collision complexes in central Indo-
on an active continental margin, in Draut, A.E., et al., eds., Lessons from nesia: Journal of Asian Earth Sciences, v. 18, p. 739–749, doi: 10.1016/
the Stratigraphic Record in Arc Collision Zones: Geological Society of S1367-9120(00)00020-1.
America Special Publication 436, doi: 10.1130/2008.2436(10). Walker, G.P.L., 1972, Crystal concentrations in ignimbrites: Contributions to
Stow, D.A.V., Cochran, J.R., and ODP Leg 116 Shipboard Scientific Party, Mineralogy and Petrology, v. 36, p. 135–146, doi: 10.1007/BF00371184.
1989, The Bengal Fan: Some preliminary results from ODP drilling: Geo- Wallace, A.R., 1869, The Malay Archipelago: Hong Kong, Periplus, 515 p.
Marine Letters, v. 9, p. 1–10, doi: 10.1007/BF02262812. Waltham, D., Hall, R., Smyth, H.R., and Ebinger, C.J., 2008, this volume, Basin
Suttner, L.J., and Dutta, P.K., 1986, Alluvial sandstone composition and paleo- formation by volcanic arc loading, in Draut, A.E., et al., eds., Lessons
climate; I. Framework mineralogy: Journal of Sedimentary Research, from the Stratigraphic Record in Arc Collision Zones: Geological Society
v. 56, p. 329–345. of America Special Publication 436, doi: 10.1130/2008.2436(02).
Taira, A., Hill, I., Firth, J., et al., 1991, Proceedings of Ocean Drilling Program, Widiyantoro, S., and van der Hilst, R.V., 1997, Mantle structure beneath Indo-
Initial Reports, v. 131: College Station, Texas, Ocean Drilling Program nesia inferred from high-resolution tomographic imaging: Geophysical
[Online]: http://www-odp.tamu.edu/publications/131_IR/VOLUME/ Journal International, v. 130, p. 167–182, doi: 10.1111/j.1365-246X.1997.
CHAPTERS/ir131_06.pdf (December 2006). tb00996.x.
Tang, J.-C., and Chemenda, A.I., 2000, Numerical modelling of arc–continent William, A.G., Lambiase, J.J., Back, S., and Jamiran, M.K., 2003, Sedimentol-
collision: Application to Taiwan: Tectonophysics, v. 325, p. 23–42, doi: ogy of the Jalan Selaiman and Bukit Melinsung outcrops, western Sabah:
10.1016/S0040-1951(00)00129-3. Is the West Crocker Formation an analogue for Neogene turbidites off-
Tang, J.C., Chemenda, A.I., Chery, J., Lallemand, S., and Hassani, R., 2002, shore?: Geological Society of Malaysia Bulletin, v. 47, p. 63–75.
Compressional subduction regime and initial arc–continent collision: Zeilinga de Boer, J.Z., and Sanders, D.T., 2002, Volcanoes in Human History:
Numerical modeling, in Byrne, T., and Liu, C.-S., eds., Geology and Geo- Princeton, New Jersey, Princeton University Press, 295 p.
physics of an Arc–Continent Collision, Taiwan: Geological Society of
America Special Paper 358, p. 177–186. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Carbonate-platform facies in volcanic-arc settings: Characteristics and controls on


deposition and stratigraphic development

Steven L. Dorobek*
Department of Geology & Geophysics, Texas A&M University, College Station, Texas 77843, USA

ABSTRACT

Shallow-marine carbonate facies from volcanic-arc settings provide an important,


but commonly overlooked, record of relative sea-level change, differential subsidence-
uplift, paleoclimate trends, and other environmental changes. Carbonate strata are
thin where volcanic eruptions are frequent and voluminous, unless shallow, bathy-
metric highs persist for long periods of time and volcaniclastic sediment and erupted
materials are trapped in adjacent depocenters. Carbonate platforms and reefs can
attain significant thickness, however, if subsidence continues after volcanic activity
ceases or the volcanic front migrates. The areal extent of shallow-marine carbonate
sedimentation is likewise affected by differential tectonic subsidence, although car-
bonate platforms are most laterally extensive during transgressive to highstand condi-
tions and when arc depocenters are filled with sediment.
Tectonic controls on shallow-marine carbonate sedimentation in arc depocenters
include (1) coseismic fault displacements and associated surface deformation; (2) long-
wavelength tectonic subsidence related to dynamic mantle flow, flexure, lithospheric
thinning, and thermal subsidence; and (3) large-scale plate deformation related to
local conditions of subduction.
Depositional controls on carbonate sedimentation in arc depocenters include
(1) the frequency, volume, and style of volcanic eruptions; (2) accumulation rates for
siliciclastic-volcaniclastic sediment; (3) the frequency, volume, and dispersal paths of
erupted material; (4) (paleo)wind direction, which influences both carbonate facies
development directly and indirectly by controlling the dispersal of volcanic ash and
other pyroclastic sediment, which can bury carbonate-producing organisms; (5) the
frequency and intensity of tsunami events; and (6) volcanically or seismically trig-
gered mass-wasting events, which can erode or bury carbonate strata.
Regarding platform morphologies in arc-related settings, (1) fringing reefs or
barrier reef systems with lagoons may develop around volcanic edifices throughout
the long-term evolution of volcanic arcs; (2) local reefs and mounds may build on
intrabasinal, fault-bounded highs within underfilled forearc, intra-arc, and backarc
basins; (3) isolated platforms with variable platform margin-to-basin transitions are

*Current address: Maersk Oil & Gas, Esplanaden 50, DK 1263 Copenhagen, Denmark.

Dorobek, S.L., 2008, Carbonate-platform facies in volcanic-arc settings: Characteristics and controls on deposition and stratigraphic development, in Draut,
A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special
Paper 436, p. 55–90, doi: 10.1130/2008.2436(04). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All
rights reserved.

55
56 Dorobek

common in “underfilled” and tectonically active depocenters; and (4) broad ramps
and rimmed carbonate shelves are typically found in tectonically mature and sedi-
ment-filled depocenters.

Keywords: carbonate platforms, reefs, carbonate facies, subsidence, basins, depocen-


ters, carbonate stratigraphy.

INTRODUCTION that only cut upper crustal levels to truly lithosphere-scale fault
systems; and (2) long-wavelength, basin-scale differential subsid-
Shallow-marine carbonate facies may constitute an impor- ence and uplift, which typically involve lithosphere-scale defor-
tant part of the stratigraphy that forms in volcanic-arc settings, mation processes (e.g., flexure, lithospheric thinning, or thermal
although they have received relatively minor attention from geo- subsidence) or dynamic mantle flow (i.e., dynamic topography-
scientists (cf. Dickinson, 1995, 2001; Smith and Landis, 1995; subsidence; cf. Gurnis, 1990a, b, 1991). Convergent margins with
Soja, 1996; Wilson and Bosence, 1997; Nunn, 1998a; Wilson, active subduction, however, are unique tectonic settings where
2002; Wilson and Lokier, 2002; Wilson and Vecsei, 2005; Bos- coseismic displacements along great subduction faults may cause
ence, 2005). Shallow-marine carbonate facies in volcanic-arc long-wavelength (>100 km) surface deformations that can sig-
settings are important recorders of sea-level changes, differen- nificantly affect depositional systems on the upper (or overriding)
tial subsidence across arc depocenters, paleoclimate, and other plate. This style of differential tectonic subsidence and uplift is
environmental conditions. Thus, careful analysis of these strata exclusive to active subduction zones and was dramatically illus-
is critical for reconstructing the complicated geological histories trated during the December 2004 Sumatran earthquake (Fig. 1).
of volcanic-arc depocenters. Significant petroleum accumula- These various scales of tectonic deformation typically inter-
tions are also stored in carbonate reservoirs in some volcanic-arc act simultaneously and in highly complex ways in volcanic-arc
depocenters (e.g., Indonesia backarc basin; Sharaf et al., 2005), settings, making the subsidence and uplift histories of arc dep-
although these rather limited examples may be associated with ocenters difficult to characterize, predict, and understand. Other
special petroleum-system conditions. tectonic characteristics of volcanic-arc settings can also affect
This paper presents a review of the tectonic and depositional carbonate sedimentation, such as large-scale aspects of subduc-
controls that influence shallow-marine carbonate systems associ- tion, plate motion, and plan-view geometries of plate margins.
ated with volcanic arcs along convergent margins. Examples of The following section first introduces large-scale tectonic con-
carbonate systems from modern and ancient settings are used to trols on arc depositional systems and progresses to consider
illustrate key principles, although it is important to note that rela- more local tectonic deformation and its effects on deposition in
tively few, thoroughly documented examples exist from which arc depocenters.
to draw conclusions or construct predictive models. Detailed
stratigraphic information for carbonate successions that formed Intra-Oceanic Island Arcs versus Continental-Margin
in a variety of arc depocenters during different times in Earth Arc Systems
history is simply not available. Instead, the main objective of this
paper is to describe how tectonic deformation, other depositional Whether a volcanic arc forms as an intraoceanic or a
controls, and siliciclastic and volcaniclastic sedimentation can continental-margin arc system is of first-order importance for
influence shallow-marine carbonate facies development in arc shallow-marine carbonate systems. Intraoceanic island arcs
settings. More complete documentation of carbonate successions form on the leading edge of an overriding oceanic plate. In
from modern arc settings and from the rock record will undoubt- contrast, continental-margin arc systems form where a slab
edly lead to refinement of the basic models presented here, but I of oceanic lithosphere is subducted beneath an upper plate of
hope this paper serves as a general guide to the analysis of arc- continental lithosphere. Arc-related basins along both types of
related carbonate systems. plate margins may be similar, although the initial topography
of the overriding plate and rheological differences between
TECTONIC CONTROLS ON ARC-RELATED oceanic and continental lithosphere of the upper plate impose
CARBONATE SYSTEMS different conditions on basin development.
Volcanism associated with intraoceanic arc systems ini-
Tectonic deformation along convergent plate margins tially occurs in submarine environments. Successive subma-
strongly influences shallow-marine carbonate depositional sys- rine eruptive events cause progressive shallowing of volcanic
tems in volcanic-arc depocenters. Differential tectonic subsid- edifices until they eventually build to within the photic zone,
ence or uplift in any tectonic setting can be attributed to (1) local, and shallow-marine carbonate deposition can begin, which
fault-controlled deformation, which can involve fault systems may take place after millions of years of eruptive activity. In
Carbonate-platform facies in volcanic-arc settings 57

14°N Extensional, Contractional, Strike-Slip, or Neutral


Arc Settings
Andaman Islands
Convergent margins can be classified according to the
12°N regional styles of deformation that affect the upper plate (cf.
Dewey, 1980). Various sectors of the upper plate can be charac-
terized by extensional, contractional, strike-slip, or very minor
(i.e., neutral) deformation. Extensional deformation can develop
10°N within (1) the thermally weakened basement of the arc massif,
(2) the backarc region, or (3) the forearc region, especially where
there is significant tectonic erosion during subduction (Clift et al.,
1998). Most deforming volcanic arcs have at least local strike-
8°N Nicobar Islands
slip fault zones, even where contractional or extensional styles
of deformation are dominant. In compressional arcs, thrust and
reverse faults typically form in the backarc region on the upper
plate; these faults are typically antithetic to the dip of the sub-
6°N ducting slab. Compressional retroarc foreland basin settings are
not considered here, although carbonate facies in extensional
TR

backarc basins are discussed in a later section.


EN

It also is important to note that some dip-trending sections


S
um
CH

4°N 24 Dec 2004 Sumatra across subduction margins show compressional deformation in
at
ra
epicenter Sumatra one segment of an arc system but other styles of deformation
Simeulue in other parts of the system. For example, the Sumatran forearc
region is generally considered to be an “accretionary” forearc
2°N Uplifted shoreface (Clift and Vannucchi, 2004), yet the arc itself is cut by the Suma-
28 March 2005
epicenter tran Fault, which is a major, through-going, strike-slip fault sys-
Subsided shoreface tem that extends along the entire length of Sumatra. Although
0 200 km Nias the Sumatran forearc region is generally thought to be accreting,
0° active zones of transtensional faulting cut at high angles across
90°E 92°E 94°E 96°E 98°E the forearc basin and partition the Sumatran forearc into multiple
depocenters. These closely juxtaposed styles of tectonic defor-
Figure 1. Distribution of coseismic uplift and subsidence across the
Andaman Islands and Western Indonesian forearc after the Decem- mation are described here simply to demonstrate that it can be
ber 2004 and March 2005 major earthquakes. Displacement along the inappropriate to make general characterizations about the over-
subduction fault between the subducting Indian Ocean plate and the all tectonic style of a particular arc system without examining
overriding Indonesian forearc propagated rapidly northward from the long-term deformation and subsidence histories across the entire
epicenters (indicated by stars) toward the Andaman Islands. Areas of system, from the trench to the backarc region, and understanding
coastal uplift are indicated by the stippled pattern. Areas of coastal
subsidence are indicated by dark shading. Dashed line represents inter- the kinematic history of the deformation. Accurate, large-scale
preted subsidence hinge line. Data obtained from SAR interferometry kinematic analyses may be impossible in strongly tectonized,
(InSAR) analyses. Modified from Japan Geographical Survey Institute ancient arc settings.
Web site.
Margin Geometries, Relative Plate Motion, and Changes
in Subduction
contrast, continental-margin arc systems may actually begin as
subaerial settings and require either a significant eustatic rise or Plate-motion vectors and the plan-view geometry of conti-
tectonic subsidence to submerge them to photic depths, when nental margins (i.e., before and during convergence and collision)
shallow-marine carbonate deposition can begin. also influence subduction or arc collision along specific margin
Although both oceanic and continental lithosphere will segments (Dewey, 1980; Thomas, 1983; Bradley, 1989). The
bend in response to the growth of surface loads like volcanic arc initiation of subduction or collision usually will be diachronous
systems, they have different flexural rigidities so their flexural along strike where the motion of one plate is highly oblique to
response will differ. Continental lithosphere also may have more another plate margin. If either plate is rotating about a vertical
inherent strength variations than oceanic lithosphere because of pole, then the rate of subduction or shortening caused by collision
its typically long and complicated geologic history. These varia- typically increases with distance from the pole of rotation. The
tions in strength and the presence of preexisting faults can influ- presence of salients or recesses along either plate margin will also
ence basin development in continental-margin arc systems. influence which segments along either margin are first to collide.
58 Dorobek

The angle of convergence or subduction also strongly influ- microplates meet along a convergent margin, some blocks might
ences the styles and distribution of deformation within the over- be “extruded,” or pushed laterally away from the zone of conver-
riding plate. Where the relative motion of both plates is nearly gence (e.g., Tapponnier et al., 1982; Coward, 1990; Maynard et
perpendicular to the trench axis, and trench rollback is slower al., 1997), although the amount of extrusion and lateral transla-
than the velocity at which the overriding plate approaches the tion of blocks away from the collision zone varies. Significant
subduction zone, compressional styles of deformation should shortening may continue to occur along the convergence zone,
dominate in the overriding plate (cf. Dewey, 1980). In contrast, but lateral translation of fault-bounded blocks may be required
oblique subduction characteristically produces some type of to accommodate additional convergence once the crust within
transtensional, transpressional, or pure strike-slip deformation in the collision zone cannot be tectonically thickened anymore.
the overriding plate. The style and geographic location of upper- Regional strike-slip-fault networks that bound the translated
plate deformation (e.g., within the arc-trench gap, the volcanic blocks accommodate the extrusion. The sense of slip on these
arc, or in the backarc region) and volcanic activity depend on faults may change over time as one block progressively plows
(1) the angle of convergence; (2) the dip of the subducted slab; into the adjoining plate and more inboard blocks move laterally
(3) whether any seamounts, active spreading-ridge segments, to accommodate additional shortening in the collision zone.
fracture zones, or other rough surface features on the subducted
slab arrive at the subduction zone; (4) the degree of mechani- Tectonic Drivers of Long-Wavelength Differential
cal coupling between the subducting and overriding plates; and Subsidence in Arc Settings
(5) the location and orientation of lateral strength variations or
preexisting fault zones in the overriding plate, which might be Volcanic-arc settings are characterized by rapidly changing
reactivated during convergence. Other mechanical or rheological and complexly distributed subsidence patterns, which reflect the
attributes of plates or microplates involved along the convergent interactions of multiple tectonic drivers that operate along sub-
margin may influence deformation in the overriding plate but duction margins. These drivers of differential long-wavelength
probably in less significant ways. subsidence generally involve deformation or long-term rheo-
The timing of collision and styles of deformation associated logical changes of the entire lithosphere, and include flexure,
with convergent margins will change if plate-motion vectors vary thermal subsidence, magmatic underplating, and thinning in
over time. Changes in the relative motion of plates also cause extensional arcs (Fig. 2). Dynamic mantle flow can also influ-
intraplate stress trajectories to be reoriented over time, and con- ence long-wavelength subsidence patterns in the upper plate (i.e.,
sequently the styles and patterns of deformation within plate inte- dynamic topography; Gurnis, 1990a, b, 1991). Long-wavelength
riors should evolve, especially where preexisting basement fault subsidence patterns will influence the distribution, plan-view
systems are reactivated by in-plane stress. dimensions, and ultimate thickness of carbonate platforms, with
Magmatic activity in the overriding plate will also be affected the most widespread development of carbonate facies generally
by changes in the nature of subduction (cf. Otsuki, 1989). Where occurring during waning stages of subduction, when tectonic
the subducted slab has a shallow dip, the location of the volcanic subsidence rates are progressively slowing and arc depocenters
front generally is farther inboard from the trench, or the margin are more likely to become filled with sediment.
may lack substantial volcanic and magmatic features altogether Deformation in the forearc region, especially within the sub-
where “flat-slab” subduction occurs. The overriding plate may duction complex, can result in other types of long-wavelength
also be deformed where the dip of the subducted plate becomes subsidence or uplift that are not related to lithospheric defor-
very shallow (nearly horizontal) and there is mechanical coupling mation. The subduction complex can deform in multiple ways,
between the subducted plate and the underside of the overrid- including complex internal thrusting, backthrusting, normal
ing plate (e.g., Neogene deformation of the Chilean-Argentinean faulting, and gravitational sliding. These different types of
foreland; Jordan et al., 1983; Jordan and Allmendinger, 1986). deformation are thought to occur as the subduction complex
If the dip of the subducted slab steepens over time, trench roll- tends to maintain a wedge shape or “critical taper” (Davis et al.,
back and extension may occur across much of the arc (Uyeda 1983; Dahlen et al., 1984). Underplating and subduction ero-
and Kanamori, 1979; Otsuki, 1989). The plan-view geometry of sion also can cause long-wavelength surface deformation across
the overriding plate may also control curvature of the subduct- the top of the subduction complex (Fig. 3). Deformation of the
ing plate, downdip from the subduction zone, which in turn may subduction complex can, in turn, cause differential subsidence
influence deformation across backarc regions in the upper plate in the outer part of the forearc basin because the subduction
(cf. Cahill and Isacks, 1992). complex commonly forms the seaward limit of this depocenter.
Where arc-continent or continent-continent collision has This differential subsidence affects shallow-marine carbonate
occurred, subduction will usually become very limited or may deposition in the outer part of the forearc basin when the basin
cease altogether. In these cases, the subduction zone may jump is filled with sediment and the basin floor is within the photic
farther outboard and also change dip (or polarity) so that contin- zone. Along convergent margins where the subduction com-
ued plate convergence can be accommodated by continued sub- plex is absent or poorly developed (e.g., with sediment-starved
duction of oceanic lithosphere. Where continental plates or arc trenches or where subduction erosion dominates), depositional
Carbonate-platform facies in volcanic-arc settings 59

DS DS
Fsc Fsc
Tva Tva Tbaoc Tbaoc

Ffbf Fbabf Ffbf


Fbabf Tra

(little or no thermal
subsidence of old
trapped oceanic crust) 0 100 km
A 0 100 km
B

DS DS
Fsc Fsc
Tva Tva

Ffbf Frfb Ffbf Tba

0 200 km
C 0 200 km D

Basin Fill Dynamic Subsidence

Continental or Arc Crust Flexural Subsidence

Oceanic Crust Thermal Subsidence

Figure 2. Causes of long-wavelength (>100 km) tectonic subsidence across the overriding plate for different types of convergent margins. Flex-
ural subsidence patterns (indicated by dashed black lines): Fsc—flexural subsidence of subducted plate caused by load of the subduction complex;
Ffbf—flexural subsidence of forearc region caused by sediment load within the forearc basin; Fbabf—flexural subsidence of backarc region caused
by sediment load within the backarc basin; Frfb—flexural subsidence of retroarc foreland basin region caused by sediment load and bordering
thrust belt. Thermal subsidence patterns (indicated by dotted lines): Tva—thermal subsidence caused by cooling magmatic arc, after arc be-
comes extinct; Tbaoc—thermal subsidence of oceanic crust created along backarc spreading center; Tra—thermal subsidence caused by cooling
of remnant magmatic arc in extensional backarc settings; Tba—thermal subsidence caused by cooling of rifted backarc continental or arc crust.
DS—long-wavelength dynamic subsidence across retroarc settings (indicated by solid black line). In-plane stress may also affect any flexural
subsidence patterns, depending on whether stress is compressive or tensile (in-plane stress effects are not shown here). Note how various tectonic
subsidence mechanisms may interfere with subsidence patterns caused by other drivers. Patterns are highly generalized. Any of these drivers
of long-wavelength subsidence can be counter-affected by shortening or tectonic inversion across the overriding plate. See text for additional
discussion. (A) Intra-oceanic volcanic arc with old oceanic crust trapped in backarc region. There may be little or no thermal subsidence of the
old, trapped oceanic crust. (B) Backarc region underlain by newly created oceanic crust formed by backarc spreading axis. (C) Compressional
backarc region (retroarc foreland basin system) underlain by continental crust. (D) Extensional backarc region underlain by rifted continental
crust. Synrift subsidence due to crustal thinning is not shown; only postrift thermal subsidence is shown for the extensional retroarc basin.

gradients from the arc to trench tend to be steep for long periods a thorough description of surface deformations caused by fault
of time, and shallow-water carbonate deposition will be limited displacements is beyond the scope of this paper, typical coseis-
mostly to fringing reefs and narrow platform systems around mic surface deformations are described in Table 1. Important
the volcanic arc. aspects of these surface deformations include the following:
1. Coseismic surface deformations are geologically instan-
Fault-Related Surface Deformation in Arc Depocenters taneous and can amount to several meters where fault
displacements are greatest, which is typically found at
Displacement along individual faults in arc settings can the midpoint of fault surfaces along which displacement
cause shorter wavelength surface deformation. This surface occurs.
deformation dramatically influences carbonate deposition if the 2. Recurrence intervals of earthquakes that cause significant
deformation occurs in shallow-water environments. Although surface deformation in arc depocenters may be several
60 Dorobek

A uplift of trench-slope break decades to several thousand years. Even with compara-
caused by back-rotation of imbricate tively long recurrence intervals, surface subsidence and
thrust sheets; local uplift patterns
controlled by individual thrust slivers
uplift can accrue to many tens or even hundreds of meters
within a few thousand or tens of thousands of years, which
will have obvious effects on carbonate facies develop-
ment. Longer earthquake recurrence intervals may allow
carbonate platforms to “recover” and fill new accommo-
dation space that was created during an earthquake.
3. Coseismic displacements on great subduction faults can
cause large areas of carbonate platforms and reefs in arc
depocenters to subside or to be instantly uplifted above
sea level (Fig. 1). Coastal and intertidal environments will
B uplift of subduction complex
over zone of underplating;
be most greatly affected. These events generally affect the
uplift geometry controlled by forearc region only, where mechanical coupling between
geometry of duplex system the subducting and overriding plates is strongest.
4. Coseismic fault displacements can generate tsunamis or
cause mass wasting that can obliterate or bury large areas
of shallow-marine carbonate deposition.
5. Some faults may creep aseismically, and although their
related surface deformations may slowly develop, sig-
nificant amounts of uplift or subsidence can still accrue
over time. Aseismic fault creep may be most common in
subduction complexes where faults cut weakly lithified,
C uplift of trench-slope break
caused by back-thrusting as highly sheared, fluid-rich materials.
subduction complex attempts to
maintain critical taper;
Active faults in arc depocenters can control carbonate-plat-
local fault-related folding form morphology, platform margin-to-basin profiles, and lateral
facies changes during deposition. Active faults can segment pre-
existing carbonate platforms or control the location of carbonate

backthrusting
Figure 3. Possible mechanisms for internal deformation and evolu-
tion of subduction complexes, with associated short- to intermediate-
wavelength surface deformation. Local surface deformations common-
ly are expressed as complex patterns of uplift, which are superimposed
D regional trench-slope
adjustment and local
on longer-wavelength patterns of tectonic subsidence. Figure modified
from Dickinson (1995). (A) Accretionary offscraping of successive
fault-controlled subsidence
imbricate thrust panels (Seely et al., 1974). For growing subduction
complexes, back-rotation of imbricate thrusts causes progressive uplift
of crest of subduction complex (i.e., trench-slope break region). There
also may be local fault-related folding, which creates antiformal sur-
face deformations. (B) Selective subduction and underplating of thrust
extensional duplexes (Sample and Fisher, 1986). Underplating at the base of the
faulting subduction complex can cause uplift over the zone of underplating.
Structural styles related to underplating are often poorly imaged on
seismic profiles but probably are most likely expressed as a duplex
zone. (C) Backthrusting of subduction complex toward forearc basin
E subsidence caused by tectonic
erosion or sediment subduction;
(Silver and Reed, 1988). Uplift of trench-slope break and distal part of
forearc basin may occur as the subduction complex maintains a critical
compaction has similar effect taper. Backthrusting in arcward side of subduction complex generally
causes uplift in trench-slope break region, with local surface deforma-
tion caused by fault-related folding. (D) Denudational normal faulting.
Both regional subsidence and local fault-controlled subsidence may
occur across trench slope as the subduction complex maintains a criti-
cal taper via extensional deformation. (E) Subduction erosion along
tectonic erosion décollement beneath subduction complex (cf. von Huene et al., 1982).
beneath Subduction erosion or sediment subduction along underside of sub-
subduction sediment duction complex creates subsidence across subduction complex. Com-
complex subduction paction of subduction complex may have similar effects.
Carbonate-platform facies in volcanic-arc settings 61

Table 1. TYPICAL COSEISMIC SURFACE DEFORMATIONS FOR


DIFFERENT FAULT TYPES
and stratal patterns (Caron et al., 2004). Elongated, tide- and storm-
Fault type Fault length Surface uplift- influenced balanid (barnacle) shoals grew on local antiformal highs
involved subsidence near the crest of the subduction complex. In contrast, arc-attached,
Normal faults 5–50 km 0–3 m low-gradient shelf and ramp profiles were built along the arc side
Reverse/thrust faults 5–100 km 0–5 m
Subduction faults 100–1000+ km 0–5 m
of the forearc basin. Carbonate strata are more sheetlike along the
Strike-slip faults: flank of the arc, are generally transgressive in character, and were
Transfer fault 1–50 km 0–3 m dominated by epifaunal bivalves. These characteristics are thought
Major intraplate or plate 50–500+ km 0–3 m to reflect a higher flux of sediment from the adjacent arc.
boundary
The transport mechanisms and volume of material involved
in sediment gravity flows will control the size of carbonate plat-
forms or reefs that might be buried during a single event. The
facies tracts wherever displacement rates exceed the ability of frequency of catastrophic depositional events within a particular
carbonate facies to fill newly generated accommodation space segment of the arc will determine whether adjacent carbonate
on either side of a fault. Carbonate lithofacies deposited around platforms and reefs can recover from a series of events. If rela-
faults that were active during deposition may contain greater tively low volumes of volcanogenic material are involved in any
amounts of gravity-flow deposits, tsunamites, or other cata- catastrophic event, and the events occur with low frequency or
strophic-event beds. are widely dispersed along strike, then carbonate platforms and
reefs may recover and even utilize lava flows or volcaniclastic
DEPOSITIONAL CONTROLS ON CARBONATE facies as substrates.
SEDIMENTATION IN ARC SETTINGS Mass wasting can also occur at any time along steep vol-
canic edifices, although these catastrophic events are triggered
The accumulation of noncarbonate sediments and eruptive more often by volcanic earthquakes or by inflation of the volca-
products has important effects on carbonate deposition in arc no’s surface as magma ascends into subsurface magma chambers
depocenters. Various depositional and erosional processes that prior to an eruption. Antecedent drainage networks on the flanks
are characteristic of arc settings can also affect carbonate deposi- of a volcano may control dispersal of low to moderate volumes
tion. The relative influence of these factors on carbonate deposi- of volcanogenic material, although dispersal of large-volume,
tion may depend on the developmental stage of the arc system. density-stratified pyroclastic flows are generally less affected by
antecedent topography. The flow paths of only the basal, high-
Erosional and Depositional Modifications to density part of pyroclastic flows may be diverted by antecedent
Bathymetric Profiles topography, such that the resultant coarse-grained deposits are
confined to ravinement floors on the flanks of a volcano. These
The bathymetry of arc depocenters can be rapidly modified high-density flows then spread laterally, where they become
by subaerial or submarine deposition of volcanic materials or unconfined in lower hillslope regions where antecedent valleys
large-scale mass-wasting events. Arc settings are also seismically are less incised and surface gradients flatten. In contrast, the low-
active during active subduction, and large-magnitude earthquakes density upper parts of pyroclastic flows are commonly stripped
commonly generate significant ground displacements, so slope from the basal, high-density part of the flows. The low-density
failure occurs frequently. Although erosional and depositional parts of pyroclastic flows can flow over topographic obstructions
modifications to shallow-water depositional profiles are relatively and lay down more areally extensive deposits. Low-density pyro-
common in arc depocenters, they are (1) episodic and thus diffi- clastic flows can also flow for several kilometers across seawater.
cult to predict and correlate in stratigraphic successions, (2) vari- Lava flow paths will also be controlled by antecedent topography,
able in lateral extent, and (3) involve highly variable amounts of unless the volume of lava exceeds the ability of valleys and hill-
rock materials. slope ravinements to contain the flows.

Sediment Flux to Arc Depocenters Underfilled versus Filled Arc Depocenters

Sediment flux from the volcanic arc or uplifted arc-massif Sedimentary basins are often described as “underfilled” or
rocks can have many effects on carbonate production in adja- “filled” (cf. Covey, 1986; Flemings and Jordan, 1990; Jordan,
cent depocenters. High sediment flux can completely overwhelm 1995), which describes whether depositional systems deposit sed-
carbonate-producing benthic organisms. Lesser amounts of sedi- iment into a depocenter or carry sediment past a former depocen-
ment flux might lead to mixed carbonate-siliciclastic-pyroclastic ter. Although a thorough discussion of this concept is beyond the
successions. The caliber of sediment supply from volcanic-arc and scope of this paper, it is useful for understanding where and when
arc-massif rocks can also determine the types of carbonate sedi- widespread carbonate platforms might form in arc depocenters.
ment produced. For example, late Pliocene carbonate facies in the Although unique conditions of subduction can determine sub-
New Zealand forearc show major lateral changes in sediment types sidence and basin-filling patterns in arc systems, arc depocenters
62 Dorobek

are generally more likely to become filled (1) during waning stages system, whereas the leeward side of the arc may be dominated
of subduction when tectonic subsidence slows down, (2) in arc by muddier, low-energy carbonate facies. Volcanic arcs can
depocenters with high sediment accumulation rates, or (3) in arc also have significant orographic influence on precipitation, with
depocenters that begin near base level at the onset of subduction. windward sides of the arc being wetter overall than the leeward
Previously deposited siliciclastic sediment and erupted sides. Higher precipitation on the windward side of an arc may
materials in filled depocenters can provide broad substrates for mean greater chemical and physical weathering, with a greater
shallow-marine carbonate deposition. Areally extensive carbonate flux of sediment, especially finer grained sediment, to adjacent
platforms most commonly form when rising sea level floods these marine depocenters. High weathering, erosion, and transport of
substrates. In addition, tectonic surface deformations become sediment to the windward side of an arc can greatly suppress
more important for shallow-water carbonate depositional systems carbonate production.
in filled depocenters. If a filled depocenter is submerged to water Prevailing winds can also greatly influence the dispersal of
depths that are within wave base and typical tidal ranges (<20 m erupted pyroclastic material, especially volcanic ash. Leeward
water depth), even subtle tectonic surface deformations will sig- depocenters may have much greater accumulations of volcanic
nificantly affect shallow-water carbonate depositional systems. ash (Sigurdsson et al., 1980), which also can suppress carbon-
ate production.
Oceanographic Circulation
Nutrient Supply
The types of carbonate sediment that accumulate in arc
depocenters will be influenced by the location of the arc system High nutrient levels in seawater suppress carbonate produc-
with respect to global climatic zones or oceanographic circula- tion (Hallock and Schlager, 1986). A high nutrient flux from
tion systems. Cool-water carbonate facies (James, 1997) will be arc systems may be related to intense chemical weathering of
more typical of high-latitude arcs or where cold, upwelling water arc rocks, eruption products, or, in the case of modern settings,
reaches shallow-water parts of an arc system. Cool-water carbon- anthropogenic input, and can have major deleterious effects on
ate deposits typically consist of the skeletal remains of heterotro- carbonate production. Upwelling of cold, nutrient-rich deep
phic and light-independent biota that are dominated by suspen- water can also suppress carbonate production or influence car-
sion feeders (e.g., bryozoans) and filter feeders (e.g., barnacles), bonate-sediment types in some arc settings (Halfar et al., 2004).
which are extremely sensitive to the amount of suspended sedi-
ments and dissolved nutrients in the water column (Henrich and CARBONATE DEPOSITION WITHIN ARC
Freiwald, 1995). Thus, extensive cool-water carbonate produc- DEPOCENTERS
tion may be difficult to establish close to high-latitude arc sys-
tems because of high sediment flux from the arc. Shallow-marine carbonate platforms and reefs can form
Some arc systems may also have unique physiography with in forearc, intra-arc, and backarc settings. Various drivers and
respect to oceanographic circulation so that anomalous water expressions of tectonic subsidence and uplift affect each setting
masses either enter oceanic passageways or upwelling cold-water and may vary as the arc system evolves. Temporal and spatial
masses impinge on shallow-marine settings and influence carbon- changes in differential tectonic subsidence have the greatest
ate facies development over large parts of the arc system. For exam- effects on carbonate sedimentation when arc depocenters are
ple, the present-day Gulf of California, Mexico, is a backarc setting nearly filled to sea level. Depositional processes and sediment
where active seafloor spreading has created an open passage at the flux from the arc system also vary between arc depocenters,
southern end of the gulf. Seasonal variations in upwelling cause which also can influence carbonate deposition.
latitudinal variations in nutrient levels within the Gulf of Califor- The carbonate-platform classification scheme of Read
nia (Halfar et al., 2004). In the southern part of the gulf, carbonate (1985) is used throughout this paper. Stratigraphic intervals that
production is oligotrophic-mesotrophic and coral reef–dominated. record progressive surface deformation during deposition are
The central gulf is dominated by mesotrophic-eutrophic, red-algal growth strata (cf. Vergés et al., 2002). Carbonate growth strata
sediments, and the northernmost parts of the gulf are dominated by are deposited around or on top of actively growing tectonic struc-
eutrophic, molluscan-bryozoan sediment assemblages. tures and can be used to reconstruct both the kinematic history of
individual structures and basin-scale patterns of tectonic subsid-
Wind Direction ence and uplift.

Paleowind direction directly influences the distribution of Carbonate Platforms and Reefs across the Forearc Region
high-energy, wave-agitated carbonate-facies belts within car-
bonate platforms (Eberli and Ginsburg, 1989). Volcanic arcs can Many forearc regions are characteristically too deep for
build significant topographic relief such that if the arc lies in a shallow-marine carbonate sedimentation, especially during early
climatic belt with strong prevailing winds, wave-agitated facies stages of subduction and basin development. Thus, extensive
tracts may preferentially build on the windward side of the arc shallow-marine carbonate platforms are not likely to develop
Carbonate-platform facies in volcanic-arc settings 63

across the forearc region until deformation or basin filling has Eratosthenes Seamount, Mediterranean Sea, Galindo-Zaldívar
provided suitable substrates for carbonate deposition. et al., 2001). Tectonic uplift and growth of the subduction com-
Using the classification scheme of Seely (1979) and Dick- plex, however, can also raise parts of the subduction complex
inson and Seely (1979) for forearc settings, and assuming that into the photic zone and possibly even above sea level, creating
the forearc region is at the proper (paleo)latitude, carbonate- substrates for shallow-marine carbonate deposition.
platform development is possible where (1) the crest of the sub- Reefs and associated carbonate facies can fringe emergent
duction complex forms a shoal water or emergent ridge (Figs. 4, parts of the subduction complex, such as in emergent islands of
5); (2) the forearc basin is largely filled with sediment so that a the western Indonesian forearc. Variably sized isolated platforms
broad, shallow, shelf-like substrate extends from the volcanic arc can also form on shallow, submerged parts of subduction com-
toward the trench-slope break (Fig. 5); (3) the arc massif in front plexes. Carbonate-platform sequences that form on subduction
of the volcanic arc is shallow and provides a suitable substrate complexes are rarely more than a hundred meters thick, and apart
for carbonate sedimentation; and (4) material flux (i.e., volca- from fringing reef systems, the platforms commonly have ramp-
niclastic sediment, ash fall, lava flows) from the volcanic arc like profiles that mimic the depositional gradients across the top
does not suppress shallow-marine carbonate deposition across of the subduction complex. Carbonate-platform sequences on
the forearc region. subduction complexes typically do not thicken or steepen over
These conditions are similar to those required for carbonate- time and then evolve into rimmed platform morphologies, which
platform development in other tectonically active basins. Silici- reflects the relatively short life of the platforms, the rapidly
clastic and volcanic flux must be sufficient to fill depocenters changing patterns of uplift and subsidence across the top of the
to sufficient levels so that regional substrates are available for subduction complex, and the deleterious effects of volcaniclastic
carbonate sedimentation under the proper sea-level conditions material supplied by the volcanic arc.
(generally high-frequency transgressive events) but not over- Carbonate platforms and reefs on top of subduction com-
whelm and terminate carbonate production. Until a forearc basin plexes may be rare in the rock record because (1) the arcward
is nearly filled with sediment and a suitable shelf-like profile is side of the subduction complex may not become shallow enough
constructed, only structurally high parts of the subduction com- for carbonate-platform development until the subduction com-
plex or frontal part of the volcanic arc will be shallow enough for plex has become tectonically uplifted or the forearc basin has
carbonate-platform development. become nearly filled with sediment, which are conditions likely
to develop only for short periods of time in the development of
Carbonate Platforms and Reefs on Subduction Complexes forearc regions with highly accretionary subduction complexes
The first area where shallow-water carbonate facies may (Clift and Vannuchi, 2004); or (2) thin carbonate-platform strata
develop across forearc settings is the subduction complex that formed on the subduction complex are misinterpreted as
(Figs. 4, 6). Most carbonate bodies associated with ancient sub- allochthonous units that have been tectonically incorporated
duction complexes consist of highly recrystallized and sheared into the subduction complex. Recognizing carbonate facies that
pelagic carbonate facies that have been tectonically accreted and formed on shallow to emergent parts of highly deformed subduc-
incorporated into the subduction complex. Carbonate-capped tion complexes requires careful age dating and identification of
seamounts on the subducting plate or carbonate strata on conti- depositional contacts (typically angular unconformities). Older
nental or arc microplates may also be obducted or accreted to the carbonate sequences that form on top of, but are subsequently
frontal part of subduction complexes (e.g., Permian Nabeyama incorporated into, actively deforming subduction complexes will
Formation, central Japan, Minoura, 1992; Daiichi-Kashima likely have highly sheared contacts with other lithotectonic units
Seamount, Japan Trench, Cadet et al., 1987; Shin Tani, 1989; in the subduction complex.

Trench
Slope Arc - Trench Gap Figure 4. Schematic cross section of
Trench the forearc region. Stippled pattern
Volcanic represents undeformed or only mildly
Slope Front
Slope Break deformed forearc basin fill. The forearc
Trench
Basins Forearc Basin Eroded basin shown here is a largely filled dep-
Fill arc rocks ocenter. Crust beneath the forearc basin
may be highly variable in composition
Pluton
but is shown in this diagram as mafic,
? ? ?
? ?? Arc Massif
probably oceanic crust. Some forearc re-
Subducting Slab ? ?? gions may not have the well-developed
of Oceanic Crust subduction complex that is depicted in
20
Subduction this cross section. Modified from Dick-
10 Arc Magma
km Complex inson (1995).
Source
10 20 30 40 50 km
64 Dorobek

UNDERFILLED FILLED rates for subduction complexes can be derived by comparing the
elevations of well-dated, subaerially exposed Quaternary reef
terraces to estimates for eustatic sea-level curves for Quaternary
Tsb Tsb time (Fig. 7).
Tr Tr “Chimneys” of precipitated carbonate minerals have also
A. Sloped B. Sloped been documented from dredge samples and by observations
(slope basin) (slope basin) made from submersible vehicles across the surface of many mod-
ern subduction complexes (Kulm et al., 1986; Kulm and Suess,
1990; Haggerty, 1991; Hattori et al., 1995). These irregular col-
umns of precipitated carbonate are associated with fluid seeps
Tsb Tsb that discharge from permeable fault zones that cut to the top of
Tr Tr the deforming subduction complex (cf. Moore and Vrolijk, 1992).
The seeps provide nutrients for chemosynthetic fauna, including a
C. Ridged D. Terraced complex microbial community. Abiogenic and biologically medi-
(submerged ridge) (deep marine) ated carbonate precipitation occurs as the cold, methane-rich pore
fluids discharge from the episodically transmissive faults. These
Tsb Tsb carbonate chimneys are not limited to photic depths and can form
anywhere along the top of the subduction complex as long as it
is above the carbonate compensation depth. Volumetrically, how-
Tr Tr
ever, carbonate chimneys represent only a small fraction of the
E. Ridged F. Shelved total sediment accumulation within the subduction complex.
(shoal-water ridge) (shallow marine) Ancient examples. Pre-Quaternary examples of shallow-
marine carbonate facies deposited on subduction complexes are
Tsb Tsb rarely described in the literature. Ancient seep-related carbonate
chimneys are reported in Eocene strata of the Barbados subduc-
tion complex (Larue and Suess, 1985).
Tr Tr One example is provided by Upper Cretaceous(?) to middle
G. Ridged H. Benched Eocene fringing reefs and carbonate ramp sequences that formed
(emergent ridge) (terrestrial) at intermittent times during development of the Central Ameri-
can forearc, Costa Rica (Lundberg, 1982; Seyfried et al., 1991;
Figure 5. Morphologic variants of forearc basins, with basin fill stip- DiMarco et al., 1995; Kolarsky et al., 1995). The Upper Creta-
pled and vertically exaggerated for clarity. Left-side diagrams show
underfilled forearc basins (most common for island arcs); right-side ceous(?) (Campanian–Maastrichtian) Barra Honda platform is a
diagrams show filled forearc basins (most common for continental- cryptic, <100-m-thick, mud-rich platform sequence with rudist
margin arc systems). Symbols: dashed lines—sea level; Tr—trench; reefs on its windward side that formed on remnants of an under-
Tsb—trench-slope break (coincides with shelf-slope break for shelved lying Cretaceous arc-forearc complex (Seyfried et al., 1991).
forearcs). Modified from Seely (1979) and Dickinson and Seely DiMarco et al. (1995) suggest that the Barra Honda platform
(1979); from Dickinson (1995).
strata may be largely Paleocene in age. Kolarsky et al. (1995)
also interpreted the basement beneath the Paleogene forearc
region of southern Costa Rica and adjacent Nicaragua and Pan-
Quaternary examples. Quaternary carbonate reefs and ama as uplifted Cretaceous oceanic crust or seamounts. These
associated facies fringe emergent parts of modern subduction contrasting interpretations are mentioned here to illustrate how
complexes in the forearc regions of Sumatra (Simeulue, Nias, difficult it can be to interpret the history of these complex set-
Siberut, Sipora, and Enggano islands), the Solomon Islands tings. Regardless, the Barra Honda platform underwent episodes
(Tetepare, Rendova, Ranongga islands), and the Lesser Antilles of shallow submergence and carbonate accumulation, followed
(Barbados). There is little published information on the thickness by intermittent subaerial exposure and karstification, which have
and lithologies of offshore Holocene reefs and related facies in been attributed to superimposed eustatic fluctuations and tectonic
these examples. uplift caused by accretionary growth of an underlying subduction
Many modern subduction complexes are partially emergent. complex (Seyfried et al., 1991). Continued development of the
Pliocene to Quaternary reefs and other shallow-water facies may Central American arc-forearc system during latest Cretaceous to
crop out as a series of emergent coastal terraces across the crest Paleocene time was associated with further structural and bathy-
of the subduction complex. These terraces record both eustatic metric differentiation of the subduction complex (or “outer arc”
sea-level changes and progressive uplift of the subduction com- of Seyfried et al., 1991), forearc basin, and volcanic arc. Fora-
plex (Mesolella et al., 1969; Wheeler and Aharon, 1991; Mann et minifer-rich ramp facies prograded from both sides of the sub-
al., 1998; Nunn, 1998a, b, 2000; Dickinson, 2001). In fact, uplift duction complex during Late Cretaceous to late Paleocene uplift
Carbonate-platform facies in volcanic-arc settings 65

SUBDUCTION COMPLEX DEPOZONE


platform types:
• isolated platforms & buildups
substrates: FOREARC BASIN DEPOZONE
• bathymetric crest of subduction complex platform types:
• may be smaller buildups within subduction complex; • small isolated buildups & reefs
probably more common in arcward parts on structural highs
• may extend arcward into forearc basin if distal • may be faulted or structurally
part of forearc basin is filled complicated by later
deformation events
substrates:
• intrabasinal structural highs
• transpressional to transtensional
structures may be typical

?
2 km ?
1 VE ~ 10 ?

10 20 30 40 50 km

Oceanic crust

‘Transitional’ forearc crust ARC-FLANK & FOREARC


BASIN DEPOZONE
Arc massif platform types:
• small isolated buildups & reefs
Subduction complex • arc-fringing platforms & reefs
• typically thin (<100 m thickness)
Trench fill • ramps to low-relief rimmed shelf profiles
• platforms tend to become wider as forearc basin fills
Forearc basin fill with sediment
substrates:
Carbonate platform/reef facies • faulted arc basement
• volcaniclastic substrates during later stages of filling

Figure 6. Schematic cross section of forearc region, showing generalized types and locations of carbonate platforms.
The forearc basin shown here is largely filled with sediment. Underfilled conditions would characterize the forearc basin
during early stages, with smaller platforms and buildups on structural highs or building from shallow, submerged parts of
the arc massif. Crust beneath forearc basin may be highly variable in composition, but is shown in this diagram as mafic,
probably oceanic crust. Note how carbonate platforms across the forearc region tend to become more areally extensive
over time as the forearc basin fills with sediment or as the subduction complex grows, both of which provide progres-
sively wider substrates for shallow carbonate sedimentation. Basin model modified from Dickinson (1995).

of the outer arc. Eocene mixed carbonate-siliciclastic strata (Mal of the arc-trench system when subduction is active and the vol-
País Formation) contain foraminiferal limestone units up to sev- canic arc is being constructed by volcanic activity. Carbonate-
eral tens of meters thick. Strata of the Mal País Formation also platform facies can accumulate to several hundred meters thick
overlie, with angular unconformity, older strata that draped base- and extend for tens to several hundred kilometers along strike.
ment rocks of the Nicoya Peninsula. Nicoya basement rocks are Volcaniclastic strata typically separate platform sequences
thought to represent an exhumed subduction complex by some within the forearc basin fill and reflect pulsed eruptive activity
workers (cf. Lundberg, 1982). or episodic reorganization of dispersal systems that deliver vol-
caniclastic sediment to the forearc basin (Beaudry and Moore,
Carbonate Platform and Reefs in Forearc Basins 1985; Dickinson, 1995).
Carbonate platforms and reefal buildups make up much of Carbonate deposition in forearc basins is influenced by pro-
the stratigraphic fill in some forearc basins. If the forearc basin is cesses that operate at different time scales. At relatively short
at the proper (paleo)latitude, tropical carbonate facies can build time scales (i.e., 105–106 yr), thin fringing reefs and narrow plat-
away from the seaward side of the volcanic arc or arc massif form systems can form around volcanic edifices at virtually any
(Fig. 6). This is commonly the bathymetrically shallowest part time in the history of a volcanic arc and forearc basin, as long as
66 Dorobek

wave-cut notches
A indicate short stillstands

relative sea level


30 m
time
20

10
Figure 7. Technique for determining
0 high-resolution rates and magnitudes
of tectonic uplift or relative sea-level
change in shallow-marine carbonate fa-
E = present cies. Wave-cut notches are a response to
B elevation A (= duration of uplift)
Meters falling sea level, resulting from either
Displacement (D)

app rate = D/A an actual eustatic fall or tectonic uplift.


are During short-term stillstands within a
=E-e

nt 40 longer-term trend of relative sea-level

Elevation (relative & real)


displacement (apparent)
sea fall, waves have time to erode coastal
re lev exposures and create notches. From
la e l Burbank and Anderson (2001).
tiv
e = elevation e
present
at time of
formation sea level se 0
a
le
ve
l
rea - 40
l
sea
an nt lev
m e em e e
isp
la c l
d
- 80
120 80 40 0
Age (ka)

submerged parts of volcanic features are within the photic zone producing organisms or by incorporating previously deposited
and sediment flux from the arc does not overwhelm carbonate carbonate strata into mass flows.
production. Development of laterally extensive and relatively Platform types. The locations and morphologies of carbon-
thick carbonate platforms, however, requires both longer periods ate platforms found in forearc basins will be strongly controlled
of time, probably on the order of several million years, long-term by the evolving topography-bathymetry across the basin, which
tectonic subsidence, and the continued availability of shallow- depends on the evolutionary stage of the basin. In immature,
water substrates. These conditions are more likely to develop underfilled forearc basins, fringing reefs and narrow, steep-sided
during later stages of forearc-basin evolution, when tectonic platforms may form only along the trench-facing side of the arc
deformation and arc volcanism are diminishing, forearc subsid- and arc massif. During intermediate stages of forearc basin evolu-
ence rates are slowing, and the forearc basin is largely filled with tion, isolated mounds and reefal buildups might form on remnant
sediment (cf. Dickinson, 1995). or actively forming structural highs across any part of the forearc
Proximity to the volcanic arc means that the forearc basin basin; intrabasinal structural highs are most common in forearc
will be greatly influenced by sediment flux from the arc. Dur- basins that are segmented by transcurrent strike-slip faults. For
ing earlier phases of forearc-basin evolution, volcanic eruptions example, transtensional deformation of the late Miocene Hel-
and seismic events may be more frequent. Catastrophic eruptions lenic forearc basin on Crete created half-grabens where footwall
or mass wasting of volcanic edifices can instantly terminate car- highs became sites for local patch-reef development within the
bonate systems in forearc basins, either by burying carbonate- basin (ten Veen and Postma, 1999).
Carbonate-platform facies in volcanic-arc settings 67

As the forearc basin is filled with sediment and the frontal they can only grow outward. Relative sea-level changes at any
part of the arc massif is eroded, regional depositional gradients given location across the Sumatran forearc region then shape the
generally flatten, and broad substrates develop where shallow- internal growth structure of these microatolls. Coral heads grow-
marine carbonate systems can develop. During mature, filled ing on stable substrates develop nearly flat upper surfaces (Sieh
stages of forearc-basin development, carbonate systems com- et al., 1999). When subjected to gradual submergence, the living
monly construct ramp or low- to moderate-relief shelf profiles coral colony along the outer parts of these microatolls develops
that prograde into the forearc basin from erosional remnants of a raised rim, whereas their older, dead interiors become depres-
the volcanic arc (cf. Beaudry and Moore, 1985; Dickinson, 1995). sions. Microatolls subjected to gradual emergence typically
Carbonate sedimentation, however, may be terminated at any develop conical shapes because growth is limited to progressively
time across a forearc basin during major eruptive events or during lower and outer parts of the expanding structure. Careful analysis
massive hillslope failure that force carbonate-producing environ- of growth bands within coral heads from the western coast of
ments to be reestablished elsewhere within the basin. Short- to Sumatra and the outer-arc islands of western Indonesia showed
long-term tectonic uplift of the arc and arc massif may cause car- temporal and regional patterns of submergence and emergence
bonate environments to become subaerially exposed and either across the Sumatran forearc over the last several hundred years,
terminate carbonate deposition or force the carbonate factory to with microatoll uplift during the giant earthquake of 1833 serving
move elsewhere. as a key correlation horizon (Sieh et al., 1999; Zachariasen et al.,
Local seep-related carbonate buildups have also been identi- 1999). The microatoll record suggests rapid submergence over
fied in some ancient forearc basins. Examples include the Juras- several decades prior to the 1833 earthquake. Rates of pre-1833
sic Fossil Bluff Group, Antarctica (Kelly et al., 1995) and the submergence ranged from 5 to 11 mm/yr and increased trench-
Eocene Joes River mélange, Barbados (Larue and Suess, 1985). ward across the Sumatran forearc (Zachariasen et al., 1999). The
Quaternary examples. Many modern forearc basins from microatolls record coseismic uplift during the 1833 earthquake
tropical settings are bordered by extensive fringing reefs and and can be observed at six sites along a 125-km length of the
carbonate platforms that built away from the volcanic arc or subduction zone, with uplift on the order of 1–2 m. Similar to the
arc massif. For example, Quaternary fringing reefs and narrow interseismic submergence patterns, coseismic uplift also appar-
rimmed platforms are found along segments of the eastern Indo- ently increased toward the trench.
nesian forearc basin, where sediment flux from the volcanic arc The great Sumatran earthquakes of December 2004 and
is relatively low. In contrast, modern carbonate facies are much March 2005 produced significant, instantaneous uplift and sub-
less extensive in forearc depocenters along the western Indone- sidence across the Sumatran forearc (Fig. 1). The differential
sian archipelago because there is greater sediment flux from the coseismic subsidence and uplift were expressed over a 300 km
volcanic arc. Carbonate deposition is suppressed in the forearc distance from the trench axis and >12° of latitude (>2000 km);
basins of Sumatra and Java because the modern volcanic arc modern coral reefs and coastal deposits were uplifted tens of
has sufficient relief to cause significant orographic effects on centimeters to >2 m locally. A regional, arc-parallel hinge line
precipitation. The winter monsoon blows eastward across the apparently separates uplifted from subsided parts of the forearc,
forearc basins of Sumatra and Java and brings heavy rainfall to with uplifted regions on the trenchward side of the forearc basin.
southwest-facing slopes of the western Indonesian volcanic arc. This uplift may be preserved in the rock record as a regional dis-
Meteoric runoff actually lowers salinity in the Java Sea to ~30 conformity within shallow-water strata. It would be difficult to
ppt during the winter monsoon season. The seasonal decrease in recognize whether these disconformable surfaces in forearc strata
salinity and increase in the amount of suspended sediment deliv- were related to coseismic surface deformation or high-frequency
ered to the shallow shelf areas around Sumatra and Java suppress sea-level falls. Regardless, the recent Sumatran earthquake pro-
carbonate sedimentation. Thus, modern fringing reefs are geo- vides an important reminder of the rapid uplift and subsidence
graphically limited to the northwestern and southeastern tips of that can occur in forearc settings.
Sumatra and a few offshore islands within the proximal forearc Other modern forearc basins where Quaternary carbonate
basin. Subsurface data suggest that volcaniclastic flux from the facies are found include (1) the forearc basin along the southeast-
arc had less influence on carbonate sedimentation during early ern side of North Island, New Zealand (Kamp and Nelson, 1988),
Miocene time, when carbonate shelves prograded away from the and (2) the island of Espiritu Santo, Vanuatu (Wells, 1988; John-
Sumatran arc massif (Beaudry and Moore, 1985), which might son and Greene, 1988; Greene et al., 1988; Cabioch et al., 1998).
reflect a weaker monsoon at that time. North Island, New Zealand. In the modern forearc basin of
Modern microatolls (i.e., massive, ring-shaped coral heads North Island, New Zealand, thin veneers of carbonate sediment
up to a few meters in diameter) provide a high-resolution record are being deposited over only ~3% of the forearc region (Kamp
of subtle uplift and subsidence across the Sumatran forearc basin and Nelson, 1988). This limited area of modern shallow-water
(Fig. 8; Sieh et al., 1999; Zachariasen et al., 1999). Modern coral carbonate sedimentation reflects the combined deleterious effects
heads in shallow-water areas surrounding the outer ridge of the of the relatively cool-water conditions of offshore eastern New
Sumatran forearc basin grow upward and outward until they reach Zealand, the high-energy wave-swept conditions of this forearc
annual lowest low tide (Zachariasen et al., 1999), at which point setting, and the flux of siliciclastic sediment from North Island.
Figure 8. (A) Schematic cross sections
that depict the effect of relative sea-level
changes on coral growth. The outermost
solid-black growth band depicts the most
recent coral growth band. Solid gray
A (a) (b) bands are dead coral growth bands that
grew under the same sea-level conditions
SL as the living coral. Dashed growth bands
Living
are dead bands that grew under differ-
coral ent, prior sea-level conditions. (a) Hemi-
SL surface SL spherical coral head below highest level
of survival (HLS), with a new skeleton
accreting in a concentric fashion on the
time Hemispherical Coral Head time Cylindrical Coral Head
outside of the head. (b) The same coral
head under the same sea-level conditions
after it has reached HLS. No additional
(c) (d) upward growth is possible, although lat-
eral accretion continues along the sides
? of the coral head that are below HLS.
(c) “Hat” morphology of a coral head
SL SL that grew up to HLS, but then underwent
a drop in HLS. The part of the coral
exposed above HLS has died. Lateral
time “Hat” Microatoll time “Cup” Microatoll growth below the new HLS develops a
lower outer rim around a higher center.
The elevation difference between the
two flats is a measure of the amount of
B 10 emergence. (d) “Cup” morphology of
a microatoll that underwent a sea-level
centimeters

UP rise after the coral had been growing at


e

em
nc

HLS. The coral grows upward toward the


e rg
ge

enc new HLS, constrained only by its growth


e HLS
er

rate. Upward and outward growth over


bm

in the old HLS surface produces a raised


su

0 1995 outer rim, indicative of submergence.


66 67 68 69 The elevation of the new HLS is not re-
64 65 7071 727374
62 63 corded by the coral until the coral grows
75 76 up to it. From Zachariasen et al. (1999).
1961 77
78 (B) Cross section of a microatoll from
79 the Indonesian forearc. The X-rayed
80 thin slab reveals a clear record of annual
81 growth bands that expanded radially out-
82
83 ward (from left to right) at ~1 cm/yr. The
84
85 HLS of the coral during the past 35 yr is
86
87 91 92 93 96
88 89 90
recorded in the topography of the coral’s
9495 97
upper surface. The arrows track the ap-
sand substrate parent rise of sea level in the 1960s and
its subsequent apparent fall. Growth pat-
terns in corals such as these can provide
high-resolution records of true sea-level
change or the effects of incremental co-
seismic subsidence or uplift. From Sieh
et al. (1999).
Carbonate-platform facies in volcanic-arc settings 69

Beyond the modern shelf, carbonate sediments (<0.5 m thick) now form the basement rocks of the present-day Vanuatu forearc
cover the tops of isolated, elliptical banks (10–50 km2 in area) region. A subduction complex is absent along this margin owing
that rise from the slope to depths of 150–500 m (e.g., Madden to its young age and the limited amount of sediment delivered to
Banks). The banks may represent the submarine expression of the trench.
surface deformation on top of the subduction complex. Carbon- The central part of the Western Belt, including Espiritu
ate sediments that form thin veneers on these submarine ridges Santo, has been actively uplifting throughout Quaternary time
consist of coarse skeletal sand (dominantly a cool-water bivalve- owing to subduction of the d’Entrecasteaux Ridge at this loca-
foraminifer assemblage with lesser coral, brachiopod, gastropod, tion. The uplift history of Espiritu Santo is recorded by a series of
and echinoid fragments). Patchy mosaics of coarse skeletal sand emergent and notched Holocene reef terraces that were deposited
are also found on the shallow shelf updip from the slope banks. on the uplifting Oligocene–Miocene volcanic substrate (Fig. 9;
These skeletal-sand accumulations on the modern shelf largely Cabioch et al., 1998). Although we know the present-day tectonic
consist of bryozoans, corals, barnacles, bivalves, and benthic setting of Espiritu Santo, and the Holocene reef terraces should
foraminifers and are associated with local patches (a few to tens be described as having formed in a forearc setting, it might be
of square kilometers in area) of gravelly sediment on rugged, cur- difficult to make the same tectono-stratigraphic interpretation if
rent-swept seafloor topography. similar carbonate terraces were found in the rock record. More
Espiritu Santo, Vanuatu. Holocene carbonate sedimentation likely, these reef terraces would be described as “intra-arc basin”
around the island of Espiritu Santo, Vanuatu, occurs above actively deposits (see discussion below about intra-arc carbonate facies),
uplifting volcanic-arc basement (cf. Cabioch et al., 1998). The simply because they are found stratigraphically above deformed,
Cenozoic tectonic history of Espiritu Santo is complex (Meffre arc-related rocks. Seismic profiles from offshore Espiritu Santo
and Crawford, 2001). The Vanuatu (or New Hebrides) archipelago also show drowned Quaternary carbonate reefs and reef caps (1–
can be divided into Western, Central, and Eastern Belts (Greene 6 km across, some reefs up to 300 m relief, reef tops at ~300 m
et al., 1988), which record discrete and geographically separated water depth) and “mounded” sediment bodies (375–450 m water
phases of volcanic activity. The Western Belt, which includes depth) (Fig. 10; Johnson and Greene, 1988; Greene et al., 1988),
Espiritu Santo, consists largely of upper Oligocene to middle Mio- which formed on submerged and complexly deformed parts of
cene calc-alkaline lavas and associated volcaniclastic rocks that the relict arc basement. The locations of some reefs and shelf
formed the volcanic arc above a westward-dipping but currently margins are probably fault controlled. These highly variable pat-
extinct subduction zone to the east. Subduction along the eastern terns of Quaternary carbonate sedimentation around Espiritu
side of Vanuatu ceased during late Miocene time and switched to Santo reflect the complex differential subsidence and uplift that
the western side of the archipelago, where an eastward-dipping can occur across young, incipiently formed forearc regions. In
subduction zone now exists. Late Miocene to Pliocene volcanic this case, Espiritu Santo represents subaerial remnants of an old
rocks of the Eastern Belt record the volcanic arc that formed above volcanic arc that were deformed and became basement for the
the east-dipping subduction zone. Late Pliocene–Holocene vol- forearc region of a younger volcanic arc.
canic rocks of the Central Chain record another westward shift It is worth noting that Quaternary carbonate facies appar-
of the volcanic arc toward the New Hebrides Trench. Remnants ently are extremely rare within forearc basins along the west-
of the older Oligocene–Miocene volcanic arc of the Western Belt facing, convergent margins of North and South America, even

40 S wave-cut
6 ka reef flat N
notches
30
Figure 9. Cross section of the uplifted
20 Tasmaloum fringing reef, Vanuatu,
coral reefs 6 showing the 6, 8, 10, 12, 14, 16, 18, and
Drill holes
10 8 20 ka time lines. Only the most char-
10 acteristic dates are reported (those in
present mean sea level
0 10,964
italics are 14C dates converted to calen-
12 dar years). A 6 ka reef flat is now 40 m
-10 above sea level, which shows the rapid
13,959 14
12,061
15,780 ? uplift rates that have characterized this
-20
14,463
18 16 ? forearc region over the last 6000 yr.
20 volcanic
20,961 ? From Cabioch et al. (1998).
substrate?
-30 21,303
>30,000 Late Pleistocene sand
-40
(m) 0 20 40 60 120 180 240m
Vanua
A 165°
Sa
170°
B Lava 0 50km

V OT
nt
a
Cr
uz
Is
la
14° S

TANDE TRO
n
Area
ds
HEBR
NORT TRENCH Vanikolo shown
N Santa
Basin
in B Maria
IDES
H NE

UGH
r
10 cm/y
Section
W

Vanua Lava Banks


E F.Z. shown
HOLM
Basin
Santa Maria HAZEL in C
SMFZ
Central FZ 15°
Basin AB
Espiritu
D’ENTRECASTEAUX Santo Aoba
PACIFIC PLATE
ZONE
Espiritu
ba
Ambrym
Malakula
AMFZ Santo Ao
EFZ
Epi

AUSTRALIA-INDIA PLATE

Pent
CO
RI

ecos
OL

Volcanic Associations VFZ

t
IS

Late Miocene to Holocene 16°


SOUT

(Central Chain)
TRO

Late Miocene to early Pliocene

NE TR
M
H

(Eastern Belt) Ambr


UG

al ym

W EN
Late Oligocene to middle Miocene ak
H

ul

HE CH
(Western Belt)
NE

BR
20°
W

Tectonic Elements

ID
E
HE

Uplifted ridge

S
BR

Rifted basin Epi


ID

167° E 168°
ES

Intra-arc basin
TR

Insular shelf
EN

Major fracture zone


CH

Trace of plate contact, dashed


where buried Insular slope
RE
0 100 200km ACTU
ER FR
Relative plate motion
HUNT
ZONE Deep basin

Quaternary
0
Quaternary
reefs C
1
UMiocene-
~3 km Pliocene
water 2 SECONDS (TWT)
Quaternary depth
East Santo slope deposits
basin fault
3
5 km Maewo
fault
4
Quaternary Eastern
North Aoba
167°E 168°E
basin fault
UMiocene-Pliocene 5
15°S
Maewo
Espiritu Aoba
Santo UOligocene-MidMiocene 6
Carbonate-platform facies in volcanic-arc settings 71

Figure 10. (A) Main geologic and tectonic features of the New Hebrides Arc. SMFZ—Santa Maria fracture zone; ABFZ—Aoba fracture zone;
AMFZ—Ambrym fracture zone; EFZ—Epi fracture zone; VFZ—Vulcan fracture zone. Modified from Greene et al. (1988). (B) Main geologic
and tectonic features of the central New Hebrides Arc, Espiritu Santo region. Basement rocks of Espiritu Santo consist of late Oligocene to
middle Miocene volcanic and volcaniclastic rocks erupted from a paleo-volcanic arc (Western Belt rocks) that formed above a now-extinct,
westward-dipping subduction zone. During late Miocene time, however, the subduction zone switched to the western side of the New Hebrides
Arc (present-day New Hebrides Trench), and an east-dipping subduction zone developed beneath arc rocks of the Western Belt, which includes
Espiritu Santo. The active volcanic arc to this subduction zone consists of volcanic centers that make up the Central Chain, to the east of the
Western Belt. Thus, Western Belt rocks now constitute the basement of the modern forearc region of the Central Chain, and Espiritu Santo is
in a transitional tectonic state. Depocenters that formed during extension of the older Oligocene–Miocene arc (Western Belt rocks) might be
considered intra-arc basins. The volcanic centers of the Western Belt are no longer active, however; so the present-day depocenters should be
considered part of the forearc of the Miocene–Holocene volcanic arc (i.e., the Central Chain). Modified from Greene et al. (1988). (C) Interpreted
seismic-reflection profile across the East Santo and North Aoba basins, New Hebrides arc system. Note Quaternary pinnacle reefs–buildups along
the outer, fault-controlled margin of the East Santo shelf. Modified from Greene et al. (1988).

those margin segments that lie within tropical climatic zones. frontal arc massif influenced the location, dimensions, and mor-
The limited amounts of modern carbonate sedimentation along phology of Neogene to Holocene reefs across the area (Fig. 11;
these margins probably reflects the strong coastal upwelling sys- Rodd, 1993).
tems that bring cold, nutrient-rich bottom water onto the shallow, Ancient examples. Ancient examples of forearc-basin car-
narrow shelf areas that border the western sides of these forearc bonate sequences include (1) Paleogene carbonate platforms on
basins. In contrast to modern continental margins, extensive transpressional ridges that transected the forearc basin of Costa
Paleozoic carbonate platforms formed along the eastern side of Rica (Krawinkel and Krawinkel, 1996); (2) Oligocene to Miocene
the paleo–Pacific Ocean because marginal oceanographic barri- reefs and platform facies (Tau, Wailotua, and Qalimare Lime-
ers (e.g., island arcs, orogenic belts, and microplates) may have stones) in the Outer Melanesian forearc basin on the island of
protected the carbonate platforms from the deleterious effects of Viti Levu, Fiji (Hathway, 1994, 1995; Rodd, 1993); (3) Miocene
upwelling (Whalen, 1995). Rethymnon Formation, eastern Crete (Pomoni-Papaioannou et
Quaternary carbonate reefs and shallow-platform facies are al., 2003); (4) Miocene mounded carbonate buildups to carbon-
also found in some relict forearc basin settings where subduction ate shelf systems in several sub-basins of the Indonesian forearc
is no longer active but where the basin’s bathymetry and structural (Beaudry and Moore, 1985; Matson and Moore, 1992; van der
features were inherited from previous times of active subduction. Werff, 1996); and (5) Pliocene “cool-water” skeletal limestones,
In some ways, these basins might be considered forearc succes- Hawke’s Bay, New Zealand (Kamp et al., 1988; Ballance, 1993;
sor basins. For example, the Indispensable Basin of the eastern Caron et al., 2004).
Solomon Islands has been interpreted as either a relict forearc The present-day Terraba belt represents the inverted Tertiary
basin or the leading edge of the partially subducted Ontong-Java forearc region of Costa Rica (Krawinkel and Krawinkel, 1996).
Plateau (Vedder and Bruns, 1989). The Ontong-Java Plateau This basin underwent three stages of basin evolution: (1) an ini-
resisted southwestward-directed subduction when it collided with tial transpressional phase during Late Cretaceous–Eocene time,
the Solomon Islands arc in middle to late Miocene time, forcing a which created the transpressional ridges that served as substrates
change in subduction polarity and initiation of a younger, north- for carbonate-platform facies; (2) a transtensional phase during
eastward-directed subduction zone beneath the southwestern Oligocene to early Miocene time, when the forearc region was
side of the Solomon Islands (Vedder and Bruns, 1989). Tectonic affected by transtensional fault systems and was subsiding rap-
structures and remnant bathymetry along the relict forearc side idly, causing the carbonate platforms to drown; and (3) a sec-
of the Solomon Islands continue to influence modern sedimenta- ond transpressional phase of deformation from middle Miocene
tion there, even ~15 m.y. after the change in subduction polarity. to Holocene time. Thus, tectonic deformation across the Costa
For example, seismic profiles along the southwestern slope of the Rican forearc basin initially created the substrates for carbonate
Indispensable Basin show several drowned Pleistocene(?) reefs sedimentation but ultimately caused the platforms to drown.
at ~300–500 m water depths along a steep, faulted(?) escarpment Upper Oligocene to Miocene carbonate intervals exposed on
just southeast of Santa Isabel Island (Niem, 1989). The steep the island of Viti Levu, Fiji, and imaged on seismic profiles from
bathymetric gradient across the forearc region and the low relief adjacent offshore areas consist of thin reefs or platform sequences
and relatively small surface area of the relict volcanic arc sug- that are interstratified with volcaniclastic strata (Figs. 11, 12;
gest that this relict forearc basin will probably remain underfilled Hathway, 1994, 1995; Rodd, 1993). The Qalimare Limestone,
for a long time into the future, thus limiting the potential surface which crops out on Viti Levu, is at least 300 m thick and consists
area for shallow-marine carbonate sedimentation. Bathymetric of coral-algal reefs and mounded facies that formed on the edge
profiles around offshore parts of Viti Levu, Fiji, are structurally of a shallow platform in front of the arc massif. Seismic profiles
controlled and record older Neogene to possibly Quaternary across the Bligh Water Basin from the northern offshore shelf
deformation. Faults that cut the older Neogene forearc basin and of Viti Levu have intrabasinal highs that served as substrates for
A 177°E 178° 179° 180°
0
1000
50

ORM

00
ATF

10
P L
EF
A RE 50
0
T SE
0

EVU
50
G REA L
UA
VAN 500
500 10
00
RM

00
17°S 10 500 WATER BASIN
FO
AT

H
PL

IG 150° 160° 170° 180°E

BL SOLOMON ISLANDS
A

500
AW

PAPUA 10°S
NEW GUINEA VANUATU
FIJI

00
50
S

10
0 CORAL SEA
YA

NEW
50 20°
0 CALEDONIA

AUSTRALIA
30°
VITI LEVU
REWA BAU TASMAN SEA
50
WATERS 0 40°
BASIN 1000km NEW
18° BASIN ZEALAND

00
0 50
100 0

10
10
00

50
BA 0 100 Pliocene sediments
0
R
AV
I Pliocene volcanics
BA
SIN
ZONE
Late Oligocene-Mid
N

500
SI

0 0 Miocene sediments
100
BA

50 & volcanics
VA

Late Eocene
SU

RE

sediments & volcanics


TU
AC

Basin outline
19°
FR

R Major fault
TE
0
50
N 0 20 40 km
100
0
HU

B
Carbonate-platform facies in volcanic-arc settings 73

Figure 11. (A) Geologic map of Fiji. Most of the platform areas are shallow-water carbonate environments, although some platform areas are
presently at subphotic depths (i.e., drowned or incipiently drowned). Also note the steep, typically fault-bounded edges of most platform areas,
as indicated by solid black lines. The islands that make up Fiji represent the eroded and faulted remnants of an older volcanic arc–forearc basin
system. Bathymetry in meters. From Rodd (1993). (B) Geologic cross section across Fiji and adjacent platform areas. This section shows various
isolated platforms and buildups in the relict Bligh Water forearc basin. Note how faults controlled platform locations, dimensions, and morphol-
ogy. From Rodd (1993).

local late Oligocene to middle Miocene reef development within packages (Kamp et al., 1988). These Pliocene skeletal sand bod-
the basin. ies formed as outer forearc basin deposits. The axes of the sand
The Miocene Rethymnon Formation (<60 m thick) in the bodies are oriented subparallel to faults and deformation fabrics
Apostoli Basin, eastern Crete, consists of skeletal-rhodalgal lime- within the subduction complex farther east, which suggests that
stone facies that were deposited on a warm, temperate ramp that the sand-body locations and orientations may have been con-
developed along the arcward side of the Neogene forearc basin trolled by complex surface deformations that developed above
in Crete. Skeletal and rhodalgal grain types in the Rethymnon parts of the subduction complex that lie beneath the outer forearc
Formation are indicative of nontropical carbonate facies, but the basin. The skeletal sand bodies were probably deposited during
presence of large benthic foraminifers suggests warm, temperate active broad wavelength folding of the forearc basin, as indicated
conditions (Pomoni-Papaioannou et al., 2003). The Rethymnon by (Fig. 13): (1) successively older Pliocene limestone strata
Formation onlaps the underlying Pandanassa Formation (allu- having progressively steeper dips and greater elevations (>300 m
vial fan and fluvial floodplain deposits) and Apostoli Formation above sea level) along the eastern flank of the basin; (2) succes-
(shoreface sandstone, Herostegina sands, and gray-marl facies), sively younger Pliocene carbonate strata, shifted laterally toward
suggesting general shallowing of the Apostoli Basin and possible the central axis of the forearc basin, especially along the more
uplift along the arcward flank of the basin during deposition of uplifted eastern flank of the basin; (3) older Pliocene carbonate
the Rethymnon Formation. sand bodies concentrated along the eastern and western flanks of
Seismic profiles across several late Oligocene–Holocene the forearc basin, grading laterally into siliciclastic facies in the
sub-basins of the western Indonesian forearc show carbonate basin axis, whereas younger carbonate intervals are thinner and
buildups and shelflike profiles of Miocene age that formed on the more sheetlike in character; and (4) unconformities within the
arcward side of these forearc depocenters (Beaudry and Moore, Pliocene section that merge, becoming composite unconformi-
1985; Matson and Moore, 1992). Most of the isolated buildups ties on the flanks of the forearc basin (Kamp and Nelson, 1987,
appear as mounded features that likely formed on structural highs 1988; Caron et al., 2004).
during transgressive events. The shelflike carbonate profiles may General tectono-stratigraphic model for carbonate facies
also record deposition during generally transgressive conditions. in forearc regions. Although detailed studies of carbonate strata
Both the carbonate buildups and shelflike platform sequences are deposited across forearc regions are relatively rare, previous stud-
overlain by progradational siliciclastic strata supplied from the ies of the tectono-geomorphologic evolution of forearc regions
arcward side of the basin during regressive events. provide insight into the possible stratigraphic evolution of car-
Pliocene strata in the forearc basin on North Island, New bonate systems that might form there. The evolution of carbonate
Zealand (Fig. 13), include coarse-grained, shallow-water, skel- platforms and reefs across the forearc region will be influenced
etal-sand sheets, which locally contain steeply dipping clinoform by (Fig. 6):

Sea level

Figure 12. Schematic tectonic and dep-


ositional model for southwestern Viti
Levu during early to middle Miocene
arc uplift = increased time. View approximately toward pres-
Isolated platform sediment flux to
on faulted high ent northeast. Note how fault-controlled
s if
forearc basin
s isolated platforms of the lower to mid-
c ma dle Miocene Qalimare Limestone were
Ar deposited on the frontal side of the vol-
canic arc, which was rising isostatically
because of crustal thickening. This in
turn caused more erosion of the arc and
greater flux of epiclastic sediment to the
forearc basin. From Hathway (1994).
74 Dorobek

Age N.Z.
(Ma) Stages LOCALITIES & LITHOSTRATIGRAPHY, HAWKE’S BAY, NEW ZEALAND

1.5 Limestone
400
Sandstone
2.0 Wn 180 Sandy mudstone
Covered
65
2.5 0 Unconformity
0
200
Wm
3.0 0
1000 200
Wp 0

3.5
0
35

4.0 Wo
0
0
0
4.5
15°SE Wn 5°SE 2°NW 8°-15°NW 18°-20°NW 10°NW
Wn Wo
Wm Wp
Wo Wm Tt
1 km Wp Wo
Forearc Basin Subduction Complex

Figure 13. Pliocene stratigraphy across the forearc basin of North Island, New Zealand. Measured stratigraphic sections at
specific locales across the forearc basin are shown in the upper part of this figure; stratigraphic thickness in meters. Note
the association of limestone with siliciclastic units and the progressive displacement of successively younger limestone
units toward the basin axis. The cross section does not show all the structures distributed across this forearc basin and
implies only broad, synformal folding over time. Most Pliocene limestone was deposited along the outer, trenchward side
of the basin, although younger carbonate units shifted progressively westward over time, toward the arc. Also note the
progressive decrease in dip of successively younger units (see lower cross-sectional view). These growth stratal patterns
reflect local surface deformation above fault-propagation folds and more regional differential subsidence/uplift across
the North Island forearc basin, which is expressed as the progressive westward shift of shallow-marine carbonate sedi-
mentation across the basin. New Zealand stage names and their relative ages include: Wn—Nukumaruan (late Pliocene-
Pleistocene); Wm—Mangapanian (late Pliocene); Wp—Waipipian (middle Pliocene); Wo—Opoitian (early Pliocene);
Tt—Tertiary. No vertical exaggeration. From Kamp and Nelson (1988).

1. Whether the subduction complex is accretionary or non- shift laterally off the flanks of the subduction complex.
accretionary. Accretionary subduction complexes grow The total area available for shallow-water carbonate
laterally as trench sediments are added to the frontal part sedimentation may actually decrease, especially if uplift
of the deforming wedge. Thus, accretionary subduction creates steep bathymetric gradients across the top of the
complexes commonly widen and shallow over time and subduction complex. If the upper surface of the subduc-
provide ever-increasing potential surface areas for shal- tion complex is below photic depths at the beginning,
low-water carbonate sedimentation. however, uplift may bring the surface into shallow-water
2. Differential uplift or subsidence across the forearc region. depths and allow a carbonate factory to develop. If the
For example, if a formerly submerged subduction com- top of the subduction complex is already emergent or
plex is uplifted and becomes emergent, any preexisting at shallow water depths, continued uplift of the subduc-
carbonate platforms or reefs across the crest of the sub- tion complex may produce progradational to downstep-
duction complex will become subaerially exposed, and ping stratal patterns within carbonate platforms or fring-
carbonate depositional environments will be forced to ing reefs that cap the subduction complex. In contrast,
Carbonate-platform facies in volcanic-arc settings 75

regional subsidence of the subduction complex (owing products or as epiclastic sediment) to the forearc. Silici-
to subduction erosion, flexure of the upper plate, or clastic deposits can serve as substrates for shallow-marine
other tectonic causes) might cause preexisting carbon- carbonate sedimentation during transgressive events if
ate platforms and reefs either to drown or to backstep sediment supply is relatively low.
toward any remaining shallow-water parts of the subduc- Shallow-marine carbonate strata may provide an excellent
tion complex. Similar complex chronostratigraphic and record of even subtle patterns of differential subsidence or uplift
facies relationships may develop above other types of across the forearc region. In many cases, dip-trending transects
actively growing intrabasinal structural highs within the through mature forearc basins show stratal patterns that indicate
forearc basin or along the proximal arcward side of the that the basins behaved more or less as broad, synformal dep-
forearc basin. ocenters, with smaller-scale, structural irregularities superim-
3. Transcurrent faults that can partition differential subsid- posed on this broader wavelength deformation. Across the area
ence or uplift across the forearc region. Transcurrent faults of the trench-slope break, deformation of the underlying subduc-
that cut the forearc region are most common where oblique tion complex typically creates a broad-wavelength (tens of kilo-
convergence occurs (e.g., Western Indonesian forearc). meters) uplift that is oriented subparallel to the trench axis. As
Transcurrent faults can partition parts of the forearc into previously described, wedge accretion and related tectonic defor-
antiformal highs or fault-bounded horst and graben seg- mation may bring the top of the subduction complex into shallow
ments, which control the location of platforms or isolated water depths and allow a carbonate factory to develop, or continue
reefs (cf. Krawinkel and Krawinkel, 1996). until the subduction complex becomes emergent and the area for
Given the complex subsidence and uplift histories of shallow-water carbonate sedimentation may actually decrease.
forearc regions, carbonate-platform and reef strata that form Uplift of the subduction complex may also force progradational
across the forearc region during active subduction will likely to downstepping stratal patterns within carbonate-platform strata
be thin (<100 m thick) and may be bounded by either subaerial- or fringing reefs that were deposited on the crest or flanks of the
exposure surfaces or drowning surfaces with complex chro- shallow subduction complex. In contrast, regional subsidence of
nostratigraphic relationships. The thickness and compositions the forearc region may cause carbonate platforms and reefs either
of interstratified noncarbonate facies will be highly variable, to drown, backstep, or shift to any remaining shallow-water parts
depending on: of the forearc region.
1. The evolving bathymetry and patterns of differential Smaller-scale seafloor irregularities across the trench-slope
subsidence or uplift across the forearc region, which break may be the surface expression of fault-propagation folds or
may change rapidly over relatively short time scales. It other surface deformations across the top of independently mov-
is possible for the forearc region to transition repeatedly ing, fault-bounded slivers within the subduction complex (e.g.,
between phases of subsidence and uplift, and regions of Pliocene forearc basin, North Island, New Zealand; Kamp and
subsidence and uplift may be closely juxtaposed in space Nelson, 1987, 1988; Kamp et al., 1988).
and time. The amount of time that lapses between epi- Reactivation of preexisting fault zones in forearc basement
sodes of uplift or subsidence is another critical factor that rocks or syndepositional deformation of older forearc-basin
determines whether episodic deformation forces carbon- strata can also create broad wavelength folds that influence facies
ate facies tracts to relocate to other shallow-water parts of patterns and growth stratal patterns within more axial parts of
the forearc region. the forearc basin. Intrabasinal forearc highs may (1) become sub-
2. The flux of volcanogenic and siliciclastic sediment into strates for shallow-marine carbonate facies if they reach shoal-
the forearc basin. Rapidly prograding mixed carbonate- water depths, (2) influence dispersal of sediment gravity flows
siliciclastic shelf facies may build from the volcanic arc shed from adjacent shallow-water platform areas if the highs are
into a forearc basin during later stages of basin evolu- more deeply submerged, or (3) be eroded by marine or subaerial
tion when the volcanic arc has built significant relief, processes as they are uplifted to progressively shallower depths
the basin is filled with sediment, and the basin is slowly or possibly even above sea level. Thus, the broad-wavelength
subsiding or uplifting. Progradational siliciclastic facies seafloor characteristics of intrabasinal fold axes would have simi-
commonly will be supplied by a series of alluvial and lar effects on sedimentation as surface deformations across the
deltaic point sources, where drainage networks reflect trench-slope break.
local climate and the irregular to ridgelike topography Lastly, transcurrent faults can partition the forearc region
of the volcanic arc. In humid settings, the arc may act into separate fault-bounded depocenters with intervening highs.
almost like a line source of sediment, with many local Transcurrent faults may strongly partition differential subsidence
drainages supplying volcaniclastic sediment to the or uplift across the forearc region and are especially common
forearc basin, especially along windward-facing slopes where oblique convergence occurs. Vertical offset on the trans-
of high-elevation archipelagos. Shallow-marine carbon- current faults may create steep bathymetric gradients that control
ate sedimentation may be suppressed where there is a the location of platform margins or isolated reefs on the interven-
large flux of volcanogenic material (either as eruptive ing highs between the faults.
76 Dorobek

Carbonate Platform and Reefs within Intra-Arc Basins that are submerged and within the photic zone, and the volume
of volcanogenic material dispersed to intra-arc basins is not great
Soja (1996) reviewed island-arc carbonate sequences, and enough to suppress carbonate production (cf. Bardintzeff et al.,
although the tectonic setting of some ancient examples cited 1985). Polymict breccias, however, dominate the stratigraphy in
therein is somewhat ambiguous, most probably formed within many intra-arc basins and may provide the only record of past
intra-arc basins or were at least built on volcanic edifices. Island- carbonate sedimentation. These breccias can be the dominant
arc carbonate sequences may characteristically have endemic lithofacies in intra-arc basin successions that are >5 km thick and
faunas, because many volcanic arcs are biogeographically iso- consist largely of volcanic and volcaniclastic clasts, with lesser
lated from more laterally extensive, land-attached carbonate amounts of sandstone, mudstone, and shallow-marine carbonate
platforms, where faunal radiation is more likely to occur (Soja, facies. Platform- and reef-derived clasts (with dimensions up to
1992, 1996). It is not clear if a tendency toward endemism has several hundred meters) within these sedimentary breccias may
any effect on carbonate-facies types, stratal architecture, or provide the only evidence for former shallow-marine carbonate
platform-reef morphology. systems in many intra-arc basins (cf. Soja, 1990).
Volcanic arcs may also serve as “refuges” for reef-building
faunas when local conditions become stressed or during global Intra-Arc Basin Classification and Tectonic Models
extinction events. For example, while most reef-building organ- Intra-arc basins are depocenters within volcanic arcs and
isms became extinct during Late Triassic time, a few coral species are underlain by arc-massif crust. Smith and Landis (1995) clas-
seemed to have survived into Early Jurassic time in volcanic-arc sified intra-arc basins as volcano-bounded, fault-bounded, or
settings that were subsequently accreted to western North Amer- hybrid types. Volcano-bounded intra-arc basins are depocenters
ica (cf. Stanley and McRoberts, 1993). found between constructional volcanic edifices. Volcano-bounded
Carbonate reefs and shallow-marine facies commonly fringe basins associated with continental-margin subduction systems
volcanic edifices within many intra-arc basins, especially if the typically have only thin sediment fills, which Smith and Landis
volcanic arc is at the proper (paleo) latitude. Carbonate reefs and (1995) suggested are due to the generally high-standing charac-
associated facies may be more common on the windward side of ter and low subsidence rates of their continental-crust substrates.
volcanic archipelagos because reef-building organisms prefer to In contrast, volcano-bounded basins from intraoceanic settings
construct frameworks on the windward side of islands and land- may accumulate thick volcanogenic strata, possibly because they
masses, and abundant ashfall deposition on the downwind side of subside more rapidly or begin as deep-water depocenters on oce-
the arc may inhibit carbonate sedimentation. Thick, areally exten- anic crust. Erosion rates, however, are generally much higher in
sive carbonate-platform sequences are apparently rare in intra-arc continental-margin arc systems than in intraoceanic arc systems,
basins because (1) these basins typically have steep bathymet- because continental-margin arc systems are largely subaerial fea-
ric profiles when they are underfilled, so there is little available tures, whereas volcanic highs and intra-arc basins may be largely
substrate for shallow-marine carbonate sedimentation; (2) there submerged in intraoceanic arc systems. Thus, erosion of volcanic
is abundant volcanic activity that inhibits carbonate sedimenta- highs also affects the amount of sediment delivered to intra-arc
tion; (3) there commonly are frequent phases of rapid subsidence basins and the rate at which they are filled.
or uplift, during which time carbonate platforms either might be Fault-bounded intra-arc basins are defined by fault networks.
drowned or terminated by subaerial exposure; and (4) frequent Fault patterns within most volcanic arcs and first-motion stud-
seismic events may trigger massive failure along the flanks of ies of intra-arc earthquakes indicate that extension or strike-slip
volcanic edifices, which can either incorporate and destroy large deformation causes most fault-bounded intra-arc basins to form.
stretches of preexisting platforms and reefs or bury shallow- Fault-bounded intra-arc basins typically have bounding faults that
marine carbonate environments with mass-flow deposits. trend parallel to the volcanic arc, although some basins and their
The thickest and most areally extensive carbonate platforms main bounding fault systems trend transverse to the volcanic arc.
and reefal buildups within intra-arc basins probably form during Fault-bounded intra-arc basins also may subside more rapidly
late stages of basin evolution or following the last phases of vol- than volcano-bounded depocenters (Smith and Landis, 1995).
canic activity when (1) volcanic edifices are likely to be subsiding Transverse oblique-slip fault systems that delineate some
rather than being uplifted or undergoing active volcanic construc- intra-arc basins may accommodate block rotations along the arc
tion; (2) the basins are more likely to be filled with volcaniclastic axis, such as in the Aleutian forearc region (Geist et al., 1988).
sediment, which will provide additional shallow-water substrates The curved trend of the Aleutian arc system, the oblique subduc-
for carbonate accumulation; (3) volcanic topography has been tion that occurs along this convergent margin, and the strong cou-
subdued by erosion so that a wider surface area is available for pling between the subducting and overriding plates may explain
carbonate sedimentation during sea-level highstands; and (4) vol- the block rotations. Differential subsidence of narrow crustal
canic eruptions and seismic events have ceased or are much less blocks between transverse faults may create submarine channels
frequent than during earlier stages of basin development. that funnel sediment gravity flows to the forearc basin or trench
Carbonate platforms and reefs can form during active phases slope. Arc-parallel strike-slip faults can also accommodate sig-
of arc volcanism as long as volcanic activity constructs features nificant lateral displacement of arc segments on either side of the
Carbonate-platform facies in volcanic-arc settings 77

fault zone (e.g., Semangko or Sumatra Fault, Beck, 1983; Philip- 1984; Bruns et al., 1989) and Vanuatu archipelagos (Carney et al.,
pine Fault, Karig et al., 1986; Sarewitz and Lewis, 1991). 1985; Greene et al., 1988).
Hybrid intra-arc basins have elements of both volcano- and The fault-bounded basement highs that delineate intra-arc
fault-bounded basins. Hybrid basins are defined by Smith and basins are parallel or oblique to arc segments. These structural
Landis (1995) as depocenters with fault-defined margins, but highs typically serve as (1) minor sediment source areas if they
where most of the topographic or bathymetric relief is defined are subaerially exposed, (2) substrates for shallow-marine sedi-
by constructional volcanic features. Arc massifs and associated mentation if they are only slightly submerged, or (3) bathymetric
intra-arc basins typically undergo complex deformation histories obstructions to sediment gravity flows if the structural highs are
and differential subsidence as a consequence of changes in sub- more deeply submerged. Although seismic data are limited from
duction parameters (dip angle, obliquity, and rate of subduction), recently rifted arcs like the Solomon Islands and Vanuatu, the
dynamic mantle flow, changes in surface loads, and magmatic arc-parallel or arc-crossing highs that separate intra-arc depocen-
activity. These large-scale tectonic and magmatic processes and ters within these arcs probably represent oblique transfer zones.
resultant patterns of uplift and subsidence within the volcanic arc Oblique dip-slip displacement on the faults that bound the intra-
change over time scales on the order of 5–20 m.y. (Smith and arc highs allows extensional strain to be transferred between nor-
Landis, 1995). As a result of the constantly changing tectonic mal fault systems within the arc. These intra-arc fault-bounded
and magmatic conditions that characterize the long-term history highs may influence sedimentation within the arc for a long time
of most volcanic arcs, intra-arc basins may transition rapidly after they form, especially if extension stops before the arc is
between volcano-bounded, fault-bounded, and hybrid types. completely rifted apart.
All styles of deformation have been recognized within volca- Transtensional, transpressional, and pure strike-slip styles of
nic arcs and their related intra-arc basins. Extensional and strike- deformation within the arc and intra-arc basins are more likely to
slip structures are generally more common within the arc massif develop where there is oblique subduction. As the convergence
than contractional structures, unless arc-continent collision has direction becomes more oblique or subparallel to the trend of
occurred. The history of deformation within intra-arc settings is the subduction system, pure strike-slip fault networks should
poorly preserved because it is obscured and overprinted by mul- become more common. Along some modern arc systems with
tiple eruptive and deformation events, thick sediment accumu- highly oblique subduction angles, regional strike-slip faults may
lations may deeply bury and obscure fault networks, or the arc cut through and translate long segments of the volcanic arc (e.g.,
and its related basins are eventually incorporated into collisional Sumatran or Semagang Fault along the island of Sumatra, west-
orogenic belts and original tectono-stratigraphic relationships ern Indonesia; Philippine Fault, Philippine archipelago).
are destroyed. Many arc systems also record flips in subduction Regional shortening within most volcanic arcs typically
polarity such that former forearc regions may become backarc or does not develop until arc-continent or continent-continent colli-
interarc settings, or vice versa. Preexisting faults within the arc sion occurs and magmatic activity within the arc has long ceased.
massif also may be reactivated by a different sense of displace- The limited amount of shortening within many volcanic arcs (and
ment during subduction-polarity flips or other changes in sub- backarc regions) may indicate that mechanical coupling between
duction characteristics. These complicated tectonic histories may the overriding and subducting plates diminishes with increasing
make it difficult to characterize the geometry, distribution, and distance from the trench (cf. Uyeda and Kanamori, 1979).
kinematics of fault networks that develop within a volcanic arc.
Extensional deformation is common where trench rollback Platform Types
causes the volcanic arc to split apart. An active seafloor-spreading Most carbonate sequences within intra-arc basins consist of
ridge may form if the arc is completely stretched apart, leaving a either fringing reefs constructed on volcanic or plutonic basement
remnant arc on the trailing side of the spreading axis and an active or barrier reefs with relatively narrow (<10 km wide) back-reef
arc on the other side. In these cases, intra-arc rift basins evolve and lagoonal facies. These fringing reefs and narrow platforms
into an extensional backarc setting. Intra-arc rifting may occur can build relatively steep depositional profiles, as indicated by
faster than in intracontinental rift settings, with correspondingly mountainside exposures of ancient examples and bathymet-
much faster rates of synrift subsidence (Yamaji, 1990). Some vol- ric profiles across modern intra-arc basins. Fringing reefs also
canic arcs, such as the Marianas arc system, have been repeatedly may form within collapsed calderas, although these probably are
split by intra-arc rifting and seafloor spreading (Hamilton, 1979). minor, short-lived accumulations that are terminated or blasted
Other arcs have been split apart during a flip in subduction polar- away by subsequent eruptions.
ity; changes in subduction polarity may not occur synchronously Carbonate platforms and reefs also form in fault-bounded
along all segments of the convergent margin. These factors may intra-arc basins where fault patterns and displacement histories
create problems with basin classification and understanding of strongly influence platform dimensions, the location of platform-
the tectonic processes that cause complex spatial and temporal margin facies tracts, and platform morphology. Small, isolated
patterns of uplift and subsidence. Examples of Cenozoic volca- platforms and pinnacle reefs are found on fault-bounded highs
nic arcs that have been rifted apart during changes in subduction within some intra-arc basins of Cenozoic island arcs in the South
polarity include the Solomon Islands (Johnson, 1979; Kroenke, Pacific. Local fault-bounded basement highs that border some of
78 Dorobek

these intra-arc basins serve as substrates for larger isolated plat- carbonate facies from the Grenadine Bank and some northwest-
forms. For example, Woodhall (1985) suggested that conjugate- ern islands, Lesser Antilles, D’Anglejan and Mountjoy, 1973;
fault patterns controlled the locations of volcanic vents and Qua- Bouysse, 1984; Dey and Smith, 1989; Quaternary Ucuna Lime-
ternary reefs on the Exploring Isles of the northern Lau Ridge. stone of the Neogene Takalau Limestone Group, Lau Ridge,
Other platform types are rare. Ramp-like to nonrimmed Woodhall, 1985; Quaternary carbonate units from islands of the
platforms and scattered mounded buildups may develop on the Western and Eastern Belts of Vanuatu, Macfarlane et al., 1988).
leeward sides of volcano-bounded intra-arc basins. Ramps may In a successor intra-arc basin the volcanic arc has been dormant
still have relatively steep depositional gradients owing to the or extinct for a long time, yet regional topography-bathymetry is
steepness of the underlying, volcanically constructed anteced- still strongly influenced by the inherited tectonic-volcanic geo-
ent topography. Leeward carbonate ramps may be relatively thin morphology. In some cases, carbonate facies may be deposited
in the rock record because volcanic ash from active volcanoes within an arc segment where the volcanoes are extinct or dor-
preferentially accumulates in leeward intra-arc depocenters and mant, even though other parts of the volcanic archipelago may be
suppresses or terminates carbonate sedimentation (cf. Eldredge active. Long periods of geologic time (10 m.y. or more) may also
and Kropp, 1985). pass between major eruptive episodes.
Intra-arc platforms and reefs exhibit complex patterns of This is the case in the Grenadine Islands, Lesser Antilles,
stratigraphic development, because intra-arc basins are character- where intra-arc carbonate facies form a submerged and incipi-
ized by complex spatial and temporal patterns of differential sub- ently drowned carbonate platform (the Grenadine Bank) that is
sidence and uplift during their evolution. During active tectonic currently at 20–40 m water depths. The Grenadine Bank prob-
deformation, platforms and reefs may rapidly expand or shrink ably formed during a Pleistocene lowstand of sea level on eroded
over short time scales (<100,000 yr), especially during icehouse remnants of the Lesser Antilles volcanic arc (D’Anglejan and
times when high-amplitude glacio-eustatic sea-level changes are Mountjoy, 1973; Dey and Smith, 1989). Active volcanoes are
more likely to occur (cf. Read, 1995). Intra-arc fault systems found within the same archipelago on the islands of Saint Vincent
commonly are steeply dipping and develop significant surface and Grenada, which are to the north and south, respectively, of
scarps (100 to >1000 m), and volcanic edifices also commonly the Grenadine Bank. Given that the Grenadine Bank is the same
have dramatic bathymetric gradients, both of which limit the area general distance from the subduction trench as are the active vol-
over which carbonate facies tracts can shift laterally during rela- canoes on Saint Vincent and Grenada, renewed volcanic activity
tive sea-level changes. During late stages of basin development, should be expected at some future time there. The lack of recent
when intra-arc depocenters are more likely to be filled, carbon- volcanic activity within the Grenadine Islands, however, and the
ate platforms and reefs may utilize volcanogenic successions as prevailing easterly wind allow extensive modern carbonate sedi-
substrates and develop into more areally extensive systems. Dif- mentation to occur on the eastern part of the Grenadine Bank
ferential subsidence and uplift across arc terranes, however, can (D’Anglejan and Mountjoy, 1973; Dey and Smith, 1989).
still be significant during later stages of basin evolution and can The Kallinago Depression in the northern Lesser Antilles
cause complex stratigraphic development within laterally exten- (Fig. 14) represents a volcano-bounded intra-arc basin that formed
sive carbonate platforms. when volcanism shifted from an outer (trenchward) Eocene
Quaternary examples. Quaternary intra-arc carbonate to early Miocene arc to an inner late Miocene to Holocene arc
platforms and fringing reefs are found in the Solomon Islands, farther west. The Kallinago Depression diminishes and merges
Fiji, and the Kadavu island group (Nunn and Omura, 1999; southward into a single arc near the island of Guadeloupe. Iso-
Nunn, 2000); Vanuatu (Macfarlane et al., 1988); Lau Ridge lated, early(?) Pleistocene buildups (e.g., Luymes Bank; Fig. 14)
(Woodhall, 1985); Tonga Ridge (Cunningham and Anscombe, have formed on the steep, volcanic substrates surrounding this
1985); the Indonesian archipelago (Sumatra, Java, Bali, islands; relatively young active arc.
Wells, 1988); the submerged Grenadine Bank and other, locally The Lau Ridge is another example where Quaternary car-
drowned pinnacles and isolated buildups on volcanic substrates bonate facies are accumulating in a remnant intra-arc setting.
elsewhere in the Lesser Antilles (D’Anglejan and Mount- The Lau Ridge represents the remains of a late Oligocene–early
joy, 1973; Bouysse, 1984; Dey and Smith, 1989); the Ryukyu Miocene volcanic arc that was left behind as a trailing, extinct arc
Islands of southwest Japan (Iryu et al., 1995; Nakamori et al., when the Lau Basin opened by seafloor spreading and separated
1995; Sagawa et al., 2001); and various parts of the Philippines. the Tonga Ridge from the Lau Ridge. Emergent to shallowly sub-
In many cases, Quaternary reefs and narrow fringing carbonate merged parts of extinct volcanic centers that compose the Lau
platforms have been uplifted to form a series of notched terraces Ridge continue to serve as substrates for Holocene carbonate
that are exposed near the modern coastlines of these islands. sedimentation (mostly skeletal sand facies; Woodhall, 1985).
These uplifted Quaternary limestone terraces attest to the rapid In the Vanuatu (New Hebrides) arc, Oligocene–Miocene
uplift rates (average rates up to 1 cm/yr) that characterize many volcanic and volcaniclastic rocks of the Western Belt of the
modern volcanic-arc settings. archipelago represent products from a volcanic arc that was
Many Quaternary reefs or shallow-marine carbonate facies rifted apart during middle Miocene (ca. 14 Ma) time. Renewed
have developed in successor intra-arc basins (e.g., Quaternary arc volcanism jumped eastward to form the mostly late Miocene
Carbonate-platform facies in volcanic-arc settings 79

A’ A A’
Luymes
1000 Anguilla Anguilla 0
3000 Bank

400
2000
1000 Shelf

0
20
Ka
0
llin Barbuda
A ag
o

20
0

0
St. Kitts
Kallinago

500
De
Depression 1
pr
Eocene-

TWT (sec)
es
Early Miocene Arc
si
on
Montserrat
0
100

Guadeloupe 2
20
00
Late Miocene-
Holocene Arc

lt?
f au
1000
200
St. Lucia 0
0 10 20 km
0 50 100 km 3
Contours in meters

Martinique

Figure 14. Map of northern Lesser Antilles and Kallinago Depression and interpreted seismic section. The Kallinago
Depression is a volcano-bounded intra-arc basin that formed between the inactive Eocene–early Miocene arc and younger
late Miocene–Holocene arc of the Lesser Antilles. Note how volcanic axes converge southward to a single volcanic arc
(i.e., Martinique). Interpreted seismic profile (location of section A–A′ shown in upper left of base map), showing slightly
deformed sedimentary fill within the Kallinago Depression (profile from Bouysse, 1984). Luymes Bank is a drowned
carbonate buildup whose top is at ~150–200 m water depth at present. Bathymetry in meters. Modified from Smith and
Landis (1995).

to late Pliocene Eastern Belt of the Vanuatu arc. Continued rift- Clifton Member of the Mount Peace Formation, Lucea Inlier,
ing between the two belts ultimately led to the youngest volcanic western Jamaica (Grippi, 1980); (4) the Oligocene(?) Paumbapa
centers within the Central Belt, which formed volcanically con- Formation on Sumba Island and offshore, Indonesia (Fortuin et
structed and fault-bounded highs during late Pliocene to Holocene al., 1997; Rutherford et al., 2001); (5) local mounded buildups
time (Macfarlane et al., 1988). Pliocene to Holocene carbonate that are scattered throughout the Tertiary section on islands and
facies form fringing reef complexes around extinct volcanoes of offshore structures of the Tonga Ridge (Cunningham and Ans-
the Western and Eastern Belts and around active volcanoes of the combe, 1985; Herzer and Exon, 1985); (6) Miocene to Pliocene
Central Belt. Subduction polarity, however, apparently changed reefal limestone and carbonate-platform facies (Tokalau Lime-
during Miocene time from west-dipping to east-dipping subduc- stone Group) on various islands of the Lau Ridge (Woodhall,
tion, and the subduction zone switched from the eastern to the 1985); (7) lower to middle Miocene skeletal limestones (Tari
western side of the arc. Given the complex tectonic history of the Formation and Qalimare Limestone) on Viti Levu, Fiji (Hath-
New Hebrides arc, it is difficult to classify some of the depocen- way, 1994, 1995); and (8) the Miocene Lelet Limestone, which
ters as either intra-arc or forearc basins because many are in a was deposited on the eroded remnants of the New Ireland vol-
protracted, transitional phase of development. canic arc, Papua New Guinea (Stewart and Sandy, 1988). These
Ancient examples. Ancient examples of intra-arc carbonate intra-arc carbonate facies typically are complexly interstratified
platforms, small banks, and reefs include (1) the Silurian Heceta with pyroclastic sands and tuffs, volcaniclastic breccias, coarse-
Formation, Alexander Terrane, southeastern Alaska (Soja, 1990); grained pebbly sands, siltstone, mudstone, and lava flows, and
(2) the Middle Devonian Kennett Formation, Lower Carbonifer- also are commonly cut by steeply dipping syn- and postdeposi-
ous Bragdon and Baird Formations, and Lower Permian McCloud tional faults.
Limestone, Eastern Klamath Mountains, California (Demirmen Other possible ancient examples of intra-arc carbonate
and Harbaugh, 1965; Watkins, 1985, 1993a, b; Watkins and sequences include a Lower Jurassic coral reef in the Telkwa
Flory, 1986); (3) Upper Cretaceous lenticular reef bodies in the Range, British Columbia (Poulton, 1989; Stanley and McRoberts,
80 Dorobek

1993); thin, Lower Jurassic coral biostromes in the Chocolate For- raised the arc substrate into the photic zone, thin carbonate pack-
mation, Arequipa massif, Peru (Wells, 1953; Stanley and McRob- ages may form rings around the volcanic substrates.
erts, 1993); Middle Jurassic coral biostromes from the Wallowa With continued arc development (in either intraoceanic or
Terrane, west-central Idaho, USA (Stanley and Beauvais, 1990); continental-margin arc systems), intra-arc basins may evolve
and Upper Triassic carbonate units (including coral reefs) of the into fault-bounded depocenters, where faults then control the
Martin Bridge Formation, Wallowa Mountains, Oregon, USA areal dimensions and locations of carbonate platforms and reefs
(Stanley, 1982; Stanley and Senowbari-Daryan, 1986), and at as well as the location of facies tracts and the overall morphol-
Lime Peak, southern Yukon Territory, Canada (Reid and Ginsburg, ogy of platform margin-to-basin transitions. Carbonate platforms
1986). Although the Mesozoic examples from North America are will likely have steep, fault-controlled margins in profile and
typically described as “island arc carbonates,” their specific paleo- have linear segments in plan view, both of which would reflect
geographic settings remain somewhat ambiguous. the influence of underlying fault zones. The trend of many lin-
Unlike Quaternary intra-arc carbonate platforms and reefs, ear platform-margin segments within fault-bounded intra-arc
for which only limited subsurface data are available, ancient basins might be predictable if the underlying fault zones develop
sequences provide better constraints on the thickness and internal with consistent orientations or as conjugate fault sets that reflect
facies of intra-arc carbonate platforms and reefs. Most intra-arc regional stress orientations.
carbonate sequences are several tens of meters thick, although As tectonic and volcanic activity wane during later stages
some are several hundred meters thick. Shallow-marine, muddy of basin development, and intra-arc basins progressively fill
carbonate facies of the Miocene Lelet Limestone on New Ireland, with sediment (largely volcaniclastic), carbonate platforms may
northeastern Papua New Guinea, may be an exception, as they become thicker and more areally extensive. Remnant volcanic-
reportedly are up to 1 km thick, with offshore equivalents of the and fault-controlled topography, however, may continue to influ-
Lelet Limestone ranging from 1 to 2 km thick (Exon et al., 1986; ence the dimensions of platforms and the locations of platform-
Stewart and Sandy, 1988). The lateral extent of ancient intra-arc margin-to-basin facies transitions until enough basin-filling sedi-
carbonate facies is usually much more difficult to constrain than ment has filled the antecedent topography.
for Quaternary examples, because ancient intra-arc basins may
be subjected to multiple phases of postdepositional deformation. Carbonate Platforms and Reefs in Backarc Basins

General Tectono-Stratigraphic Model for Intra-Arc Carbonate sedimentation also occurs in backarc settings on
Carbonate Platforms and Reefs a variety of tectonic structures and bathymetric highs constructed
The different types of intra-arc basins influence the types of by depositional processes. As in other arc depocenters, the areal
carbonate platforms and reefs that form in them. In general, the extent, thickness, morphology, and internal stratigraphy of car-
rapid rates of tectonic deformation in intra-arc settings, the likeli- bonate platforms and reefs will be determined by the tectonic
hood of catastrophic eruptions, and the close proximity of high- history of the backarc and whether it is filled with sediment.
relief source areas for siliciclastic sediment (i.e., the volcanic arc)
cause carbonate sedimentation to be only intermittent and short- Backarc-Basin Types and Tectonic Models
lived within intra-arc basins. Backarc basins form on the back side of subduction-related
Fringing reefs and narrow platforms are the likely domi- volcanic arcs (Fig. 16). Extensive geophysical surveys of many
nant type of carbonate systems found in most intra-arc settings modern backarc regions over the last 20 yr or so have led to
(Fig. 15). These rarely attain significant thickness (i.e., >100 m the recognition of three main types of backarc basins (Dick-
thick) before a catastrophic eruption or mass-wasting event inson, 1974; Karig, 1983; Ingersoll, 1988; Marsaglia, 1995):
buries the carbonate system. Mass-wasting events also might (1) extensional backarc basins that form by rifting and seafloor
incorporate carbonate strata, causing large-scale disruption or spreading within or behind continental-margin or intraoceanic
transport of shallow-marine carbonate facies into deeper-water arcs, (2) remnant-ocean backarc basins that form by entrap-
settings. Displaced blocks of reef and shallow-water carbonate ment of old oceanic crust and are associated with intraoceanic
facies within deep-water shale and volcaniclastic deposits may subduction zones, and (3) compressional backarc basins, which
provide the only evidence of the former intra-arc platforms. are more commonly classified as retro-arc foreland basins (see
There may be differences between intra-arc carbonate suc- discussion below). Marine backarc basins may be underlain
cessions from intraoceanic and continental-margin arc systems. by continental to transitional crust if they form by extension
In intraoceanic arcs, shallow-marine carbonate facies will likely behind a continental-margin arc system. In contrast, backarc
not appear within intra-arc basins until the volcanic arc has built basins will be underlain by highly subsided oceanic crust if the
itself into the photic zone. Volcano-bounded and hybrid intra-arc basin forms along an intraoceanic subduction zone, and rem-
basins may dominate over fault-bounded basins in young intra- nant oceanic crust is trapped behind the newly formed volca-
oceanic volcanic arcs. The basal part of the basin-filling sequence nic arc. Backarc basins also can be underlain by young oceanic
in these basins will likely consist of deep-water volcaniclastic crust if they form where rifting in the backarc region ultimately
and pelagic-hemipelagic deposits. After volcanic activity has leads to seafloor spreading.
Carbonate-platform facies in volcanic-arc settings 81

VOLCANO-BOUNDED INTRA-ARC BASINS


platform types: FAULT-BOUNDED INTRA-ARC BASINS
• isolated platforms & buildups, commonly platform types:
steep-sided • isolated platforms & buildups
• thin fringing platforms & reefs during early • similar controls & relationships as for
stages; complexly interstratified with volcano-bounded basins (windward side,
volcanics, volcaniclastics complex chronostratigraphy, etc.)
• may be preferential development on • extensional deformation common; post-rift
windward side of arc thermal subsidence will affect
• complex chronostratigraphic relationships stratigraphic development
between individual carbonate sequences • internal growth strata possible if platforms
• may be drowned platforms adjacent to on actively deforming fault-bounded highs
actively growing platforms substrates:
• platforms become larger & thicker as • volcanics, volcaniclastics, faulted arc
depocenters fill, but eruptions can have basement or volcanics/volcaniclastics
catastrophic effects on platforms • platforms become larger & thicker as
substrates: depocenters fill or as deformation wanes
• volcanic edifices & volcaniclastic fill

Volcano-bounded Fault-bounded
intra-arc basins intra-arc basins
Figure 15. Types of carbonate platforms and buildups within intra-arc basins. Platform thicknesses, widths, and relief are
not to scale. Only generalized platform types and buildups are shown. Isolated platforms and fringing platforms-reefs
are most common. Stratigraphic relationships and depositional contacts with structural features and various geomorpho-
logic surfaces across intra-arc basins are also highly generalized. Modified from Smith and Landis (1995) to emphasize
carbonate-platform facies.

Extensional backarc basins are commonly associated with (commonly a continental margin) that defines the distal side of
subduction zones where trench rollback occurs. Many exten- the basin (see below).
sional backarc basins actually begin as fault-bounded intra-arc Backarc regions typically are areas of overall crustal exten-
basins, when a volcanic arc begins to rift apart. An extensional sion or strike-slip deformation. Crustal shortening is rare in most
backarc basin might come into existence when there is the backarc regions except for some continental-margin arc systems
first evidence for a topographically expressed, but volcanically where the subducting and overriding plates are strongly coupled
extinct, remnant arc that is separated from an active volcanic (e.g., segments of the Andean and western Indonesian backarc
arc by a rift zone. There likely will be a distinct bathymetric regions and the Sea of Japan backarc region). In cases where sig-
or topographic axis to the newly formed extensional backarc nificant crustal shortening occurs in the backarc region of con-
basin at this point in its tectonic evolution. This definition for tinental-margin arc systems, the flexural basins that develop are
an extensional backarc basin does not depend on the amount of more appropriately described as retro-arc foreland basins, which
stretching across the rift zone or on whether rifting has ceased are not considered further in this paper.
or seafloor spreading has begun. The remnant arc also can be The styles and intensity of deformation in the backarc
both a source area for siliciclastic sediment and a substrate for region depend largely on (1) the tectono-evolutionary stage (or
shallow-marine carbonate sedimentation, although carbonate maturity) of the arc system; (2) the relative motions (e.g., angle
facies may be deposited only during early stages of the basin’s of convergence) between the subducting and overriding plates;
history, before the remnant arc has undergone significant ther- (3) the age, thickness, dip, and crustal type of the subducting slab;
mal subsidence. (4) possible changes in subduction polarity during convergence;
Backarc basins that are underlain by trapped fragments of (5) whether trench rollback is possible; (6) whether a spreading
older oceanic crust have different tectonic and subsidence histo- ridge, seamount, or other large-scale seafloor features have been
ries than those that form by backarc extension. Remnant-ocean subducted; or (7) whether arc-continent or continent-continent
backarc basins are associated with intraoceanic subduction collision has begun.
zones, and their dimensions depend on where subduction is initi- Extensional backarc regions are dominated by normal-fault
ated within the overriding plate and the nearest bathymetric high networks that are subparallel or slightly oblique to the trend of
82 Dorobek

old oceanic crust


trapped by intraplate
concentrated along bends in the rift system that are interpreted
subduction zone as accommodation zones.
Modern examples of backarc extension behind continen-
A tal-margin volcanic arcs include the Andaman Sea, the Okinawa
0 100 km Trough, and the Japan Sea. Highly oblique subduction from west-
ernmost Indonesia northward to the continental margin of Myan-
mar, strong coupling between the subducting Indian-Australian
plate and overriding Eurasian plate, and subduction of an aseis-
new oceanic crust mic ridge may explain the backarc rifting and seafloor spreading
created by in the Andaman Sea (Eguchi et al., 1980). The Okinawa Trough
back-arc spreading
is an incipient continental backarc basin and displays similar
B structural features as the Sumisu rift zone (Letouzey and Kimura,
1986). The Japan Sea is a fully developed continental-margin
0 100 km
backarc basin, where contraction or transpression has occurred
along its eastern side, and subduction may be in its beginning
stages (Kikuchi et al., 1991; Tamaki, 1995; Okamura et al., 1995;
continental crust beneath
retro-arc foreland basin
Takano, 2002).
(compressional arc system) Retro-arc foreland basins are flexural depocenters that result
from crustal shortening on the continental side of the volcanic
C arc; examples include the Andean and western Indonesian “back-
arc” settings. In comparison, crustal shortening within backarc
0 200 km basins (sensu stricto) along intraoceanic convergent margins is
apparently rare. The limited amount of shortening in the backarc
region of most intraoceanic volcanic arcs probably reflects weak
rifted continental crust beneath
extensional back-arc basin
coupling between the subducting and overriding plates along
these types of convergent margins.
D
Platform Types
0 200 km The lateral extent and cumulative thickness of carbonate
facies within backarc basins will vary, depending on the tectonic
origin(s) and evolutionary stage of the basin (Fig. 17). For back-
Figure 16. Cross sections through different types of convergent mar- arc basins that form by entrapment of old oceanic crust behind
gins, showing crustal types of backarc regions. (A) Intraoceanic volca- the arc, carbonate facies may be entirely limited to shallow sub-
nic arc, with old oceanic crust trapped in backarc region. (B) Backarc
region underlain by newly created oceanic crust created by spreading
merged areas on the backarc side of the arc massif, and there
axis. (C) Compressional backarc region (retro-arc foreland basin sys- may be no other shallow-water substrates for carbonate platforms
tem) underlain by continental crust. (D) Extensional backarc region across the backarc region. Most emergent or shallow-water sub-
underlain by rifted continental crust. strates in these backarc basins will be constructed by volcanic
activity. Continued growth of carbonate platforms and reefs in
these backarc settings may be frequently interrupted by volcanic
eruptions or mass-wasting events so that reefs and other carbon-
the volcanic arc. As in other extensional settings, oblique-slip ate facies must be reestablished after a catastrophic event destroys
transfer faults trend at high angles to basin-bounding normal or buries them (Eldredge and Kropp, 1985).
faults and accommodate displacement transfer between offset or In contrast, extensional backarc basins will have bathymet-
nonparallel normal faults. This style of extensional deformation ric profiles that reflect stretching of the arc massif or other types
is characteristic of the 700-km-long Izu-Bonin intraoceanic arc of backarc lithosphere. During initial rifting of the volcanic arc,
system, which is considered to be an example of an extensional carbonate sequences may be deposited in intra-arc basins (see
arc system where an incipient backarc basin(s) is forming (Taylor above). Where stretching continues to the point of whole litho-
et al., 1990, 1991; Klaus et al., 1992). The extensional sub-basins sphere failure and a seafloor-spreading ridge develops, a backarc
along the Isu-Bonin arc are typically asymmetric, 25–110-km- basin is clearly identifiable, and trailing rifted margins develop
long and 25–40-km-wide features. Most basin-bounding normal on either side of the basin. There is no consensus, however, as to
faults trend subparallel to the axis of the arc system. Similar to when an intra-arc basin structurally, bathymetrically, or sedimen-
continental rift systems, oblique transfer faults and accommo- tologically transitions into a backarc basin.
dation zones link oppositely dipping normal faults that bound Apparently there are no preserved examples of thick, long-
the rift depocenters. Submarine volcanic centers appear to be lived carbonate platforms within backarc basins. Backarc basins
INTRA-OCEANIC BACKARC SETTINGS COMPRESSIONAL BACKARC SETTINGS
A pinnacle reef or B smaller isolated ramp sequence ramp sequence
outer reef tract platforms, shoals, and reefs constructed on older constructed on foreland
shelf/ramp sequence
may be on antiformal highs above foreland basin strata side of basin
constructed on older
fault-controlled back-arc thrust belt (commonly post-orogenic)
volcaniclastic strata

TRAPPED OCEANIC CRUST


RETRO-ARC FORELAND BASIN
0 50 km 0 200 km

more regional ramp/shelf sequences


during post-rift stages; flexural

EXTENSIONAL BACKARC SETTINGS: C may contain syn-rift platforms


on fault-bounded highs
onlap and post-rift thermal
subsidence; substrates include
C. Backarc intracontinental rifts if depocenters are marine post-rift fill and continental basement
D. Backarc intracontinental rifts leading
to spreading and conjugate
passive margins
E. Intra-arc rifting leading to spreading
0 200 km
EXTENSIONAL BACK-ARC BASINS

may contain syn-rift platforms


on fault-bounded highs thermally subsiding may contain syn-rift platforms
if depocenters are marine, conjugate passive margin;
D but more extensive ramps/shelves
form during post-rift phase
similar controls on regional platform E on fault-bounded highs
if depocenters are marine,
thermally subsiding remnant
volcanic arc provides substrate
development as passive margins but more extensive ramps/shelves for short-lived isolated
during post-rift phase platforms and reefs

BACK-ARC SPREADING RIDGE


BACK-ARC SPREADING RIDGE
0 100 km
0 100 km

Figure 17. Different types of backarc settings and regional carbonate platforms that may form in them. In all examples shown, the subduction
zone and forearc region (not shown) are left of the volcanic arc (indicated by triangle). Horizontal scales are generalized; no vertical scale implied.
Only general patterns of platform development are shown; internal stratigraphy and details about platform morphology will vary significantly in
each backarc setting. (A) Intraoceanic volcanic arc, with old oceanic crust trapped in backarc region. Carbonate platforms and reefs can form on
the backarc side of intraoceanic arcs. In general, substrates for these backarc carbonate systems will be constructional volcanic edifices, eroded
arc-massif basement, or volcanic-volcaniclastic deposits. Local fault-controlled bathymetry may control platform dimensions and facies patterns.
(B) Retro-arc foreland-basin system. Extensive carbonate sedimentation can occur in retro-arc foreland basins if they are slightly underfilled
and the continental foreland is below sea level. Thus, carbonate sedimentation is probably more common during underfilled stages of retro-arc
foreland-basin development. Carbonate ramps are common, especially along the distal (foreland) side of the basin. Various styles and intensities
of deformation across the foreland can significantly affect carbonate sedimentation. Foreland-basin carbonate systems are discussed in more
detail in later sections. (C) Extensional backarc region underlain by rifted continental crust. If rifted continental basement is flooded early, synrift
carbonate platforms may form on fault-bounded basement highs. More extensive platforms may develop during late post-rift stages, when the
extensional backarc basin is nearly filled and thermal subsidence dominates. (D) Backarc region underlain by newly created oceanic crust created
by spreading axis. Where stretching in the backarc region progresses to the point of whole lithosphere failure, a seafloor spreading axis will de-
velop. If stretching began within an intra-arc setting, the trailing remnant arc becomes extinct and undergoes relatively rapid thermal subsidence.
The remnant arc can serve as a substrate for carbonate-platform development, although these are usually short-lived platforms. (E) If stretching
began within an intra-arc setting, conjugate passive margins would develop on both sides of the newly created backarc ocean basin. Both conju-
gate margins should behave like other passive margins. Thus, carbonate platforms and reefs should become more areally extensive over time as
the margins undergo thermal subsidence and progressively wider potential substrates become available for carbonate sedimentation.
84 Dorobek

that are underlain by trapped oceanic crust are deep-water set- locations reflect the ongoing faulting and underfilled character
tings almost from their inception, so shallow-marine carbon- of many Quaternary backarc basins, which typically are in early
ate facies will not be deposited until the volcanic arc shoals to stages of development (i.e., <5 m.y. old). Carbonate platforms
shallow-water conditions. Even after the volcanic arc has built and reefs are rare in more tectonically mature backarc basins,
itself up to shallow-water depths, backarc carbonate facies gen- which have undergone more tectonic subsidence, and potential
erally will consist only of fringing reefs or narrow rimmed-shelf substrates for carbonate sedimentation are below photic depths.
platforms that are constructed on the steep-sided volcanic edifice. Ancient examples. Ancient examples of backarc carbonate
More areally extensive carbonate platforms can be constructed platforms and reefs include (1) thin intervals of skeletal limestone
only after appropriate substrates for shallow-marine carbonate in the dominantly siliciclastic lower member (150 m total thick-
sedimentation have been constructed, which requires some com- ness) of the middle Cretaceous Olvidada Formation, Baja Cali-
bination of erosion of the volcanic edifice (by both subaerial and fornia Norte, Mexico (Phillips, 1993); (2) Upper Triassic (Norian)
submarine processes) and deposition of volcaniclastic sediment coral buildups and carbonate foreslope(?) “conglomerates” (70 m
and lava flows. total thickness), Puale Bay, Alaska Peninsula (Wang et al., 1988);
Similarly, most carbonate sequences within extensional and (3) possibly various Triassic carbonate-platform sequences
backarc basins consist of fringing reefs or relatively narrow from Western North America (Stanley, 1982). The interpretation
ramps and rimmed shelves that are constructed (1) on the backarc of many of these examples as backarc-basin carbonate sequences
side of the active volcanic arc and arc massif, (2) for a short time is based largely on paleogeographic reconstructions where the
on the remnant arc as it drifts away from the active arc, or (3) on relative position of the volcanic arc and the direction of subduc-
fault-bounded basement highs that partition extensional backarc tion are known. Future work may ultimately show, however, that
regions into sub-basins. Postrift thermal subsidence ultimately some of these examples formed in forearc or intra-arc basins.
submerges the drifting remnant arc to subphotic water depths so
that any productive carbonate factories on the remnant-arc crust General Tectono-Stratigraphic Model for Carbonate Facies
are soon terminated. Scattered pinnacle reefs and mounded build- Deposited across Backarc Regions
ups that are typically less than a few hundred meters thick may General tectono-stratigraphic models for carbonate platforms
be constructed on synrift topography in backarc regions. These and reefs that form in backarc settings depend on whether the
buildups apparently are drowned soon after they form because of backarc basin is extensional or nonextensional (Fig. 17). Given
the combined effects of rapid subsidence and short-term eustatic the relative randomness of eruptive events and shifting dispersal
sea-level rise. patterns for volcaniclastic sediment, there may not be a general-
Compressional backarc settings are more like retro-arc fore- ized and recurrent stratigraphic succession that develops in back-
land basins, so they are not discussed here. Regardless, carbonate arc regions. Carbonate sedimentation, however, is probably more
platforms and reefs are common in these settings, forming on common, long-lived, and areally extensive during later stages of
fault-bounded highs that result from inversion tectonics. Isolated backarc basin evolution, when subsidence rates should be slower,
platforms and buildups are common morphologies. volcanic activity is waning, deposition of volcaniclastic and epi-
Quaternary examples. Quaternary carbonate platforms and clastic sediment has created broad substrates for carbonate plat-
reefs are found in the backarc basins of the eastern Indonesian form development, and arc topography could be highly eroded so
archipelago, the Marianas, and Fiji. Quaternary carbonate reefs that less sediment is supplied to the backarc region.
and platforms from modern backarc settings may overlap, either In remnant-ocean backarc basins that form where older oce-
geographically or conceptually, with intra-arc carbonate systems, anic crust is trapped behind an intraoceanic volcanic arc, carbon-
especially where fault-bounded intra-arc basins evolve into back- ate facies will likely develop only after volcanic edifices have built
arc basins during progressive stretching of a volcanic arc. themselves up to within photic depths (Fig. 17A). Narrow shelves
Most Quaternary carbonate systems in backarc basins consist or fringing reef systems may form along the backarc side of the
of (1) predominantly fringing reefs or barrier reef tracts with nar- newly formed volcanic arc. Carbonate platforms can become
row back-reef “lagoons” that are constructed on the backarc side more areally extensive on the backarc side of these intraoceanic
of the volcanic arc, (2) drowned pinnacle reefs or mounded car- arcs when deposition of volcaniclastic sediment and lava flows
bonate buildups that also typically form on fault-bounded highs have built broad, shallow-water substrates. Active volcanism
in the backarc region, or (3) relatively narrow platforms that are and tectonic deformation of the volcanic arc will prevent thick
constructed on faulted backarc basement highs that either extend carbonate accumulations from developing, but as volcanism and
from the volcanic arc or serve as accommodation zones that par- deformation wane, thicker platform sequences might develop.
tition extensional backarc regions into separate sub-basins. The Thus, the oldest carbonate sediments in remnant-ocean backarc
present-day abundance of fringing reef tracts and drowned build- basins likely record when the growing volcanic arc first shallowed
ups reflects both the high-amplitude (i.e., >100 m) character of into photic depths. These initial deposits might be followed by
the post-Pleistocene sea-level rise and the typically steep struc- complexly interlayered sequences of thin carbonate facies with
tural or volcanic topography that was flooded by this rise in sea thicker volcaniclastic deposits and lava flows. The thickest and
level. The obvious structural controls on platform dimensions and perhaps most areally extensive carbonate sequences record the
Carbonate-platform facies in volcanic-arc settings 85

end of volcanic and tectonic activity in remnant-ocean backarc Tectonically Influenced Unconformities in Arc Depocenters
basins. Rare(?) crustal shortening might occur at any evolutionary
stage in remnant-ocean backarc basins and could generate surface Unconformity development across carbonate platforms in
deformations (probably related to fault-propagation folding) or arc depocenters can affect the distribution of porosity and per-
cause regional flexural subsidence across the backarc region that meability within these platforms. Tectonically enhanced uncon-
might also influence carbonate-sedimentation patterns. formities will be best developed within carbonate platforms that
In contrast, carbonate sedimentation in extensional backarc form where tectonic uplift may occur, such as:
basins will likely be influenced by the fault systems that accom- 1. The crest of accretionary subduction complexes, where
modate stretching (Fig. 17C–E). Initial stretching will likely the internal deformation history of the subduction com-
occur within the hot and weak lithosphere that underlies the vol- plex controls the character of surface deformations that
canic arc, so arc-related basement rocks may provide initial sub- affect unconformity development. Disconformities that
strates for the earliest carbonate strata. Most carbonate platforms might cap shallowing-upward carbonate sequences on
in modern extensional backarc basins consist of fringing reefs or the crest of the subduction complex can grade laterally
narrow rimmed shelves that build from the steep tectonic, volca- into conformable contacts toward more axial parts of the
nic, or erosional topography on either side of the basin. There are forearc basin (Caron et al., 2004).
no reported examples of synrift platforms or reefs within more 2. Intra-arc basins, which might be uplifted as a volcanic arc
axial, deep-water parts of modern extensional backarc basins, but grows. Entire isolated platforms or only minor segments
they should be recognizable in ancient backarc basins by their of barrier- or fringing-reef systems may be capped by tec-
association with extensional fault systems and by their internal tonically enhanced unconformities in intra-arc settings,
growth stratal patterns. As stretching proceeds, the evolving depending on the scale of arc uplift and whether uplift
synrift topography must influence the location, dimensions, and patterns are affected by local faulting.
internal growth stratigraphy of carbonate platforms and reefs, Carbonate platforms in extensional and neutral backarc
just as in marine intracontinental rift settings. If extension pro- basin settings are not likely to have major, tectonically enhanced
gresses to the point at which a spreading axis develops, the rifted unconformities because these settings are generally undergoing
basement of the tectonically separated remnant and still-active subsidence, with little uplift throughout much of their history. In
arc terranes may begin to subside independently. The remnant contrast, compressional backarc settings (i.e., retro-arc foreland
arc will likely undergo faster rates of postrift subsidence than the basins) may have tectonically enhanced unconformities in the
active-arc side of the extensional backarc basin, because there wedge-top, forebulge, and backbulge depocenters (cf. Dorobek,
will not be a nearby heat source that will prevent the remnant 1995; Jordan, 1995; DeCelles and Giles, 1996).
arc from thermally subsiding. In fact, continued volcanic activ-
ity, changes in the properties of the subducting slab, or various SUMMARY
subduction-related geodynamic processes may even cause uplift
along the active-arc side of the extensional backarc region. Carbonate platforms and reefs provide sensitive records
Carbonate sedimentation, however, may be most extensive in of differential subsidence and uplift, environmental conditions,
extensional backarc basins during their late synrift to early postrift and interactions with other depositional systems in arc settings.
(i.e., “drift”) stages of development when there are numerous In general, detailed stratigraphic studies are lacking for all arc-
potential substrates across the backarc region where fringing and related depocenters because original stratal relationships are usu-
isolated platforms can develop (cf. Dorsey and Kidwell, 1999). ally poorly preserved, physical access in many arc settings (both
These isolated platforms will reflect the strong influence of rem- ancient and modern) is difficult, high-quality seismic profiles and
nant synrift topography, where fault-bounded highs are substrates extensive well data are not widely available, and actual strati-
for carbonate platforms and structural lows trap siliciclastic and graphic relationships are highly complex in three dimensions.
volcaniclastic sediment from the adjacent active and remnant arcs. Thus, it is difficult to construct general stratigraphic models for
Later in the postrift history, when sediments have filled the rem- any arc depocenter.
nant synrift topography, broad depositional surfaces may serve as Long-wavelength tectonic-subsidence patterns in arc sys-
substrates for carbonate platforms during relative sea-level rises. tems will control the spatial distribution of shallow-water car-
Along the remnant-arc side of the extensional backarc basin, the bonate facies over long time scales and ultimately will control
inactive arc will be erosionally beveled and continue to undergo the thickness of carbonate strata. Local, fault-controlled dif-
thermal subsidence. Sediment supply from the remnant arc and ferential subsidence will influence the sizes, morphology, and
subsidence rates should progressively decrease over time, which internal facies patterns of carbonate platforms that develop in
may also enhance development of broad carbonate platforms any arc depocenter.
during later postrift phases. In contrast, carbonate sedimentation Siliciclastic and volcaniclastic flux will come largely from
along the active-arc side of the basin may never become areally the volcanic arc or arc-massif rocks. Carbonate facies may be bet-
extensive or long-lived as long as there is continued flux of volca- ter developed with greater distance from a volcanic arc, although
nogenic material to the backarc region. carbonate facies may even accumulate along the flanks of the arc
86 Dorobek

if drainage patterns divert siliciclastic diment elsewhere, volcanic Bradley, D.C., 1989, Taconic plate kinematics as revealed by foredeep stratigra-
eruptions are infrequent or are of low volume, or major wind- phy, Appalachian orogen: Tectonics, v. 8, p. 1037–1049.
Bruns, T.R., Vedder, J.G., and Culotta, R.C., 1989, Structure and tectonics
ward-leeward effects are caused by arc topography. along the Kilinailau Trench, Bougainville–Buka Island region, Papua
Carbonate-platform morphology can be highly variable New Guinea, in Vedder, J.G., and Bruns, T.R., eds., Geology and Offshore
across arc depocenters and is largely dependent on the availability Resources of Pacific Island Arcs—Solomon Islands and Bougainville,
Papua New Guinea Regions: Houston, Circum-Pacific Council for Energy
of tectonic or depositionally constructed substrates. For tropical and Mineral Resources, Earth Science Series, v. 12, p. 93–123.
carbonate systems, these substrates generally need to be at photic Burbank, D.W., and Anderson, R.S., 2001, Tectonic Geomorphology: Cam-
depths. Cool-water carbonate facies in arc depocenters, however, bridge, Massachusetts, Blackwell Science, 274 p.
Cabioch, G., Taylor, F.W., Récy, J., Lawrence Edwards, R., Gray, S.C., Faure,
might form over a broader range of water depths, but they will G., Burr, G.S., and Corrège, T., 1998, Environmental and tectonic influ-
still be limited by the availability of suitable substrates. Other ence on growth and internal structure of a fringing reef at Tasmaloum
factors, such as nutrient levels and wind direction, also influence (SW Espiritu Santo, New Hebrides island arc, SW Pacific), in Camoin,
G.F., and Davies, P.J., eds., Reefs and Carbonate Platforms in the Pacific
carbonate-facies development in arc depocenters, although the and Indian Oceans: International Association of Sedimentologists Special
relative influence of these controls is highly specific to any par- Publication 25, p. 261–277.
ticular arc setting. Cadet, J.P., Kobayashi, K., Aubouin, J., Bouleque, J., Deplus, C., Dubois, J.,
I hope this review paper serves as a starting point for analy- von Huene, R., Jolivet, L., Kanazawa, T., Kasahara, J., Koizurni, K., Lal-
lemand, S., Nakamura, Y., Pautot, G., Suyehiro, S., Tokuyama, H., and
sis of these important systems, although more detailed studies, Yamazaki, T., 1987, The Japan trench and its juncture with the Kuril
especially of relatively young or modern arc settings, are needed. trench: Cruise results of the Kaiko project leg 3: Earth and Planetary Sci-
Unraveling the history of more highly deformed, ancient exam- ence Letters, v. 83, p. 267–284, doi: 10.1016/0012-821X(87)90071-9.
Cahill, T., and Isacks, B.L., 1992, Seismicity and shape of the subducted Nazca
ples will require the insight provided by studies from younger arc Plate: Journal of Geophysical Research, v. 97, p. 17,503–17,529.
settings as well as careful tectonic reconstructions, structural res- Carney, J.N., Macfarlane, A., and Mallick, D.I.J., 1985, The Vanuatu island
torations, accurate age dating, and detailed sedimentologic and arc—An outline of the stratigraphy, structure, and petrology, in Nairn,
A.E.M., et al., eds., The Ocean Basins and Margins—The Pacific Ocean:
stratigraphic analysis of the ancient examples. New York, Plenum Press, v. 7a, p. 685–718.
Caron, V., Nelson, C.S., and Kamp, P.J.J., 2004, Contrasting carbonate deposi-
ACKNOWLEDGMENTS tional systems for Pliocene cool-water limestones cropping out in central
Hawke’s Bay, New Zealand: New Zealand Journal of Geology and Geo-
physics, v. 47, p. 697–717.
Much of the research for this paper was completed while I was Clift, P., and Vannucchi, P., 2004, Controls on tectonic accretion versus ero-
on leave from Texas A&M University during the 2000–2001 sion in subduction zones: Implications for the origin and recycling of
academic year. ExxonMobil Upstream Research Company the continental crust: Reviews of Geophysics, v. 42, p. RG2001, doi:
10.1029/2003RG000127.
(EMURC) supported the research during my leave and provided Clift, P.D., MacLeod, C.J., Tappin, D.R., Wright, D., and Bloomer, S.H.,
an excellent work environment. I am indebted to colleagues at 1998, Tectonic controls on sedimentation and diagenesis in the Tonga
EMURC, both past and present, but most notably Jim Markello, trench and forearc, SW Pacific: Geological Society of America Bul-
letin, v. 110, p. 483–496, doi: 10.1130/0016-7606(1998)110<0483:
who was a major proponent for the study. Comments and sug- TCOSAD>2.3.CO;2.
gestions by Dan Bosence, John Reijmer, Sam Rice, Moyra Wil- Covey, M., 1986, The evolution of foreland basins to steady state: Evidence
son, and co-editor Peter Clift improved the manuscript and are from the western Taiwan foreland basin, in Allen, P., and Homewood,
P., eds., Foreland Basins: International Association of Sedimentologists
most appreciated. Special Publication 8, p. 77–90.
Coward, M.P., 1990, The Precambrian, Caledonian and Variscan framework to
REFERENCES CITED NW Europe, in Hardman, R.F.P., and Brooks, J., eds., Tectonic Events
Responsible for Britain’s Oil and Gas Reserves: Geological Society [Lon-
don] Special Publication 19, p. 3–36.
Ballance, P.F., 1993, The New Zealand Neogene forearc basins, in Ballance,
P.F., ed., South Pacific Sedimentary Basins, Sedimentary Basins of the Cunningham, J.K., and Anscombe, K.J., 1985, Geology of Eua and other islands,
World 2: Amsterdam, Elsevier Science, p. 177–193. Kingdom of Tonga, in Scholl, D.W., and Vallier, T.L., eds., Geology and
Bardintzeff, J.M., Brousse, R., and Gachon, A., 1985, Conditions of building Offshore Resources of Pacific Island Arcs—Tonga Region: Houston, Cir-
coral reefs on a volcano: Mururoa in Tuamotu and Rurutu in Australes cum-Pacific Council for Energy and Mineral Resources, Earth Science
(French Polynesia), in Gabrie, C., and Salvat, B., eds., Proceedings of Fifth Series, v. 2, p. 221–257.
International Coral Reef Congress: Miami, Florida, Rosenstiel School of Dahlen, F.A., Suppe, J., and Davis, D., 1984, Mechanics of fold-and-thrust belts
Marine and Atmospheric Science, University of Miami, p. 401–405. and accretionary wedges: Cohesive Coulomb theory: Journal of Geophys-
Beaudry, D., and Moore, G.F., 1985, Seismic stratigraphy and Cenozoic evolu- ical Research, v. 89, p. 10,087–10,101.
tion of West Sumatra forearc basin: American Association of Petroleum D’Anglejan, B., and Mountjoy, E.W., 1973, Submerged reefs of the east-
Geologists Bulletin, v. 69, p. 742–759. ern Grenadines shelf margin: Geological Society of America Bul-
Beck, M.E., Jr., 1983, On the mechanism of tectonic transport in zones of letin, v. 84, p. 2445–2454, doi: 10.1130/0016-7606(1973)84<2445:
oblique subduction: Tectonophysics, v. 93, p. 1–11, doi: 10.1016/0040- SROTEG>2.0.CO;2.
1951(83)90230-5. Davis, D., Suppe, J., and Dahlen, F.A., 1983, Mechanics of fold-and-thrust
Bosence, D.W.J., 2005, A new, genetic classification of carbonate platforms belts and accretionary wedges: Journal of Geophysical Research, v. 88,
based on their basinal and tectonic setting in the Cenozoic: Sedimentary p. 1153–1172.
Geology, v. 175, p. 49–72, doi: 10.1016/j.sedgeo.2004.12.030. DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems: Basin Research,
Bouysse, P., 1984, The Lesser Antilles island arc: Structure and geodynamic v. 8, p. 105–123, doi: 10.1046/j.1365-2117.1996.01491.x.
evolution, in Biju-Duval, B., Moore, J.C., et al., Initial Reports of the Demirmen, F., and Harbaugh, J.W., 1965, Petrography and origin of Perm-
Deep Sea Drilling Project: Washington, U.S. Government Printing Office, ian McCloud Limestone of northern California: Journal of Sedimentary
v. 78, p. 83–103. Petrology, v. 35, p. 136–154.
Carbonate-platform facies in volcanic-arc settings 87

Dewey, J.F., 1980, Episodicity, sequence, and style at convergent plate boundar- Gurnis, M., 1990b, Plate-mantle coupling and continental flooding: Geophysi-
ies, in Strangway, D.W., ed., The Continental Crust and Its Mineral Depos- cal Research Letters, v. 17, p. 623–626.
its: Geological Association of Canada Special Paper 20, p. 533–573. Gurnis, M., 1991, Continental flooding and mantle-lithosphere dynamics, in
Dey, S., and Smith, L., 1989, Carbonate and volcanic sediment distribution pat- Sabadini, R., et al., eds., Glacial Isostasy, Sea Level and Mantle Rheol-
terns on the Grenadines Bank, Lesser Antilles Island Arc, Eastern Carib- ogy: Dordrecht, Netherlands, Kluwer Academic Publishers, p. 446.
bean: Bulletin of Canadian Petroleum Geology, v. 37, p. 18–30. Haggerty, J.A., 1991, Evidence from fluid seeps atop serpentine seamounts in
Dickinson, W.R., 1974, Plate tectonics and sedimentation, in Dickinson, W.R., the Mariana Forearc; clues for emplacement of the seamounts and their
ed., Tectonics and Sedimentation: Society of Economic Paleontologists relationship to forearc tectonics, in Meyer, A.W., et al., eds., Evolution of
and Mineralogists Special Publication 22, p. 1–27. Mesozoic and Cenozoic Continental Margins: Marine Geology, v. 102,
Dickinson, W.R., 1995, Forearc basins, in Busby, C.J., and Ingersoll, R.V., eds., p. 293–309, doi: 10.1016/0025-3227(91)90013-T.
Tectonics of Sedimentary Basins: Cambridge, Massachusetts, Blackwell Halfar, H., Godinez-Orta, L., Mutti, M., and Valdez-Holguin, B., 2004, Nutri-
Science, p. 221–261. ents and temperature controls on modern carbonate production: An exam-
Dickinson, W.R., 2001, Paleoshoreline record of relative Holocene sea levels on ple from the Gulf of California, Mexico: Geology, v. 32, p. 213–216, doi:
Pacific islands: Earth-Science Reviews, v. 55, p. 191–234, doi: 10.1016/ 10.1130/G20298.1.
S0012-8252(01)00063-0. Hallock, P., and Schlager, W., 1986, Nutrient excess and the demise of
Dickinson, W.R., and Seely, D.R., 1979, Structure and stratigraphy of forearc coral reefs and carbonate platforms: Palaios, v. 1, p. 389–398, doi:
regions: American Association of Petroleum Geologists Bulletin, v. 63, 10.2307/3514476.
p. 2–31. Hamilton, W., 1979, Tectonics of the Indonesian region: U.S. Geological Sur-
DiMarco, G., Baumgartner, P.O., and Channell, J.E.T., 1995, Late Cretaceous– vey Professional Paper 1078, 345 p.
early Tertiary paleomagnetic data and a revised tectonostratigraphic subdi- Hathway, B., 1994, Sedimentation and volcanism in an Oligocene-Miocene
vision of Costa Rica and western Panama, in Mann, P., ed., Geologic and intra-oceanic arc and fore-arc, southwestern Viti Levu, Fiji: Geological
Tectonic Development of the Caribbean Plate Boundary in Southern Cen- Society [London] Journal, v. 151, p. 499–514.
tral America: Geological Society of America Special Paper 295, p. 1–27. Hathway, B., 1995, Deposition and diagenesis of Miocene arc-fringing plat-
Dorobek, S.L., 1995, Synorogenic carbonate platforms and reefs in foreland form and debris-apron carbonates, southwestern Viti Levu, Fiji: Sedimen-
basins: Controls on stratigraphic evolution and platform/reef morphology, tary Geology, v. 94, p. 187–208, doi: 10.1016/0037-0738(94)00086-A.
in Dorobek, S.L., and Ross, G.M., eds., Stratigraphic Evolution of Fore- Hattori, M., Kanie, Y., Oba, T., and Akimoto, K., 1995, Environmental condi-
land Basins: SEPM (Society for Sedimentary Geology) Special Publica- tions of carbonates and chemosynthetic animal communities associated
tion 52, p. 127–147. with cold seepage zones along the subduction zone in Sagami Bay, central
Dorsey, R.J., and Kidwell, S.M., 1999, Mixed carbonate-siliciclastic sedimen- Japan: Fossils, v. 60, p. 13–22.
tation on a tectonically active margin: Example from the Pliocene of Baja Henrich, R., and Freiwald, A., 1995, Controls on modern carbonate sedimenta-
California Sur, Mexico: Geology, v. 27, p. 935–938, doi: 10.1130/0091- tion on warm-temperate to arctic coasts, shelves and seamounts in the
7613(1999)027<0935:MCSSOA>2.3.CO;2. northern hemisphere: Implications for fossil counterparts: Facies, v. 32,
Eberli, G.P., and Ginsburg, R.N., 1989, Cenozoic progradation of northwestern p. 71–108, doi: 10.1007/BF02536865.
Great Bahama Bank, a record of lateral platform growth and seal-level Herzer, R.H., and Exon, N.F., 1985, Structure and basin analysis of the southern
fluctuations, in Crevello, P.D., Wilson, J.L., Sarg, J.F., and Read, J.F., eds., Tonga forearc, in Schole, D.W., and Valier, T.L., eds., Geology and off-
Controls on Carbonate Platform and Basin Development: SEPM Special shore resources of Pacific island arcs—Tonga Region: Houston, Circum-
Publication 44, p. 339–352. Pacific Council for Energy and Mineral Resources, Earth Science Series,
Eguchi, T., Uyeda, S., and Maki, T., 1980, Seismotectonics and tectonic history v. 2, p. 55–73.
of the Andaman Sea, in Toksöz, M.N., et al., eds., Oceanic Ridges and Ingersoll, R.V., 1988, Tectonics of sedimentary basins: Geological Soci-
Arcs: New York, Elsevier, p. 425–441. ety of America Bulletin, v. 100, p. 1704–1719, doi: 10.1130/0016-
Eldredge, L.C., and Kropp, R.K., 1985, Volcanic ashfall effects on intertidal 7606(1988)100<1704:TOSB>2.3.CO;2.
and shallow-water coral reef zones at Pagan, Mariana Islands: Tahiti, Pro- Iryu, Y., Nakamori, T., Matsuda, S., and Abe, O., 1995, Distribution of marine
ceedings of 5th International Coral Reef Congress, p. 195–200. organisms and its geological significance in the modern reef complex
Exon, N.F., Stewart, W.D., Sandy, M.J., and Tiffin, D.L., 1986, Geology of the of the Ryukyu Islands: Sedimentary Geology, v. 99, p. 243–258, doi:
eastern New Ireland Basin, northeastern Papua New Guinea: Bureau of 10.1016/0037-0738(95)00047-C.
Mineral Resources, Journal of Australian Geology and Geophysics, v. 10, James, N.P., 1997, The cool-water carbonate depositional realm, in James, N.P.,
p. 39–51. and Clarke, J.A.D., eds., Cool-water Carbonates: SEPM (Society for Sed-
Flemings, P.B., and Jordan, T.E., 1990, Stratigraphic modeling of fore- imentary Geology) Special Publication 56, p. 1–20.
land basins: Interpreting thrust deformation and lithosphere rheology: Johnson, D.P., and Greene, H.G., 1988, Modern depositional regimes, offshore
Geology, v. 18, p. 430–434, doi: 10.1130/0091-7613(1990)018<0430: Vanuatu, in Greene, H.G., and Wong, F.L., eds., Geology and Offshore
SMOFBI>2.3.CO;2. Resources of Pacific Island Arcs—Vanuatu Region: Houston, Circum-
Fortuin, A.R., Van der Werff, W., and Wensink, H., 1997, Neogene basin history Pacific Council for Energy and Mineral Resources, Earth Science Series,
and paleomagnetism of a rifted and inverted forearc region, on- and off- v. 8, p. 287–299.
shore Sumba, Eastern Indonesia: Journal of Asian Earth Sciences, v. 15, Johnson, R.W., 1979, Geotectonics and volcanism in Papua New Guinea: A
p. 61–88. review of the late Cainozoic: BMR Journal of Australian Geology and
Galindo-Zaldívar, J., Nieto, L.M., Robertson, A.H.F., and Woodside, J.M., Geophysics, v. 4, p. 181–207.
2001, Recent tectonics of Eratosthenes Seamount: An example of sea- Jordan, T.E., 1995, Retroarc foreland and related basins, in Busby, C.J., and
mount deformation during incipient continental collision: Geo-Marine Ingersoll, R.V., eds., Tectonics of Sedimentary Basins: Cambridge, Mas-
Letters, v. 20, p. 233–242, doi: 10.1007/s003670000059. sachusetts, Blackwell Science, p. 331–362.
Geist, E.L., Childs, J.R., and Scholl, D.W., 1988, The origin of summit basins Jordan, T.E., and Allmendinger, R.W., 1986, The Sierra Pampeanas of Argen-
of the Aleutian Ridge: Implications for block rotation of an arc massif: tina: A modern analogue of Rocky Mountain foreland deformation:
Tectonics, v. 7, p. 327–341. American Journal of Science, v. 286, p. 737–764.
Greene, H.G., Macfarlane, A., and Wong, F.L., 1988, Geology and offshore Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A.,
resources of Vanuatu—Introduction and summary, in Greene, H.G., and and Ando, C.J., 1983, Andean tectonics related to geometry of subducted
Wong, F.L., eds., Geology and Offshore Resources of Pacific Island Nazca plate: Geological Society of America Bulletin, v. 94, p. 341–361,
Arcs—Vanuatu Region: Houston, Circum-Pacific Council for Energy and doi: 10.1130/0016-7606(1983)94<341:ATRTGO>2.0.CO;2.
Mineral Resources, Earth Science Series, v. 8, p. 1–25. Kamp, P.J.J., and Nelson, C.S., 1987, Tectonic and sea level controls on non-
Grippi, J., 1980, Geology of the Lucea Inlier, western Jamaica: Journal of the tropical Neogene limestones in New Zealand: Geology, v. 15, p. 610–613,
Geological Society of Jamaica, v. 19, p. 1–24. doi: 10.1130/0091-7613(1987)15<610:TASCON>2.0.CO;2.
Gurnis, M., 1990a, Bounds on global dynamic topography from Phanero- Kamp, P.J.J., and Nelson, C.S., 1988, Nature and occurrence of modern and
zoic flooding of continental platforms: Nature, v. 344, p. 754–756, doi: Neogene active margin limestones in New Zealand: New Zealand Journal
10.1038/344754a0. of Geology and Geophysics, v. 31, p. 1–20.
88 Dorobek

Kamp, P.J.J., Harmsen, F.J., Nelson, C.S., and Boyle, S.F., 1988, Barnacle- Geology and Geophysics of Continental Margins: American Association
dominated limestone with giant cross-beds in a non-tropical, tide-swept, of Petroleum Geologists Memoir 53, p. 157–181.
Pliocene forearc seaway, Hawke’s Bay, New Zealand: Sedimentary Geol- Maynard, J.R., Hofmann, W., Dunay, R.E., Bentham, P.N., Dean, K.P., and Wat-
ogy, v. 60, p. 173–195, doi: 10.1016/0037-0738(88)90118-2. son, I., 1997, The Carboniferous of western Europe: The development of a
Karig, D.E., 1983, Temporal relationships between back arc basin formation petroleum system: Petroleum Geoscience, v. 3, p. 97–115.
and arc volcanism with special reference to the Philippine Sea, in Hayes, Meffre, S., and Crawford, A.J., 2001, Collision tectonics in the New Hebri-
D.E., ed., The Tectonic and Geologic Evolution of Southeast Asian Seas des arc (Vanuatu): Island Arc, v. 10, p. 33–50, doi: 10.1046/j.1440-
and Islands (Pt. II): American Geophysical Union, AGU Monograph 1738.2001.00292.x.
Series, no. 27, P. 318–325. Mesolella, K.J., Matthews, R.K., Broecker, W.S., and Thurber, D.L., 1969, The
Karig, D.E., Sarewitz, D.R., and Haeck, G.D., 1986, Role of strike-slip astronomical theory of climate change: Barbados data: Journal of Geol-
faulting in the evolution of allochthonous terranes in the Philippines: ogy, v. 77, p. 250–274.
Geology, v. 14, p. 852–855, doi: 10.1130/0091-7613(1986)14<852: Minoura, K., 1992, Dolomitization of Permian reef carbonates in a Jurassic
ROSFIT>2.0.CO;2. subduction complex, central Japan: Sedimentary Geology, v. 80, p. 41–
Kelly, S.R.A., Ditchfield, P.W., Doubleday, P.A., and Marshall, J.D., 1995, 52, doi: 10.1016/0037-0738(92)90030-U.
An Upper Jurassic methane-seep limestone from the Fossil Bluff Group Moore, J.C., and Vrolijk, P., 1992, Fluids in accretionary prisms: Reviews of
forearc basin of Alexander Island, Antarctica: Journal of Sedimentary Geophysics, v. 30, p. 113–135.
Research, v. 65, p. 274–282. Nakamori, T., Iryu, Y., and Yamada, T., 1995, Development of coral reefs of
Kikuchi, Y., Tono, S., and Funayama, M., 1991, Petroleum resources in the the Ryukyu Islands (southwest Japan, East China Sea) during Pleisto-
Japanese island-arc setting: Episodes, v. 14, p. 236–241. cene sea-level change: Sedimentary Geology, v. 99, p. 215–231, doi:
Klaus, A., Taylor, B., Moore, G.F., Murakami, F., and Okamura, Y., 1992, Back- 10.1016/0037-0738(95)00045-A.
arc rifting in the Izu-Bonin island arc: Structural evolution of Hachijo Niem, A.R., 1989, Coral and limestone, carbonate cements in siliciclastic rocks,
and Aoga Shima rifts: Island Arc, v. 1, p. 16–31, doi: 10.1111/j.1440- and chalk from dredge samples, Solomon Islands and Bougainville, Papua
1738.1992.tb00054.x. New Guinea, in Vedder, J.G., and Bruns, T.R., eds., Geology and Offshore
Kolarsky, R.A., Mann, P., and Monechi, S., 1995, Stratigraphic development of Resources of Pacific Island Arcs—Solomon Islands and Bougainville,
southwestern Panama as determined from integration of marine seismic Papua New Guinea Regions: Houston, Circum-Pacific Council for Energy
data and onshore geology, in Mann, P., ed., Geologic and Tectonic Devel- and Mineral Resources, Earth Science Series, v. 12, p. 175–202.
opment of the Caribbean Plate Boundary in Southern Central America: Nunn, P.D., 1998a, Pacific Island Landscapes: Suva, Fiji, University of the
Geological Society of America Special Paper 295, p. 159–200. South Pacific, Institute of Pacific Studies, 318 p.
Krawinkel, H.E., and Krawinkel, J.J., 1996, From strike-slip depositional set- Nunn, P.D., 1998b, Late Cenozoic emergence of the islands of the northern
ting to fold and thrust belt; Tertiary to Recent structural and sedimentary Lau-Colville Ridge, southwest Pacific, in Stewart, I.S., and Vita-Finzi, C.,
development of a forearc basin (Terraba Basin, southern Costa Rica): eds., Coastal Tectonics: Geological Society [London] Special Publication
American Association of Petroleum Geologists annual convention, San 146, p. 269–278.
Diego, California, May 19–22, 1996, Annual Meeting Abstracts, v. 5, Nunn, P.D., 2000, Significance of emerged Holocene corals around Ovalau and
p. 78–79. Motuiki islands, Fiji, southwest Pacific: Marine Geology, v. 163, p. 345–
Kroenke, L.W., 1984, Cenozoic tectonic development of the southwest Pacific: 351, doi: 10.1016/S0025-3227(99)00114-0.
United Nations ESCAP, CCOP/SOPAC, Technical Bulletin 6, 122 p. Nunn, P.D., and Omura, A., 1999, Penultimate Interglacial emergent reef
Kulm, L.D., and Suess, E., 1990, Relationship between carbonate deposits and around Kadavu Island, Southwest Pacific: Implications for late Quater-
fluid venting; Oregon accretionary prism, in Langseth, M.G., and Moore, nary island-arc tectonics and sea-level history: New Zealand Journal of
J.C., eds., Special section on the Role of fluids in sediment accretion, Geology and Geophysics, v. 42, p. 219–227.
deformation, diagenesis, and metamorphism in subduction zones: Journal Okamura, Y., Watanabe, M., Morijiri, R., and Satoh, M., 1995, Rifting and
of Geophysical Research, v. 95, p. 8899–8915. basin inversion in the eastern margin of the Japan Sea: Island Arc, v. 4,
Kulm, L.D., Suess, E., Moore, J.C., Carson, B., Lewis, B.T., Ritger, S.D., p. 166–181, doi: 10.1111/j.1440-1738.1995.tb00141.x.
Kadko, D.C., Thornburg, T.M., Embley, R.W., Rugh, W.D., Massoth, Otsuki, K., 1989, Empirical relationships among the convergence rate of plates,
G.J., Langseth, M.G., Cochrane, G.R., and Scamman, R.L., 1986, Oregon rollback rate of trench axis and island-arc tectonics: Laws of convergence
subduction zone; venting, fauna, and carbonates: Science, v. 231, p. 561– rate of plates: Tectonophysics, v. 159, p. 73–94, doi: 10.1016/0040-
566, doi: 10.1126/science.231.4738.561. 1951(89)90171-6.
Larue, D.K., and Suess, E., 1985, Eocene subduction-driven vent community, Phillips, J.R., 1993, Stratigraphy and structural setting of the mid-Cretaceous
Joes River melange, Barbados: Eos (Transactions, American Geophysical Olvidada Formation, Baja California Norte, Mexico, in Gastil, R.G., and
Union), v. 66, p. 1097. Miller, R.H., eds., The Prebatholithic Stratigraphy of Peninsular Califor-
Letouzey, J., and Kimura, M., 1986, The Okinawa Trough: Genesis of a back- nia: Geological Society of America Special Paper 279, p. 97–106.
arc basin developing along a continental margin: Tectonophysics, v. 125, Pomoni-Papaioannou, F., Drinia, H., and Dermitzakis, M.D., 2003, Neogene
p. 209–230, doi: 10.1016/0040-1951(86)90015-6. non-tropical carbonate sedimentation in a warm temperate biogeographic
Lundberg, N., 1982, Evolution of the slope landward of the Middle America province (Rethymnon Formation, Eastern Crete, Greece): Sedimentary
Trench, Nicoya Peninsula, Costa Rica, in Leggett, J.K., ed., Trench- Geology, v. 154, p. 147–157, doi: 10.1016/S0037-0738(02)00127-6.
Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Poulton, T.P., 1989, A Lower Jurassic coral reef, Telkwa Range, British Colum-
Active Plate Margins: Geological Society [London] Special Publication bia: Canadian Society of Petroleum Geologists Memoir 13, p. 754–757.
10, p. 131–147. Read, J.F., 1985, Carbonate platform facies models: American Association of
Macfarlane, A., Carney, J.N., Crawford, A.J., and Greene, H.G., 1988, Van- Petroleum Geologists Bulletin, v. 69, p. 1–21.
uatu—A review of onshore geology, in Greene, H.G., and Wong, F.L., Read, J.F., 1995, Overview of carbonate platform sequences, cycle stratigraphy
eds., Geology and Offshore Resources of Pacific Island Arcs—Vanu- and reservoirs in greenhouse and ice-house worlds, in Read, J.F., et al.,
atu Region: Houston, Circum-Pacific Council for Energy and Mineral eds., Milankovitch Sea Level Changes, Cycles and Reservoirs on Carbon-
Resources, Earth Science Series, v. 8, p. 45–92. ate Platforms in Greenhouse and Icehouse Worlds: SEPM (Society for
Mann, P., Taylor, F.W., Lagoe, M.B., Quarles, A., and Burr, G., 1998, Acceler- Sedimentary Geology) Short Course Notes 35, p. 1–102.
ating late Quaternary uplift of the New Georgia island group (Solomon Reid, R.P., and Ginsburg, R.N., 1986, The role of framework in Upper Tri-
island arc) in response to subduction of recently active Woodlark spread- assic patch reefs in the Yukon (Canada): Palaios, v. 1, p. 590–600, doi:
ing center and Coleman seamount: Tectonophysics, v. 295, p. 259–306, 10.2307/3514709.
doi: 10.1016/S0040-1951(98)00129-2. Rodd, J.A., 1993, New reef targets for oil and gas exploration in Fiji, Southwest
Marsaglia, K.M., 1995, Interarc and backarc basins, in Busby, C.J., and Inger- Pacific: Geological Society of Malaysia Bulletin, v. 33, p. 313–330.
soll, R.V., eds., Tectonics of Sedimentary Basins: Cambridge, Massachu- Rutherford, E., Burke, K., and Lytwyn, J., 2001, Tectonic history of Sumba
setts, Blackwell Science, p. 299–329. Island, Indonesia, since the Late Cretaceous and its rapid escape into the
Matson, R.G., and Moore, G.F., 1992, Structural influences on Neogene subsid- forearc in the Miocene: Journal of Asian Earth Sciences, v. 19, p. 453–
ence in the central Sumatra fore-arc basin, in Watkins, J.S., et al., eds., 479, doi: 10.1016/S1367-9120(00)00032-8.
Carbonate-platform facies in volcanic-arc settings 89

Sagawa, N., Nakamori, T., and Iryu, Y., 2001, Pleistocene reef development Tamaki, K., 1995, Opening tectonics of the Japan Sea, in Taylor, B., ed., Backarc
in the southwest Ryukyu Islands, Japan: Palaeogeography, Palaeocli- Basins: Tectonics and Magmatism: New York, Plenum Press, p. 407–420.
matology, Palaeoecology, v. 175, p. 303–323, doi: 10.1016/S0031- Tapponnier, P., Peltzer, G., LeDain, A.Y., Armijo, R., and Cobbold, P., 1982,
0182(01)00377-7. Propagation extrusion tectonics in Asia: New insights from simple experi-
Sample, J.C., and Fisher, D.M., 1986, Duplexes and underplating in an ancient ments with plasticene: Geology, v. 10, p. 611–616, doi: 10.1130/0091-
accretionary complex, Kodiak Islands, Alaska: Geology, v. 14, p. 160– 7613(1982)10<611:PETIAN>2.0.CO;2.
163, doi: 10.1130/0091-7613(1986)14<160:DAAUIA>2.0.CO;2. Taylor, B., Brown, G., Fryer, P., Gill, J.B., Hochstaedter, A.G., Hotta, H., Lang-
Sarewitz, D.R., and Lewis, S.D., 1991, The Marinduque intra-arc basin, Phil- muir, C.H., Leinen, M., Nishimura, A., and Urabe, T., 1990, ALVIN–Sea
ippines: Basin genesis and in situ ophiolite development in a strike-slip Beam studies of the Sumisu Rift, Izu-Bonin Arc: Earth and Planetary Sci-
setting: Geological Society of America Bulletin, v. 103, p. 597–614, doi: ence Letters, v. 100, p. 127–147, doi: 10.1016/0012-821X(90)90181-V.
10.1130/0016-7606(1991)103<0597:TMIABP>2.3.CO;2. Taylor, B., Klaus, A., Brown, G.R., and Moore, G.F., 1991, Active tectonics of
Seely, D.R., 1979, The evolution of structural highs bordering major forearc the north and central Aegean Sea: Journal of Geophysical Research, v. 96,
basins: American Association of Petroleum Geologists Memoir 29, p. 19,611–19,622.
p. 245–260. ten Veen, J.H., and Postma, G., 1999, Neogene tectonics and basin fill patterns
Seely, D.R., Vail, P.R., and Walton, G.G., 1974, Trench slope model, in Burk, in the Hellenic outer-arc (Crete, Greece): Basin Research, v. 11, p. 223–
C.A., and Drake, C.L., eds., The Geology of Continental Margins: New 241, doi: 10.1046/j.1365-2117.1999.00097.x.
York, Springer-Verlag, p. 249–260. Thomas, W.A., 1983, Continental margins, orogenic belts, and intracra-
Seyfried, H., Astorga, A., Amann, H., Calvo, C., Kolb, W., Schmidt, H., and tonic structures: Geology, v. 11, p. 270–272, doi: 10.1130/0091-
Winsemann, J., 1991, Anatomy of an evolving island arc: Tectonic and 7613(1983)11<270:CMOBAI>2.0.CO;2.
eustatic control in the south Central American fore-arc area, in Mac- Uyeda, S., and Kanamori, H., 1979, Back-arc opening and the mode of subduc-
donald, D.I.M., ed., Sedimentation, Tectonics and Eustasy—Sea-level tion: Journal of Geophysical Research, v. 84, p. 1049–1061.
Changes at Active Margins: International Association of Sedimentolo- van der Werff, W., 1996, Variation in forearc basin development along the
gists Special Publication 12, p. 151–163. Sunda Arc, Indonesia: Journal of Southeast Asian Earth Sciences, v. 14,
Sharaf, E., Simo, J.A., Carroll, A.R., and Shields, M., 2005, Stratigraphic p. 331–349, doi: 10.1016/S0743-9547(96)00068-2.
evolution of Oligocene–Miocene carbonates and siliciclastics, East Java Vedder, J.G., and Bruns, T.R., 1989, Geologic setting and petroleum prospects
basin, Indonesia: American Association of Petroleum Geologists Bulletin, of basin sequences, offshore Solomon Islands and eastern Papua New
v. 89, p. 799–819. Guinea, in Vedder, J.G., and Bruns, T.R., eds., Geology and Offshore
Shin Tani, 1989, Detailed topographic study of the Daiichi-Kashima Seamount: Resources of Pacific Island Arcs: Solomon Islands and Bougainville,
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 71, p. 31–47, doi: Papua New Guinea Regions: Houston, Circum-Pacific Council for Energy
10.1016/0031-0182(89)90028-X. and Mineral Resources, Earth Science Series, v. 12, p. 287–322.
Sieh, K., Ward, S.N., Natawidjaja, D., and Suwargadi, B.W., 1999, Crustal defor- Vergés, J., Marzo, M., and Muñoz, J.A., 2002, Growth strata in foreland
mation at the Sumatran subduction zone revealed by coral rings: Geophys- settings: Sedimentary Geology, v. 146, p. 1–9, doi: 10.1016/S0037-
ical Research Letters, v. 26, p. 3141–3144, doi: 10.1029/1999GL005409. 0738(01)00162-2.
Sigurdsson, H., Sparks, R.S.J., Carey, S.N., and Huang, T.C., 1980, Volcano- von Huene, R., Langseth, M., Nasu, N., and Okada, H., 1982, A summary of
genic sedimentation in the Lesser Antilles arc: Journal of Geology, v. 88, Cenozoic tectonic history along the IPOD Japan Trench transect: Geolog-
p. 523–540. ical Society of America Bulletin, v. 93, p. 829–846, doi: 10.1130/0016-
Silver, E.A., and Reed, D.L., 1988, Backthrusting in accretionary wedges: Jour- 7606(1982)93<829:ASOCTH>2.0.CO;2.
nal of Geophysical Research, v. 93, p. 3116–3126. Wang, J., Newton, C.R., and Dunne, L., 1988, Late Triassic transition from bio-
Smith, G.A., and Landis, C.A., 1995, Intra-arc basins, in Busby, C.J., and Inger- genic to arc sedimentation on the Peninsular terrane: Puale Bay, Alaska
soll, R.V., eds., Tectonics of Sedimentary Basins: Cambridge, Massachu- Peninsula: Geological Society of America Bulletin, v. 100, p. 1466–1478,
setts, Blackwell Science, p. 263–298. doi: 10.1130/0016-7606(1988)100<1466:LTTFBT>2.3.CO;2.
Soja, C.M., 1990, Island arc carbonates from the Silurian Heceta Formation of Watkins, R., 1985, Volcaniclastic and carbonate sedimentation in late Paleo-
southeastern Alaska (Alexander terrane): Journal of Sedimentary Petrol- zoic island-arc deposits, Eastern Klamath Mountains, California:
ogy, v. 60, p. 235–249. Geology, v. 13, p. 709–713, doi: 10.1130/0091-7613(1985)13<709:
Soja, C.M., 1992, Potential contributions of ancient oceanic islands to evolu- VACSIL>2.0.CO;2.
tionary theory: Journal of Geology, v. 100, p. 125–134. Watkins, R., 1993a, Carbonate bank sedimentation in a volcaniclastic arc set-
Soja, C.M., 1996, Island-arc carbonates: Characterization and recognition in ting: Lower Carboniferous limestones of the Eastern Klamath terrane,
the ancient geologic record: Earth-Science Reviews, v. 41, p. 31–65, doi: California: Journal of Sedimentary Petrology, v. 63, p. 966–973.
10.1016/0012-8252(96)00029-3. Watkins, R., 1993b, Permian carbonate platform development in an island-
Stanley, G.D., Jr., 1982, Triassic carbonate development and reefbuilding in arc setting, Eastern Klamath Mountains, California: Journal of Geology,
western North America: Geologische Rundschau, v. 71, p. 1057–1075, v. 101, p. 659–666.
doi: 10.1007/BF01821118. Watkins, R., and Flory, R.A., 1986, Island arc sedimentation in the Middle
Stanley, G.D., Jr., and Beauvais, L., 1990, Middle Jurassic corals from the Wal- Devonian Kennett Formation, Eastern Klamath Mountains, California:
lowa Terrane, west-central Idaho: Journal of Paleontology, v. 64, p. 352– Journal of Geology, v. 94, p. 753–761.
362. Wells, J.W., 1953, Mesozoic invertebrate faunas of Peru, part 3, Lower Juras-
Stanley, G.D., Jr., and McRoberts, C.A., 1993, A coral reef in the Telkwa sic corals from the Arequipa region: American Museum Novitiates 1631,
Range, British Columbia: The earliest Jurassic example: Canadian Jour- 14 p.
nal of Earth Sciences, v. 30, p. 819–831. Wells, S.M., 1988, Coral Reefs of the World, Volume 3: Central and Western
Stanley, G.D., and Senowbari-Daryan, B., 1986, Upper Triassic, Dachstein- Pacific: Nairobi, Kenya/IUCN, Gland, Switzerland, and Cambridge, UK,
type, reef limestone from the Wallowa Mountains, Oregon; first UNEP (United Nations Environment Programme), Regional Seas Direc-
reported occurrence in the United States: Palaios, v. 1, p. 172–177, doi: tories and Bibliographies, 329 p.
10.2307/3514511. Whalen, M.T., 1995, Barred basins: A model for eastern ocean basin car-
Stewart, W.D., and Sandy, M.J., 1988, Geology of New Ireland and Djaul bonate platforms: Geology, v. 23, p. 625–628, doi: 10.1130/0091-
Islands, northeastern Papua New Guinea, in Marlow, M.S., et al., eds., 7613(1995)023<0625:BBAMFE>2.3.CO;2.
Geology and Offshore Resources of Pacific Island Arcs—New Ireland and Wheeler, C.W., and Aharon, P., 1991, Mid-oceanic carbonate platforms as oce-
Manus Regions, Papua New Guinea: Houston, Circum-Pacific Council anic dipsticks: Examples from the Pacific: Coral Reefs, v. 10, p. 101–114,
for Energy and Mineral Resources, Earth Science Series, v. 9, p. 13–30. doi: 10.1007/BF00571828.
Takano, O., 2002, Changes in depositional systems and sequences in response Wilson, M.E.J., 2002, Cenozoic carbonates in SE Asia: Implications for equa-
to basin evolution in a rifted and inverted basin: An example from the torial carbonate development: Sedimentary Geology, v. 147, p. 295–428,
Neogene Niigata-Shin’etsu basin, Northern Fossa Magna, central doi: 10.1016/S0037-0738(01)00228-7.
Japan: Sedimentary Geology, v. 152, p. 79–87, doi: 10.1016/S0037- Wilson, M.E.J., and Bosence, D.W.J., 1997, Platform-top and ramp deposits of
0738(01)00286-X. the Tonasa Carbonate Platform, Sulawesi, Indonesia, in Fraser, A.J., et al.,
90 Dorobek

eds., Petroleum Geology of Southeast Asia: Geological Society [London] Tonga region: Houston, Circum-Pacific Council for Energy and Mineral
Publication 126, p. 247–279. Resources, Earth Science Series, v. 2, p. 351–378.
Wilson, M.E.J., and Lokier, S.J., 2002, Siliciclastic and volcaniclastic influences Yamaji, A., 1990, Rapid intra-arc rifting in Miocene, Northeast Japan: Tecton-
on equatorial carbonates; insights from the Neogene of Indonesia: Sedi- ics, v. 9, p. 365–378.
mentology, v. 49, p. 583–601, doi: 10.1046/j.1365-3091.2002.00463.x. Zachariasen, J., Sieh, K., Taylor, F.W., Edwards, R.L., and Hantoro, W.S., 1999,
Wilson, M.E.J., and Vecsei, A., 2005, The apparent paradox of abundant foramol Submergence and uplift associated with the giant 1833 Sumatran subduc-
facies in low latitudes: Their environmental significance and effect on tion earthquake: Evidence from coral microatolls: Journal of Geophysical
platform development: Earth-Science Reviews, v. 69, p. 133–168, doi: Research, v. 104, p. 895–919, doi: 10.1029/1998JB900050.
10.1016/j.earscirev.2004.08.003.
Woodhall, D., 1985, Geology of the Lau Ridge, in Scholl, D.W., and Vallier,
T.L., eds., Geology and Offshore Resources of Pacific Island Arcs— MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Sediment waves in the Bismarck Volcanic Arc, Papua New Guinea

Gary Hoffmann
Eli Silver
Simon Day
Eugene Morgan
Earth Sciences Department, University of California, Santa Cruz, California 95064, USA

Neal Driscoll
Scripps Institution of Oceanography, La Jolla, California 92093, USA

Daniel Orange*
AOA Geophysics, Castroville, California 95012, USA

ABSTRACT

In the Bismarck Volcanic Arc in Papua New Guinea, six fields of sediment waves
were imaged with sonar. Sediment structures observed in seismic data and swath
bathymetry are not unique and can result from predominantly continuous (bottom)
currents, or episodic (turbidity) currents, or from deformation of sediment. Two of
these wave fields overlap and appear to be of turbidity-current origin and modified by
bottom currents, with one field unconformably overlying the other field. A field off the
coast of Dakataua caldera displays an arcuate morphology, and a series of enclosed
depressions within the field suggests creation by extensional deformation of rapidly
deposited sediment. Scour features in side-scan imagery suggest turbidity-current
activity, which also likely modifies the sediment waves. The wave field is isolated from
hyperpycnal currents, however, suggesting that in the absence of a shelf, coastal ero-
sion and small landslides can produce semiregular gravity-driven sediment flows that
deposit in deep (>1400 m) water. In Kimbe Bay a fourth sediment-wave field also dis-
plays arcuate morphology and enclosed depressions within the field. This wave field
is found within a bay >40 km from shore and also appears to have been formed by a
combination of extensional deformation of sediment and energetic current activity.
Two additional fields in Hixon Bay are fed by small and medium rivers (<~450 m3/s
mean annual discharge) draining volcanoes and mountainous regions. One small field
appears within a slide scar, suggesting that the initial topography of the scar provided
the conditions for early sediment-wave growth. A much larger field is best explained
by repeated hyperpycnal currents originating from the Pandi River. We cored a series
of upward-fining, graded sequences consistent with a turbidity-current origin. Ages
from these cores and measurements of relative thickness in sub-bottom imagery of the

*Present address: Black Gold Energy, Plaza Kemang Timor 22, Jakarta 12510, Indonesia

Hoffmann, G., Silver, E., Day, S., Morgan, E., Driscoll, N., and Orange, D., 2008, Sediment waves in the Bismarck Volcanic Arc, Papua New Guinea, in Draut,
A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special
Paper 436, p. 91–126, doi: 10.1130/2008.2436(05). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All
rights reserved.

91
92 Hoffmann et al.

field constrain deposition rates for the field and suggest that a large part of the Pandi
River discharge must be bypassing the shelf and depositing on the sediment-wave
field in deep water (>1200 m). These findings suggest that the sedimentary record
in arc collision zones will be dominated by mass-wasting deposits very close to vol-
canoes, and by river discharge depositing in select, extent regions far from shore.
Because sedimentation rates can vary by a factor of 2 between the two flanks of a
sediment wave, care must be taken when comparing bed thickness across an entire
sedimentary section.

Keywords: sediment waves, Papua New Guinea, Bismarck Sea, marine geology,
geomorphology.

INTRODUCTION giving us the rare opportunity to examine the distinguishing cri-


teria using multiple wave fields within the same data set. We can
More than half of the total suspended sediment supplied by also test if these criteria can be consistently applied to differenti-
rivers to the sea originates from small to medium mountain rivers ate wave-forming processes, and what sediment waves formed by
(<~450 m3/s mean annual discharge; Mulder et al., 2003), espe- these processes tell us about dispersal of sediment.
cially those in tectonically active regions such as the Bismarck We begin with a discussion of the processes that create sedi-
Volcanic Arc in Papua New Guinea (Milliman and Syvitski, ment waves and morphologically similar features. One of these
1992). At another active arc-continent collision, Taiwan, as much processes, turbidity currents, is surrounded by some contro-
as 42% of the sediment discharge to the ocean occurs by hyper- versy, and so we proceed with an examination of the term and
pycnal flows (>40 g/L; Warrick and Milliman, 2003), typically its relation to hyperpycnal currents. We follow the discussion of
during flood events (Dadson et al., 2005). Large earthquakes can sediment-wave-generation processes and turbidity currents with
cause an increase in the occurrence of hyperpycnal flows and the our specific observations and discuss them in terms of their dis-
percentage of sediment discharged at hyperpycnal concentrations tinguishing features and how they relate to climate and sediment
for some rivers for short intervals, owing to landslides (Warrick transport in the Bismarck Sea.
and Milliman, 2003; Dadson et al., 2005). Under certain condi-
tions, these hyperpycnal turbidity currents can result in deep-sea STUDY LOCATION
sediment waves (H.J. Lee et al., 2002; Wynn and Stow, 2002;
Schwehr et al., 2007). Papua New Guinea has two seasons, the summer monsoon
The Bismarck Volcanic Arc differs from arc-collisional set- season and the winter trade-wind season. Although cyclonic
tings such as Taiwan and other regions where turbidity-current events are uncommon, near-continuous tropical rain produces
sediment waves have been identified (for example, the Var field frequent landslides and consequent large sediment discharge to
southeast of France; Migeon et al., 2001) by having active vol- the coastal ocean from the mountains, as well as by floods any
canism. Widely dispersed tephra may increase the frequency of time of year (McAlpine and Keig, 1983; Pickup, 1984; Walsh
hyperpycnal discharges (Hayes et al., 2002), and ashfall from and Nittrouer, 2003). The sediment load for the entire island of
volcanic eruptions may also directly produce turbidity currents New Guinea is high, at ~1.7 × 109 t yr–1, roughly the estimated
(Fiske et al., 1998). Bottom currents, rapid slope failure, and slow load of all North American rivers combined (Milliman, 1995).
soft-sediment deformation, for instance caused by repeated earth- The Bismarck Volcanic Arc is located in northern Papua
quake loading, are also processes that have been suggested as New Guinea and forms the southern boundary of the Bismarck
being capable of producing or greatly modifying sediment-wave Sea (Fig. 1). In the eastern half of the arc the Solomon Sea plate
fields (Gardner et al., 1999; Lee and Chough, 2001; H.J. Lee et is subducting beneath the Bismarck Sea plate at the New Britain
al., 2002; Wynn and Stow, 2002). O’Leary and Laine (1996), H.J. Trench (Johnson, 1979; Taylor, 1979). In the west, subduction
Lee et al. (2002), and Wynn and Stow (2002) provide criteria to has ceased as the Finisterre and Adelbert terranes, parts of a rem-
distinguish among these various processes. nant Paleogene volcanic arc that lies in the Bismarck forearc, col-
Because large undulations (hundreds of meters to a few kilo- lide with the Australian plate and are being uplifted to form the
meters) are easily identifiable in sonar surveys of the seafloor, and Finisterre and Adelbert Mountains (Abbott et al., 1994; Pigram
hyperpycnal turbidity currents are greatly affected by climate, and Davies, 1987).
uplift, and seismicity (Mulder et al., 2003), turbidity-current- The volcanism in the arc is related to subduction of the
generated sediment-wave fields should be ideal targets for study- Solomon Sea plate and occurs behind the remnant Paleogene
ing these factors in arc and arc-collisional environments. In the arc, which also includes the island of New Britain. The subma-
Bismarck Sea we imaged multiple sets of large sediment waves, rine environment surrounding the Bismarck Volcanic Arc was
Sediment waves in the Bismarck Volcanic Arc 93

144˚ 146˚ 148˚ 150˚ 152˚


-3˚ -3˚
ismic Zone
Bismarck Sea Se
Schouten Islands Bismarck Sea

-4˚ k R.
Manam Garove R -4˚
Sepi Ritter debris
TE
Karkar Dakataua KBE Fig 14
Tolokiwa flow GP
HB

WB
Fig 5 Fig 9
Adelbert Fig 2 L

F
-5˚ Mts. U Pandi -5˚
Ne Sakar WP KB B ts.
w Long G ai M River
Gu P K kan
ine Umboi Na
aF
old
Ra New Britain
mu
-6˚ an
dT
-M
ark Finisterre Ritter -6˚
hr ha Mts.
us m
tB Fa
elt ult ench
in Tr
100 km ew B
rita
-7˚ N -7˚
Solomon Sea
-8˚ -8˚
144˚ 146˚ 148˚ 150˚ 152˚
Figure 1. Location map showing regional setting of the wave fields of the Bismarck Volcanic Arc. The New Britain Trench accommodates
subduction of the Solomon Sea plate. New Britain and the Finisterre and Adelbert terranes are parts of a remnant Paleogene arc. The modern
subduction-related volcanic arc is offset to the north of the remnant arc. From west to east, some of the modern volcanoes include the Schouten
Islands, Manam, Karkar, Long, Tolokiwa, Umboi, Sakar, Ritter, Garove, Dakataua and the Willaumez Peninsula (WP), Pago (P), Karai (K), Gal-
losfulo (G), Bamus (B), Lolobau (L), Ulawun (U), and Rabaul (R), which lies on the northeast tip of the Gazelle Peninsula (GP) on the island of
New Britain. The Willaumez Peninsula is composed of a north-south string of volcanoes. The locations of several other volcanoes are shown as
small black circles. Major faults in the region include the Ramu-Markham Fault and the Wide Bay Fault (WBF). Two new faults were mapped
during this survey: the Kimbe Bay Escarpment (KBE) in Kimbe Bay (KB), east of the Willaumez Peninsula, and the Torkoro Escarpment (TE),
in the Hixon Bay region (HB) north of Lolobau. The outline of the debris flow from the 1888 collapse of Ritter Island is shown. The Sepik River
is the largest river on the north coast of Papua New Guinea. It drains a large part of the New Guinea Fold and Thrust Belt. The Pandi River drains
a large part of the Nakanai Mountains in New Britain, and the eastern flanks of Ulawun and Bamus volcanoes. The locations of Figures 2, 5, 9,
and 14 are shown.

the target of this study. We mapped submarine volcanic collapse that juts northward from New Britain into the Bismarck Sea,
features, including deposits of the 1888 Ritter Island collapse ending in Dakataua caldera. To the east of the Willaumez Pen-
(Johnson, 1987), for input into tsunami models (e.g., Ward, insula is Kimbe Bay, which is bounded on three sides by large
2001). The locations of a selection of volcanoes in the arc are volcanoes. It is also the location of a newly mapped fault, the
shown in Figure 1. The arc extends westward to the Schouten Kimbe Bay Escarpment.
Islands between 143° E and 145° E, where it meets the Bismarck Collapse blocks were found near several volcanoes dur-
Sea seismic zone. It extends east to ~152° E at Rabaul caldera ing the survey, including off Tolokiwa Island and off Dakataua
on the northeast tip of the Gazelle Peninsula. The study area caldera. At both these locations, sediment-wave fields were also
includes the Schouten Islands in the west but goes only so far mapped nearby. Sediment-wave fields were also mapped in
as the Hixon Bay area just west of the Gazelle Peninsula. The Kimbe Bay and the Hixon Bay region (boxed regions of Fig. 1).
Gazelle Peninsula is moving northwest relative to the rest of These sediment-wave fields form the basis of the current paper.
the island of New Britain along the left-lateral Wide Bay Fault
(Fig. 1; Madsen and Lindley, 1994). The Wide Bay Fault defines SEDIMENT-WAVE FORMATION AND DISTINCTIVE
the eastern edge of the Hixon Bay region, where we mapped a FEATURES
new fault, the Torkoro Escarpment.
The study area includes the deep-water (>500 m depth) Sediment waves, in the most common use of the term, refers
environment around the volcanoes in the arc, including on to any undulating pattern that is caused by differential sediment
either side of the Willaumez Peninsula, a string of volcanoes deposition and erosion over time, analogous to dunes or antidunes,
94 Hoffmann et al.

as opposed to undulating patterns that are caused by slope failure The generally accepted concept of bottom-current sediment-
such as slow deformation or slide events (H.J. Lee et al., 2002; wave generation is that internal waves induced within a current
Wynn and Stow, 2002). Large undulations with wavelengths of in a density-stratified medium flowing over topography will dif-
100 m to >10 km are observed in a variety of settings, including ferentially deposit and erode sediment on the seafloor (Flood,
on channel levees, axes, and mouths or in canyons (Normark et 1988). Sediment waves formed by bottom currents have simi-
al., 1980; Kidd et al., 1998; Nakajima and Satoh, 2001; Migeon lar dimensions as turbidity-current-generated waves (Wynn and
et al., 2001; Lewis and Pantin, 2002); on the flanks of volcanic Stow, 2002). Wave crests tend to be linear or sinuous with bifurca-
islands (Wynn et al., 2000); in troughs (Howe, 1996); and on pla- tions (Wynn and Stow, 2002). Bedforms often migrate upcurrent
teaus and continental slopes (Lee and Chough, 2001; O’Leary but can also migrate downcurrent (Wynn and Stow, 2002). They
and Laine, 1996). Short-wavelength (0.1–1 m) sediment waves are distinct from turbidity-current-generated sediment waves in
have been observed in settings such as terrestrial rivers (Mohrig that they can form on very low slopes or flat seafloor, and wave
and Smith, 1996; Jerolmack and Mohrig, 2005a, 2005b; Cole- crests are often aligned obliquely to the slope because the cur-
man et al., 2005). Owing to imaging limitations in the deep-sea rents are contour-parallel (Wynn and Stow, 2002). Both bottom-
environment, however, we confine our current discussion to large and turbidity-current-generated waves tend to display asymmetry
sediment waves. between flanks in backscatter imagery (H.J. Lee et al., 2002).
Two processes are recognized as producing sediment Undulating patterns can also arise from deformation, how-
waves as defined here, namely turbidity currents and bottom ever. These can occur from extension (O’Leary and Laine, 1996)
currents (Wynn and Stow, 2002; Schwehr et al., 2007). Waves or compression (Lee and Chough, 2001; Hill et al., 2004). These
of turbidity-current origin are created by numerous episodic undulating patterns may be related to rapid slope failure, such as
events preferentially depositing sediment on the upstream flank slumps (H.J. Lee et al., 2002), or to slow deformation caused by
of the waves in antidune fashion when the flow is in the Froude intermittent earthquake loading (Lee and Chough, 2001). These
number regime of ~0.5–1.9 (Hand, 1974; Bowen et al., 1984; features can be up to 10 km in wavelength, and up to 100 m in
Wynn et al., 2000; H.J. Lee et al., 2002). The Froude number is wave height (Wynn and Stow, 2002), and they can occur on steep
the ratio between current speed and the speed of gravity waves slopes or very shallow slopes (<0.5°; Lee and Chough, 2001).
in a density-stratified fluid, such as a sediment-laden-current In profile, deformational features are commonly associated with
layer in seawater. Recent research suggests that initiation of faulting (O’Leary and Laine, 1996; H.J. Lee et al., 2002). Plan-
a turbidity-current sediment-wave field may require breaks in form imagery of these features is rare, but wave crests may be
slope or preexisting undulations, but once started, these wave arcuate, without bifurcation (Wynn and Stow, 2002).
fields may grow on their own (Ercilla et al., 2002; H.J. Lee et Bottom currents, turbidity currents, and deformation can all
al., 2002; Kubo and Nakajima, 2002). interact in the creation of sediment-wave fields (Howe, 1996;
Deep-sea sediment waves caused by turbidity currents gen- Faugères et al., 2002; H.J. Lee et al., 2002). In this manuscript
erally have wavelengths of 200 m to 7 km, and wave heights of we will try to tease out the dominant and subordinate processes
2–70 m, and occur on slopes ranging from 0.1° to 0.7° (Wynn and responsible for creating these morphologic “sediment-wave”
Stow, 2002; H.J. Lee et al., 2002). In profile, the waves migrate features.
upstream and upslope because of preferential deposition on the
upstream flank (Wynn and Stow, 2002; H.J. Lee et al., 2002). TURBIDITY FLOWS AND HYPERPYCNAL
Internal reflections are commonly continuous between waves CURRENTS
(H.J. Lee et al., 2002). Wave crests tend to be oriented perpen-
dicular to the local slope, as turbidity currents travel downslope Considerable controversy exists regarding the nature of
(Wynn and Stow, 2002; H.J. Lee et al., 2002; Schwehr et al., flow and deposits of turbidity currents, as well as the terms
2007). Wave sequences progressively thin downslope, wave used in the literature that apply to various sediment transport
heights can decrease downslope, and some reflections pinch out regimes (Kneller and Buckee, 2000; Shanmugam, 2000; Mul-
downslope in profile, since sediment concentration and depo- der and Alexander, 2001; Mulder et al., 2001; Shanmugam,
sition in turbidity currents progressively decrease downslope 2002; Mulder et al., 2002). For the sake of clarity in our present
(Mulder et al., 1998). Thus deposition in the distal end will be discussion, we note an overlap between the terms hyperpycnal
from large events that occur only rarely (H.J. Lee et al., 2002). current and turbidity flow, and we define our particular usage
Wave crests tend to be linear or sinuous in plan view, with some in this paper.
bifurcation (Wynn and Stow, 2002; H.J. Lee et al., 2002). Waves Following Mulder and Alexander (2001), hyperpycnal cur-
formed by turbidity currents are often associated with other evi- rent refers to sediment-laden river discharge that is denser than
dence of turbidity-current activity, such as channels, levees, and the body of water into which it enters. The flow can travel long
scour features. In their review, H.J. Lee et al. (2002) note that the distances in a semidiscrete layer, entraining fluid and possibly
overall cross-sectional shape of a sediment-wave field is com- additional sediment if it is exerting shear stresses on the bed
monly concave upward. high enough to cause erosion (Skene et al., 1997; Mulder and
Sediment waves in the Bismarck Volcanic Arc 95

Alexander, 2001; Mulder et al., 2003). For river discharge enter- turbidity currents. We begin in the central part of the arc north
ing seawater, the sediment concentration needs to be at least of Tolokiwa, and then continue eastward to look at Dakataua,
~40 kg m–3 for the flow to be negatively buoyant (Mulder and Kimbe Bay, and Hixon Bay (Fig. 1).
Syvitski, 1995; Warrick and Milliman, 2003). Depending on the
river, hyperpycnal-discharge events can occur about once every TOLOKIWA WAVE FIELDS
100 yr, or as often as once per year (Mulder et al., 2003).
Turbidity flow refers to sediment suspended primarily by the Tolokiwa Island volcano, near the center of the Bismarck
upward component of turbulence (Lowe, 1982; Middleton, 1993; Volcanic Arc (Fig. 1), has undergone at least one lateral collapse
Stow et al., 1996; Shanmugam, 2000; Mulder and Alexander, in the Holocene. The sediment waves occur in a field north of the
2001). Such flows contain suspended sediment concentrations island on the northern side of a broad channel ~8 km wide and
up to 10% by volume (Shanmugam, 2000; Mulder and Alexan- 40 m deep that grades at ~0.2° to 0.5° from west to east into the
der, 2001). They can be generated by a number of mechanisms, basin (Fig. 2).
such as entrainment of sediment into surrounding seawater at Two sediment-wave fields overlap in this region. The first is
the upper surface of a debris flow. Other mechanisms include composed of large sediment waves with wavelengths of 600 m
hyperpycnal river discharge, and concentration processes in a to 1 km and wave heights of ~10 m to 40 m (Fig. 3). The second
low-density layer (Lowe, 1982; Middleton and Hampton, 1976; field is composed of smaller sediment waves with wavelengths
Parsons et al., 2001) such as a layer of tephra fallout during erup- of 100 m to 400 m and wave heights of a few meters (Fig. 3;
tions (Fiske et al., 1998). Coastal erosion during storm events Fig. 4C, E). Waves in both sets are identifiable in side-scan
that affect easily eroded sequences could also generate turbidity images as having a distinctly higher backscatter along the eastern
currents (Hayes et. al., 2002). Scully et al. (2002), for instance, flanks (Fig. 3).
showed that energetic water waves increase the capacity for crit- The field of large sediment waves is overlapped by the field
ically stratified gravity flows to transport sediment. Combined of small sediment waves, with some small waves appearing
with direct turbidity-current generation from ashfall and indirect between and overlying large waves (Fig. 3; Fig. 4C). The crests
formation from reworking of tephra, the presence of volcanoes of the small waves are oriented slightly oblique to their local
may significantly increase the occurrence of turbidity-current slope. This obliquity varies from wave to wave but is <30°. The
generation in this setting over other settings. We will therefore crests of the large waves are perpendicular to the local slope.
use turbidity current to refer to a sustained turbidity flow gen- The high-backscatter eastern, downslope flanks of the waves
erated by any mechanism, and hyperpycnal current to refer to correspond to the thinner, more steeply dipping beds of the waves
turbidity currents generated owing to hyperpycnal discharge at (Fig. 4). The large sediment waves commonly show differential
river mouths. deposition, with the upslope side receiving a factor of 2 or more
accumulated sediment than the downslope side. The crests of the
METHODS sediment waves migrate upslope through time (Fig. 4A, B). The
small waves lap onto the large waves (Fig. 4C, D), and they over-
Multibeam bathymetric data were collected aboard the R/V lie parallel reflectors that begin at a depth of 5 m (Fig. 4E).
Kilo Moana on cruise KM0419, using a hull-mounted SIMRAD
EM-120 echo sounder. Backscatter data were collected using a DAKATAUA WAVE FIELD
towed MR1 side-scan sonar system. Side-scan data were grid-
ded in a 16 m grid. Multibeam bathymetry and backscatter data The Dakataua wave field resides on the northern flank of
were collected along the 1000 km length of the arc in a swath Dakataua caldera. The field is 24 km by 18 km and elongated
up to 100 km wide. Bathymetric data were median filtered, using N-S (Fig. 5). The average north-dipping slope of the field is 1.4°,
a 500 m kernel. High-resolution Compressed High Intensity with the upslope flanks of the waves dipping south between 1°
“Radar” Pulse (CHIRP) sub-bottom sonar data were collected and 1.5°, and the downslope flanks dipping north ~10°. The wave
with an Edgetech 1–6 kHz swept frequency with a 50 ms dura- field as a whole forms a topographic high of up to 200 m relative
tion. The CHIRP data are unmigrated. This system transmitted to the flatter seafloor that surrounds the wave field. The seafloor
approximately every 5–10 m along selected transects. Penetration to the east of the wave field is deeper than that to the west. This
was up to a few tens of meters. Twelve short (<2 m) gravity and difference decreases northward, although north of the 240-m-
piston cores were collected, primarily targeting volcanic-collapse high ridge at the north end of the field the seafloor maintains an
debris flows identified in side-scan imagery. eastward-dipping component of ~0.15° (Fig. 5).
We examine four sediment-wave fields that were identified Between the caldera and the wave field, the seafloor dips
in the Bismarck Arc in order to determine the various genera- smoothly at ~5°. In the side-scan data the waves generally do
tion mechanisms of the sediment-wave fields. Doing so will help not exhibit asymmetry; however, the downslope flanks nearest
us understand how material is transported through the arc sys- the caldera display high backscatter (Fig. 6). Scour features are
tem and how important slope failure is compared to bottom and apparent on the 5° slope leading into the wave field (Fig. 6).
Fig 4

Ritter debris flow


es
Large Sed Waves av
dW
Se
all
Sm

Collapse
Blocks Fig 3

5 km

No Tolokiwa
Data Island

Figure 2. Bathymetry of Tolokiwa sediment-wave fields and surroundings. Two distinct fields overlap—one with wavelengths of 0.6–1 km,
labeled large sed waves, and one with wavelengths of 200–400 m, labeled small sed waves. Locations of Figures 3 and 4 are shown. Bathymetric
contour interval is 25 m, with annotated contours in bold every 200 m.
Ritter
debris
flow
Fig 4

Large waves
Small waves

147˚ 36' 147˚ 39' 147˚ 42' 147˚ 45' 147˚ 48'
-5˚ 02' -5˚ 02'
0
17860
17
0

40
172

17

Ritter
1820
1800

2 km debris
0

flow
170

-5˚ 04' -5˚ 04'


Fig 4

-5˚ 06' -5˚ 06'


Large waves
Small waves

-5˚ 08' -5˚ 08'

147˚ 36' 147˚ 39' 147˚ 42' 147˚ 45' 147˚ 48'
Figure 3. Side-scan mosaic (top) and bathymetry (bottom), contoured at 2 m intervals of Tolokiwa sediment-wave fields. Location of Figure 4
shown. The waves are apparent in the side-scan imagery as features with higher acoustic backscatter (darker shades) in the western flanks than
eastern flanks. Insonification direction is north-south. Nadir lines run east-west at 5°02′ S and 5°06′ S. Black east-west lines are sea-surface
reflection artifacts. In both parts of the figure, short dashed lines show local trend of slope. In side-scan image, several wave crests are denoted
with short, thin lines to highlight the slight obliquity of the crest-normal to the slope. Dotted lines outline the sediment-wave fields. Note in the
side-scan image that some small waves appear within the large waves in the region of overlap.
upslope migration (Left) The upslope flank is 9.2 m thick, measured to the lowest strong
reflection; the downslope flank, measured to the same reflection, is as thin
as 4.5 m. These measurements assume a sound velocity of 1500 m/s.

(Below) Small sediment waves onlapping


large sediment waves

W
A E

2.38
15 m VE = 17
2.40
500 m
2.42 D
TWT (s)

2.44

2.46
2.48
2.50
2.52

5m

125 m

B E
Note that reflections on the upslope sides Parallel reflections beneath small wavelength
of the waves thin or pinch out downslope sediment waves

(Left) Small sediment waves appearing


in between two large sediment waves,
and onlapping them.

Figure 4. CHIRP profile of eastern end of sediment-wave field north of Tolokiwa Island where the field merges with the ripple field. Location
shown in Figure 2. VE—vertical exaggeration; TWT—two-way traveltime. Enlarged boxes demonstrate particular features of note. All enlarged
boxes are at the same scale, which is 4 times larger than and at the same vertical exaggeration as the base figure.
150 00© 150 05© 150 10© 150 15©
-4 40© -4 40©
Dakataua Lineation

20
00 10 km
Dakataua
Wave Field

20
4 45© -4 45©

00
Fig. 7

Fig. 8

20
00
4 50© -4 50©

1800
Collapse Block
Field
0

16
180

00

4 55© -4 55©
1400
00
16
northwest
Wave ridge
Field

5 00© Fig. 6
-5 00©
No
Data ua
ata
ak
L. D

Dakataua Caldera

150 00© 150 05© 150 10© 150 15©


Figure 5. Location map showing wave field in relation to Dakataua caldera and bathymetry (in meters). Contour interval is 25 m, with annotated
contours in bold every 200 m. Locations of Figures 6, 7, and 8 are shown. A lineation, shown as the dash-dot line, runs through the field approxi-
mately N-S. The lineation ends slightly west of the 240-m-high ridge at the north end of the field.
150˚ 00' 150˚ 05' 150˚ 10'

4 km
Dakataua Lineation

-4˚ 45' -4˚ 45'

Fig. 7

Fig. 8
-4˚ 50' -4˚ 50'

collapse
blocks

High backscatter
downslope
flanks
-4˚ 55' -4˚ 55'

northwest ridge High backscatter scour streaks

150˚ 00' 150˚ 05' 150˚ 10'


Figure 6. Side-scan mosaic of Dakataua sediment wave field. Locations of Figures 7 and 8 shown. Scale is 1.6 times that of Figure 5. High
backscatter is dark. The two elongated regions of high backscatter in the field (~4°52′ S and 4°53′ S) are in the troughs between waves. Solid
lines denote the slope break between wave crests and steep downslope flanks of waves. Dashed lines denote enclosed depressions. The Dakataua
lineation is shown as the dash-dot line.
Sediment waves in the Bismarck Volcanic Arc 101

The waves themselves are 1–4 km in length, with wave 10° (Fig. 10). The high backscatter of the downslope sides of
heights of ~50 m, as measured from the upslope trough to the the waves is associated in the CHIRP profiles with thinning
crest. In plan view they are irregular, although generally arcu- sequences and upslope migration (Fig. 13). But in most cases
ate, with crests largely continuous across the width of the field (Fig. 13A, B, D) the upslope migration is apparent only on the
(Fig. 6). The crests are sinuous, and some are bifurcate (Figs. 6 steep downslope flanks, and only in a few areas can upslope
and 7), and they also tend to be wider than the troughs between migration be seen clearly across a whole wave, which might be
waves (Fig. 7). Many of the troughs between waves contain due to cutting across some waves obliquely (Fig. 13C). Addition-
enclosed depressions (Figs. 6 and 7). Some of the depressions are ally, the character of the reflections varies from wave to wave,
associated upslope with steep slopes that display high backscat- especially in terms of penetration depth. Some wave crests are
ter, but many are not (Fig. 6), and the depressions are not associ- imaged to a depth of ~8 m (Fig. 13B, C), whereas wave crests
ated with high backscatter scour. both upslope and downslope of these crests are imaged only to a
A lineation runs N-S through the field (Fig. 5), characterized depth of 2–3 m (Fig. 13A, D).
by eastward-dipping slopes, widened depressions, and pinched-
out wave crests (Fig. 7). The CHIRP failed to penetrate more HIXON BAY
than 1 m throughout this field (Fig. 8).
Hixon Bay allows us to test whether fluvial sedimentation is
KIMBE BAY WAVE FIELD an important process far from shore in this system. It is several
kilometers wide and the receptacle for the Pandi River (Fig. 14).
Whereas the Tolokiwa and Dakataua waves are both iso- This river drains ~800 km2 of the Nakanai Mountains and the
lated from large inputs of fluvial sediment, Kimbe Bay is fed on eastern flanks of Ulawun and Bamus volcanoes (Fig. 1).
three sides by a multitude of small rivers that drain volcanoes.
The Kimbe Bay wave field lies near the center of the bay, almost Backscatter, Morphology, and Sub-bottom Profiles of the
50 km from shore, adjacent to the Kimbe Bay Escarpment. Hixon Bay Sediment-Wave Field
The Kimbe Bay wave field is >30 km long and 15 km wide
(Fig. 9). In plan view the waves are generally arcuate, although A large sediment-wave field is found in the Tokoro Trough
irregular, with wavelengths varying from 1 to 4 km and wave of Hixon Bay (Figs. 14–16). This trough attains depths >1900 m,
heights reaching >80 m in some areas (Fig. 9). The average slope and two channels feed into the trough from the Pandi River
of the wave field is 1°. Within the field, many enclosed depres- (Fig. 14). The sediment-wave field (Fig. 15) begins shortly below
sions attain a depth of ~100 m and are found between wave crests the point at which channel scour diminishes between Torkoro and
(Fig. 10). Some low-wave-height features (~10 m) are found Mele Reefs. A slide scar is present east of Lolobau Ridge and is
west of the main sediment-wave field. The main wave field lies shown in more detail in Figure 17.
on a topographic high of up to 250 m relative to the sea floor In side-scan images (Fig. 18), high backscatter scour fea-
to the west (Fig. 9). To the east, a crevice ~2 km wide and up tures highlight the Pandi River channels. The slide scar shows
to ~100 m deep separates the wave field from the Kimbe Bay a stippled pattern of high backscatter in the upper part, and the
Escarpment. To the south of the wave field is a topographic ridge edges are highlighted as high backscatter streaks in the lower
at least 200 m high (Fig. 9). Several channels lead into the crev- half of the slide scar. The sediment-wave field (Fig. 18) appears
ice between the wave field and the escarpment. One channel is in backscatter as a pattern of alternating higher and lower back-
associated with high backscatter as it cuts through the ridge east scatter, with regions of varying wavelengths. We divide the wave
of the Kimbe Bay Escarpment (Fig. 11), then drops 200 m into field into six groups on the basis of morphology (Fig. 16), and we
a plunge pool (S.E. Lee et al., 2002), which is >100 m deep as summarize their characteristics in Table 1.
measured from the downslope side of the channel (Fig. 9). The six groups of waves are also distinct from one another
The crevice increases in width northward, to nearly 6 km, in sub-bottom imagery (Table 2). We correlate a transparent
where depressions appear (Fig. 9). The depressions and the layer between stronger reflections across group 2 and into group
escarpment appear as very high backscatter features in side-scan 4 (Fig. 19). Reflection thickness variations above this layer
imagery (Fig. 11). The high backscatter regions within these indicate that wave-crest deposition rates decrease monotoni-
depressions occur on the southern slopes, and the depressions cally downslope. The thickness varies by a factor of 1.5 across
attain depths up to 75 m (Fig. 12). A lineation is contrasted against the group 2 waves, and by nearly a factor of 2 from group 2 to
the background sediment of the wave field, appearing as a higher group 4 (Figs. 19–21). We also measured the relative thickness
backscatter region nearly 10 km long and 1 km wide (Fig. 11). of reflector sequences at the distal end of the wave field, tak-
The steeper, downslope sides of the sediment waves pro- ing advantage of continuous reflectors in group 4 (Fig. 22D).
duce higher backscatter than the rest of the wave field (Fig. 11). From the upslope side to the downslope side of Figure 22D the
The slopes of the high backscatter faces of the waves approach sequence thins by ~20%.
150˚ 04' 150˚ 06' 150˚ 08'
-4˚ 46' -4˚ 46'
19
20

1 km 19
0 80
196
1940 1960
1900 1940
1880 192
0
1860
bifurcation
1880
-4˚ 48' -4˚ 48'
1900

1860 1880

Fig. 8
18
20
0
184 1900
1860
18
0 40
1860 180
-4˚ 50' 1820 -4˚ 50'
17

19
Dakataua Lineation

80

00
182
bifurcation 0
180
0
80

bifurcation
17

182
0
0
176
80

-4˚ 52' 1780 1740 -4˚ 52'


16

0
172

1700

150˚ 04' 150˚ 06' 150˚ 08'


Figure 7. Bathymetry of Dakataua wave field, contoured at a 4 m interval, with annotated contours in bold every 20 m.
Location of CHIRP line depicted in Figure 8 shown. The downslope sides are consistently steeper than the upslope sides
of the features. Note the bifurcation near the center of the CHIRP line shown in Figure 8 and elsewhere. The Dakataua
lineation is characterized by eastward-dipping slopes, widened depressions, and pinched-out wave crests.
5m
CHIRP penetration is about 1 m.
250 m Diffractions give the illusory
appearance of deeper penetration.

S N

30 m VE = 13.9
2.40
1 km
2.44

2.48
T W T (s)

2.52

2.56
Data Gap

2.60

2.64

Downslope, the subsurface remains


largely impenetrable.

Figure 8. CHIRP profile of Dakataua wave field. Location shown in Figure 7. VE—vertical exaggeration; TWT—two-way traveltime. Detailed
enlargements are at the same vertical exaggeration and 3.5 times the scale as the base figure. Both demonstrate the extremely limited penetration
(~1 m) that characterizes this wave field.
Fig 12

Depressions

K imb e Bay Escarp me nt


Kimbe Wave Field

Plunge Pool
Fig 10

Crevice Channels
Fig 13

10 km
No Data

Figure 9. Bathymetry of the Kimbe Bay sediment-wave field. Contour interval is 25 m, with annotated contours in bold every 200 m. Side-scan
mosaic of the same region at the same scale is shown in Figure 11. Motion along the newly discovered Kimbe Bay Escarpment is unknown. In
the region between the escarpment and the wave field are several depressions associated with high backscatter. The outlines of these depressions
correspond to regions of high backscatter. Detailed bathymetry and backscatter imagery of the pull-apart basins are shown in Figure 12. Loca-
tions of Figures 10 and 12 are shown. CHIRP line depicted in Figure 13 is shown.
150˚ 38' 150˚ 40' 150˚ 42' 150˚ 44'

0
180
1 km
-5˚ 00' -5˚ 00'

18
1825

00
00
18
5
1775 180
0 177
00
175
0 18

Fig 13

17
-5˚ 02' 1725
-5˚ 02'

75
1700 1750
175
25

0
17

1700

17
1675

25
1675

1725 1675
1700

50
16

-5˚ 04' 00 -5˚ 04'


16 17
00
1625
16
75

16
1625
50
00
16

1675
16
00

-5˚ 06' -5˚ 06'


172
5

150˚ 38' 150˚ 40' 150˚ 42' 150˚ 44'


Figure 10. Detailed bathymetry of Kimbe Bay sediment-wave field. Contour interval is 5 m, with annotated contours in bold every 25 m. CHIRP
line depicted in Figure 13 is shown; the scale of this line is 2.9 times that of Figures 9 and 11. Note the pronounced irregular morphology of
the waves, with an overall arcuate pattern of alternating crests and troughs; superimposed on this is a chaotic assortment of peaks and valleys,
resulting in highly sinuous structures.
150˚ 30' 150˚ 35' 150˚ 40' 150˚ 45' 150˚ 50'
-4˚ 50' -4˚ 50'
Fig 12

Lin
ea
tio
n
-4˚ 55'

K imb e Bay Escarp me nt


-4˚ 55'

Fig 10 Plunge Pool


-5˚ 00' -5˚ 00'

Crevice
Fig 13

Channel
Scour
-5˚ 05' -5˚ 05'
10 km

-5˚ 10' -5˚ 10'


150˚ 30' 150˚ 35' 150˚ 40' 150˚ 45' 150˚ 50'
Figure 11. Side-scan mosaic of the same region depicted in Figure 9, and at the same scale. Dotted outline of field same as in Figure 9. Loca-
tions of Figures 10 and 12 are shown. CHIRP line depicted in Figure 13 is shown. Low backscatter is light, and high backscatter is dark. High
backscatter pattern within the wave field is associated with troughs. High-backscatter lineation is oriented 33.7° with respect to the general trend
of the Kimbe Bay Escarpment. Solid lines denote the slope break between wave crests and steep downslope flanks of waves. Dashed lines denote
enclosed depressions.
150˚ 38' 150˚ 40' 150˚ 42' 150˚ 44'
-4˚ 50' -4˚ 50'

-4˚ 51' -4˚ 51'

-4˚ 52' -4˚ 52'

-4˚ 53'
Lin

-4˚ 53'
ea
tio
n

-4˚ 54' -4˚ 54'


1 km

150˚ 38'
150˚ 38' 150˚ 40'
150˚ 40' 150˚ 42'
150˚ 42' 150˚ 44'
150˚ 44'
-4˚ 50' -4˚ 50'

00
20
19
2050 75 165
19 0
50
-4˚ 51' 50 -4˚ 51'
20

2025

20
202 75
5
-4˚ 52' 50 -4˚ 52'
20 20
25

-4˚ 53' 20 -4˚ 53'


Lin

00
ea

25
tio

20
n

-4˚ 54' -4˚ 54'


1 km

150˚ 38' 150˚ 40' 150˚ 42' 150˚ 44'


Figure 12. Close-up side-scan mosaic and bathymetry of enclosed depressions in the crevice, and the backscatter lineation noted in Figure 11.
Contour interval is 5 m, with annotated contours in bold every 25 m. Scale is 2.8 times that of Figures 9 and 11.
The waves display some downslope thinning (B) and
apparent upslope migration on their downslope B
flanks (A and B). The irregularity of the morphology
(see Figure 12) results in the CHIRP profiles imaging 10 m
reflections from adjacent surfaces at different 500 m
depths (A).
seafloor reflections
from different locations

upslope
A migration

sequence thins
upslope migration downslope

S N

2.20
30 m VE = 15
1 km
2.30
T W T (s)

2.40

2.50

upslope
C migration D 10 m
500 m

upslope
migration

Apparent upslope migration is present in most of the waves (C and D). In some cases this is true for the whole wave (C), but for
most of the waves, upslope migration is only apparent on the steep, downslope flanks where it becomes difficult to distinguish
between internal structure and diffractions (A, B, and D). Penetration varies widely from wave to wave, with some crests imaged
to a depth of up to 6 to 8 m (B and C), and others only imaged to a depth of 2 to 3 m (A and D), suggesting sedimentation
conditions are not consistent across the wave field.
Figure 13. CHIRP profile across a part of the Kimbe Bay wave field. Location shown in Figures 9, 10, and 11. VE—vertical exaggeration;
TWT—two-way traveltime. Detailed enlargements are at 4.2 times the scale and at the same vertical exaggeration as the base figure.
151˚ 10' 151˚ 15' 151˚ 20' 151˚ 25' 151˚ 30' 151˚ 35'
-4˚ 35' No Data
-4˚ 35'
Lolobau
ent

1 800
00
Trough
rpm

16
No Data
Esca

Torkoro
Trough
e
oro

idg

-4˚ 40' -4˚ 40'


R
Tork

oro

Eas
ork

00
12

t To
st T

rko
14
We

00

ro R
sediment waves

idg
e
No Data

00
-4˚ 45' Mele -4˚ 45'
16
e Rf
slid Mele
12
00 Seamount

Fig. 17

-4˚ 50' Fig. 16 -4˚ 50'


Torkoro Reef
Lolobau Ridge 5 km
Channels No Data

-4˚ 55' Lolobau -4˚ 55'


Island Hixon Bay

Pandi New Britain


Ulawun Volcano River
-5˚ 00' -5˚ 00'
151˚ 10' 151˚ 15' 151˚ 20' 151˚ 25' 151˚ 30' 151˚ 35'
Figure 14. Bathymetry of the Hixon Bay region. Contour interval is 25 m, with annotated contours in bold every 200 m. Several channels feed
into the imaged region, including those from the Pandi River, which drains part of the Nakanai Mountains (see Fig. 1) and Ulawun volcano. The
Pandi River channels fade out as they enter the region between the Mele Seamount and Torkoro Reef. This is where a major sediment-wave field
begins. This field is shown in more detail in Figure 16, the location of which is shown here. A slide scar is observed on the slope east of West
Torkoro Ridge and Lolobau Ridge. This scar is shown in more detail in Figure 17, the location of which is shown here.
151˚ 10' 151˚ 15' 151˚ 20' 151˚ 25' 151˚ 30' 151˚ 35'
-4˚ 35' -4˚ 35'
Ch
an
ne
ls
Lolobau co
ent

ur
Trough
m
carp

Torkoro
Trough
Es

e
oro

idg

-4˚ 40' -4˚ 40'


R
Tork

Large sediment waves


oro

Cha
Lower slide nne

Eas
l sco
ork

scar ur

t To
st T

rko
We

Small sed wvs

ro R
idg
Upper slide

e
scar
-4˚ 45' Mele -4˚ 45'
lower Rf
backscatter

Ch
Mele

an
Seamount

ne
stippled pattern

ls
co
u r
Fig. 17

-4˚ 50' Fig. 16 -4˚ 50'


Torkoro Reef
Lolobau Ridge Channel scour 5 km

-4˚ 55' Lolobau -4˚ 55'


Island Hixon Bay

Pandi New Britain


Ulawun Volcano River
-5˚ 00' -5˚ 00'
151˚ 10' 151˚ 15' 151˚ 20' 151˚ 25' 151˚ 30' 151˚ 35'
Figure 15. Hixon Bay area, with the location and scale the same as for Figure 14. Low backscatter is light, and high backscatter is dark. Regions
without data are white. The slope of the Torkoro Escarpment is notably high in backscatter. The slide scar shows up clearly as a pattern of high
backscatter. The backscatter pattern is distinctly different between the upper slide scar and the lower slide scar. The wave field within the box
locating Figure 16 can be seen as a repetitive pattern of higher and lower backscatter. Both large sediment waves and small sediment waves can
be seen. The large sediment waves are seen as a longer wavelength, lower contrast pattern, whereas the small sediment waves are seen as a shorter
wavelength, higher contrast pattern. Channels are associated with high-backscatter scour.
151˚ 18' 151˚ 20' 151˚ 22' 151˚ 24' 151˚ 26'

1650
-4˚ 38' -4˚ 38'

Fig 22
18

BB08
50
17
75

-4˚ 40' -4˚ 40'


1825

BB09
17

1800
50
17

1725 1675
00

1675 1775
-4˚ 42' 50 -4˚ 42'
17

1700
1725
0
165

Group 5
5

Group 6
162

75
1700 16

-4˚ 44' Group 4 1650 -4˚ 44'


r
sca 0
de 160 21
Sli Fig
15

1625
1550 25
75

4
1525
0 1

1600
Gr

1 km
140
ou

75
15 Seamount
p3
1500

50
-4˚ 46' 15 -4˚ 46'
1500

1475
1350

145
0

Group 2
13
50

1300 25
14

50
12
Fig

00
-4˚ 48' 12 -4˚ 48'
Fig
19

0 142
115 5
20

0 140
110 0
0 1375 Group 1
100

Torkoro Reef

151˚ 18' 151˚ 20' 151˚ 22' 151˚ 24' 151˚ 26'
Figure 16. Bathymetry of the Hixon Bay main sediment-wave field. Contour interval is 5 m, with annotated contours in bold every 25 m. Sedi-
ment waves are divided into five groups, based on morphology. Locations of CHIRP profiles through sediment waves in Figures 19 through 22
are shown. Arrows highlight the channels that lead to the sediment-wave field. See text for details on the groups of waves. Core holes BB08 and
BB09 are shown.
112 Hoffmann et al.

151˚ 14' 151˚ 15' 151˚ 16' 151˚ 17'


14
125 14 25
0 00
13
75
-4˚ 44' -4˚ 44'

13
0.5 km

50
region of undulating bathymetry

25
1175
13
-4˚ 45' -4˚ 45'
1300
Figure 17. Detailed bathymetry of the slide
scar in the Hixon Bay region. Contour inter-
1275

val is 5 m, with annotated contours in bold


p
12

ar every 25 m. The slide scar and scarp are de-


50

-4˚ 46' Sc -4˚ 46' lineated with the dotted line. In the upslope
portion of the scar, an undulating pattern is
122

seen. A CHIRP line that cuts across these


5

features is shown in Figure 25.

-4˚ 47' 26 -4˚ 47'


Fig 0
120

-4˚ 48' -4˚ 48'

151˚ 14' 151˚ 15' 151˚ 16' 151˚ 17'

Gravity Cores from the Hixon Bay Sediment-Wave Field lower in core BB09 than in core BB08. Subsequently, however,
sedimentation would have been more rapid by nearly an order of
We collected two gravity cores to examine the nature of sedi- magnitude in core BB09 than in core BB08, because of the cor-
ment being deposited and its rate of deposition. The core locations relation of the well-sorted fine white and black sand unit at 65 cm
are shown in plan view in Figure 16 and in cross section in Fig- depth in core BB09 and at 42 cm depth in core BB08 (Fig. 23).
ure 22. Core BB09 was taken in the trough between two waves, In both cases, this white and black sand unit is topped with a pale
just upslope of a 1-km-wavelength, 20-m-high wave. The slope gray clay unit that is slightly thicker in core BB08. Radiocarbon
on which it is located dips <0.05° to the south. Core hole BB08 is dating of planktonic foraminifers in core BB08 indicates that the
3.5 km downslope from core hole BB09 on the downslope flank white and black sand unit was deposited between 373 and 242 yr
of a wave near the very end of the field. The wave has a length of ago. The pale gray clay unit above it then took more than 200 yr
1 km and a height of 10 m. The north-dipping flank of the wave to emplace, at a rate of roughly 0.5 mm yr–1.
on which core hole BB08 is located has a slope of 1°. Core BB09 Calibrated radiocarbon dates of planktonic foraminifers from
is 164 cm long, and core BB08 is 85 cm long (Fig. 23). Depths in both cores are shown plotted in Figure 24. The depths of the dates
the cores are not corrected for compaction during collection. from each core have been normalized to the top of the well-sorted
Core BB09 penetrated through a unit of gray clay containing white and black sand layer we assume is from a single, discrete
pumice dropstones (~130 cm; Fig. 23). Core BB08 penetrated event affecting both sites. Ages in both Figure 23 and Figure 24
into, but not through, this same unit (~80 cm). In both cores a are given in years before 2004, when the cores were collected.
series of silty and sandy tephras is interbedded with fine silt, silty Ages are taken from median probability dates, with upper and
clay, and clay units (Fig. 23). The well-sorted fine-black-sand lower ranges shown.
tephra at 105 cm depth in core BB09 correlates physically with Sedimentation patterns are broadly similar at the sites of
the well-sorted fine-black-sand tephra at 53 cm depth in core cores BB08 and BB09, but in detail there are important differ-
BB08 (Fig. 23). If this correlation holds, then the sedimentation ences, in particular the thickness of the pale clay layer above the
rate between the clay-pumice unit and this tephra unit was slightly white and black sand layer, the relative thicknesses between the
151˚ 18' 151˚ 20' 151˚ 22' 151˚ 24' 151˚ 26'

-4˚ 38' -4˚ 38'

Fig 22
-4˚ 40' -4˚ 40'

high backscatter
spots
-4˚ 42' -4˚ 42'

Group 5
Group 6

-4˚ 44' -4˚ 44'


r Group 4
sca
de
Sli 21
Fig
1 km
Seamount
Gr
ou
p3

-4˚ 46' -4˚ 46'

Group 2
Fig
Fig

-4˚ 48' -4˚ 48'


20
19

scour
Group 1

Torkoro Reef

151˚ 18' 151˚ 20' 151˚ 22' 151˚ 24' 151˚ 26'
Figure 18. Side-scan mosaic of the central Hixon Bay region sediment-wave field. Dark tones are high backscatter. Region and scale are the same
as for Figure 16. Locations of the CHIRP lines depicted in Figures 19 through 22 are shown. The different groups of waves in the sediment-wave
field display different acoustic characteristics, with some groups displaying strong asymmetry between upslope and downslope flanks (groups 3
and 6), and others displaying weak or no asymmetry. High backscatter spots mark the slopes of a few group 4 waves nearest the slide scar.
114 Hoffmann et al.

TABLE 1. BACKSCATTER AND MORPHOLOGICAL CHARACTERISTICS OF HIXON BAY SEDIMENT WAVES


Group number Wave length Wave height Backscatter contrast Wave crests Channel
(m) (m)
1 600–800 20–30 None Slightly sinuous Yes
2 ~1000 30–80 Faint Somewhat sinuous No
3 500, D.D. 20, D.D. High Linear or somewhat sinuous Yes
4 1000–2000 30–40 High contrast spots near slide scar Slightly sinuous No
5 ~1000 15–20 Faint Slightly sinuous No
6 100–500 10–20 Faint Linear or slightly sinuous No
Note: Backscatter contrast refers to the flanks of individual waves. Channel denotes whether the group is associated with a
channel directly upslope. D.D.—decreasing downslope.

TABLE 2. SUB-BOTTOM PROFILE CHARACTERISTICS OF HIXON BAY SEDIMENT WAVES


Group number Penetration depth Reflector character Upslope Continuous reflectors Figure reference
(m) migration
1 <5 N.D. N.D. N.D. 19
2 Up to 20 Well stratified, D.T. Yes Between some waves 19, 20
3 <5 Primarily N.D. N.D. 20
diffractions
4 Up to 20, Well stratified, D.T. Yes Between some waves, 19, 21, 22
less at distal end at distal end
5 ~10 Well stratified, D.T. Yes Between some waves 21, 22
6 <5 Primarily N.D. N.D. 21
diffractions
Note: D.T.—downslope thinning, meaning the sequence is observed to thin downslope; N.D.—no data, meaning the group was not
imaged clearly in sub-bottom profiles.

well-sorted fine-black-tephra layer and the white and black sand In profile, one wave shows clear upslope migration (Fig. 25).
layer, and the number of upward-fining units above the white The complex morphology and slopes of up to 4° within the slide
and black sand layer (Fig. 23). The heterogeneity between the scar prevent clear imaging of the field with a shallow-towed
cores is not surprising because the wave field as a whole is het- CHIRP system. The downslope end of the slide scar also appears
erogeneous, and sedimentation patterns vary across the field as to have been buried (Fig. 16).
a whole.
DISCUSSION
Hixon Bay Region Slide
Based solely on the criteria given in O’Leary and Laine
The upper part of the Hixon Bay slide contains undulating (1996), H.J. Lee et al. (2002), and Wynn and Stow (2002), all
features, and sediment dispersed from slope failures within the of the imaged wave fields fail simple classification. Nonetheless,
slide scar may deposit on the distal end of the sediment-wave they provide important clues to the origins of sediment-wave
field (Fig. 17). The slide scar is apparent but diverges from a fields. Sediment waves, although well-recognized features, are
half-circle shape in its southern end. A clearly defined chute runs not ubiquitous but form only under very specific conditions.
downslope toward Torkoro Trough, although part of the northern Once the origin of a given wave field has been established, it
boundary of this chute is obscured by undulations that compose yields important information on broader processes.
most of the upper part of the slide scar. These undulations are
reminiscent of the sediment waves discussed above, with a wave- Tolokiwa Sediment-Wave Fields
length of roughly 1 km and wave heights of up to 40 m. Either
side of the slide chute appears as a channel ~5 m deeper than the The upslope migration and acoustic asymmetry observed in
center of the chute (Fig. 17). the large sediment waves northeast of Tolokiwa suggest a current
In side-scan imagery these channels are associated with high origin. The dimensions of the waves and the average slope of
backscatter streaks (Fig. 15). The waveforms appear as alternat- the field are consistent with either a turbidity-current origin or
ing high and low backscatter regions, with the steeper, downslope a bottom-current origin. The region is seismically active, and so
flanks associated with high backscatter. Where the overall slope the low slopes (<0.5°) are also consistent with a deformational
is flattest the contrast in backscatter is much less. origin, but the fields are not immediately associated with a fault.
sequence thins
downslope

SE NW

1.75
VE = 8.2
50 m
Group 1 1 km
1.85
Levee
TWT (S)

1.95

2.05
Group 2
2.15
Group 4

2.25

upslope migration

10 m
200 m
continuous reflections
between features

Figure 19. CHIRP line across Hixon Bay sediment waves. Location shown in Figure 18. Detailed enlargements are at the same vertical exaggera-
tion and at 5.1 times the scale of the central base figure. The black tick marks indicate a transparent layer between a series of reflections that is
correlated throughout this part of the wave field. Angled lines highlight upslope migration in the wave field. This CHIRP line transects a small
part of group 1, a small levee, group 2, and a small part of group 4. VE—vertical exaggeration; TWT—two-way traveltime. The thickness of
deposits above the transparent layer increases upslope from ~5 m in group 4 to 6 m at the bottom of group 2 to 9 m at the top of group 2. This
measurement is for the thickness at the wave crest and assumes a sound velocity of 1500 m s–1.
erosional surface
10 m 200 m

SE NW

1.60

1.70 VE = 6.6
channel 50 m 1 km
1.80
levee
TWT (s)

1.90

2.00
Group 2
2.10 Group 3
2.20

Figure 20. CHIRP line across the Hixon Bay sediment-wave field. Location shown in Figure 18. The line begins within the channel leading to
group 3. It then crosses the levee of the channel at a highly oblique angle. The line then crosses the largest waves of group 3. Because the group
3 waves are small, diffractions dominate the CHIRP image. Penetration in the group 3 waves is also <5 m. The group 2 waves display upslope
migration, as highlighted by the angled lines. VE—vertical exaggeration; TWT—two-way traveltime.
10 m
250 m

2.10
W E

2.20
TWT (s)

outside of
wave field Group 6
2.30

50 m VE = 13.9
Group 4 Group 5
1 km
2.40

10 m
250 m

Figure 21. CHIRP line across the Hixon Bay sediment waves. Location shown in Figure 18. This line cuts obliquely across the southern end
of group 4, group 5, and the central part of group 6, and slightly beyond. The tick marks indicate a transparent layer that we interpret to be the
same transparent layer seen in Figure 19. VE—vertical exaggeration; TWT—two-way traveltime. The group 4 waves appear to display upslope
migration, as highlighted by the angled lines. Note the decreasing penetration from group 4 to group 5 to group 6.
B Upslope migration
C BB09

S N

2.30
50 m VE = 16
TWT (s)

2 km
2.40
Group 4
BB09 Torkoro
Group 5 BB08 Trough
2.50

continuous reflections

A D BB08
10 m 500 m

Upslope migration

Figure 22. CHIRP profile of Hixon Bay sediment wave field. Location shown in Figure 18. VE—vertical exaggeration; TWT—two-way travel-
time. Core holes BB08 and BB09 are shown. This line crosses the group 5 waves perpendicular to their crests. Each of these waves displays clear
upslope migration, as shown in the two enlargements (A and B). In the group 4 waves shown here, upslope migration can be seen (angled line
in panels C and D). Also apparent is a downslope thinning of the sequence. Reflections in panel D are continuous and grade into parallel reflec-
tions in Torkoro Trough, and thin by ~20% from the upslope-most wave in panel D to the downslope-most wave. This thinning is also seen when
comparing cores BB08 and BB09 (see Fig. 23).
Depth (cm)
age before BB09 description description
BB08 age before
2004
2004
0 faintly bedded brownish-gray
0
fine silt fining up to clay upward-fining silty-clay
to clay

faintly bedded brownish-gray thin dark tephra


fine silt fining up to clay bearing silts and clay 97
upward-fining silt to clay
faintly bedded brownish-gray
fine silt fining up to clay
25 upward-fining silt to 25
basal mixed fine sand poorly sorted silty clay
faintly bedded brownish-gray
90 fine silt fining up to clay pale gray clay 187
faintly bedded brownish-gray sorted fine white and
fine silt fining up to clay 242
black sand
silty clay 373
pale clay 437
upward-fining fine-silt to
50 homogeneous brownish-gray silty clay
50
silty clay well sorted fine black sand tephra
119 sand injection upward-fining fine-silt to 437
silty clay
pale gray clay silty clay 573
sorted fine white and black sand fine silt
tephra silt/fine sand
poorly sorted upward-fining
clay and silt brown-gray silty clay
75 75
gray clay with pumice
brown mud dropstones

laminated silt and silty-clay

black sandy tephra


gray clay
100 tephra sand injection 100
black well sorted fine
sand tephra

interbedded clays and tephras

125 125
gray clay with pumice dropstones

pale gray well sorted silt


tephra

150 pale gray homogenous clay 150

silty tephra

Figure 23. Cores BB09 and BB08 from the Hixon Bay region sediment-wave field, with calibrated radiocarbon ages of planktonic foraminifers.
Both cores penetrated a gray clay unit containing pumice dropstones. Above this unit is a series of tephra deposits and silt, silty clay, and clay
deposits. Both units contain a thin layer of well-sorted white sand with black grains (65 cm depth in core BB09, and 42 cm depth in core BB08),
topped by a unit of pale gray clay. Ages (in years before 2004) from this pale gray clay unit indicate that it was deposited slowly in comparison with
the upward-fining silt and clay units found above it in both cores. Depths are uncorrected for compaction of sediment during core collection.
0

-10
4 mm
6 mm y

yr (
-1 left

-20
r-1 (righ

axis)
t axis)

Depth in BB09 (cm)


Depth in BB08 (cm)

-30
Top of pale clay layer (BB08)

-40 Top of well sorted white and black sand layer

-50

-60

-70

Data from BB08 Data from BB09


-80
0 100 200 300 400 500 600 700
Age (years before 2004)

Figure 24. Depth-age plot of radiocarbon dates from planktonic foraminifers from cores BB08 and BB09 in the Hixon Bay sediment-wave field.
Age is in years before 2004, when the cores were collected. Depths have been normalized to the top of the well-sorted white and black sand
layer we observe in both cores, and which we assume is from a single, discrete event affecting both sites. Depth within core BB08 is shown on
left axis. Depth within core BB09 is shown on right axis. Median ages are shown, with upper and lower ranges shown. For comparison, lines
corresponding to example sedimentation rates for each axis are drawn. The sedimentation rate for core BB09 since the late nineteenth century is
~6 mm yr–1. The average sedimentation rate for core BB08 since the event that produced the white and black sand layer is about two-thirds the
average sedimentation rate in core BB09 since the same event. Thus, 4 mm yr–1, using the scale of the left axis, is shown.
One wave in the field (right)
shows apparent upslope-
migration

1.50
Scarp

1.60
VE = 11.9
30 m
TWT (s)

1 km
1.70

1.80

1.90

5m
200 m

Figure 25. CHIRP line through Hixon Bay slide. Detailed enlargements are at the same vertical exaggeration and at 5 times the scale of
the base figure. VE—vertical exaggeration; TWT—two-way traveltime. Upslope migration is clear only in one wave in this field. Location
of figure shown in Figure 17.
122 Hoffmann et al.

1.9

1.8

1.7
relative deposition thickness

1.6

1.5

1.4

1.3

1.2

1.1

1
0 1 2 3 4 5 6 7
distance (km)
distance (km)

Figure 26. Plot of relative deposition thickness on wave crests in the Hixon Bay sediment wave field as a function of
relative distance. Measurements are taken from seven wave crests shown in Figure 19, normalized to the thickness of the
most distal wave crest measured. These are plotted as diamonds. Three measurements are taken from wave crests in the
distal end of the waves shown in Figure 22, plotted as crosses. These measurements are also normalized to the most distal
wave crest in the set. Distance is kilometers upslope from the most distal measurement in the set. The proximal measure-
ment from Figure 22 (the cross at 5.5 km) is extrapolated to the wave crest from the site of core hole BB09 and assumes
a relative sedimentation rate at the site of this core hole of 1.55 times the sedimentation rate at the site of core hole BB08.
The plot shows that these two sets of measurements, which are separated by 10 km, are consistent with a downslope ex-
ponential decay of sedimentation rate from wave crest to wave crest. Short-dashed line shows an exponential curve with
a characteristic decay length of 7.6 km.

We interpret that some of the small waves are younger features velocity and direction, and possibly a change in density stratifi-
than the large waves, owing to the onlapping contact that the cation within the ambient seawater (Flood, 1988). However, the
small waves have with the large waves. Nevertheless, other small crests of the large waves are oriented perpendicular to the local
waves interspersed with the larger waves could have formed con- slope, suggesting that they were formed by turbidity currents
comitantly. Because of their acoustic asymmetry, and the oblique derived from shallow regions to the west. The small waves thus
orientation of their wave crests with respect to local gradients, may have formed subsequent to or concomitantly with the large
we interpret the small waves to be of bottom-current origin. The waves as they were modified by bottom currents.
lack of evidence for turbidity-current activity in the vicinity of
the wave fields is consistent with this interpretation. Dakataua Sediment-Wave Field
The origin of the large waves remains inconclusive. We sug-
gest, however, that the consistent upslope migration more likely Features in the Dakataua wave field suggest both a deforma-
indicates a current origin than a deformational origin. The near- tional origin and a turbidity-current origin. The decreasing wave
est possible source of turbidity currents to the field is debris flows height with distance from the caldera (Fig. 5), combined with the
from Tolokiwa, but the orientation of the waves is inconsistent decreasing high backscatter scour features in side-scan imagery
with currents flowing from south to north. One alternative is that (Fig. 6), suggest that turbidity currents may play a role. The field
the large waves were also generated by bottom currents. In this is isolated from hyperpycnal currents, however. The wave field
case, the differences in morphology between the large waves begins ~5 km from shore, south of a steep slope that drops more
and small waves would imply a difference in bottom-current than 1200 m in this space. We suggest that coastal erosion by
Sediment waves in the Bismarck Volcanic Arc 123

energetic waves during large storms could result directly in tur- entirely distinct from one another, with some wave crests continu-
bidity currents as sediment is concentrated in the surf zone during ous through the boundaries of two or more groups (Fig. 18). This
peaks of storm activity and then released during lulls in wave overlap suggests that either the morphology predates deposition
action (Scully et al., 2002). Energetic waves could also result in by currents or that currents interact with topography, depend-
turbidity currents by destabilizing the slope of the caldera, result- ing on their magnitude, and that a continuum of magnitudes and
ing in numerous small slope failures of weakly consolidated sedi- slightly different flow paths will result in indistinct boundaries
ment that rapidly entrains seawater to become a turbidity flow between flow conditions, giving rise to the variability in the mor-
(Marr et al., 2001). Additionally, ashfall from eruptions can result phology of the field. Hyperpycnal currents that originate from the
in a hyperconcentrated sediment layer on the sea surface, which Pandi River will tend to follow the channels leading toward the
can result in turbidity currents (Fiske et al., 1998; Parsons et al., wave field, as seen by the high backscatter scour in these chan-
2001). Lastly, turbidity flows could be triggered by earthquakes. nels (Fig. 18).
Turbidity-current-generated sediment waves proximal to volca- A turbidity-current origin for the Hixon Bay sediment
nic sources but lacking fluvial input are recognized elsewhere waves is strongly supported by sub-bottom data. The similarity
(Wynn et al., 2000). of reflections from wave to wave (Figs. 19–22), the progressive
The arcuate and irregular morphology of the waves, as well downslope thinning of reflection sequences (Figs. 19, 22), the
as the large enclosed depressions in the field, suggest that defor- continuity of reflections between waves (Figs. 19, 22), and the
mation of some kind may play a role in the formation of these ubiquitous upslope migration of waves (Figs. 19–22) all argue
sediment waves. The Dakataua lineation (Fig. 7) may be a fault against a deformational origin and in favor of a current origin.
with some degree of downthrow to the east. We suggest that the Torkoro Trough is enclosed on three sides (Fig. 14). Cores taken
Dakataua wave field is formed by some combination of turbidity from the sediment-wave field show a series of upward-fining,
currents and deformation. Lacking sub-bottom data, however, the graded units consistent with turbidites (Fig. 23). These obser-
origin of the features remains speculative. vations argue against a bottom-current origin and in favor of a
turbidity-current origin of the wave field. The heterogeneous
Kimbe Bay Sediment-Wave Field flow conditions across the wave field may be largely the result
of focused flow down the channels and different magnitudes of
The Kimbe Bay wave field is far from a potential source of flow events.
hyperpycnal currents. The highly irregular and relatively arcu- The high-backscatter regions within the slide scar (Fig. 25)
ate plan-view morphology and the many enclosed depressions in and the high-backscatter spots on the upslope flanks of three
the field suggest a deformational origin. The waves display some sediment waves near the base of the slide scar (Fig. 18) suggest
subsurface characteristics of current activity, however (Fig. 13), active sediment transport through the slide scar. Some of this
in upslope migration of the wave crests. The presence of nearby sediment may reach the distal portion of the sediment-wave field.
channels and plunge pools indicates that turbidity currents are However, the undulating topography in the upslope part of the
active in the vicinity of this field. The crevice to the east of the slide scar (Fig. 17) implies a site of active deposition. This depo-
field may be related to motion along the Kimbe Bay Escarpment sition, together with the large, mountainous drainage area of the
if this is an active fault, or the crevice may be the site of focused Pandi River in comparison with the catchments of rivers drain-
channel flow. The depressions in the crevice may be tectonically ing the east flank of Lolobau or the west flank of Ulawun, and
controlled, although without subsurface imagery we cannot con- also the absence of channels or turbidity scour leading toward
firm this interpretation. The lineation in the field (Fig. 11) may be the slide scar, all suggest that hyperpycnal discharge from the
the site of a splay fault related to the Kimbe Bay Escarpment. Pandi River is the dominant source for the upward-fining units in
Like the Dakataua field, the Kimbe Bay field forms a topo- the cores. Slope failure and turbidity flows caused by rapid sedi-
graphic high relative to the surrounding seafloor, suggesting mentation and triggered by earthquakes also may explain some
higher sedimentation rates within the field. The two fields are also of the heterogeneity of the field. We conclude that the Hixon
similar in terms of wave height, wavelength, the presence of deep Bay sediment-wave field is of hyperpycnal-current origin, at
enclosed depressions within the fields, and irregular morphology. least insofar as active growth is concerned. We cannot determine
Both fields are closely related to signs of turbidity-current activ- whether or not the field grew from initial topography created by
ity and possible faults. We suggest that the Kimbe Bay field is some other process.
also formed by a combination of turbidity currents and deforma-
tion, although we cannot determine which process dominates. Sediment Supply to the Hixon Bay Sediment-Wave Field

Hixon Bay Sediment-Wave Field The Hixon Bay sediment-wave field demonstrates that sig-
nificant quantities of sediment are capable of reaching the Kimbe
The morphological heterogeneity of this sediment-wave Bay wave field, >40 km from shore. The Hixon Bay sediment-
field suggests that different parts of the field formed under dif- wave field covers ~70 km2 (Fig. 16). The waves display a pattern
ferent conditions. However, the groups we have defined are not of decreasing deposition downslope. Using a transparent layer
124 Hoffmann et al.

that can be correlated through much of the field, we measured the some of this may be related to currents flowing down the Hixon
relative thicknesses of deposits on seven wave crests (Fig. 19). slide scar (Fig. 25). Thus, we conservatively halve this estimate of
Farther downslope in the distal end of the field, we measured sediment flux arriving from the Pandi River. Integrating, the aver-
three more by tracing continuous reflections (Fig. 22D) and by age deposition rate for the whole field is ~5 mm yr–1. Multiplying
extrapolating the difference in sedimentation rates observed in by the area of the field, this yields a volumetric deposition of 3.5
cores BB08 and BB09 to the wave crest nearest core hole BB09 × 105 m3 yr–1. Assuming a bulk density of the sediment of 2200 kg
(Fig. 22C), taking advantage of continuous reflections between m–3 results in a mass deposition rate of 7.7 × 105 t yr–1. If this
this core and the wave crest. estimate represents a significant portion of Pandi River discharge,
Although flow conditions vary throughout the field, we then it places a testable lower bound on the sediment supply of
assume that sediment is primarily, although not exclusively, this system.
originating as discharge from the Pandi River. The similarity of We cannot characterize the uncertainty in this estimation, but
cores BB08 and BB09 (Fig. 23) and the similarity of reflection it suggests that a large part of the Pandi River sediment bypasses
sequences from wave to wave in groups 2, 4, and 5 (Figs. 19–22) the Papua New Guinea shelf and is deposited in deep water.
are consistent with this assumption. We assume no discontinuities Because we observe sedimentation rates of >4 mm yr–1 nearly
in sedimentation rate between groups. That reflector thickness 40 km from shore, we conclude that fluvial sediment entering
decreases steadily downslope from wave to wave, even between Kimbe Bay can easily account for the deposition we observe in
groups (Figs. 19, 22) is consistent with this assumption. The the Kimbe Bay wave field, although probably not the morphol-
waves display upslope migration, with thicker upslope sediment ogy itself. The rivers entering Kimbe Bay are smaller than the
deposits and thinner downslope sediment deposits. We assume Pandi River, on the whole, and 40 km from shore the Hixon Bay
that the thickness of sediment at a wave crest is approximately sediment-wave field is dying out, not just beginning, as is the
the average sediment thickness for a given wave. We normalize case in Kimbe Bay. Thus, although the Kimbe Bay field may
the seven measurements of thickness from Figure 19 to the thick- be dominantly of turbidity-current origin, it is probably not of
ness measurement of the wave crest farthest downslope in this hyperpycnal-current origin.
set. We similarly normalize the three measurements of thickness We note that previous work has suggested that as much as
from the distal end of the wave field (Fig. 22). Lastly, we assume 90% of sediment from the Sepik River may bypass the shelf as
a power-law distribution of hyperpycnal-current events from the sediment gravity flows (Walsh and Nittrouer, 2003). This infer-
Pandi River (Warrick and Milliman, 2003; Dadson, et al., 2005). ence was based on a comparison of sediment discharge from the
Thus, deposition rates are greater in the upslope part of the wave Sepik and measurements of depositional thickness on the shelf.
field because this region receives sediment from many small, fre- The measured thickness of deposits only accounted for 10% of
quent events as well as from large, infrequent events, whereas the expected discharge. A similar argument was made by Dadson
downslope part of the wave field receives sediment only from the et al. (2005) regarding sediment discharged at hyperpycnal con-
large, infrequent events (H.J. Lee et al., 2002). centrations from the Choshui River in Taiwan. The Hixon Bay
A power-law distribution of events is expected to produce an sediment waves present the opposite side of the story for the
exponential decay of sedimentation rates away from the source. Pandi River. Results from this field suggest that on active tec-
We find that the two sets of normalized relative-thickness mea- tonic margins with narrow shelves, hyperpycnal flows play an
surements are consistent with a single exponential decay rate important role in delivering sediment to the ocean. This may be
(Fig. 26). Linear functions will also fit both sets of measure- capable of explaining missing shelf deposits elsewhere, such as
ments. However, linear functions are not invariant to scalar multi- off the Sepik River, the Choshui River, and the Eel River (Walsh
plication. Because we are normalizing each set of measurements, and Nittrouer, 2003; Dadson, et al., 2005; Sommerfield and Nit-
a single linear function cannot fit both sets. The characteristic trouer, 1999).
decay rate of an exponential function, however, is invariant to
scalar multiplication. The characteristic decay length of this curve CONCLUSIONS
is ~7.6 km. That is, from wave crest to wave crest, sedimenta-
tion rates are inferred to decrease by half 7.6 km downslope from We image six sediment-wave fields in the Bismarck Volca-
any given point in the field. Based on the above assumptions, we nic Arc. The two Tolokiwa fields appear to have formed by tur-
extrapolate this decay rate to the remainder of the field and cali- bidity currents and were subsequently or concomitantly modified
brate it with sedimentation rates inferred from radiocarbon dating by bottom currents. The Kimbe Bay and Dakataua wave fields
of cores BB08 and BB09 (Fig. 24). appear to be of deformational and turbidity-current origin. In
Because we are interested in modern discharge of the Pandi Hixon Bay, two hyperpycnal-current-generated sediment-wave
River, we use only the upper parts of the cores, estimating sedi- fields are observed. One is a small, apparently young field in its
ment accumulation in the distal end of the field at 4 mm yr–1 initial stages of growth. The other field is an older, more stable
(Fig. 24). Note that this estimate does not account for compression feature related to discharge from the Pandi River. Measurements
of the cores during collection and so is a lower bound estimate. from this wave field indicate that a large part of sediment dis-
However, some of this is likely to be pelagic sedimentation, and charge from the Pandi River is deposited in deep water.
Sediment waves in the Bismarck Volcanic Arc 125

The evidence for turbidity currents in the vicinity of the Coleman, S.E., Zhang, M.H., and Clunie, T.M., 2005, Sediment-wave develop-
Dakataua and Kimbe Bay wave fields, and that both fields are ment in subcritical water flow: Journal of Hydraulic Engineering, v. 131,
p. 106–111, doi: 10.1061/(ASCE)0733-9429(2005)131:2(106).
topographic highs relative to the surrounding seafloor, suggest Dadson, S., Hovius, N., Pegg, S., Dade, W.B., Horng, M.J., and Chen, H., 2005,
focused deposition within the area of these wave fields. Between Hyperpycnal river flows from an active mountain belt: Journal of Geo-
these fields and the Hixon Bay wave fields, sediment waves with physical Research, v. 110, p. F04016, doi: 10.1029/2004JF000244.
Ercilla, G., Belén, A., Wynn, R.B., and Baraza, J., 2002, Turbidity current sedi-
turbidity-current components of formation appear to be a com- ment waves on irregular slopes: Observations from the Orinoco sediment-
mon feature in the Bismarck Volcanic Arc, particularly where wave field: Marine Geology, v. 192, p. 171–187, doi: 10.1016/S0025-
sedimentation rates are relatively high. In these locations, depo- 3227(02)00554-6.
Faugères, J.-C., Gonthier, E., Mulder, T., Kenyon, N., Cirac, P., Griboulard,
sition rates can vary greatly in the space of a few kilometers, both R., Berné, S., and Lesuavé, R., 2002, Multi-process generated sediment
along the strike of the arc because of focused sedimentation, and waves on the Landes Plateau (Bay of Biscay, North Atlantic): Marine
across the strike because of the variation of sedimentation rates Geology, v. 182, p. 279–302, doi: 10.1016/S0025-3227(01)00242-0.
Fiske, R.S., Cashman, K.V., Shibata, A., and Watanabe, K., 1998, Tephra dis-
between the upslope and downslope flanks of a sediment wave. persal from Myojinsho, Japan, during its shallow submarine eruption of
Thus, care must be taken when applying interpretations of a sedi- 1952–1953: Bulletin of Volcanology, v. 59, p. 262–275, doi: 10.1007/
mentary section to the remainder of the arc environment. s004450050190.
Flood, R.D., 1988, A lee wave model for deep-sea mudwave activity: Deep-Sea
Between these fields and the Tolokiwa fields, we do not Research, v. 35, p. 973–983, doi: 10.1016/0198-0149(88)90071-4.
observe large sediment waves. Topographic relief in this region Gardner, J.V., Prior, D.B., and Field, M.E., 1999, Humboldt slide—A large
is lower than in eastern New Britain, and so we expect that sedi- shear-dominated retrogressive slope failure: Marine Geology, v. 154,
p. 323–338, doi: 10.1016/S0025-3227(98)00121-2.
mentation rates are significantly lower. Thus, sediment waves Hand, B.M., 1974, Supercritical flow in density currents: Journal of Sedimen-
with turbidity-current contributions to formation may be an indi- tary Petrology, v. 44, p. 637–648.
cator of relatively high sedimentation rates, although the inverse Hayes, K., Montgomery, D.R., and Newhall, C.G., 2002, Fluvial sediment trans-
port and deposition following the 1991 eruption of Mount Pinatubo: Geo-
may not be true. We are not aware of large turbidity-current sedi- morphology, v. 45, p. 211–224, doi: 10.1016/S0169-555X(01)00155-6.
ment waves that are recognized in the sedimentary record. This is Hill, J.C., Driscoll, N., Weissel, J.K., and Goff, J.A., 2004, Large-scale elon-
probably due largely to an inherent difficulty in recognizing such gated gas blowouts along the U.S. continental margin: Journal of Geo-
physical Research, v. 109, p. B09101, doi: 10.1029/2004JB002969.
large, subtle features from outcrops, particularly if the sequence Howe, J.A., 1996, Turbidite and contourite sediment waves in the northern Rock-
is at all deformed. Sections in which high sedimentation rates all Trough: North Atlantic Ocean: Sedimentology, v. 43, p. 219–234.
have been recognized thus may be good places for reexamining Jerolmack, D., and Mohrig, D., 2005a, Interactions between bed forms: Topog-
raphy, turbulence, and transport: Journal of Geophysical Research, v. 110,
the sedimentary record for sediment-wave fields. p. F02014, doi: 10.1029/2004JF000126.
Jerolmack, D.J., and Mohrig, D., 2005b, A unified model for subaqueous
ACKNOWLEDGMENTS bed form dynamics: Water Resources Research, v. 41, p. W12421, doi:
10.1029/2005WR004329.
Johnson, R.W., 1979, Geotectonics and volcanism in Papua New Guinea: A
Funding for this research was provided by the Marine Geology review of the late Cainozoic: Bureau of Mineral Resources, Journal of
and Geophysics section of the U.S. National Science Founda- Australian Geology and Geophysics, v. 4, p. 181–207.
Johnson, R.W., 1987, Large-scale volcanic cone collapse: The 1888 slope fail-
tion, grant OCE-0327004 to E. Silver and S. Ward, and grant ure of Ritter volcano, and other examples from Papua New Guinea: Bul-
OCE-0328278 to N. Driscoll. We are grateful for the advice and letin of Volcanology, v. 49, p. 669–679, doi: 10.1007/BF01080358.
support of many people in carrying out this research, including Kidd, R.B., Lucchi, R.G., Gee, M., and Woodside, J.M., 1998, Sedimentary
processes in the Stromboli Canyon and the Marsili Basin, SE Tyrrhenian
Hugh Davies, James Robins, Bruce Appelgate and the Hawaii Sea: Results from sidescan sonar surveys: Geo-Marine Letters, v. 18,
Mapping Research Group, Russell Perembo and Susan John of p. 146–154, doi: 10.1007/s003670050062.
the University of Papua New Guinea for picking foraminifers, Kneller, B., and Buckee, C., 2000, The structure and fluid mechanics of tur-
bidity currents: A review of some recent studies and their geological
Michaele Kashgarian of the radiocarbon facility at Lawrence implications: Sedimentology, v. 47, p. 62–94, doi: 10.1046/j.1365-
Livermore National Laboratory for assistance with dating, the 3091.2000.047s1062.x.
Oregon State University coring team, and the captain and crew Kubo, Y., and Nakajima, T., 2002, Laboratory experiments and numerical simu-
lations of sediment-wave formation by turbidity currents: Marine Geol-
of the research vessel Kilo Moana. We thank A. Draut (Associ- ogy, v. 192, p. 105–121, doi: 10.1016/S0025-3227(02)00551-0.
ate Editor), H. Lee, J. Sample, and K. Straub for their thorough Lee, H.J., Syvitski, J.P.M., Parker, G., Orange, D., Locat, J., Hutton, E.W.H.,
and insightful reviews of this manuscript. Most of the figures for and Imran, J., 2002, Distinguishing sediment waves from slope failure
deposits: Field examples, including the ‘Humboldt slide,’ and model-
this paper were generated in part using Generic Mapping Tools. ing results: Marine Geology, v. 192, p. 79–104, doi: 10.1016/S0025-
3227(02)00550-9.
REFERENCES CITED Lee, S.E., Talling, P.J., Ernst, G.G.J., and Hogg, A.J., 2002, Occurrence and origin
of submarine plunge pools at the base of the US continental slope: Marine
Geology, v. 185, p. 363–377, doi: 10.1016/S0025-3227(01)00298-5.
Abbott, L.D., Silver, E.A., Thompson, P.R., Filewicz, M.V., Schneider, C., and Lee, S.H., and Chough, S.K., 2001, High-resolution (2–7 kHz) acoustic and
Abdoerrias, 1994, Stratigraphic constraints on the development and tim- geometric characters of submarine creep deposits in the South Korea Pla-
ing of arc-continent collision in northern Papua New Guinea: Journal of teau: East Sea: Sedimentology, v. 48, p. 629–644, doi: 10.1046/j.1365-
Sedimentary Research, v. 64, p. 169–183. 3091.2001.00383.x.
Bowen, A.J., Normark, W.R., and Piper, D.J.W., 1984, Modeling of turbidity Lewis, K.B., and Pantin, H.M., 2002, Channel-axis, overbank and drift sedi-
currents on Navy submarine fan, California Continental Borderland: Sedi- ment waves in the southern Hikurangi Trough, New Zealand: Marine
mentology, v. 31, p. 169–185, doi: 10.1111/j.1365-3091.1984.tb01957.x. Geology, v. 192, p. 123–151, doi: 10.1016/S0025-3227(02)00552-2.
126 Hoffmann et al.

Lowe, D.R., 1982, Sediment gravity flows. II. Depositional models with spe- O’Leary, D.W., and Laine, E., 1996, Proposed criteria for recognizing intrastratal
cial reference to the deposits of high-density turbidity currents: Journal of deformation features in marine high resolution seismic reflection profiles:
Sedimentary Petrology, v. 52, p. 279–297. Geo-Marine Letters, v. 16, p. 305–312, doi: 10.1007/BF01245561.
Madsen, J.A., and Lindley, I.D., 1994, Large-scale structures on Gazelle Penin- Parsons, J.D., Bush, J.W.M., and Syvitski, J.P.M., 2001, Hyperpycnal plume for-
sula, New Britain: Implications for the evolution of the New Britain Arc: mation from riverine outflows with small sediment concentrations: Sedi-
Australian Journal of Earth Sciences, v. 41, p. 561–569. mentology, v. 48, p. 465–478, doi: 10.1046/j.1365-3091.2001.00384.x.
Marr, J.G., Harff, P.A., Shanmugam, G., and Parker, G., 2001, Experiments on Pickup, G., 1984, Geomorphology of tropical rivers: I. Landforms, hydrology,
subaqueous sandy gravity flows: The role of clay and water content in and sedimentation in the Fly and lower Purari, Papua New Guinea, in
flow dynamics and depositional structures: Geological Society of America Schick, A.P., ed., Channel Processes—Water, Sediment, Catchment Con-
Bulletin, v. 113, p. 1377–1386, doi: 10.1130/0016-7606(2001)113<1377: trols: Braunschweig, Catena Supplement 5, p. 1–17.
EOSSGF>2.0.CO;2. Pigram, C.J., and Davies, H.L., 1987, Terranes and the accretion history of the
McAlpine, J.R., and Keig, G., 1983, Climate of Papua New Guinea: Canberra, New Guinea orogen: Bureau of Mineral Resources, Journal of Australian
Australian National University Press, 200 p. Geology and Geophysics, v. 10, p. 193–211.
Middleton, G.V., 1993, Sediment deposition from turbidity currents: Annual Schwehr, K., Driscoll, N., Tauxe, L., 2007, Origin of continental margin mor-
Review of Earth and Planetary Sciences, v. 21, p. 89–114, doi: 10.1146/ phology; submarine-slide or downslope current-controlled bedforms, a
annurev.ea.21.050193.000513. rock magnetic approach: Marine Geology, v. 240, p. 19–41.
Middleton, G.V., and Hampton, M.A., 1976, Subaqueous sediment transport Scully, M.E., Friedrichs, C.T., and Wright, L.D., 2002, Application of an ana-
and deposition by sediment gravity flows, in Stanley, D.J., and Swift, lytical model of critically stratified gravity-driven sediment transport and
D.J.P., eds., Marine Sediment Transport and Environmental Management: deposition to observations from the Eel River continental shelf, North-
New York, Wiley, p. 197–218. ern California: Continental Shelf Research, v. 22, p. 1951–1974, doi:
Migeon, S., Savoye, B., Zanella, E., Mulder, T., Faugères, J.-C., and Weber, O., 10.1016/S0278-4343(02)00047-X.
2001, Detailed seismic-reflection and sedimentary study of turbidite sedi- Shanmugam, G., 2000, 50 years of the turbidite paradigm (1950s–1990s):
ment waves on the Var Sedimentary Ridge (SE France): Significance for Deep-water processes and facies models—A critical perspective:
sediment transport and deposition and for the mechanisms of sediment- Marine and Petroleum Geology, v. 17, p. 285–342, doi: 10.1016/S0264-
wave construction: Marine and Petroleum Geology, v. 18, p. 179–208, 8172(99)00011-2.
doi: 10.1016/S0264-8172(00)00060-X. Shanmugam, G., 2002, Discussion on Mulder et al. (2001, Geo-Marine Letters,
Milliman, J.D., 1995, Sediment discharge to the ocean from small mountainous v. 21, p. 86–93), Inversely graded turbidite sequences in the deep Medi-
rivers: The New Guinea example: Geo-Marine Letters, v. 15, p. 127–133, terranean: A record of deposits from flood-generated turbidity currents?:
doi: 10.1007/BF01204453. Geo-Marine Letters, v. 22, p. 108–111, doi: 10.1007/s00367-002-0100-3.
Milliman, J.D., and Syvitski, J.P.M., 1992, Geomorphic/tectonic control of sed- Skene, K.I., Mulder, T., and Syvitski, J.P.M., 1997, INFLO1: A model predicting
iment discharge to the ocean: The importance of small mountain rivers: the behaviour of turbidity currents generated at river mouths: Computers
Journal of Geology, v. 100, p. 525–544. & Geosciences, v. 23, p. 975–991, doi: 10.1016/S0098-3004(97)00064-2.
Mohrig, D., and Smith, J.D., 1996, Predicting the migration rates of sub- Sommerfield, C.K., and Nittrouer, C.A., 1999, Modern accumulation rates
aqueous dunes: Water Resources Research, v. 32, p. 3207–3217, doi: and a sediment budget for the Eel shelf: A flood-dominated depositional
10.1029/96WR01129. environment: Marine Geology, v. 154, p. 227–241, doi: 10.1016/S0025-
Mulder, T., and Alexander, J., 2001, The physical character of subaqueous sedi- 3227(98)00115-7.
mentary density flows and their deposits: Sedimentology, v. 48, p. 269– Stow, D.A.V., Reading, H.G., and Collinson, J.D., 1996, Deep seas, in Reading,
299, doi: 10.1046/j.1365-3091.2001.00360.x. H.G., ed., Sedimentary Environments; Processes, Facies, and Stratigra-
Mulder, T., and Syvitski, J.P.M., 1995, Turbidity currents generated at river phy: Oxford, UK, Blackwell Science, p. 395–453.
mouths during exceptional discharges to the world oceans: Journal of Taylor, B., 1979, Bismarck Sea: Evolution of a backarc basin: Geology, v. 7,
Geology, v. 103, p. 285–299. p. 171–174, doi: 10.1130/0091-7613(1979)7<171:BSEOAB>2.0.CO;2.
Mulder, T., Syvitski, J.P.M., and Skene, K.I., 1998, Modeling of erosion and Walsh, J.P., and Nittrouer, C.A., 2003, Contrasting styles of off-shelf sediment
deposition by turbidity currents generated at river mouths: Journal of accumulation in New Guinea: Marine Geology, v. 196, p. 105–125, doi:
Sedimentary Research, v. 68, p. 124–137. 10.1016/S0025-3227(03)00069-0.
Mulder, T., Migeon, S., Savoye, B., and Faugères, J.-C., 2001, Inversely graded Ward, S.N., 2001, Landslide tsunami: Journal of Geophysical Research, v. 106,
turbidite sequences in the deep Mediterranean: A record of deposits from p. 11,201–11,215, doi: 10.1029/2000JB900450.
flood-generated turbidity currents?: Geo-Marine Letters, v. 21, p. 86–93, Warrick, J.A., and Milliman, J.D., 2003, Hyperpycnal sediment discharge from
doi: 10.1007/s003670100071. semiarid southern California rivers: Implications for coastal sediment
Mulder, T., Migeon, S., Savoye, B., and Faugères, J.-C., 2002, Reply to dis- budgets: Geology, v. 31, p. 781–784, doi: 10.1130/G19671.1.
cussion by Shanmugam on Mulder et al. (2001, Geo-Marine Letters, Wynn, R.B., and Stow, D.A.V., 2002, Classification and characterization
v. 21, p. 86–93), Inversely graded turbidite sequences in the deep Medi- of deep-water sediment waves: Marine Geology, v. 192, p. 7–22, doi:
terranean: A record of deposits from flood-generated turbidity currents?: 10.1016/S0025-3227(02)00547-9.
Geo-Marine Letters, v. 22, p. 112–120, doi: 10.1007/s00367-002-0096-8. Wynn, R.B., Masson, D.G., Stow, D.A.V., and Weaver, P.P.E., 2000, Turbidity
Mulder, T., Syvitski, J.P.M., Migeon, S., Faugères, J.-C., and Savoye, B., 2003, current sediment waves on the submarine slopes of the western Canary
Marine hyperpycnal flows: Initiation, behavior, and related deposits. A Islands: Marine Geology, v. 163, p. 185–198, doi: 10.1016/S0025-
review: Marine and Petroleum Geology, v. 20, p. 861–882, doi: 10.1016/ 3227(99)00101-2.
j.marpetgeo.2003.01.003.
Nakajima, T., and Satoh, M., 2001, The formation of large mudwaves by turbid-
ity currents on the levees of the Toyama deep-sea channel, Japan Sea: Sed-
imentology, v. 48, p. 435–463, doi: 10.1046/j.1365-3091.2001.00373.x.
Normark, W.R., Hess, G.R., Stow, D.A.V., and Bowen, A.J., 1980, Sediment
waves on the Monterey Fan levee: A preliminary physical interpretation:
Marine Geology, v. 37, p. 1–18, doi: 10.1016/0025-3227(80)90009-2. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

The Lichi Mélange: A collision mélange formation along early arcward backthrusts
during forearc basin closure, Taiwan arc-continent collision

Chi-Yue Huang
Department of Earth Sciences, National Cheng Kung University, Tainan, Taiwan, and Research Center of Ocean Environment
and Technology, National Cheng Kung University, Tainan, Taiwan

Chih-Wei Chien
Department of Earth Sciences, National Cheng Kung University, Tainan, Taiwan

Bochu Yao
Guangzhou Marine Geological Survey, Ministry of Land and Resource, P.R.China, Guangzhou 510075

Chung-Pai Chang
Center for Space and Remote Sensing Research, National Central University, Chungli 320, Taiwan

ABSTRACT

Marine surveys show that the submarine Huatung Ridge extends northward to
the Lichi Mélange in the southwestern Coastal Range, suggesting that formation of
the Lichi Mélange is related to arcward thrusting of the forearc strata in the western
part of the North Luzon Trough during active arc-continent collision off southern Tai-
wan. A new seismic survey along the 21° N transect across the North Luzon Trough in
the incipient arc-continent collision zone further reveals that deformation of the Hua-
tung Ridge occurred soon after sedimentation in the western forearc basin, whereas
sedimentation was continuous in the eastern part of the remnant North Luzon Trough
until the complete closure of the forearc basin approaching SE Taiwan. This suggests
that the sequence in the Huatung Ridge can be coeval with just the lower sequence of
the remnant-forearc-basin strata. Multiple lines of new evidence, including micropa-
leontology, clay mineralogy, and fission track analyses along the Mukeng River and
its tributary key sections, are used to test this thrusting-forearc-origin hypothesis of
the Lichi Mélange.
In the SW Coastal Range the Lichi Mélange lies between the collision suture of
Longitudinal Valley to the west and the Taiyuan remnant forearc basin to the east. A
field survey indicates that the Taiyuan forearc-basin sequence and its volcanic base-
ment were thrust westward over the Lichi Mélange along the east-dipping Tuluan-
shan Fault. The Lichi Mélange shows varying degrees of fragmentation of strata,
mixing, and shearing. An apparently wide range of facies is present, from the weakly
sheared broken formation facies, with discernible relict sedimentary structures, to
the intensely sheared block-in-matrix mélange facies, with pervasively scaly foliation

Huang, C.Y., Chien, C.W., Yao, B., and Chang, C.P., 2008, The Lichi Mélange: A collision mélange formation along early arcward backthrusts during forearc basin
closure, Taiwan arc-continent collision, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Colli-
sion Zones: Geological Society of America Special Paper 436, p. 127–154, doi: 10.1130/2008.2436(06). For permission to copy, contact editing@geosociety.org.
©2008 The Geological Society of America. All rights reserved.

127
128 Huang et al.

dipping to the SE. Sedimentological study reveals that the subangular to subrounded,
fractured, matrix-supported metasandstone conglomerates in the pebbly mudstone
layers are repeatedly found in the broken formation facies of the Lichi Mélange. Their
composition and occurrence are identical to the deep-sea-fan conglomerate beds in
the Taiyuan remnant-forearc-basin strata to the east. Benthic foraminiferal faunas
are similar in the Lichi Mélange, regardless of the varying intensity of shearing and
strata disruptions, and are compatible with the benthic foraminiferal fauna in the
Taiyuan remnant-forearc-basin turbidites, supporting the interpretation that the pro-
tolith of the Lichi Mélange was originally deposited in the North Luzon Trough. Age
determination of planktic microfossils further demonstrates that the Lichi Mélange is
early Pliocene (3.5–3.7 Ma), implying that this mélange was deposited in a short time
and that deformation occurred soon after its deposition. The early Pliocene age of the
Lichi Mélange is coeval with just the lower part of the Taiyuan remnant forearc stra-
ta, and is much younger than the upper forearc sequence (3–1 Ma). Thus the Taiyuan
coherent-forearc-basin strata (3.7–1 Ma) were deposited continuously in the remnant
North Luzon Trough regardless of the deformation in its western part (the proto-
mélange). This scenario is an analogue for the modern configuration of the Huatung
Ridge–remnant North Luzon Trough off the southern Coastal Range in the active
arc-continent collision zone north of lat 21° N.
In addition to its kaolinite content (11–15%), the clay mineral composition of
the Lichi Mélange is compatible with the Taiyuan remnant forearc turbidites. In the
Coastal Range, kaolinites are found only in the volcanic rocks of the Tuluanshan For-
mation. This additional kaolinite in the Lichi Mélange could not have been derived
from the exposed accretionary prism to the North Luzon Trough by sedimentary mass
slumping, because no such volcanic rocks are now exposed in the accretionary prism
west of the Coastal Range. Instead, they could have been derived from the Tuluan-
shan Formation when it was emplaced into the Lichi Mélange by thrusting during
the last 1 Ma when the Luzon arc-forearc was accreted to form the southern Coastal
Range. Thus the kaolinites of the volcanic arc rocks were redistributed into the Lichi
Mélange by fluid flows along the ubiquitous geological fractures in the mélange, con-
sistent with the field occurrences of the large, rootless, fault-bounded volcanic rocks
of andesitic breccia, tuff, and agglomerates that were floating in the intensely sheared
block-in-matrix mélange facies of the Lichi Mélange.
Mélange is commonly considered to develop in the accretionary prism of a sub-
duction zone. However, the Lichi Mélange in the SW Coastal Range originated from
the thrust forearc strata, representing a unique forearc mélange for orogenic belts
worldwide. The young age and wide distribution—especially the continuous offshore-
onshore connection—of the Lichi Mélange provides a unique example for further
research into active modern mélange-forming processes by forearc thrusting during
progressive closure of the forearc basin in this active region of arc-continent collision.

Keywords: Lichi Mélange, coastal range in eastern Taiwan, forearc collision mélange,
initial arc-continent collision, forearc closure.

INTRODUCTION 1992; Reed et al., 1992; Lundberg et al., 1997; Malavieille et al.,
2002). Stratigraphic and geochemical studies have shown that the
The Coastal Range in eastern Taiwan has resulted from the Coastal Range is composed of three accreted Miocene–Pliocene
accretion of the Luzon Arc-forearc to the exhumed metamorphic volcanic islands, three Plio-Pleistocene remnant forearc basins,
basement of eastern Taiwan (Fig. 1) during the arc-continent col- two intra-arc basins, and the Pliocene Lichi Mélange (Fig. 2).
lision in the last 2 m.y. (Chai, 1972; Biq, 1973; Huang et al., 2000, The accreted Miocene–Pliocene volcanic islands (from north to
2006). Now the location of initial arc-continent collision lies off south: Yuehmei, Chimei, and Chengkuangao) are composed of
southern Taiwan (Fig. 1; Huang and Yin, 1990; Huang et al., andesite, agglomerates, and tuff of the Tuluanshan Formation,
119 o E 120 o E 121 o E 122 o E 123 o E

China V
V V
V
o
25 o N 25 N
Okinawa

ge
an
Trough

R
Taiwan Taiwan

n
hil ls
Strait

sh a
r al
ent

ueh
24o N

F o ot
24o N
C

Hs

Ra n .
ge
V
L.
Ry
uk
n
yu
er

ge

stal
Ran III
est

Tr e n
c h
C oa
W

23o N 23o N
ch

TT

Arc
n
Tr e

Philippine
SLT
Fig. 10

HTR

Sea
HP

II
K a lope
ing

22o N 22o N
Luzon
p
S
o

Hengchun

South
Ridge
a

N L T
l
ani

China
M

21o N I 21o N
North

Sea

o o o o o
119 E 120 E 121 E 122 E 123 E

Figure 1. Shaded relief map showing regional tectonics inland and off the shore of Taiwan and lines of three E-W seismic profiles. The South
China Sea oceanic crust is subducting eastward along the Manila Trench, developing the North Luzon Arc. The accretionary prism of the Manila
subduction system extends from the submarine Hengchun Ridge northward to the Miocene deep-sea sediments in the Central Range–Hengchun
Peninsula. The ophiolite-bearing Kenting Mélange in the Hengchun peninsula represents the subduction complex within the accretionary prism.
Subsequently, the North Luzon Arc collided obliquely with the underthrusting Eurasian continent north of 21° N. The collision resulted in de-
formation of the western part of the North Luzon Trough as the Huatung Ridge backthrusted eastward, thereby closing the North Luzon Trough
from south to north. The Luzon Arc and forearc are accreting on eastern Taiwan, as the Coastal Range and the Huatung Ridge further extend
directly northward as the Lichi Mélange in the southernmost Coastal Range. I—intraoceanic subduction zone; II—initial arc-continent collision
zone; III—advanced arc-continent collision zone; LM—Lichi Mélange; KM—Kenting Mélange; NLT— North Luzon Trough; HTR—Huatung
Ridge; TT—Taitung Trough; SLT—Southern Longitudinal Trough; HP—Hengchun Peninsula; L.V.—Longitudinal Valley.
130 Huang et al.

Chen et al., 1990; Huang et al., 1992, 1995, 2000). In the Coastal
TECTONOSTRATIGRAPHIC Range the chaotic Lichi Mélange, lying between the remnant Loho
o
MAP OF 24 and Taiyuan forearc basins to the east and the collision suture of
THE COASTAL RANGE Hualien the Longitudinal Valley to the west (Fig. 2), has long been con-
sidered the key to understanding the arc-continent tectonic pro-
Longitudinal Valley v
cesses in Taiwan (Hsu, 1956; Biq, 1971, 1973; Wang, 1976; Page
v

E
Post-collision Sequence v and Suppe, 1981). The Lichi Mélange occurs in a zone ~2 km

NG
in the suture basin v
v
wide with characteristics of block-in-matrix fabrics, preferred
Forearc Basin Sequence foliation in scaly argillaceous matrix, and extensional web and
RA v
v
Shuilien Basin Yuehmeiv Shuilien boudinage structures in the sandstone blocks (Fig. 4; Chen, 1997;
v Chang et al., 2000). The variously sized blocks (millimeters to
L

Loho Basin v
RA

kilometers) of different lithologies in the Lichi Mélange include


v
Taiyuan Basin the andesite suite (andesite, volcanic breccias, tuffs, and volca-
NT

v
v
v
Collision Complex niclastic turbidites), the ophiolitic suite (serpentine, gabbro, and
CE

Lichi Melange
pillow basalt), and the sedimentary suite (sandstone, sandstone-
v
shale interbeds, shale, and limestone; Biq, 1971; Wang, 1976;
Luzon Volcanic Arc v
Hsu, 1976; Liou et al., 1977; Page and Suppe, 1981). The andesite
lt
au

v
v v Yuehmei
ei F

Volcanic island v v suite is derived from the Luzon Volcanic Arc, whereas the origin
im

+ v
of the dismembered ophiolite blocks may represent the oceanic
Ch

v
KLS
Chimei +
t

v v volcanic island + +
aul

+
+ +
+ crust of either the South China Sea (Suppe et al., 1981; Chung
y f

v Chengkuangao Chimei v
TLS + and Sun, 1992) or the Philippine Sea plate beneath the Luzon
alle

v v
v +
volcanic island v +
al V

v v v forearc-Arc (Juan et al., 1980; Malavieille et al., 2002). On the


Intra-Arc v v v
din

v
v v
v v other hand, the sedimentary blocks include the weakly lithified
gitu

Chingpu Basin v v
v
v Pliocene turbidite units (tens of meters to a kilometer in size) with
Lon

v v
Chengkung Basin v
similar lithology, age, and sedimentary turbidite structures as the
Loho v
v
remnant-coherent-forearc-basin strata of the Coastal Range, and
the angular, well-lithified, whitish quartz-rich, feldspathic sand-
lt

SEA
fau

v
v
stones of late Miocene age (meters to kilometers in size). The lat-
an

Antung v v
sh

v v
an

v
v
ter unit is similar to the deep-sea-fan sandstones in the upper part
lu

Fuli v
v v
Tu

v
Chengkuangao of the accretionary prism in the Hengchun Peninsula (southern
NE

v v
v
v
v
v
v
extension of the Central Range; Cheng et al., 1984; Huang et al.,
Chengkung 1997) and is not observed in the coherent-forearc-basin turbidites
PPI

v
v
v
v
v (Pliocene–Pleistocene age) of the Coastal Range.
Kuanshan v
Fig. 9B
I LI

v
Taiyuan The origin of the Lichi Mélange has been variously inter-
o o
23 Fig. 8 23 preted (Fig. 5). It was first considered to be a subduction complex
Tungho
PH

v
v v
v v
that developed in the former Manila Trench during subduction of
v
v v the South China Sea oceanic crust in the Pliocene–Pleistocene
v
v
(Fig. 5A; Biq, 1971, 1973) before the arc-continent collision.
ange

v
Then a sedimentary slumping olistostrome model was proposed
an ge

N
ange

(Fig. 5B; Wang, 1976; Ernst, 1977; Page and Suppe, 1981; Lin
t al lRR
C en tr al R
CCooaasst a .V.

Lichi Fig. 9A
L

0 10 km and Chen, 1986). The Lichi Mélange was regarded as a sedi-


mentary olistostrome deposited in the western part of the North
Lutao
Luzon Trough forearc basin by eastward mass wasting from
Taitung o
121 30' the exposed accretionary prism (the Central Range) during the
Pliocene–Pleistocene arc-continent collision, thus displaying a
Figure 2. Tectonostratigraphic map of the Coastal Range (modified
from Huang et al., 1995). Rectangles denote locations of Figures 8 facies change with the host non-sheared coherent-forearc-basin
and 9A, B. strata to the east (Fig. 5B). These two disparate models were
proposed on the basis of on-land field surveys and mineralogical
studies of the dismembered ophiolite blocks and large sandstone
whereas the remnant forearc basins (Shuilien, Loho, and Taiyuan) blocks in the Lichi Mélange without any constraint from regional
and the intra-arc basins (Pliocene Chingpu basin on Chimei vol- marine geological considerations. Intensive marine surveys off
canic island and Pleistocene Chengkung basin on Chengkuangao southern Taiwan in the early 1990s began to provide detailed
volcanic island) are filled with turbidites (Takangkou and Chimei bathymetric maps and regional tectonic structures. These investi-
Formations) derived from the accretionary prism. (Figs. 2 and 3; gations do not support the tectonic subduction origin of the Lichi
Biostratigraphy
LITHOSTRATIGRAPHY
EPOCH

TIME (Ma)
Longitudinal COASTAL RANGE
Valley

NORTH MIDDLE SOUTH


Suture Collision
Basin Melange Shuillien Chingpu Loho Taiyuan Chengkung
Forearc Intra-arc Forearc Forearc Intra-arc
Sequence Basin Sequence Sequence Basin
Pinanshan
PLEISTOCENE

0.7 Conglomerate

NN19 1.1 Chimei


N22 Chimei Formation
1.5 Formation
Takangkou
CN13 1.9
Formation
NN18 Chimei
NN17 2.3 Shuilien Takangkou
§Ê
Shuilien
Takangkou (Arc collapse)
Conglo- Conglomerate (Intra-arc)
N21 merate Takangkou
PLIOCENE

CN12 2.7
Formation Formation Formation Tungho Ls.
NN16
3.1
§¥§º
Formation
Takangkou
Tungho Ls.
Takangkou Fomation
Formation
Formation
3.5
NN15 CN11 Lichi Melange
N19/20

NN14 3.9
C (Arc collapse)
NN13 4.3
(Intra-arc)
CN10

4.7
(Chengkuangao Volcanic Island
B
NN12 5.1 Tuluanshan Formation
A
N18 ¥§•¶©Ls.
Kangkou
MIOCENE

5.5
NN11 ( Yuehmei ( Chimei
6 Volcanic
Volcanic
Island ) Island )
15 Tuluanshan Formation

Figure 3. Tectonostratigraphy of the Coastal Range.

Figure 4. Occurrence of angular, sheared,


whitish quartz-rich, feldspathic sand-
stone blocks in the intensely sheared
block-in-matrix mélange facies (δ-grade)
of the Lichi Mélange at sample location
LM-4 (location shown in Fig. 9A). The
large sandstone block with a black dot in
the center is ~8 m wide.
A Subduction Model (Plio-Pleistocene)
Figure 5. Models proposed for the origin of the
Subduction Complex North Luzon Trough Luzon Arc
Lichi Mélange. (A) Tectonic subduction com-
Lichi Melange forearc Basin Tuluanshan Fm.
plex model (Biq, 1971, 1973; after Chang et al.,
South China Sea
Takangkou & 2001). (B) Sedimentary slumping olistostrome
Chimei Fms.
~ V Huatung Basin model (Page and Suppe, 1981). (C) Tectonic col-
~ V V
~ ~~ ~
V
V
V V
V
Arc volcanics V lision complex model (modified from Chang et
~ V
al., 2001): (1) continuous turbidite sedimentation
in the whole North Luzon Trough forearc basin
(3.5–3.7 Ma); (2) arcward (eastward) backthrust-
Eurasian Plate Philippine Sea Plate ing, resulting in formation of the proto–Lichi
Mélange, similar to the present-day Huatung
Ridge in the western part of the forearc basin;
the backthrusting also resulted in emplacement
of the deep-sea-fan sandstone from the accre-
tionary prism in the Hengchun Peninsula to the
B Slumping Olistostrome Model (Pliocene) proto–Lichi Mélange (3–2 Ma); (3) westward
accretion and thrusting of the Luzon Arc-forearc
Uplifted accretion ary prism
to form the southern Coastal Range (<1 Ma) and
(Central Range) Coastal Range formation of the modern Lichi Mélange (for de-
North Luzon Trough
tails, see text). L—lower part of remnant forearc
NW Erosion Forearc Basin SE strata; U—upper part of forearc strata; 1—ac-
Sub
mar
cretionary prism; 2—Lichi Mélange; 3—Luzon
ine s
lide Arc; 4—Pinanshan Conglomerate in suture basin
flow ~ 2000 m
Mixed Terrain of of the Longitudinal Valley. Blocks in the Lichi
coh ern t
fore arc tur bid ites Mélange: a—deep-sea sandstone in accretionary
Continent -derived Turbiditic Rocks Lichi a t io n prism of Hengchun Peninsula; b—oceanic crust
ne

Melange Form
Ta ka ng ka ou
Zo

and Ophiolitic Rocks V V


Formation V of Luzon Arc-forearc; c—volcanic rocks of Lu-
ult

V V V
V
Tuluanshan
V + +
V V V
V
V
V
V
zon Arc.
Fa

V V V
+Chimei+
V V V V V V V V
+ + + + +
V V V V V Igneous complex V V

C Tectonic Collision Model (Plio-Pleistocene)


Legend
(3) <1 Ma
Longitudinal

Coastal Range (southern part)

(Taiyuan Basin)

Volcanic Island)
(Chengkuangao
W

Forearc Basin
E
Accretionary

Longitudinal Che
n

Luzon Arc
Volc gkuang
Valley

Valley Lichi Melange Taiyuan


anic a
Basin
Islan o
Prism

d
Lichi
Centr
U Melange
L
~ ~~
al Ra ~ ~ ~~ 4
~

~
~~

~~~~~
~ ~

nge
~
~

~ ~ ~~ ~ ~ ~
~

~ c U
North Luzon Arc 2 b L 3
~~

Taiyuan
Trough Ch 1 a
(2) 3 ~ 2 Ma (Remnant North Vo engk
Proto-Lichi Melange Luzon Trough) lca ua
Hengchun nic nga
Peninsula
Southern Isla o
e)
(Huatung Ridg
Remnant Forearc Basin
Longitudinal nd
Trough

Accre Suture L
U

tiona basin

ry Pri L

sm North
Luzon Arc

(1) 3.5 ~ 3.7 Ma on Trough


Hengchun
Hengchun North Luz
Peninsula
Peninsula

Acc Forearc Basin


retio L
nary canic bas
ement

Pr
st ~ Vol
anic cru
ism Oce
The Lichi Mélange 133

Mélange, because the Lichi Mélange lies to the east off the accre- the southwestern Coastal Range (Fig. 2), where the best sections
tionary wedge instead of within the accretionary prism (Huang of the Lichi Mélange are exposed, showing different grades of
and Yin, 1990; Reed et al., 1992; Huang et al., 1992, 2000; Liu strata fragmentation, mixing, and shearing. The exposures here
et al., 1998; Malavieille et al., 2002). The marine surveys show led to various interpretations of the origin of the Lichi Mélange
a direct connection between the submarine Huatung Ridge and in previous studies.
the Lichi Mélange in the southernmost Coastal Range (Fig. 6;
Huang et al., 1992; Huang, 1993). This led to the proposal of a STUDY METHODS
tectonic collision model, which suggests that the Lichi Mélange
originated from the thrusting and shearing of the forearc strata, Marine Surveys
similar to the modern Huatung Ridge in the western part of the
North Luzon Trough during initial arc-continent collision in the A new multichannel seismic survey along a 21° N transect
early Pliocene (Huang et al., 2000; Chang et al., 2000, 2001). across the North Luzon Trough off southern Taiwan (Figs. 1 and
The purpose of this study is to integrate independent lines 7) was conducted in 1999, using the R/V Tanbao. The capabili-
of new evidence, including a new marine 240-fold multichannel ties of this vessel included 240 channels, an I/O digital streamer,
seismic profile across the North Luzon Trough as well as on-land 12.5 m of the interval between channels, a 50 m shot distance, a
field observations and studies of the planktic and benthic fora- 250 m offset, 2 ms of sample, and 10 s of recording. The energy
miniferal fauna, clay mineral composition, and fission track pat- source was a TI sleeve gun–airgun array. The total size was 3000
terns of zircon grains in the Lichi Mélange to compare with the C.I., and the pressure was 2000 psi. The results of the survey were
Taiyuan remnant-forearc-basin strata to the east to constrain the then integrated with two 6-channel seismic profiles (MW9006–
origin of the Lichi Mélange. Most study samples were collected 26 and MW9006–31, Fig. 1; Chang et al., 2001) previously con-
from the Mukeng River and its tributary south of Kuanshan in ducted by the R/V Moana Wave in 1990 to show development of

Figure 6. Geological continuation from


the Lichi Mélange in the southernmost
Coastal Range to the submarine Hua-
tung Ridge (deformed forearc unit by
eastward thrusting) in the initial arc-
continent collision zone (marine image
adapted from Malavieille et al., 2002).
PC—Pinanshan Conglomerate in Lon-
gitudinal Valley.
MW 9006-31 10km
Luzon
Southern
Two-way travel time (sec)

0.0 W Huatung Volcanic E 0.0


Longitudinal Ridge Arc
1.0 Trough 1.0
Taitung
2.0 Trough 2.0
Lutao
3.0 Volcanic 3.0
4.0 Island 4.0

5.0 5.0

MW 9006-26 Southern
0.0
W Longitudinal Huatung Luzon E
Trough Ridge Volcanic
1.0 Arc 1.0
Two-way travel time (sec)

2.0
North 2.0
Luzon
3.0 3.0
Trough Lanhsu
4.0 4.0
Volcanic
5.0 5.0
Island
6.0 6.0
7.0 7.0
8.0 8.0
GMGS 973
Hengchun Ridge
Accretionary Prism
W 8921 8961 9001 9041 9081 9121 9161 9201 9241 9281 9321 9361 9401 9441 9481 9521 9561 9601 9641 9681 9721 E

Luzon Volcanic Arc


Two-way travel time (sec)

3.0 Proto-Huatung 3.0


Ridge North Luzon Trough
4.0 6 Forearc Basin 4.0
5
4
5.0 3 5.0
2
1
6.0 6.0
Batan
Volcanic
7.0 Island 7.0

8.0 8.0

Figure 7. Three seismic profiles, from north to south (see Fig. 1 for location), showing progressive formation of the Huatung Ridge and closure of
the North Luzon Trough forearc basin. Line GMGS-973 further revealed syndeformational sedimentation in the North Luzon Trough, regardless
of deformation of the proto–Huatung Ridge in the western part of the trough filled with six sequences of strata separated by five unconformities
resulting from episodic arc-continent collisions.
The Lichi Mélange 135

the Huatung Ridge along progressive closure of the North Luzon Mukeng River sections (Fig. 8) and supplementary localities in
Trough by arc-continent collision. the southernmost Coastal Range (Fig. 9). Both planktic and ben-
thic foraminifers in 200 g of scaly argillaceous matrix (γ- to δ-
On-Land Study grade shearing) or the Bouma Te part of weakly sheared turbidite
layers (β-grade) were identified. Distributions of the main ben-
Raymond’s (1984) classification of different grades of strata thic foraminifers were semiquantitatively plotted against stratig-
fragmentation, mixing, and shearing—from coherent facies (α- raphy according to their relative abundance. The results were then
grade), weakly sheared broken formation facies (β-grade) to compared with the benthic foraminifers in the coherent-forearc-
strongly sheared dismembered facies (γ-grade), and intensely basin turbidites in the Taiyuan Basin (Chang, 1967). Multivari-
sheared mélange facies (δ-grade)—was followed for mapping ate cluster analysis was used to compare the similarity of benthic
the distribution and occurrence of the Lichi Mélange along the foraminifers within the Lichi Mélange. Benthic foraminifers that

Sector 1 Sector 2 Sector 3


(Fig. 12 A, B) (Fig. 13 A, B) (Fig. 14 A, B, C)
1 N

N
500 m 19
18 16
D

Fault
17 15
A B C 14
2 N
I IIa IIb
50
60
r
ve
4
85
Ri 5
04
65
Mushi Taiyuan
13
River

65 85
2 Bridge
11 Forearc
06 10
03 n g 12 Basin
40 u ke 07 09 IV
3 N 1 01 M 08 20
3 Nanshi
05 22 Bridge
21 65 24 H
Peinan

02

nge
23
I 25

Ra
85

27 26

al
II III
han
ntr
4 N
A Sector 3

nge
Ce
28
(Fig. 14 D, E)
s

Ra
C
Tu l u a n

IV
al
31 K
29 ast
30
32 33
35 D
Co

50
5 N 55 L
34 39 Taitung
36
60
37 40 Raymond (1984)
65 45

38
1 ~5 Stress study sites of Chang et al. (2001)

Sector 1 Sector 2
(Fig. 12 C, D) (Fig. 13 C,D)
- grade - grade - grade Sandstone Block Conglomerates Igneous Block

Figure 8. Detailed geological map and sampling locations of the study area along the Mukeng River and its southern tributary sections. Ray-
mond’s (1984) classification system in the right bottom corner is followed to map the distribution of various sheared units of the Lichi Mélange
(modified from Chang et al., 2000). The section is divided into three sectors according to their exposure sequences from west to east. Field oc-
currence and distributions of major benthic foraminifers are shown in Figures 12–14. K—Kuanshan; L—Lichi; H—Hualien.
o o
121 15'E 121 17'30''E

lt
N Fuli V
BV

f au
g
V V o

fon
23 10'N

e
ng
0 1 km 1.0

Frequency
V V

Yu
40.5 Ma TF-6

Ra
12.5 Ma
V V 0.5 64.5 Ma
LM-10

al
2.5 Ma
TF-9 V V

n tr
0.0
TF-8 0 50 100 150 200 250 300
Ce 23 V V 1.0
3.0 Ma

Frequency
ey
TF-7
V
ll 70 V V
Va
0.5 15.6 Ma
V

lt
100 Ma
V V

fa u
al

V 0.0
d in

V V 0 50 100 150 200 250 300


i tu

2.72 Ma
V

an
76 1.0

Frequency
V V 5.8 Ma
ng

n sh V TF-9
Lo

l ua
0.5
V TF-7 92 Ma
Tu

g V V V
0.0
0 50 100 150 200 250 300
1.0 2.08 Ma

Frequency
V V 23 V V
TF-8
0.5
V
V V 0.0
g V V 0 50 100 150 200 250 300

L egend V V Age (Ma)


Gravel and sand TF-6
(Holocene) Takangkou Formation
Lichi Melange V V
g
With ophiolite 62 Taiyuan 1 .0 88 Ma

Frequency
Tungho Limestone o 18 Ma
(Late Pliocene) 23 N
Takangkou Formation
0 .5
LM-10
(Pliocene-Pleistocene)
V Tuluanshan Formation
(Miocene-Pliocene)
Slate and schist V V 0 .0
(Eocene-Miocene) Tungho 0 50 100 150 200 250 300
o o
121 15'E 121 17'30''E Age (Ma)
Lichi Melange

N A 1 .0 120 Ma
F requency

60 Ma
30 58 LM-1
42 33 65
0 .5 12.8 Ma
40 50
40 65
0 .0
0 100 200 300 400 500
58 1 .0
Frequency

0 1 2 Km 9.6 Ma
120 Ma LM-2
PEI NAN RIVER

0 .5
12
48 20
72 0 .0
42 LM-5 0 100 200 300 400 500
60 1 .0
F requency

50 9.76 Ma LM-3
Lichi 48 0 .5
110 Ma
LM-2 0 .0
0 50 100 150 200 250 300
50
1 .0
Frequency

22 66 Ma
LM-3 LM-4
LM-1 g g
0 .5 14.8 Ma

0 .0
LM-4 0 50 100 150 200 250 300
g Legend 1 .0
11 Ma
Frequency

Pin an sh an Con glome r st e


LM-5
L ic h i Me lan ge 130 Ma
0 .5
S an d st on e
Blocks in

L ime st on e
melange

Agglome r at ic an d e sit e
B asalt 0 .0
G ab b ro 0 50 100 150 200 250 300
S e r p e n t in e , Pe r id ot it e
Tak an gk ou For mat ion Age (Ma)
Lichi Melange

Figure 9. Fission track records of zircon grains separated from whitish quartz-rich feldspathic sandstone blocks in
the Lichi Mélange (A) and from the coherent forearc turbidite layers of the Taiyuan remnant forearc basin (B).
The Lichi Mélange 137

occur in <5 samples, and the samples that contain <50 specimens Taiyuan remnant-forearc-basin turbidites were then compared
of benthic foraminifers in 200 g of sediments, were deleted dur- with the samples collected from the nonmetamorphosed deep-sea
ing computation of the cluster analysis. sandstones in the accretionary prism of the Hengchun Peninsula
Mineral compositions of the clay fraction (<2µm) of the col- (Fig. 10) to constrain the origin of the whitish sandstone blocks
lected samples were also identified by X-ray powder diffraction in the Lichi Mélange.
following the method described by Lin and Chen (1986). The
results were then compared with those of Lin and Chen (1986), RESULTS
who had systematically documented the clay mineral composi-
tions in the Lichi Mélange and the coherent-forearc-basin strata Marine Surveys off the Southern Coastal Range
in the southern Coastal Range.
In addition, fission tracks of the zircon grains separated The new 240-fold multichannel seismic profile (GMGS-973
from the Bouma Ta part of the Taiyuan remnant-forearc-basin in Fig. 1) across the North Luzon Trough (forearc basin) along
turbidites were counted to compare with the results obtained the 21° N transect is shown in Figure 7. The profile shows syn-
from the whitish quartz-rich feldspathic sandstone blocks in deformational sedimentation in the North Luzon Trough. The
the sheared Lichi Mélange near Fuli and the type locality of the forearc basin is filled by six sequences of strata separated by five
Lichi Mélange along the Peinan River (Fig. 9). The detection and unconformities. All of these six sequences in the western part of
counting methods of the fission track age analysis described by the North Luzon Trough are tilted upward and deformed gently
Liu (1982) and Liu et al. (2001) were followed. The results of the by eastward thrust, whereas the strata in the eastern part of the
zircon fission track records from both the Lichi Mélange and the trough remain flat, suggesting no deformation at all (Fig. 7).

O
22 10'N 1.0
50-65 Ma

Frequency
HP-3
HP-4 5 km 0.5
HP-4
Shallow - marine slope basin

LEGEND
Mt
Holocene

0.0
Recent 0 50 100 150 200 250 300
Alluvium 1.0
HP-3
HP-2 15-30 Ma
Taiping Fm. Frequency 85-100 Ma
Pleistocene

Szekou Fm.
Ls
Hengchun 0.5
Limestone HP-3
~~ unconformity ~~
Pliocene

5 Kt Ma Maanshan Fm.
0.0
~~ unconformity ~~ 0 50 100 150 200 250 300
15-25 Ma
1.0
Slope and trench-fill turbidites

55-65 Ma
Frequency

Hengchun
HP-2 Kenting Melange
in accretionary prism

Kt
Middle - Late Miocene

O
22 00'N (subduction complex)
0.5 HP-2
Mt Mutan Fm.
Ma HP-1
Mt lower fan turbidites
7
upper- middle fan
conglomerates and
sandstones 0.0
0 50 100 150 200 250
Loshui Fm.
Ls 1.0
middle fan sandstones 15-30 Ma
Frequency

Hengchun Peninsula 70-85 Ma


0.5
HP-1

O O
120 40'E 120 50'E
0.0
0 50 100 150 200 250
Age (Ma)
Figure 10. Fission track records of zircon grains separated from deep-sea-fan sandstones in the accretionary prism of Hengchun Peninsula. The
fission-track age pattern of these deep-sea sandstones is similar to that obtained from the whitish quartz-rich feldspathic sandstone blocks in the
Lichi Mélange (Fig. 9A).
138 Huang et al.

Geology of the Lichi Mélange Exposed along the Figure 8 shows a detailed geological map of the Lichi
Mukeng River Mélange exposed along the Mukeng River and its southern trib-
utary in the southwestern Coastal Range (Chang et al., 2001).
The Lichi Mélange is exposed in the southwestern and south- The coherent turbidite sequence (Takangkou Formation) of the
ernmost parts of the Coastal Range (Figs. 1 and 2). The Lichi Taiyuan remnant forearc basin is thrust westward over the Lichi
Mélange lies between the Longitudinal Valley to the west and the Mélange along the Tuluanshan Fault (Fig. 8). West of this fault
remnant forearc basins (Loho Basin and Taiyuan Basin) and vol- the study section shows various degrees (β-, γ-, and δ-grades of
canic rocks to the east (Fig. 2). In the northern Coastal Range the Raymond, 1984) of strata fragmentation, mixing, and shearing.
Lichi Mélange could have been eroded away by the river flowing The sequences with γ-grade and δ-grade shearing (dismembered
northward along the eastern side of the Longitudinal Valley. The and mélange facies) exhibit characteristics of block-in-matrix
Longitudinal Valley, a 2–3-km-wide collision suture between the features (Fig. 4), with pervasively scaly foliation dipping to the
Central Range (accretionary prism and exhumed metamorphic southeast (measurements of foliations at five localities are shown
basement) and the Coastal Range (accreted arc-forearc; Fig. 2) in Figure 8; Chang et al., 2000) without discernible bedding (units
accommodates a part of the shortening across southeastern Tai- I, II, III, and IV in Fig. 8), whereas the weakly disturbed (β-grade
wan (Yu et al., 1997). Between Antung and Kuanshan the volca- shearing or broken formation facies) sequences (units A, B, C,
nic basement beneath the Taiyuan forearc strata has been thrust and D in Fig. 8) still preserve distinct Bouma turbidite sequences
westward over the Lichi Mélange along the Tuluanshan Fault and clear bedding, which generally strike NE and dip SE. How-
(Fig. 2). Consequently, the Tuluanshan volcanics are exposed ever, faults for various thickness are commonly observed in the
between the Lichi Mélange to the west and the remnant Taiyuan intervals of different grades (centimeters thick in β-grade, several
forearc basin to the southeast (Fig. 2). In the Coastal Range this meters thick in γ-grade, and hundred meters thick in δ-grade) of
is the only place where the volcanic rocks appear west of the shearing units and also along the boundaries between the dif-
remnant forearc basins, and they lie between the Lichi Mélange ferent units (Fig. 11; Chang et al., 2001). From west to east the
and the Taiyuan remnant forearc basin. Otherwise, the volca- study profiles are divided into three sectors (Fig. 8) for convenient
nic andesitic rocks are always observed to the east of remnant- comparisons and discussion. Sector 1 includes a γ-grade unit and
forearc-basin strata (Fig. 2; Hsu, 1956, 1976; Huang et al., 1995). a β-grade unit (units I and A, respectively), whereas sector 2 is
The east-dipping Tuluanshan Fault runs in a NNE-SSW direc- composed of two γ-grade units (units IIa and IIb) and two β-grade
tion, representing the last accretion of the Luzon Arc-forearc to units (units B and C) in the north (Fig. 8). Units IIa and IIb are
the exhumed Eurasian continent (eastern Central Range) as the combined together as the thick δ-γ-grade unit (unit II) in the south.
southern part of the Coastal Range in the last 1 m.y. (Chang et al., Sector 3 displays a γ-grade unit III, a thick and less disturbed (β-
2000, 2001; Huang et al., 2000, 2006). grade) unit D, and a thick (~500 m along the river), highly sheared

Figure 11. Ubiquitous shearing planes


with fault gouges in the Lichi Mélange,
especially in mélange facies unit IV (δ-
grade of Raymond, 1984; Fig. 8).
The Lichi Mélange 139

(γ- and δ-grades) unit IV directly west of the Tuluanshan Fault DISCUSSION
(Fig. 8). Angular ophiolite blocks, predominantly of gabbro and
whitish sandstone blocks, are found restrictedly in units I, II, III, Significance of Marine Geology off the Southern Coastal
and IV with dismembered facies or mélange facies. In contrast, Range in the Active Arc-Continent Collision Region
pebbly mudstone layers composed of subangular to subrounded
metasandstone conglomerates are found only in the less disturbed, Marine seismic investigations off southeastern Taiwan reveal
broken formation facies of units A, C, and D (Fig. 8). the progressive closure of the forearc basin owing to the east-
ward (arcward) thrusting of the forearc basin strata to develop
Planktic Foraminifers and Age Determination of the the Huatung Ridge (Fig. 7; Reed et al., 1992). Profile GMGS-
Lichi Mélange 973 along the 21° N transect across the North Luzon Trough
(Fig. 7), where the Luzon arc collides incipiently against the Eur-
Planktic foraminifers, including Sphaeroidinella dehiscens, asian continent, further shows that (1) active syndeformational
Globorotalia tumida, G. multicamerata, Neogloboquadrina sedimentation occurs in the forearc basin; once the sediments are
altispira, and Sphaeroidnellopsis seminulina, are commonly deposited in the trough, the sequence in the western part of the
found in most of the samples from the Lichi Mélange along the forearc basin is deformed and then covered unconformably by
Mukeng River and its tributaries. They show a persistent early the overlying sequence; (2) folding and thrusting in the western
Pliocene age (upper Zone N19-20), consistent with results part of the forearc basin have occurred since the early history
obtained previously by calcareous nannoplankton (Zone NN15; of forearc sedimentation; (3) sedimentation is continuous in the
Chi et al., 1981; Chi, 1982; Barrier and Müller, 1984). eastern part of the forearc basin regardless of active deforma-
tion in the west; and (4) the eastward thrusting accommodates
Benthic Foraminifers in Various Degrees of Sheared Units the space between the backstop of the accretionary prism and
of the Lichi Mélange the Luzon Arc. In compiling two 6-channel seismic profiles in
the north (Fig. 7; MW9006–26 and MW9006–31; Chang et al.,
Distributions of benthic foraminifers in the various degrees of 2001), regional marine geology shows that the deformed Hua-
sheared strata in the Lichi Mélange along the Mukeng River and tung Ridge forearc becomes larger and larger so that finally the
its tributary are shown in Figures 12–14. As a result of turbidite remnant North Luzon Trough is closed at its northern terminus,
deposition, the benthic foraminifers in the Lichi Mélange are the approaching the southern Coastal Range (Fig. 6). Meanwhile, the
mixed fauna of the indigenous deep-water taxa and the displaced deformed ridge crest is thrust upward so that sediments can no
shallow-marine species. The deep-marine fauna include the cos- longer be deposited on the Huatung Ridge.
mopolitan species of the genera Bulimina, Uvigerina, Fontbotia, Because the Huatung Ridge connects northward with the
Pullenia, Hyalinea, Melonis, Gyroidinoides, and Cyclammina, Lichi Mélange in the southernmost Coastal Range (Fig. 6), it is
living on the continental slope to abyssal plain of the present conceivable to consider that the Lichi Mélange has originated
oceans. The shallow-marine fauna include species of the genera from the deformed forearc basin strata like the modern Huatung
Amphistegina, Ammonia, Operculina, Asterorotalia, Pseudorota- Ridge and to predict that the Lichi Mélange could have the fol-
lia, Calcarina, Elphidium, Lenticulina, and Cellanthus, living in lowing geological characteristics:
warm waters of the shallow shelf (<200 m). In each study sample, A. The Lichi Mélange should lie west of the remnant forearc
individuals of these shallow- and deep-marine genera make up basin;
60%–80% of the total benthic foraminiferal specimens (Chien, B. The Lichi Mélange would contact the coherent forearc
2003). For convenient discussion, the species with similar ranges basin strata by faulting;
of water depth or ecology are shown together in Figures 12–14. C. The depositional time interval of the Lichi Mélange would
be less than that of the entire forearc basin sequences, and
Clay Mineral Composition the depositional age of the Lichi Mélange can be equiva-
lent only with the lower sequence, but older than the upper
Clay mineral compositions of the sheared mudstone in the sequence, of the remnant-coherent-forearc-basin strata;
broken formation facies or scaly argillaceous matrix and the D. Regardless of shearing intensity and facies, the microfos-
quartz-rich sandstone blocks in the mélange facies of the Lichi sils within the Lichi Mélange should be similar to and
Mélange along the Mukeng River and its southern tributary are compatible with the fauna in the remnant forearc basin to
listed in the Appendix. The samples are all characterized by bear- the east;
ing illite (average, 51.5%), chlorite (average, 19.5%), mixed- E. The depositional bathymetry of the Lichi Mélange would
layer clay minerals (average, 11.7%), and kaolinite (average, be similar to the present North Luzon Trough (1500–
11.1%). Smectite is minor (<3%). Kaolinite is always found in 3000 m);
the samples of the Lichi Mélange regardless of the mudstones in F. The relict sedimentary structures and the host sediment
the broken formation facies, the scaly argillaceous matrix, or the composition of the Lichi Mélange would be similar to
angular whitish sandstone blocks in the mélange facies. those of the coherent-forearc-basin turbidites;
140 Huang et al.

A
SHALLOW MARINE TAXA DEEP-SEA TAXA
Sheared

B
grades
Age

Sector Ammonia spp.


Amphistegina spp. Asterorotalia spp. Calcarina calcar Fontbotia wuellerstorfi
Lenticulina spp. Hyalinea balthica Pullenia bulloides Uvigerina spp. Gyroidinoides spp. Cyclammina spp.
1 Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Elphidium spp. Cibicidoides spp.
Melonis spp.

Unit Sample
4 00 m

Number
N19 /20 ; NN15 (3.7 ~ 3.5 Ma)

04-0
300 m

04-1
A
04-2

04-3
200 m

04-4

03
I
100 m

02

01
0m

Figure 12. (A) Examples of field occurrence and sampling localities of units I and A, respectively, in part of sector 1. (B) Distribution of major
benthic foraminifers along the Mukeng River section.

G. The early eastward thrusting in the forearc basin in the example, sedimentary slumping or mud diapirism—responsible
initial arc-continent collision stage and the final west- for formation of the Lichi Mélange.
ward thrusting of the Luzon Arc-forearc during the
advanced arc-continent collision accreted to the Coastal Significance of Field Surveys and Age Determination
Range could have resulted in accretions of the accre-
tionary wedge materials west of the forearc basin and Field surveys along the Mukeng River and its tributary show
the volcanic arc rocks beneath the forearc basin to the that the Lichi Mélange lies west of the Taiyuan remnant forearc
highly sheared, block-in-matrix, dismembered facies basin and is thrust over by the coherent-forearc-basin strata along
and mélange facies of the Lichi Mélange; however, these the Tuluanshan Fault (Fig. 8). Neither the Lichi Mélange nor the
blocks would not appear in the weakly sheared broken volcanic rocks are observed in the Taiyuan remnant forearc basin
formation facies of the Lichi Mélange or the remnant- east of the Tuluanshan Fault, consistent with predictions A and B
coherent-forearc-basin strata. in the preceding subsection.
If our different lines of new data agree well with these geo- Biostratigraphic study reveals that the protolith of the Lichi
logical characteristics or predictions listed above, we would Mélange was deposited in a narrow age range within Zone N19-20
prefer that the Lichi Mélange originated along the arcward of planktic foraminifers or Zone NN15 (3.5–3.7 Ma) of calcare-
thrusts during forearc basin closure by active arc-continent col- ous nannoplankton (Figs. 12–14; Chi et al., 1981; Chi, 1982; Bar-
lision. On the other hand, if the new data do not fit well with the rier and Müller, 1984). This age is coeval with the lower remnant
predictions, we would consider the other possible origins—for forearc basin sequence (Fig. 3) but is older than the upper forearc
The Lichi Mélange 141

29

C
SHALLOW MARINE TAXA DEEP-SEA TAXA
D
Sheared
grades

Sector
Age

Ammonia spp.
Amphistegina spp. Asterorotalia spp. Calcarina calcar Fontbotia wuellerstorfi
1 Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Elphidium spp.
Lenticulina spp. Hyalinea balthica
Cibicidoides spp.
Pullenia bulloides
Melonis spp.
Uvigerina spp.Gyroidinoides spp. Cyclammina spp.

Unit Sample
300 m

Number
N19; / 20 NN15 (3.7 ~ 3.5 Ma)

29

A
20 0 m

28
1 00 m

I
27
0m

Figure 12. (C) Occurrence and imbricate structure of the fractured, subangular to subrounded metasandstone conglomerates in two pebbly mud-
stone layers of broken formation facies in unit A. (D) Distribution of major benthic foraminifers in part of sector 1 along the southern tributary
section (detailed position shown in Fig. 8).

sequence (Zones N21–22 of planktic foraminifers, Chang, 1967; Significance of the Benthic Foraminiferal Study
Zones NN16–19 of calcareous nannoplankton, <3.5–1.15 Ma,
Horng and Shea, 1996). This indicates that (1) the turbidites in Figure 15 shows the result of cluster analysis of the benthic
the proto–Lichi Mélange were deposited in a short time (~0.2 foraminifers in the Lichi Mélange. No samples of the mélange
m.y.), (2) deformation (at ca. 3.5 Ma) of the proto–Lichi Mélange facies (δ-grade of shearing) appear in Figure 15 because of their
occurred as soon as these turbidites were deposited (3.5–3.7 Ma), strong depression or their broken or incomplete preservation
and (3) after deformation of the proto–Lichi Mélange the young from intensive compression or shearing, which prevents precise
turbidites (3.5–1 Ma) were deposited continuously in the upper taxonomic determinations from the few fossil remains. There-
part of the Taiyuan remnant forearc basin, a scenario analogue fore, the samples collected from mélange facies are omitted from
with the modern configuration of the Huatung Ridge–remnant computation.
North Luzon Trough off the southern Coastal Range in the mod- The samples are clustered into two main groups. Up to 79%
ern active arc-continent collision zone (Fig. 7). The results of bio- of the samples are clustered together as group A, whereas 21% of
stratigraphic study are consistent with prediction C in the preced- the samples (only 8 samples) are clustered as group B. Samples
ing subsection. of group B were collected primarily from the weakly sheared,
142 Huang et al.

A
A
B SHALLOW MARINE TAXA DEEP-SEA TAXA
Sheared

Sector
grades
Age

Ammonia spp.
Amphistegina spp. Asterorotalia spp. Calcarina calcar Fontbotia wuellerstorfi
Pullenia bulloides
2 Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Elphidium spp.
Lenticulina spp. Hyalinea balthica
Cibicidoides spp.
Melonis spp.
Uvigerina spp. Gyroidinoides spp. Cyclammina spp.

Unit Sample
Number
11
500 m

10-1
N19 / 20 ; NN15 (3.7 ~ 3.5 Ma)

10-2
C
400 m

10-3
10-4
300 m

09-1

09-2
20 0 m

09-3
IIb
08-1
08-2
07-1
B 07-2
1 00 m

06-2
06-1
05
IIa
0m

Figure 13. (A) Examples of field occurrence and sampling localities of unit C. (B) Distribution of major benthic foraminifers along the Mukeng
River section in the north.

broken formation facies (β-grade) turbidites of unit C (6 sam- The result implies that the benthic foraminifers in the Lichi
ples) and two individual samples from the dismembered facies Mélange are similar to each other regardless of different degrees
(γ-grade) of unit IIb (sample 6–2) in sector 2 and unit III (sample of shearing intensity, consistent with prediction D in the first sub-
38–2) of sector 3 (Fig. 8). Foraminifers in group B predominate section under “Discussion.”
in the indigenous deep-water species without significant shal- Foraminifers in the Taiyuan remnant forearc sequence in
low-marine assemblages (Fig. 13B), whereas samples of group the southern Coastal Range were studied by Chang (1967).
A contain both the indigenous deep-water and the displaced Unfortunately, the abundance of each benthic foraminifer was
shallow-water fauna (Figs. 12–14). Moreover, each group A and reported semiquantitatively (R, <6; F, 6–10; C, >11–20; A, >21–
group B contains the samples collected from the units of the 50). A comparison of the present work with the result of Chang
dismembered facies and the broken formation facies (Fig. 15). (1967) can be made by cross-checking the faunal lists and
The Lichi Mélange 143

D SHALLOW MARINE TAXA DEEP-SEA TAXA


Sheared

Age
grades

Sector Ammonia spp.


Amphistegina spp. Asterorotalia spp. Calcarina calcar Fontbotia wuellerstorfi
Lenticulina spp. Hyalinea balthica Pullenia bulloides Uvigerina spp. Gyroidinoides spp. Cyclammina spp.
2 Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Elphidium spp. Cibicidoides spp.
Melonis spp.

Sample
570 m

Unit Number
37-1
37-2
37-3
500 m

37-4

37-5
N19 / 20; NN15 (3.7 ~ 3.5 Ma)

C 36
400 m

35
33-1
33-2
33-3
33-4
30 0 m

34
200 m

32

II
100 m

31

30
0m

Figure 13. (C) Examples of field occurrence and sampling localities of unit C. (D) Distribution of major benthic foraminifers in part of sector 2
along the southern tributary section (detailed position shown in Fig. 8).

relative abundance of the major taxa in both studies (Table 1). used. Our study uses a newer classification of benthic
The results suggest that the benthic foraminifers in the Lichi foraminifers by Loeblich and Tappan (1988), whereas
Mélange (this study) and the Taiyuan remnant forearc basin tur- Chang (1967) followed an older classification of Loe-
bidites (Chang, 1967) are similar: blich and Tappan (1964).
1. The benthic foraminifers are all characterized by high 2. Planktic foraminifers are abundant (>85% of total fora-
diversity (Taiyuan forearc basin turbidites: 40 genera, miniferal individuals), and benthic foraminifers are pre-
68 species, in Chang, 1967; Lichi Mélange: 86 genera dominantly calcareous genera (>75%) in both Taiyuan
and 167 species, this study) and low abundance of each forearc turbidites and the Lichi Mélange.
species (85% species < 6 individuals/200 g of sediments 3. Foraminifers in both Taiyuan forearc turbidites and the
in the Lichi Mélange and coherent forearc basin turbi- Lichi Mélange are composed of indigenous deep-marine
dites). The difference in species number between these fauna and shallow-marine species (Table 1). The deeper
two studies is due to the different classification system marine taxa include Bulimina striata d’Orbigny, B. cf.
A

C SHALLOW MARINE TAXA DEEP-SEA TAXA


Sheared
grades

Sector
Age

Ammonia spp.
Amphistegina spp. Asterorotalia spp. Calcarina calcar Fontbotia wuellerstorfi
Pullenia bulloides
Lenticulina spp. Hyalinea balthica Uvigerina spp.Gyroidinoides spp. Cyclammina spp.
3 Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Elphidium spp. Cibicidoides spp.
Melonis spp.

Unit Sample
Number
4 80 m

19
40 0 m

15-2
N19 / 20 ; NN15 (3.7 ~ 3.5 Ma)

D 15-1

16-1
300 m

16-2
16-3
16-4
16-5
18
17
2 00 m

14
14-1
III 13-1
13-2
12-1
1 00 m

12-2
12-3
12-4
12-5
12-6
0m

Figure 14. (A, B) Examples of field occurrence and sampling localities of units III and D, respectively. (C) Distribution of major benthic fora-
minifers in part of sector 3 along the Mukeng River section in the north.
26
D

E Sector
SHALLOW MARINE TAXA DEEP-SEA TAXA
Sheared
grades
Age

Ammonia spp.
3 Amphistegina spp. Asterorotalia spp. Calcarina calcar
Operculina spp. Pseudorotalia spp. Cellanthus craticulatus
Fontbotia wuellerstorfi
Lenticulina spp. Hyalinea balthica Pullenia bulloides
Melonis spp.
Uvigerina spp.Gyroidinoides spp. Cyclammina spp.
Elphidium spp. Cibicidoides spp.

Unit Sample
7 00 m

Number
600 m

IV 26
500 m

N19 /20 ; NN15 (3.7 ~ 3.5 Ma)

25
400 m

24
30 0 m

D
200 m

23
100 m
0m

22

Figure 14. (D) Examples of field occurrence and sampling locality 26 in mélange facies of unit IV. (E) Distribution of major benthic foramini-
fers in part of sector 3 along the southern tributary of the Mukeng River (detailed position shown in Fig. 8).
146 Huang et al.

Rescaled Distance Cluster Combine


0 5 10 15 20 25
+ + + + + +
8-1 Mudstone
8-2 Shale
9-1 Mudstone Group B
10-4 Mudstone
6-2 Shale
10-3 Mudstone
38-2 Mudstone
10-1 Mudstone
12-2 Mudstone
15-1 Sandstone
11 Mudstone
13-2 Sandstone
4-0 Siltstone
4-1 Sandstone Figure 15. Multivariate cluster analysis
4-2 Sandstone of the samples collected from the Lichi
37-4 Shale Mélange, based on benthic foraminiferal
37-5
Group A assemblage. The study samples are clus-
Mudstone
5 Mudstone tered into two main groups. Each group
16-1 Mudstone includes samples with various degrees of
16-2 Mudstone shearing. For details of both groups, see
15-1 Shale the text.
18 Mudstone
37-3 Shale
2 Mudstone
4-4 Mudstone
28 Mudstone
27 Mudstone
16-4 Mudstone
38-5 Mudstone
31 Mudstone
16-5 Mudstone
17 Mudstone
19 Mudstone
25 Mudstone
36 Mudstone
23 Mudstone
24 Mudstone
38-1 Mudstone
14 Mudstone

rostrata Brady, Uvigerina probocidea Schwager, U. his- foraminifers are typical of the Kuroshio Current fauna
pida Schwager, U. subperegrina Cushman and Klein- living in the outer shelf facing the open ocean (Wang,
pell, Fontbotia wuellerstorfi (Schwager), Gyroidinoides 1985; Huang, 1989).
orbicularis (d’Orbigny), Pullenia bulloides (d’Orbigny), 4. The indigenous deep-marine benthic foraminifers in the
Melonis nicobarensis (Cushman), and M. pompilioides Lichi Mélange and the Taiyuan remnant forearc basin tur-
(Fichtel and Moll). These deep-marine benthic foramin- bidites are also the major fauna recovered in the present
ifers are cosmopolitan species living in water depths of hemipelagic muds of the North Luzon Trough (Huang,
1500–3000 m of modern deep oceans (Van Morkhoven 1993), whereas the shallow-marine species are commonly
et al., 1986). The shallow-marine species are Amphi- found ubiquitously in the turbidite layers off southern and
stegina radiata (Fichtel and Moll); A. lessonii d’Orbigny; northeastern Taiwan, including the North Luzon Trough
Operculina complanata (Defrance); O. ammonoides forearc basin and the Okinawa Trough backarc basin.
(Gronovius); Asterorotalia gaimardii inermis Billman, They were displaced from the shallow shelf by turbidity
Hottinger and Oesterle; A. subtrispinosa (Ishizaki); currents (Huang et al., 2005; Chiu, 2005).
A. yabei (Ishizaki); Pseudorotalia indopacifica (Thal- The conclusions of this benthic foraminiferal study thus
mann); P. schroeteriana (Parker and Jones); Calcarina indicate that the host rocks of the Lichi Mélange were part of
calcar d’Orbigny; Cellanthus craticulatus (Fichtel and the North Luzon Trough forearc sequences before deformation.
Moll); Elphidium crispum (Linné); and Lenticulina The conclusions agree well with predictions D and E in the first
limbosa (Reuss). Today, these shallow-marine benthic subsection under “Discussion.”
The Lichi Mélange 147

TABLE 1: COMPARISON OF BENTHIC FORAMINIFERAL TAXA AND THEIR ABUNDANCE


IN THE LICHI MÉLANGE (THIS STUDY) AND THE TAIYUAN FOREARC TURBIDITES (CHANG, 1967)
Lichi Mélange Taiyuan Forearc Basin
(this study) (Chang, 1967)
Shallow-marine taxa: Shallow-marine taxa:
Amphistegina radiata (Fichtel and Moll) Amphistegina radiata (Fichtel and Moll)
Amphistegina lessonii d’Orbigny Operculina ammonoides (Gronovius)
Operculina complanata (Defrance) Pseudorotalia gaimardii compressiuscula (Brady)
Operculina ammonoides (Gronovius) Pseudorotalia schroeteriana (Parker & Jones)
Ammonia takanabensis (Ishizaki) Elphidium sp.
Asterorotalia gaimardii inermis Billman, Lenticulina atlantica (Barker)
Hottinger and Oesterle Lenticulina calcar (Linné)
Asterorotalia subtrispinosa (Ishizaki) Lenticulina iota (Cushman)
Asterorotalia yabei (Ishizaki) Lenticulina nikobarensis (Schwager)
Pseudorotalia indopacifica (Thalmann) Lenticulina spp.
P. cf. schroeteriana (Parker & Jones)
Calcarina calcar d’Orbigny
Cellanthus craticulatus (Fichtel and Moll)
Elphidium crispum (Linné)
Lenticulina limbosa (Reuss)
Lenticulina orbicularis (d’Orbigny)
Lenticulina iota (Cushman)
Lenticulina cf. cushmani (Galloway and Wissler)
Lenticulina spp.

Deep-marine taxa: Deep-marine taxa:


Bulimina striata d’Orbigny Bulimina striata d’Orbigny
Fontbotia wuellerstorfi (Schwager) Fontbotia wuellerstorfi (Schwager)
Hyalinea balthica (Gmelin) Gyroidinoides orbicularis (d’Orbigny)
Cibicidoides cf. crebbsi (Hedberg) Pullenia bulloides (d’Orbigny)
Cibicidoides subhaidingerii (Parr) Melonis pompilioides (Fichtel and Moll)
Gyroidinoides neosoldanii (Brotzen) Uvigerina (Siphouvigerina) ampullacea Brady
Gyroidinoides orbicularis (d’Orbigny) Uvigerina hispida Schwager
Pullenia bulloides (d’Orbigny) Uvigerina crassicostata Schwager
Melonis nicobarensis (Cushman) Uvigerina excellens Todd
Melonis pompilioides (Fichtel and Moll) Uvigerina proboscidea Schwager
Uvigerina asperula Czjzek Uvigerina subperegrina Cushman and Kleinpell
Uvigerina hispida Schwager Cyclammina orbicularis Brady
Uvigerina peregrina Cushman Cyclammina pusilla Brady
Uvigerina proboscidea Schwager Cyclammina tani Ishizaki
Uvigerina subperegrina Cushman and Kleinpell
Cyclammina japonica Asano kaiensis Fukuta and
Shinoki
Cyclammina tani Ishizaki
Cyclammina cf. trullissata (Brady)

Sediment Composition the upper coherent forearc basin turbidites (<3 Ma; Teng, 1982),
are never observed significantly in the Lichi Mélange. However,
Our field surveys recognize that three distinct end-member lenses of pebbly mudstones (each layer 1–2 m thick; Fig. 12C)
facies are present in the Lichi Mélange: weakly sheared broken are found in the weakly sheared broken formation facies (units
formation facies (β-grade), strongly sheared dismembered facies A, C, and D in Fig. 8). These subangular to subrounded, matrix-
(γ-grade), and highly sheared mélange facies (δ-grade). Like the supported conglomerates (each pebble 5–30 cm in size) are
coherent forearc basin turbidites east of the Tuluanshan Fault, composed primarily of metasandstone and are commonly frac-
the weakly sheared, broken formation facies still preserves dis- tured (Fig. 12C). The orientation of these conglomerates shows
tinct turbidite sedimentary structures like the Bouma turbidite imbricate structure (Fig. 12C), representing typical channel fill-
sequence Ta–Te, and bedding without any angular sheared ophi- ings derived from the exposed Eocene metasandstone beds in the
olite or sedimentary blocks. The turbidite layers of the broken eastern Central Range. Similar pebbly mudstone layers with a
formation facies are primarily of the classical sandy flysch type composite thickness of 150 m can be traced in the lower part of
(Figs. 12–14; C and D facies of Mutti and Lucchi, 1978). The the entire Taiyuan remnant forearc basin (Yao et al., 1988). Such
basal part of these classical turbidite layers (Bouma Ta–Tc) is feeder-channel conglomerate beds, generally called the Shuil-
composed primarily of quartz, like the lower part of the coher- ien Conglomerate, occur in stratigraphic levels in the remnant-
ent forearc sequences. Slate chips, which are commonly found in forearc-basin sequences of the Coastal Range (Fig. 3).
148 Huang et al.

Therefore, sedimentological features, including occurrence last 1 m.y. (Huang et al., 2006). At that time the configuration
and composition of conglomerates in the pebbly mudstone lay- of the Central Range–Coastal Range thus should not have dif-
ers and the turbidite structures preserved in the broken formation fered much from how it looks today. This implies that it may be
facies, support the view of the Lichi Mélange as part of the North improbable to consider that the kaolinite in the Lichi Mélange
Luzon Trough forearc basin sequences in agreement with predic- was derived from the volcanic arc in the eastern Central Range
tion F in the first subsection under “Discussion.” by eastward sedimentary mass slumping.
In the Coastal Range, kaolinite is found only in the volcanic
Significance of Clay Mineral Composition rocks of the Tuluanshan Formation (Wang and Yang, 1975), sug-
gesting that the source of kaolinite in the Lichi Mélange could
Clay mineral compositions in the Coastal Range were stud- have been related to the tectonic disturbance of the Tuluanshan
ied by Wang and Yang (1975), Lin and Chen (1986), Buchovecky volcanic rocks, as follows:
and Lundberg (1988), and Yao et al. (1988). Lin and Chen (1986) 1. Field surveys show that the volcanic rocks of the Tulu-
documented systematically that the clay mineral compositions of anshan Formation beneath the Taiyuan forearc basin are
the muddy matrix and sedimentary blocks in various facies of thrust westward over the Lichi Mélange along the Tulu-
the Lichi Mélange were similar to each other in containing illite, anshan Fault (Fig. 8), and this is the only region in which
chlorite, kaolinite, and a little semectite, which are similar to the the Tuluanshan volcanics are exposed between the Lichi
clay mineral assemblages in the Taiyuan remnant-forearc-basin Mélange and the Loho-Taiyuan remnant forearc basins;
turbidites except for kaolinite. This was not expected, because 2. The volcanic sequence of the Tuluanshan Formation never
sedimentological and micropaleontological studies all show that appears within the remnant forearc basins;
the Taiyuan forearc basin turbidites and the Lichi Mélange were 3. Large, isolated, rootless volcanic blocks of volcaniclas-
derived from the same source and were all deposited in the same tics, andesitic agglomerates, and tuffs are present in the
North Luzon Trough. Consequently, they should contain simi- Lichi Mélange from Loho to Lichi (Fig. 2; Hsu, 1956,
lar clay mineral composition just as they contain similar pebble 1976). These large blocks are fault bounded in the dis-
composition and the benthic foraminiferal fauna. membered or mélange facies of the Lichi Mélange. Near
Lin and Chen (1986) claimed that the Lichi Mélange Antung, hot springs are present along the fault lines where
could have two sediment sources: a continental, low-grade- the fault-bounded tuff and andesitic agglomerates occur
metamorphosed sediment source (accretionary prism of the Cen- in the Lichi Mélange.
tral Range) to provide abundant illite and chlorite, and a weath- We believe that such thrusting, fragmentation, and mixing of
ered, low-latitude, high-rainfall archipelagic volcanic source for the volcanic rocks within the mélange facies have allowed incor-
derivation of kaolinite by eastward sedimentary mass slumping. poration of the kaolinites from the Tuluanshan Formation into the
In contrast, the Taiyuan coherent-forearc turbidite sequences entire Lichi Mélange, but not to the overlying coherent-forearc-
could have only a continental low-grade-metamorphosed sedi- basin strata, by fluid flow (Fisher, 1996) along the ubiquitously
ment source (Yao et al., 1988; Buchovecky and Lundberg, 1988). sheared planes or fractures within the Lichi Mélange.
Therefore, there is no kaolinite in the Taiyuan coherent-forearc- Therefore, characteristic occurrence of clay mineral compo-
basin sequences. Our present work confirms the results of Lin sition in the Lichi Mélange does not conflict with the conclu-
and Chen (1986) that there is significant kaolinite in the Lichi sion that the protolith of the Lichi Mélange was deposited in the
Mélange, regardless of shearing intensity, and agrees that the Lichi western part of the North Luzon Trough forearc basin. Instead,
Mélange and the Taiyuan remnant-forearc-basin strata were all occurrence of characteristic kaolinite in the Lichi Mélange is
deposited in the North Luzon Trough. However, marine geology strongly related to the structures in the southern Coastal Range,
off the Coastal Range shows that the North Luzon Trough forearc showing tectonic involvement of the volcanic basement beneath
basin lies west of the Luzon Volcanic Arc (Figs. 1 and 2). If the the forearc basin in the formation of the Lichi Mélange during
volcanic arc has provided the weathered kaolinite to the forearc closure of the North Luzon Trough, when the Luzon arc-forearc
basin, as Lin and Chen (1986) claimed, the Taiyuan forearc basin was accreted to the southern Coastal Range during the last 1 m.y.
turbidites in the eastern part of the North Luzon Trough would Consequently, the conclusion of this clay-mineral-composition
have received more kaolinite than the Lichi Mélange in the west- study is consistent with field surveys and prediction G in the first
ern part of the North Luzon Trough, unless part of the volcanic subsection under “Discussion.”
arc had been thrust to exhume the metamorphic Central Range in
eastern Taiwan and exposed as the source for providing kaolin- Significance of the Whitish Quartz-Rich Feldspathic
ite to the Lichi Mélange in the western part of the North Luzon Sandstone Blocks in the Lichi Mélange
Trough but not to the remnant Taiyuan forearc basin strata in the
east. However, no such Neogene volcanic arc rocks are found in In the Lichi Mélange, some allochthonous blocks—such as
the eastern Central Range today. Moreover, south of lat 23° N the the angular, whitish, well-lithified, quartz-rich, feldspathic sand-
Taiyuan remnant forearc basin and the Chengkuangao volcanic stone blocks and ophiolite blocks—occur in the block-in-matrix
island were accreted as the southern Coastal Range during the dismembered facies and mélange facies. These blocks have not
The Lichi Mélange 149

been observed in the weakly sheared broken formation facies of accretionary prism (the western Central Range; Fig. 1) were par-
the Lichi Mélange or in the remnant-forearc-basin sequences. tially reset, and the zircons separated from the exhumed metamor-
The origin of these allochthonous sandstone blocks is problem- phic basement (eastern Central Range; Fig. 1) were totally reset.
atic. Like the kaolinite, these whitish angular sandstone blocks Our samples all show partial annealing characteristics, indicating
have been interpreted as being derived from the exposed accre- that they were all derived from the accretionary prism. But the
tionary prism by eastward sedimentary slumping toward the young peak age of the Taiyuan remnant-forearc-basin strata (2.7–
western North Luzon Trough, the site of the Lichi Mélange (Page 2 Ma) is much younger than that of the sheared whitish sandstone
and Suppe, 1981; Lin and Chen, 1986). However, no independent blocks of the Lichi Mélange (9–18 Ma), suggesting that not only
evidence supports this slumping hypothesis. the source, but also the mechanism, for emplacing these sheared
Similar angular, sheared, quartz-rich sandstone blocks (1– angular sandstone blocks into the Lichi Mélange may have sig-
15 cm in size), embedded in the late Pliocene sheared mudstone nificantly differed from the turbidite sedimentation in the Taiyuan
(ca. 3 Ma), have been dredged from the modern Huatung Ridge remnant forearc basin. Moreover, these angular, sheared, whitish
but have never been collected from the remnant North Luzon quartz-rich sandstone blocks have not been observed in the Tai-
Trough off the Coastal Range (Huang et al., 1992; Huang, 1993). yuan remnant-forearc-basin strata, indicating that they may not
The sandstone blocks and the associated mudstone in the Hua- have been delivered to the Lichi Mélange by turbidity currents as
tung Ridge contain not only illite and chlorite but also significant Taiyuan remnant-forearc-basin turbidite layers.
kaolinite similar to what we found in the Lichi Mélange (Huang To identify the source of these quartz-rich, feldspathic sand-
et al., 1992). stone blocks in the Lichi Mélange, we also analyzed four samples
To understand the significance of these allochthonous sand- collected from the deep-sea sandstones in the accretionary prism
stone blocks in the Lichi Mélange, we studied fission tracks of of the Hengchun Peninsula for comparison (Fig. 10). These
zircon grains, separated from the whitish quartz-rich feldspathic samples also show a partial reset feature with two age peaks:
sandstone blocks in the Lichi Mélange, and also separated from a young Neogene peak (15–30 Ma) and an old Mesozoic peak
the coarse grains in the Taiyuan remnant-forearc-basin turbi- (50–100 Ma; Fig. 10). The frequency of the Mesozoic peak is as
dite layers. high as the Neogene peak, similar to that found in the whitish
Fission track analyses of the zircon grains, separated from sandstone blocks of the Lichi Mélange (Fig. 9A). These sand-
the angular, whitish sandstone blocks (>2 m) in the dismembered stone blocks in the Lichi Mélange were determined to be late
facies and the mélange facies in Fuli, and the type locality of the Miocene in age (Chi, 1982; Barrier and Müller, 1984), consistent
Lichi Mélange (Fig. 8), show a partial annealing feature with two with the deep-sea-fan sandstones in the accretionary prism of the
age peaks concentrated at the early–middle Miocene (9–18 Ma) Hengchun Peninsula (Chang, 1966; Huang et al., 1997) but older
and the late Mesozoic (66–130 Ma; Fig. 9A). However, the zir- than the Pliocene–Pleistocene Taiyuan remnant-forearc-basin
con grains of the Taiyuan remnant-forearc-basin turbidites show strata and the Pliocene protolith of the Lichi Mélange.
a distinctly different fission track pattern: one that is completely Three routes are possible through which the late Miocene
reset in a tuff sample in the lower part of the forearc sequence deep-sea-fan sandstones in the late Miocene accretionary prism
(2.08 Ma in sample TF-8; Fig. 9B) or partially reset with a young could have been transported to the Pliocene dismembered facies
peak of Pliocene age (2.5 Ma in sample TF-6; 3.0 Ma in sample and mélange facies of the Lichi Mélange: (1) by erosion from
TF-7; 2.7 Ma in sample TF-9) and an old tail of Mesozoic age the exposed accretionary prism (now the Central Range, before
(64–100 Ma; Fig. 9B). Although the fission track records of the exhumation of the underthrust Eurasian continental materials
zircon grains all show a similar partial annealing feature, there during the last 2.5 m.y.; Huang et al., 2006) and then transported
is a significant difference between them: a higher frequency of a to the North Luzon Trough by turbidity flow longitudinally from
young peak (<7 Ma) in the Taiyuan remnant-forearc-basin turbi- north to south along the trough axis; (2) by erosion as (1) but
dites (Fig. 9B), but a higher frequency of an old peak (>66 Ma, transported eastward by mass slumping into the western part
Mesozoic) in the angular whitish sandstone blocks of the Lichi of the North Luzon Trough (Page and Suppe, 1981); or (3) by
Mélange (Fig. 9A). The frequency for the Mesozoic age in the eastward thrusting to move the non-exposed late Miocene accre-
Lichi Mélange samples is as high as, or even higher than, the tionary prism materials upward into the deformed forearc strata
frequency for the Neogene age. On the contrary, the coherent (proto–Lichi Mélange, like the modern Huatung Ridge) in the
forearc turbidites show a distinct late Neogene age peak but a western part of the North Luzon Trough.
very low frequency for a Mesozoic age. It is improbable that the first and second routes could have
Taiwan has undergone subduction and collision tectonics caused the difference in the fission-track age patterns of zircons
since the Miocene. The fission track study of zircons has been between the Lichi Mélange and the Taiyuan remnant forearc
used successfully to constrain the source of sediment provenance turbidites, because both processes would have provided similar
and geothermal history during the subduction and collision tec- zircon grains from the same source (the exposed accretionary
tonics of Taiwan (Liu, 1982; Liu et al., 2001; Huang et al., 2006; prism) to the North Luzon Trough. Therefore, they should have
Fuller et al., 2006). These studies have shown that fission tracks similar fission-track age patterns in the eastern or the western part
of zircon grains from the Miocene deep-marine turbidites of the of the trough either by longitudinal turbidity flow or by eastward
150 Huang et al.

mass slumping. In contrast, the third route could have provided by sedimentary slumping, because in the Lichi area the ophiol-
the zircon grains from different sources by different mechanisms ite blocks occur with the whitish sandstone blocks, which were
independently to the Lichi Mélange and to the Taiyuan remnant- improbably emplaced into the Lichi Mélange by mass slumping
forearc-basin turbidites. As the marine seismic profiles (Fig. 7) (discussed previously). Therefore, like the concurrent late Mio-
demonstrate, we believe that the blocks of whitish sandstone could cene angular sandstone blocks, the East Taiwan Ophiolite also
have been thrust eastward from the non-exposed accretionary could have been thrust eastward from the accretionary prism of
prism unconformably beneath the western North Luzon Trough the Hengchun Peninsula during the early stage of arc-continent
forearc basin to the deformed proto–Lichi Mélange (Huatung collision. However, in the accretionary prism the ophiolite blocks
Ridge) in the western part of the North Luzon Trough, instead are observed only in the Kenting Mélange of the Hengchun Pen-
of having been derived from the exposed accretionary prism by insula (Fig. 10), and they differ from the East Taiwan Ophiolite
turbidity flows or mass slumping downward to the entire North in composition. For example, the chromatite that is commonly
Luzon Trough. By such eastward backthrusting, the late Miocene found in the Kenting Mélange of the Hengchun Peninsula (Chu
deep-sea sandstone is found restricted to the strongly sheared, et al., 1988) has never been observed in the East Taiwan Ophio-
block-in-matrix, dismembered facies and the mélange facies of lite. Unlike the deep-sea-fan sandstones that occur ubiquitously
the Lichi Mélange but not to the weakly sheared relict, broken in the Hengchun Peninsula (Huang et al., 1997), the ophiolite
formation facies of the Lichi Mélange or the Taiyuan remnant- blocks in the Kenting Mélange are present in a highly limited
forearc-basin strata. This eastward thrusting mechanism is exactly area (Fig. 10). Therefore, it would have been improbable for these
like that of the Huatung Ridge, where we found similar angular, ophiolite blocks to have been thrust eastward from the accretion-
whitish sandstone blocks with kaolinite, or where the proto–Lichi ary prism of the Hengchun Peninsula.
Mélange is developing (Fig. 7), and is consistent with prediction From a regional tectonic view, we prefer the interpretation
G in the first subsection under “Discussion.” that the East Taiwan Ophiolite may represent the oceanic crust of
the Philippine Sea Plate (Juan et al., 1980; Malavieille et al., 2002).
Source of the East Taiwan Ophiolite These ophiolite blocks could have been westward thrust from the
oceanic crust beneath the North Luzon Trough forearc basin or
Dismembered ophiolite blocks are well known in the Lichi the Luzon Arc basement during the final westward accretion of
Mélange in the Coastal Range, eastern Taiwan, and are com- the Luzon Arc-forearc to eastern Taiwan during the last 1 m.y.
monly known as the East Taiwan Ophiolite. However, the source (Fig. 5C, 3). The age of the Luzon Arc is no later than middle
and the mechanism for emplacing these oceanic-affinity blocks Miocene, consistent with the age of the East Taiwan Ophiolite.
within the Lichi Mélange are still the subject of debate. Along The occurrence of ophiolitic rocks in the Banda forearc basement
the Mukeng River the number and size of these blocks (usually has been documented from the Bobonaro Mélange of the Timor
0.1–2 m) are much less common and are smaller than those in arc-continent collision orogen by thrusting emplacement (Harris
the Kuanshan and Lichi areas (Fig. 1). Along the Mukeng River et al., 1998). It is also possible that the East Taiwan Ophiolite
sections the East Taiwan Ophiolite blocks do not occur with could have been developed by suprasubduction (Pearce et al.,
the whitish sandstone blocks. However, they both appear to be 1984) by which the lower crustal materials, such as serpentinite
restricted to the intensely sheared, dismembered facies or the and peridotite, could be found in a forearc basin setting, similar
mélange facies of units I, II, III, and IV (Fig. 8). In the Lichi area to the modern Mariana forearc (Fryer et al., 1985).
the larger East Taiwan Ophiolite blocks (several to hundreds of
meters) do occur commonly with the whitish sandstone blocks Global Significance of the Lichi Mélange: Comparison
(Liou et al., 1977; Page and Suppe, 1981). There are two pos- with the Bobonaro Mélange in the Timor Arc-Continent
sibilities for the source of the East Taiwan Ophiolite blocks: the Collision Orogen
South China Sea oceanic crust or the oceanic basement beneath
the Luzon Arc-forearc of the Philippine Sea plate. Mélange is generally considered to be developed within the
If the East Taiwan Ophiolite was part of the South China accretionary prism by scraping off the underthrusting oceanic
Sea oceanic crust, its age would be younger than late Oligocene crust and the overlying deep-sea marine sediments along the
because this sea opened in 32–17 Ma (Taylor and Hayes, 1980). décollement fault in the subduction zone (Silver and Beutner,
Microfossils recovered from a thin red shale intercalated within 1980; Cowan, 1985; Cloos and Shreve, 1988), and the mechanism
this ophiolite sequence indicated an age of 15 Ma (Zone NN5 of responsible for formation of mélange could be tectonic thrusting,
calcareous nannoplanktons; Huang et al., 1979), which would be sedimentary slumping, or mud diapirism (Cowan, 1985; Barber
close to the last phase of the opening of the South China Sea. This et al., 1986). However, the Pliocene Lichi Mélange was devel-
conclusion may favor an oceanic-crust source for the South China oped in the forearc basin rather in the accretionary prism. The
Sea (Suppe et al., 1981). However, no East Taiwan Ophiolite–like sedimentary mass-slumping process is unlikely as the major
rock occurs in the present Central Range. These ophiolite blocks, mechanism responsible for formation of the Lichi Mélange, as
presumably of a South China Sea source, could not have been discussed previously. Today, two small mud volcanoes, each
derived from the exposed accretionary prism (Central Range) 100–200 wide, are found within the Lichi Mélange (Yang et al.,
The Lichi Mélange 151

2004). However, in comparison with the wide distribution of the Mélange, but they played an important role in the development of
Lichi Mélange (50 km long, 2 km wide), such small areas of mud the Bobonaro Mélange.
volcanoes could not have played an important role in the devel-
opment of the Lichi Mélange. CONCLUSIONS
Today, mélange can be found also in the Timor arc-continent
collision region. The Bobonaro Mélange in Timor Island is char- New seismic surveys along the 21° N transect across the
acterized by active development, long extension from offshore North Luzon Trough show a syndeformational sedimentation
to onshore for more than 2000 km, a young age (since 3 Ma), in the forearc basin. Episodic deformation has occurred in the
and good exposures. Other characteristics include a wide variety western part of the North Luzon Trough since the early history of
of rock types in block-in-matrix facies caused by polymecha- forearc sedimentation by the active collision between the Luzon
nisms of tectonic detachment, mud diapirism, and sedimentary Arc and the Eurasian continent. The deformation resulted in for-
slumping (Barber et al., 1986; Harris et al., 1998). In comparing mation of the Huatung Ridge, which continues northward to the
the Lichi Mélange in the Coastal Range, eastern Taiwan, with Lichi Mélange in the southernmost Coastal Range, suggesting
the Bobonaro Mélange in Timor, of the Banda Arc region, it is a direct geological relationship between formation of the Lichi
interesting to note similarities and differences. For example, the Mélange and the submarine Huatung Ridge. Multiple lines of
Lichi Mélange and the Bobonaro Mélange are both of a young new evidence, including on-land field surveys, planktic and ben-
age (ca. 3 Ma), forming mélanges today in active arc-continent thic foraminiferal study, clay mineral composition, and zircon
collision tectonics, traceable from onshore to offshore, involving fission-track analyses, confirm that the Lichi Mélange was part
forearc basement tectonics, and showing variable degrees of rock of the North Luzon Trough forearc sequences (Fig. 5C, 1), rep-
shearing, fragmentation, and mixing from weakly sheared bro- resenting a collision mélange formed along arcward (eastward)
ken formation facies to block-in-clay mélange facies. However, backthrusting during forearc closure when the Luzon arc collided
the Lichi Mélange developed in a forearc basin setting rather than incipiently with the Eurasian continent.
in an accretionary prism setting, as did the Bobonaro Mélange. The formation of the Lichi Mélange involved two thrust
Therefore, no continental fragment of the lower plate, like the events: (1) early arcward backthrusting in the western part of the
Australian pre- and post-rifting sequences in the Bobonaro North Luzon Trough to form the proto–Lichi Mélange, like the
Mélange, was involved during formation of the Lichi Mélange. Huatung Ridge, during progressive closure of the forearc basin
On the other hand, the Luzon Volcanic Arc was involved during in the initial arc-continent collision stage (ca. 3 Ma; Figure 5C,
formation of the Lichi Mélange, but the Banda Volcanic Arc was 2), followed by (2) westward thrusting of the Luzon Arc-forearc
not involved in formation of the Bobonaro Mélange. The micro- onto eastern Taiwan to form the present southern Coastal Range
fossil composition of the Lichi Mélange (Pliocene) is much sim- during the advanced arc-continent collision stage during the last
pler than that of the Bobonaro Mélange (Jurassic to Pleistocene). 1 m.y. (Fig. 5C, 3).
Deep-sea pelagic sediments such as radiolarian ooze are pres- The Lichi Mélange provides a unique example in compari-
ent in the Bobonaro Mélange, but only hemipelagic muds and son with worldwide orogenic belts in that the mélange has devel-
turbidites in the Lichi Mélange. Mud diapirism and sedimentary oped by tectonic thrusting along closure of the forearc basin in
mass slumping were insignificant in the development of the Lichi the region of active arc-continent collision.
152 Huang et al.

APPENDIX

APPENDIX. CLAY MINERAL COMPOSITION (IN PERCENTAGES) OF CLAY FRACTIONS ALONG THE MUKENG RIVER
AND ITS SOUTHERN TRIBUTARY SECTIONS IN THE LICHI MÉLANGE
No. Sample Lithology Illite Chlorite Kaolinite Smectite Mixed-layer
E MK01-1- Sheared mudstone 47 18 13 Trace to absent 22
E MK02- Sheared mudstone 43 20 9 Trace to absent 28
E MK03-0- Non-sheared silty mudstone 56 11 6 Trace to absent 27
E MK03-3- Slightly sheared mudstone 57 14 8 Trace to absent 21
E MK03-4- Sheared mudstone 48 19 12 Trace to absent 21
E MK04- Sheared mudstone 52 22 10 Trace to absent 16
C MK06-1- Sheared mudstone 47 14 11 Trace to absent 28
C MK10- Sheared mudstone 51 21 10 Trace to absent 18
B MK12- Non-sheared mudstone 56 18 10 Trace to absent 16
B MK13-1- Sheared mudstone 52 19 11 Trace to absent 18
D MKN01- Non-sheared mudstone 48 22 11 Trace to absent 19
A MKN02-2- Slightly sheared mudstone 47 19 12 Trace to absent 22
A MKN02-4- Slightly sheared mudstone 49 19 11 Trace to absent 21
A MKN02-6 Sheared mudstone 59 12 11 Trace to absent 18
A MKN03-1- Sheared mudstone 53 21 11 Trace to absent 15
A MKN04- Sheared mudstone 56 21 10 Trace to absent 13
A MKN05-2- Non-sheared shale 53 16 10 Trace to absent 21
A MKN05-4- Non-sheared mudstone 44 23 14 Trace to absent 19
A MKN05-6- Sheared mudstone 47 20 10 Trace to absent 23
A MKN07- Sheared mudstone 39 17 15 Trace to absent 29
A MKN08- Non-sheared mudstone 63 19 14 Trace to absent 4
F MKS01- Non-sheared shale 67 21 8 Trace to absent 4
F MKS02- Non-sheared mudstone 52 14 13 21
F MKS03- Sheared mudstone 53 15 14 18
F MKS03S- Non-sheared siltstone 46 19 14 21
F MKS04-2- Non-sheared mudstone 64 23 7 6
F MKS04-4- Sheared mudstone 63 19 12 6
F MKS05-1- Non-sheared mudstone 53 18 10 19
F MKS05-2- Non-sheared mudstone 55 18 10 17
F MKS06- Non-sheared mudstone 52 19 12 17
G MKS08- Sheared mudstone 52 16 9 23
G MKS09- Mud in conglomerate layer 52 19 9 20
G MKS10- Sheared mudstone 53 18 11 18
TF-06a Mudstone 54 24 6 16
TF-07 Mudstone 60 23 5 12
TF09 Mudstone 50 21 8 21
LM-1 mud Mudstone 51 19 14 16
YF02C Mudstone 53 18 6 23
YF03C Mudstone 54 16 15 15
YF01C Sandstone 28 50 15 7
LM-1 Sandstone 42 23 22 13
LM01 Sandstone 38 27 14 21
FK01C Sandstone 42 15 17 16

Average: 51.5 1 9.5 11.1 17.8

For comparison, average clay mineral composition of


43.7 9.8 13.7 3.7 28.9
Lin and Chen (1986):
The Lichi Mélange 153

ACKNOWLEDGMENTS Chiu, Y.L., 2005, Sedimentological study of Site 1202, Leg 195, in southern
Okinawa Trough off NE Taiwan [M.S. thesis]: Department of Earth Sci-
ences, National Cheng Kung University, 80 p.
This study was financially supported by grants from the National Chu, H.T., Shen, P., and Jeng, R.C., 1988, The origin of chromatite from the
Science Council (NSC93–2116-M-006–001; NSC94–2116- Kenting Mélange, southern Taiwan: Proceedings of the Geological Soci-
M006–005) and the Research Center of Ocean Environment and ety of China, v. 31, p. 33–52.
Chung, S.L., and Sun, S.S., 1992, A new genetic model for the East Taiwan
Technology, NCKU, to C.Y. Huang, and grant G2000046705 Ophiolite and its implications for Dupal domains in the Northern Hemi-
from the National Important Basic Research and Development sphere: Earth and Planetary Science Letters, v. 109, p. 133–145, doi:
Projects to B. Yao. We appreciate the crew members of the R/V 10.1016/0012-821X(92)90079-B.
Cloos, M., and Shreve, R.L., 1988, Subduction-channel model of prism accre-
Tanbao for their great efforts during the seismic surveys across tion, mélange formation, sediment subduction, and subduction erosion
the North Luzon Trough off southern Taiwan. The authors also at convergent plate margins: 1. Background and description: Pageoph,
appreciate the constructive discussions with R. Harris and D. v. 128, p. 455–500, doi: 10.1007/BF00874548.
Cowan, D., 1985, Structural styles in Mesozoic and Cenozoic mélanges in
Reed in comparing the Lichi Mélange in eastern Taiwan with the western Cordillera of North America: Geological Society of Amer-
the Bobonaro Mélange in Timor. The manuscript was improved ica Bulletin, v. 96, p. 451–462, doi: 10.1130/0016-7606(1985)96<451:
by comments from A. Basu and an anonymous reviewer. SSIMAC>2.0.CO;2.
Ernst, W.G., 1977, Olistostromes and included ophiolite debris from the
Coastal Range of eastern Taiwan: Geological Society of China Memoir,
REFERENCES CITED v. 2, p. 97–114.
Fisher, D.M., 1996, Fabrics and veins in the forearc: A record of cyclic fluid
Barber, A.J., Tjokrosapoetro, S., and Charlton, T.R., 1986, Mud volcanoes, flow at depth of 15 km, in Bebout, G.E., et al., eds., Subduction: Top to
shale diapirs, wrench faults and mélanges in accretionary complex, east- Bottom: American Geophysical Union Monograph 96, p. 75–89.
ern Indonesia: American Association of Petroleum Geologists Bulletin, Fryer, P., Ambos, E.L., and Hussong, D.M., 1985, Origin and emplace-
v. 70, p. 1729–1741. ment of Mariana forearc seamounts: Geology, v. 13, p. 774–777, doi:
Barrier, E., and Müller, C., 1984, New observations and discussions on the ori- 10.1130/0091-7613(1985)13<774:OAEOMF>2.0.CO;2.
gin and age of the Lichi Mélange: Geological Society of China Memoir, Fuller, C.W., Willett, S.D., Fisher, D., and Lu, C.Y., 2006, A thermomechanical
v. 6, p. 303–325. wedge model of Taiwan constrained by fission-track thermochronometry:
Biq, C., 1971, Comparison of mélange tectonics in Taiwan and in some other Tectonophysics, v. 425, p. 1–24, doi: 10.1016/j.tecto.2006.05.018.
mountain belts: Petroleum Geology of Taiwan, no. 9, p. 79–106. Harris, R.A., Sawyer, R.K., and Audley-Charles, M.G., 1998, Collision mélange
Biq, C., 1973, Kinematic pattern of Taiwan as an example of actual continent- development: Geologic associations of active mélange-forming processes
arc collision: Taipei, Report of the Seminar on Seismology: US-ROC with exhumed mélange facies in the western Banda orogen, Indonesia:
Cooperative Science Program, abstract, p. 21–26. Tectonics, v. 17, p. 458–479, doi: 10.1029/97TC03083.
Buchovecky, E.J., and Lundberg, N., 1988, Clay mineralogy of mudstones from Horng, C.S., and Shea, K.S., 1996, Dating of the Plio-Pleistocene rapidly
the southern Coastal Range, eastern Taiwan: Unroofing of the orogen ver- deposited sequence based on integrated magneto-biostratigraphy: A case
sus in-situ diagenesis: Acta Geologica Taiwanica, no. 26, p. 247–261. study of the Madagida-Chi section, Coastal Range, eastern Taiwan: Jour-
Chai, B.H.T., 1972, Structure and tectonic evolution of Taiwan: American Jour- nal of the Geological Society of China, v. 39, p. 31–58.
nal of Science, v. 272, p. 389–422. Hsu, T.L., 1956, Geology of the Coastal Range, eastern Taiwan: Geological
Chang, C.P., Angelier, J., and Huang, C.Y., 2000, Origin and evolution of a Survey of Taiwan Bulletin, v. 8, p. 39–64.
mélange: The active plate boundary and suture zone of the Longitudinal Hsu, T.L., 1976, The Lichi Mélange in the Coastal Range framework: Taiwan
Valley, Taiwan: Tectonophysics, v. 325, p. 43–62, doi: 10.1016/S0040- Geological Survey Bulletin, v. 25, p. 87–95.
1951(00)00130-X. Huang, C.Y., 1989, Implication of the Post-Lushanian faunal change for the
Chang, C.P., Angelier, J., Huang, C.Y., and Liu, C.S., 2001, Structural evolu- occurrence of Kuroshio Current in the early late Miocene: Foraminiferal
tion and significance of a mélange in a collision belt: The Lichi Mélange evidence from the Chuhuangkeng section, northern Taiwan: Proceedings
and the Taiwan arc-continent collision: Geological Magazine, v. 138, of the Geological Society of China, v. 32, p. 21–45.
p. 633–651. Huang, C.Y., 1993, Bathymetric ridge and trough in the active arc-continent
Chang, L.S., 1966, A biostratigraphic study of the Tertiary in the Hengchun collision region off southeastern Taiwan: Reply and discussion: Journal
peninsula, Taiwan, based on smaller foraminifera (III: Southern Part): of the Geological Society of China, v. 36, p. 91–109.
Proceedings of the Geological Society of China, no. 9, p. 53–63. Huang, C.Y., and Yin, Y.C., 1990, Bathymetric ridges and troughs in the active
Chang, L.S., 1967, A biostratigraphic study of the Tertiary in the Coastal arc-continent collision region off southeastern Taiwan: Proceedings of the
Range, eastern Taiwan, based on smaller foraminifera (I: Southern Part): Geological Society of China, v. 33, p. 351–372.
Proceedings of the Geological Society of China, no. 10, p. 64–76. Huang, C.Y., Shyu, C.T., Lin, S.B., Lee, T.Q., and Sheu, D.D., 1992, Marine
Chen, C.H., Shieh, Y.N., Lee, T., Chen, C.H., and Mertzman, S.A., 1990, Nd- geology in the arc-continent collision zone off southeastern Taiwan:
Sr-O isotopic evidence for source contamination and an unusual mantle Implications for late Neogene evolution of the Coastal Range: Marine
component under Luzon arc: Geochimica et Cosmochimica Acta, v. 54, Geology, v. 107, p. 183–212, doi: 10.1016/0025-3227(92)90167-G.
p. 2473–2483, doi: 10.1016/0016-7037(90)90234-C. Huang, C.Y., Yuan, P.B., Song, S.R., Lin, C.W., Wang, C., Chen, M.T., Shyu,
Chen, W.S., 1997, Mesoscopic structures developed in the Lichi Mélange dur- C.T., and Karp, B., 1995, Tectonics of short-lived intra-arc basins in the
ing the arc-continent collision in the Taiwan region: Journal of the Geo- arc-continent collision terrane of the Coastal Range, eastern Taiwan: Tec-
logical Society of China, v. 40, p. 415–434. tonics, v. 14, p. 19–38, doi: 10.1029/94TC02452.
Cheng, Y.M., Huang, C.Y., Yeh, J.J., and Chen, W.S., 1984: The Loshui For- Huang, C.Y., Wu, W.Y., Chang, C.P., Tsao, S., Yuan, P.B., Lin, C.W., and Xia,
mation: Deeper water sandstones in the Hengchun Peninsula, southern K.Y., 1997, Tectonic evolution of accretionary prism in the arc-conti-
Taiwan: Acta Geologica Taiwanica, no. 22, p. 35–52. nent collision terrane of Taiwan: Tectonophysics, v. 281, p. 31–51, doi:
Chi, W.R., 1982, The calcareous nannofossils of the Lichi mélange and the 10.1016/S0040-1951(97)00157-1.
Kenting mélange and their significance in the interpretation of plate tec- Huang, C.Y., Yuan, P.B., Lin, C.W., and Wang, T.K., 2000, Geodynamic pro-
tonics of the Taiwan region: Ti-Chih (Geology), v. 4, p. 99–112. cesses of Taiwan arc-continent collision and comparison with analogs in
Chi, W.R., Namson, J., and Suppe, J., 1981, Stratigraphic record of plate inter- Timor, Papua New Guinea, Urals and Corsica: Tectonophysics, v. 325,
actions in the Coastal Range of eastern Taiwan: Geological Society of p. 1–21, doi: 10.1016/S0040-1951(00)00128-1.
China Memoir, v. 4, p. 155–194. Huang, C.Y., Chiu, Y.L., and Zhao, M., 2005, Core description and preliminary
Chien, C.W., 2003, Forearc origin of the Lichi Mélange in Coastal Range, sedimentology features of silt and sand depositions of ODP Site 1202D,
eastern Taiwan: Micropaleontological evidence [M.S. thesis]: Institute of Leg, 195, southern Okinawa Trough: Terrestrial, Atmospheric and Oce-
Geosciences, National Taiwan University, 74 p. anic Sciences, v. 16, p. 19–44.
154 Huang et al.

Huang, C.Y., Yuan, P.B., and Tsao, S.J., 2006, Temporal and spatial records of Howell, M.F., eds., Marginal Basin Geology: Geological Society [Lon-
active arc-continent collision in Taiwan: A synthesis: Geological Society don] Special Publication 16, p. 77–94.
of America Bulletin, v. 118, p. 274–288, doi: 10.1130/B25527.1. Raymond, L.A., 1984, Classification of mélanges, in Raymond, L.A., ed.,
Huang, T.C., Chen, M.P., and Chi, W.R., 1979, Calcareous nannofossils from Melanges: Their Nature, Origin, and Significance: Geological Society of
the red shale of the ophiolite-mélange complex, eastern Taiwan: Geologi- America Special Paper 198, p. 7–20.
cal Society of China Memoir, v. 3, p. 131–138. Reed, D.L., Lundberg, N., Liu, C.S., Luo, B.Y., 1992, Structural relations along
Juan, V.C., Lo, H.J., and Chen, C.C., 1980, Genetic relationships and emplace- the margins of the offshore Taiwan accretionary wedge: Implications for
ment of the exotic basic rocks enclosed in the Lichi mélange, east Coastal accretion and crustal kinematics: Acta Geologica Taiwanica, no. 30, p.
Range, Taiwan: Proceedings of the Geological Society of China, no. 23, 105–122.
p. 56–68. Silver, E.A., and Beutner, E.C., 1980, Mélanges: Geology, v. 8, p. 32–34, doi:
Lin, S.B., and Chen, G.T., 1986, Clay minerals from the Lichi Mélange and its 10.1130/0091-7613(1980)8<32:M>2.0.CO;2.
adjacent formations in the Coastal Range, eastern Taiwan: Acta Geologica Suppe, J., Liou, J.G., and Ernst, W.G., 1981, Paleogeographic origins of the
Taiwanica, no. 24, p. 319–356. Miocene East Taiwan Ophiolite: American Journal of Science, v. 281,
Liou, J.G., Suppe, J., and Ernst, W.G., 1977, Conglomerates and pebbly mud- p. 228–246.
stones in the Lichi Mélange, eastern Taiwan: Geological Society of China Taylor, B., and Hayes, D.E., 1980, The tectonic evolution of the South China
Memoir, v. 2, p. 115–128. Sea Basin, in Hayes, D.E., ed., The Tectonic and Geologic Evolution of
Liu, C.S., Liu, S.Y., Lallemand, S., Lundberg, N., and Reed, D.L., 1998, Digital the Southeast Asian Seas and Islands (Part I): American Geophysical
elevation model offshore Taiwan and its tectonic implications: Terrestrial, Union Monograph 23, p. 89–104.
Atmospheric and Oceanic Sciences, v. 9, p. 705–738. Teng, L.S., 1982, Stratigraphy and sedimentation of the Shuilien Conglomer-
Liu, T.K., Hsieh, S., Chen, Y.G., and Chen, W.S., 2001, Thermo-kinematic evo- ate, northern Coastal Range, eastern Taiwan: Acta Geologica Taiwanica,
lution of the Taiwan oblique-collision mountain belt as revealed by zircon no. 21, p. 201–220.
fission track dating: Earth and Planetary Science Letters, v. 186, p. 45–56, Van Morkhoven, F.P.C.M., Berggren, W.A., and Edwards, A.S., 1986, Cenozoic
doi: 10.1016/S0012-821X(01)00232-1. cosmopolitan deep-water benthic foraminifera: Bulletin des Centres de
Liu, T.W., 1982, Tectonic implication of fission track ages from the Central Recherches Exploration-Production, Elf-Aquitaine Memoir, v. 11, 406 p.
Range, Taiwan: Proceedings of the Geological Society of China, no. 25, Wang, C.S., 1976, The Lichi Formation of the Coastal Range and arc-continent
p. 22–37. collision in eastern Taiwan: Geological Survey of Taiwan Bulletin, v. 25,
Loeblich, A., Jr., and Tappan, H., 1964, Treatise on Invertebrate Paleontology, p. 73–86.
Part C, Protista 2, Sarcodina, Chiefly “Thecamoebians” and Foraminifer- Wang, P.X., Zhang, J., and Min, Q., 1985, Distribution of foraminifera in surface
ida: Geological Society of America and University of Kansas Press, 900 sediments of the East China Sea, in Wang, P., ed., Marine Micropaleontol-
p. ogy of China, China Ocean Press: Berlin, Springer-Verlag, p. 34–69.
Loeblich, A., Jr., and Tappan, H., 1988, Foraminiferal Genera and Their Clas- Wang, Y., and Yang, C.N., 1975, Expandable clay deposits and X-ray diffraction
sification: New York, Van Nostrand Reinhold, 970 p. study of montmorillonite from Coastal Range, Taiwan: Acta Geologica
Lundberg, N., Reed, D.L., Liu, C.S., and Lieske, J., Jr., 1997, Forearc-basin Taiwanica, no. 18, p. 14–25.
closure and arc accretion in the submarine suture zone south of Taiwan: Yang, T.F., Yeh, G.H., Fu, C.C., Wang, C.C., Lan, T.F., Lee, H.F., Chen, C.H.,
Tectonophysics, v. 274, p. 5–23, doi: 10.1016/S0040-1951(96)00295-8. Walia, V., and Sung, Q.C., 2004, Composition and exhalation flux of
Malavieille, J., Lallemand, S.E., Dominguez, S., Deschamps, A., Lu, C.Y., Liu, gases from mud volcanoes in Taiwan: Environmental Geology, v. 46,
C.S., and Schnurle, P., and the ACT Scientific Crew, 2002, Arc-continent p. 1003–1011, doi: 10.1007/s00254-004-1086-0.
collision in Taiwan: New marine observations and tectonic evolution: Yao, T.M., Tien, P.L., and Wang-Lee, C., 1988, Clay mineralogical studies on
Geological Society of America Special Paper 358, p. 187–211. the Neogene formations, Taiyuan basin, southern Coastal Range, eastern
Mutti, E., and Lucchi, F.R., 1978, Turbidites of the Northern Apennines: Intro- Taiwan: Acta Geologica Taiwanica, no. 26, p. 263–277.
duction to facies analysis: International Geology Review, v. 44, p. 125– Yu, S.B., Chen, H.Y., and Kuo, L.C., 1997, Velocity field of GPS stations in
166. the Taiwan area: Tectonophysics, v. 274, p. 41–59, doi: 10.1016/S0040-
Page, B.M., and Suppe, J., 1981, The Pliocene Lichi Mélange of Taiwan: Its 1951(96)00297-1.
plate tectonic and olistostromal origin: American Journal of Science,
v. 281, p. 193–227.
Pearce, J.A., Lippard, S.J., and Robert, S., 1984, Characteristics and tectonic
significance of suprasubduction zone ophiolites, in Kokelaar, B.P., and MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Oblique subduction in an island arc collision setting: Unique sedimentation, accretion,


and deformation processes in the Boso TTT-type triple junction area, NW Pacific

Yujiro Ogawa*
Doctoral Program in Earth Evolution Sciences, University of Tsukuba, Tsukuba 305-8572, Japan

Yoshihiro Takami
College of Natural Science, University of Tsukuba, Tsukuba 305-8571, Japan

Sakiko Takazawa*
College of Natural Science, University of Tsukuba, Tsukuba 305-8571, Japan

ABSTRACT

The NW corner of the Pacific Ocean is a place of unique Tertiary tectonism,


which provides one of the clearest examples of arc-arc collision. Voluminous Creta-
ceous rhyolitic-granitic magmatism along the continental margin continues into the
Paleogene. In contrast, Miocene island arc volcanism follows Eocene boninitic mag-
matism in the Izu-Mariana Arc, in association with the opening of backarc basins,
including those in the Philippine and Japan Seas. The triple junction between the
Eurasian, Philippine Sea, and Pacific plates arrived in the area south of Tokyo dur-
ing the Miocene, just as the Japan Sea was opening. After the beginning of Philip-
pine Sea plate subduction to the north, the Izu Island Arc began to collide obliquely
with the Honshu Arc. As a result, this unique tectonic setting in the NW Pacific has
produced a miniature Alpine-type orogenic belt (Tanzawa) in the collisional center,
whereas in the eastern part of the Izu Arc sediment has been actively accreting in that
forearc. Such settings have resulted in systematic accretionary prism formation from
the early Miocene in the Boso-Miura peninsular area to the present in the Sagami
Trough area. We modeled the tectonics by a simple sandbox experiment. Systematic
fault and fracture patterns of the oblique subduction type are predicted to occur dur-
ing arc-arc collision.

Keywords: island arc collision, triple junction, forearc sedimentation, forearc sliver,
accretionary prism, analogue modeling.

*Corresponding author, Ogawa: fyogawa45@yahoo.co.jp. Present address, Takazawa: Ministry of Economy, Trade and Industry of Japan, Kasumigaseki, Tokyo
100-8901, Japan.

Ogawa, Y., Takami, Y., and Takazawa, S., 2008, Oblique subduction in an island arc collision setting: Unique sedimentation, accretion, and deformation processes
in the Boso TTT-type triple junction area, NW Pacific, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record
in Arc Collision Zones: Geological Society of America Special Paper 436, p. 155–170, doi: 10.1130/2008.2436(07). For permission to copy, contact editing@
geosociety.org. ©2008 The Geological Society of America. All rights reserved.

155
156 Ogawa et al.

INTRODUCTION by the collision of the Izu Arc with the Honshu forearc. However,
the geology and tectonics have not yet been fully explained with
Plate tectonic triple junctions provide special geologic and reference to the oblique subduction setting.
tectonic conditions in which unique topography, sedimentation, In this paper we first review the submarine topography and
and deformation can form. Among the various types of triple geology along the present Sagami Trough area on the basis of the
junctions (McKenzie and Morgan, 1969), a trench-trench-trench marine work of Kato et al. (1983), Kong et al. (1984), Ohkouchi
(TTT) type of triple junction has been observed only off the Boso (1990), and Iwabuchi et al. (1990), and many other works related
Peninsula, central Japan, in the northwestern corner of the Pacific to the KAIKO project as well as the terrestrial geology of Ogawa
Ocean. This is called the Boso triple junction (Ogawa et al., and Horiuchi (1978), Ogawa (1982), and Hanamura and Ogawa
1989; Seno et al., 1989) (Fig. 1). At present, another similar type (1993), and others. We then present new evidence, including sim-
of convergent triple junction is known nearby just north of Mount ple two-dimensional (2D) and three-dimensional (3D) analogue
Fuji, which is called the Fuji triple junction. Thus, in a narrow shear-box experiments for oblique subduction (Takami, 1999;
area beneath the Tokyo region, two converging triple junctions Takazawa, 2003) in order to explain the uniqueness of the TTT
make a very complicated tectonic setting (Sato et al., 2005). type of the triple junction area in terms of the present sedimenta-
The geometry and tectonics of the Boso triple junction were tion and deformation in an oblique subduction arc-arc collisional
discussed by Seno et al. (1989) and Huchon and Labaume (1989). plate boundary.
The triple junction was originally called the off-Boso triple junc-
tion by Ogawa et al. (1989). The “central Japan triple junction” TECTONIC SETTING
of Lallemant et al. (1996) is not an adequate term, because there
are two triple junctions in central Japan. Seismic profiling and The northwestern Pacific Rim has a long tectonic his-
related submarine researches were performed during the Japan- tory, dating from the Mesozoic and characterized by Jurassic-
France KAIKO project (Renard et al., 1987; Nakamura et al., Cretaceous Pacific plate subduction under NE Asia. The Japa-
1987; Pautot et al., 1987). After this project the principal results nese main islands are the products of such long-term subduction
were published by Ogawa et al. (1989), Seno et al. (1989), Soh et (Seno and Maruyama, 1984; Maruyama and Seno, 1986; Taira
al. (1988, 1990), and Lallemant et al. (1996), and many submarine et al., 1989; Ogawa et al., 1997; Isozaki, 1996). Since the Oligo-
and on-land research projects were conducted. Most important is cene and early to middle Miocene, when the Shikoku Basin, the
the recognition that the northwestern part of the triple junction in easternmost backarc basin of the Philippine Sea plate, opened,
the Sagami Trough is a place of active sedimentation, deforma- the present Izu-Mariana Island Arc front has been moving from
tion, and accretion along the oblique subduction boundary caused west to east just south off the shore of Honshu (Kobayashi, 1984,

Figure 1. Index map of the Izu Arc collision zone with the Honshu Arc since the middle Miocene (left) to the present
(right). Izu forearc accretion occurs on the NW of the Boso triple junction. Plates A, B, and C are the Pacific, Philippine
Sea, and Eurasian plates, respectively. Plate C, just on the NW side of the triple junction, was converted to the North
American plate during the Quaternary, when the boundary between the North American and Eurasian plates extended to
central Honshu. After Hanamura and Ogawa (1993).
Oblique subduction in an island arc collision setting 157

1995). After seafloor spreading in the Shikoku Basin ceased, the 2006). A similar stratigraphy is preserved as the remnant of the
Izu Island Arc and its NE edge, coincident with the triple junc- “Mineoka plate” and exposed onshore as the Mineoka Ophiolite
tion, arrived close to its present position. Subsequently the Shi- in the Boso Peninsula (Fig. 4) (Ogawa and Taniguchi, 1988; Sato
koku Basin lithosphere began to subduct to the NE under Honshu et al., 1999; Sato and Ogawa, 2000; Ogawa and Takahashi, 2004).
(Seno and Maruyama, 1984). Consequently, the Izu Island Arc Similarly, parts of the Philippine Sea plate are found in northern
is now connected to central Honshu, but it also extends as far New Guinea as the North New Guinea plate (Seno, 1984; Lus et
south as the Mariana Arc. The ridge of the Izu-Mariana Arc itself al., 2004), comprising Cretaceous to Eocene basaltic rocks and
has collided with the Honshu Arc in central Honshu in several their pelagic cover (Ogawa and Taniguchi, 1987, 1988; Mohiud-
stages (Matsuda, 1978; Ogawa, 1985; Ogawa et al., 1985; Soh din and Ogawa, 1998; Hirano et al., 2003; Ogawa and Sashida,
et al., 1988; Aoike, 1999; Sato et al., 1999). The timing of the 2005). Such oceanic plate history is now preserved in the Boso
first collision is synchronous with the opening of the Japan Sea Peninsula and has been documented by the detailed structural
at 15 Ma (Niitsuma, 1988; Takahashi and Saito, 1997). After that and stratigraphic analysis of the Mineoka Ophiolite Belt (Hirano
time the northeastern corner of the Philippine Sea plate has sup- et al., 2003; Takahashi et al., 2003; Ogawa and Takahashi, 2004;
plied Izu-Mariana forearc materials to the Honshu forearc that Mori and Ogawa, 2005).
have been accreted to form the Miura-Boso accretionary prisms The Mineoka Ophiolite Belt has lain in a forearc sliver fault
(Figs. 2 and 3) (Ogawa, 1982, 1985; Ogawa and Taniguchi, 1988; zone since the middle Miocene, after which time there has been a
Hanamura and Ogawa, 1993; Takahashi and Saito, 1997; Yama- systematic development of accretionary prisms south of the Boso-
moto et al., 2000, 2005). Miura Peninsulas (Ogawa, 1983; Ogawa and Taniguchi, 1988;
The age of the Izu Arc extends at least to the Eocene Lallemant et al., 1996). These prisms are formed of Izu forearc
(ca. 45 Ma; Seno and Maruyama, 1984). According to the age and paleo–Sagami Trough sediments derived from an area east of
determination and spreading history of the Philippine Sea plate the Izu collision zone (Fig. 2) (Ogawa, 1982, 1985; Ogawa and
(Kinoshita, 1980; Okino et al., 1994; Kobayashi et al., 1995), the Taniguchi, 1987, 1988; Hanamura and Ogawa, 1993; Yamamoto
arc was once far to the south, in the Southern Hemisphere, and it et al., 2000, 2005; Yamamoto and Kawakami, 2005). Thus the
rotated clockwise during its northward movement to its present Sagami Trough is now the modern oblique subduction boundary
position (Seno and Maruyama, 1984; Hall et al., 1995; Takahashi at the northeasternmost corner of the Philippine Sea plate and has
and Saito, 1997). Because of spreading in the backarc basins fol- been a place of sedimentation and deformation between the Izu
lowing rifting of the volcanic-island-arc ridges, the oldest parts volcanic and Honshu continental arcs (Figs. 2 and 4). The Sag-
of the Philippine Sea plate are now in both the westernmost and ami Trough is also a place of large hazardous earthquakes in the
easternmost parts of the present Philippine Sea plate (Kobayashi, historic and present ages, as typified by the Kanto earthquakes
1984, 1995; Kobayashi et al., 1995). (Shishikura, 2003; Sato et al., 2005). The relative present motion
The easternmost part of the Philippine Sea plate, the remnant of the Philippine Sea plate to the North American plate is toward
of the leading edge of the Izu forearc at the trenches, is composed N35°W at a rate of 2.5 cm/yr (Fig. 4; Seno, 1993).
of ophiolitic rocks, ranging from Cretaceous to Miocene, but is The Miura-Boso Peninsulas contain good outcrops of Mio-
dominated by Eocene boninitic rocks and younger island arc vol- cene to Pliocene Izu forearc, deep-sea, volcaniclastic deposits
canic materials (Fujioka and Sakamoto, 1999; Ishiwatari et al., that are now incorporated within accretionary prisms and covered

Figure 2. Schematic view of the model for the accretion of Izu forearc sediments to the Honshu forearc along the proto-
Sagami Trough. View from the south. A voluminous amount of Izu volcaniclastic sediments plus some trench axial chan-
nel sediments from the Honshu Arc were deposited both in the Izu forearc and the proto–Sagami Trough, finally to be
accreted to the Honshu forearc. Adapted from Hanamura and Ogawa (1993).
158 Ogawa et al.

Figure 3. Simplified trace map of the Tertiary formations on the Miura (M, west) and Boso (B, east) Peninsulas. Shaded areas represent lower
Miocene and older sedimentary rocks. Dashed doubled lines approximately trace the present plate subduction–related faults. Neogene sediments
to the north of the Mineoka Ophiolite Belt are composed of forearc basin sediments, whereas those to the south are mostly of accretionary prism
material. After Ogawa and Taniguchi (1988).

unconformably with shallow-water, coarser grained, volcanicla- Ogawa, 1998), to the Miura Group (Ogawa et al., 1989; Hanamura
stic sedimentary rocks of the forearc basin or the trench slope and Ogawa, 1993; Yamamoto and Kawakami, 2005; Yamamoto
type (Hanamura and Ogawa, 1993; Yamamoto and Kawakami, et al., 2005; Yamamoto, 2006), and finally to the Chikura Group
2005) (Figs. 2 and 3). The former Izu forearc sediments were (part of which may be forearc basin deposits) (Saito, 1992).
generally deep-water deposits, formed in several kilometers of Such accretionary prisms are highly deformed by domi-
water depth (Kitazato, 1997), mostly hemipelagic (Soh et al., nant thrusts and folds, although most of the original sedimen-
1989, 1991; Stow et al., 1998, 2002), partly contourite (Lee and tary structures are preserved (Figs. 2 and 3). The reason for the
Ogawa, 1998), and include many intercalations of turbiditic or good preservation is mostly due to the fact that the deforma-
fall-type volcaniclastic deposits (tuff and lapilli, conglomerate, tion occurred under semilithified conditions, which involved the
and sandstone) (Soh et al., 1989). All these sedimentary rocks most significant flow or shear along bedding planes and the con-
are composed of Izu Arc volcanic rocks or their derivatives. Such centration of strain in faulted or folded regions (Ogawa, 1982;
sequences are divided into three prisms from north to south and Yamamoto et al., 2005; Yamamoto, 2006). As a result, the sedi-
from older to younger, passing from the Emi Group (Hirono and ments between the faults and the deformed sections preserve
Oblique subduction in an island arc collision setting 159

35 N

140 E

Figure 4. Tectonic map of the Sagami Trough. MP—Miura Peninsula; BP—Boso Peninsula; SB—Sagami Basin; MSTB—Middle Sagami
Trough Basin; OBC—Okinoyama Banka Chain; SOT—So-oh Trough; BC—Boso Canyon; NB—North Basin; MF—Mogi Fan at the triple junc-
tion. Triangle shows the relative plate motion at present between NAM (North American plate), PHS (Philippine Sea plate), and PAC (Pacific
plate). Relative motion of the Eurasian plate took place in the early Quaternary. After Seno and Maruyama (1984). Line A is the seismic profile
shown in Figure 5, and the other lines are those of the KAIKO project, some of which are shown in Figure 6.

their original structures. Therefore, along the coast, exposures 2. Sedimentary rocks in the accretionary prisms are domi-
on the uplifted benches permit clear observation of sedimen- nated by volcaniclastic sedimentary rocks (tuff, lapilli,
tary structures, including grading, lamination, bioturbation, and and their derivatives) together with hemipelagic, muddy
other features. sedimentary rocks or contourite background deposits.
Ogawa and his students have published descriptions of 3. Sedimentary rocks in the accretionary prisms are deformed
the terrestrial geology along the Miura-Boso Peninsula coasts by many thrust faults but still preserve original sedimen-
(Hanamura and Ogawa, 1993; Yamamoto et al., 2000, 2005; tary textures and structures because of their rapid uplift
Yamamoto and Kawakami, 2005; Yamamoto, 2006). The land after accretion, although there has locally been much liq-
geology is summarized as follows: uefaction, injection, slumping, and faulting.
1. Thick forearc basin sedimentary rocks have accumulated 4. Rotations of the accretionary prisms around a vertical axis
since the early Miocene over the Mineoka Belt (in the are strong, and mostly on the NW side of the Miura-Boso
north) and in the middle Boso Peninsula, whereas far- Peninsulas and reaching on average 30°–50° clockwise
ther south several accretionary prisms have developed rotation. Locally, however, rotation has been as much as
between the Izu forearc and Honshu forearc. The lat- 80°, owing to the collision between the Izu and Honshu
ter sediments were originally deposited in the deep sea, Arcs (Kanamatsu et al., 1996). Collision has occurred
partly in the Izu forearc and partly along the boundary (6–3 Ma) in the southern Boso Peninsula between the
between the Izu and Honshu Arcs, in settings similar to accretionary prism and the forearc or the overlying trench
the present Sagami Trough (Figs. 2–4). slope sedimentation (Yamamoto and Kawakami, 2005).
160 Ogawa et al.

TOPOGRAPHIC AND GEOLOGIC SUMMARY OF THE formations, ranging in age from early Miocene to Pleistocene,
SAGAMI TROUGH are overlain by multiple unconformities. Terrestrial geology and
stratigraphy show that NW-trending folds and thrusts are more
The Sagami Trough is the present day plate boundary strongly deformed in comparison with the older formations
between the North American plate to the NE and the Philip- (Kimura, 1976; Ogawa, 1982; Ohkouchi, 1990).
pine Sea plate to the SW. The latter is now subducting obliquely Systematic younging of these formations toward the SW is
under the former at a rate of 2.5 cm/yr. The SE end of the Sagami recognized in the Miura and Boso Peninsulas, where accretionary
Trough is coincident with the Boso triple junction on the Izu- prism development toward the SW is well documented (Ogawa
Ogasawara (Bonin) trench floor (Fig. 4). The topography of the and Taniguchi, 1988; Ogawa et al., 1989). The accreted sedimen-
Sagami Trough is divided approximately into three segments, tary rocks either were originally deposited on the flanks of the Izu
from the NW in the Sagami Bay area (Sagami Basin), to the Arc or in the trench (Fig. 2) (Hanamura and Ogawa, 1993). The
middle Sagami Trough area (Middle Sagami Trough Basin), and only structural difference in the Sagami Bay area is the forma-
finally the Boso triple junction area in the SE (Fig. 4). Each area tion of a dissecting fault pattern subperpendicular to the dominant
has characteristic topography and internal structure (Nakamura NW lineament, which we suggest to be a systematic Riedel shear
et al., 1984, 1987; Pautot et al., 1987). Seismic profiling has been system, as described in the section “Riedel Shear Experiments.”
acquired, but direct observations by submersible are limited, and The northeastern part of the Sagami Bay features the
no drilling has yet been carried out. Okinoyama Bank Belt, trending from NW to SE as a trench
slope break. A nearly parallel, slightly oblique en echelon pat-
Sagami Bay Area tern of NW-SE faults and subperpendicular NE-SW faults cut the
entire trench slope area in the NE part of the basin (Fig. 6). The
The Sagami Bay area is occupied by a wide basin, the Sag- Okinoyama Bank Belt has steep cliffs along its SW boundary, and
ami Basin, which is filled with a thick wedge of trench-fill sedi- many Calyptogena (representative chemosynthetic clam) colonies
ments in the Sagami Trough (Kato et al., 1983; Kong et al., 1984; have been known at the foot of these cliffs (Kanie, 1999). These
Ohkouchi, 1990) (Fig. 5). Pleistocene to Holocene sediments en echelon faults are thought to be active and coincident with the
in the basin are not highly deformed but are thrust-bounded to Sagami Tectonic Line (Kimura, 1976) (Fig. 5), corresponding to
the NE. A fold and thrust belt has developed on the NE side, the deformation front or splay fault of the Philippine Sea plate
where Miocene sedimentary rocks are deformed. Such Neogene subduction (Kato et al., 1983; Sato et al., 2005).

Figure 5. Seismic profile of the Sagami


Basin and its interpretation along line
A in Figure 4, showing a large amount
of trench-wedge movement in the Sa-
gami Trough (Sagami Basin). The Sa-
gami Tectonic Line (S.T.L.) SW of the
Okinoyama Bank Chain (O.B.C.) is an
offscraping or splay thrust. Adapted from
Kato et al. (1983). A—Holocene sedi-
ments; B—Pleistocene; C—Pliocene;
D—middle to upper Miocene; E—lower
Miocene.
Oblique subduction in an island arc collision setting 161

Figure 6. Topographic features of the Sagami Basin and a new tectonic interpretation as the Riedel shears of each fault system (after Takami,
1999). Line A is the seismic profile shown in Figure 5. OIS—Oiso Spur; SMK—Sagami Knoll; MZK—Manazuru Knoll; MUK—Miu-
ra Knoll; MSK—Misaki Knoll; OYB—Okinoyama Bank; TKC—Tokyo Canyon; HIIs—Hatsushima Island; KMF—Kozu-Matsuda Fault;
AGC—Ashigara Canyon.

Nakamura et al. (1984) attempted to explain the en echelon Taisho-Kanto earthquake in 1923 (M = 7.9), indicate that this
pattern with subperpendicular faults by proposing a special type boundary is not undergoing eduction but subduction (Ando, 1974;
of plate boundary. These workers proposed that “eduction” is Aita, 1993; Sato et al., 2005). If so, this prompts the question of
occurring in this region, i.e., that the underlying plate (the Phil- why dissection of the topography occurs in the Sagami Bay area
ippine Sea plate) does not subduct but is instead pulled away and whether this is different from other areas. The 2D effects of
from the overriding plate (North American plate) with an exten- oblique subduction can be analyzed by a simple sandbox experi-
sional sense of motion. However, Kato et al. (1983) and Ohk- ment described in the section “Riedel Shear Experiments.”
ouchi (1990) interpreted seismic profiles to indicate SW-vergent
thrust faults in the area of the Okinoyama Bank Belt, forming Middle Sagami Trough Area
an accretionary prism (Fig. 5). Ohkouchi (1990) suggests that all
of Sagami Bay is characterized by a continuous anticline, based The middle Sagami Trough area is situated south off the
on seismic profile interpretation, and concludes that horizontal shore of the Boso Peninsula as a NW-trending longitudinal basin
compression has occurred around this area. (Middle Sagami Trough Basin in Fig. 4), demarcated by two
In addition, all the mechanistic solutions for earthquakes faults, coincident with the NE Boso Escarpment and the SW Awa
and tsunamis, and the resultant crustal movement from the Canyon (Figs. 4 and 7). The interpretation of topography and
162 Ogawa et al.

seismic profiles obtained during the KAIKO project indicates basin, oblique faults with superficial fractures are being system-
that the convex topography of the basin-fill sediments between atically developed (Fig. 8). This set of fractures may be inter-
the two faults represents a modern accretionary prism within an preted as R1 shears, as discussed in the section “Riedel Shear
oblique subduction setting (Tanahashi, 1986; Ogawa et al., 1989) Experiments.” The 3D effect of oblique subduction can be suc-
(Figs. 7 and 8). This part of the present accretionary prism is col- cessfully analyzed by a simple sandbox experiment.
liding along its NW side with the Boso Peninsula, where a sharp
thrust-fault scarp has developed in the So-oh Trough (Fig. 8). The Boso Triple Junction Area
Genroku-Kanto earthquake in 1703 (M = 8.2) occurred beneath
this scarp, just off the Boso Peninsula (Shishikura, 2003). Many The Boso triple junction area has a unique setting both in
thrust faults in en echelon fashion were developed in the southern topography and internal structure (Ogawa et al., 1989; Seno et
fault zone, whereas almost straight faults with dominant strike- al., 1989; Iwabuchi et al., 1990). This area is not only tectonically
slip motion were developed along the northern boundary of the but also topographically unstable. In a large sense the triple junc-
basin (Fig. 8). A forearc basin lies on the NE side of the accre- tion area is a large, deep triangular basin (Figs. 4 and 9), divided
tionary complex. The Sagami Trough canyon flows in a meander- into three parts, one on the NW side, where a wide basin is up to
ing fashion at the foot of the fault scarp (Boso Escarpment) along 7400 m deep (North Basin in Fig. 4); another is the trench floor at
the northeastern boundary (Boso Canyon) (Figs. 4 and 8). 9400 m water depth, and the third is surrounded by steep slopes.
The sediments in the basin were accreted from Izu forearc The first basin has some deep sea terraces along which mud vol-
basin sediments. Some are slide deposits from the increase in the canoes and dissecting channels have developed (Ogawa et al.,
slope angle, causing slope instability, as the Izu Arc is colliding 1989). Canyons connecting to land on the northern side of the
with the island of Honshu (Tokuyama et al., 1988). Within the wide basin are channeled toward this basin but show significant

30 km

(twtsec) 10 km

Figure 7. Seismic profiles of the middle Sagami Trough Basin, showing the Izu forearc sediments (right) accreted to the
left in the Sagami Trough (SGT). BOC—Boso Canyon; AWC—Awa Canyon; MYC—Miyake Canyon; twt—two-way
traveltime. The Miyake Canyon is within the Izu Arc, and the Boso and Awa Canyons are in the Sagami Trough. Adapted
from Ogawa et al. (1989). The bottom profile is from the KAIKO I Research Group (1986).
Oblique subduction in an island arc collision setting 163

140°E 141°E

55
ne
35°N

Li
Line 49

34°30'N

Normal fault
Reverse fault
Fracture

Figure 8. Topographic division of the Middle Sagami Trough Basin area, adapted from Tanahashi (1986). Thrust faults and normal faults and
fractures exhibit en echelon patterns of a Riedel shear system.

meandering (Fig. 4). The canyons appear to be deformed and surrounding three plate (Fig. 4) indicates that the present triple
dragged by the downwarping of the basin basement (Soh et al., junction is unstable, the triple point is moving to the NW, chang-
1990), as they are not directly oriented toward the deepest part ing the system of the junction. The topography (Figs. 9 and 10)
but show clockwise rotation. We suggest that older canyons were indicates such instability, resulting in the unique sedimentation,
rotated by such downwarping (Fig. 9). accretion, and resultant erosion by gravitational instability.
The main channel in the negative basin flows out through a
narrow gorge into the Izu-Ogasawara (Bonin) trench floor at a RIEDEL SHEAR EXPERIMENTS
depth of 9400 m. At this point it forms the deep-sea Mogi Fan
(Nakamura et al., 1987; Soh et al., 1988; Ogawa et al., 1989; In order to better understand the tectonics around the Sagami
Seno et al., 1989) (Figs. 4 and 9). The fan deposits were thrust Trough, we performed two types of simple sandbox tests: one in
and folded locally prior to accretion to the Izu forearc (Geologi- 2D to model Riedel shears caused by horizontal displacement,
cal Survey of Japan, 1982; Soh et al., 1988; Ogawa et al., 1989). and another in 3D to simulate oblique subduction.
Huge deep-sea submarine slide bodies are recognized (Fig. 10).
Surface topographies have many unstable slopes, shown by large- 2D Riedel Shear-Box Test
scale concave and convex cliffs and scarps by submarine sliding.
This was observed in the form of rough ridges and slides dur- For 2D tests, Bartlett et al. (1981), Naylor et al. (1986), and
ing unmanned submersible dives (KR99–11 Shipboard Scientific many others have run simple shear-box tests and recognized
Party, 2000). systematic development of Riedel shears. Unfortunately, these
In total the triple junction area is a wide, 100-km-scale tri- experiments had no free zone on the displacement basement.
angular area of very unstable topography that is still in a state We prepared a new experiment, using a rubber sheet, which
of active collapse. Because the present relative motion of the was dislocated along a displacement basement, as shown in
164 Ogawa et al.

140˚E 141˚E 142˚E 143˚E

Boso Peninsula

35˚N 35˚N

M. S. T. B.
Sa
gam
i Tr
oug
h Boso Triple Junction

34˚N 34˚N

33˚N 33˚N
140˚E 141˚E 142˚E 143˚E

-9000 -6500 -6000 -5500 -5000 -4500 -4000 -3500 -3000 -1000

Figure 9. Bathymetric map showing the topography in and around the Boso triple junction, courtesy of JAM-
STEC (JAMSTEC R/V Mirai and other data edited by M. Nakanishi, 2006, personal commun.). Strong lin-
eaments from NW to SE from the Boso Peninsula side (B) continue to the middle part of the trench slope,
gradually becoming obscure in the triangle area near the triple junction. Notice a trench floor submarine fan in
the center, which is deformed by frontal thrust. M—Miura Peninsula; B—Boso Peninsula; I—Izu Peninsula.

Figure 11 (left), relative to the superjacent powder. The differ- displacement occurred, which was followed by R2 (antithetic)
ences and characteristics of our experiment in comparison with shear with a left-lateral sense. Finally, P-shear (thrust shear) with
conventional methods are as follows: (1) A rubber sheet is used right-lateral displacement occurred (Fig. 11, right).
as a cover for the dislocating board, and the center of the rub- Theoretically R1 and R2 are associated as conjugate shear
ber sheet, 5–15 cm wide, is mobile (freely sheared) while the planes, but in fact the former is more strongly developed, and in
margins are settled by resin. (2) Quartz powder sediment, with a general R2 rarely or only seldom occurs. The P-plane is not a the-
grain size averaging 8 µm in diameter, is set on this rubber in a oretically resultant fracture but experimentally occurs as a result
layer varying from 1 to 4 cm thick. (3) The board is dislocated in of the accumulation of material to produce horizontally compres-
a right-lateral sense (dextrally) by a continuously moving motor sive pressure into a thrust-component–bearing shear plane (Nay-
at a constant speed of 2 mm/s. The strain generated by this stress lor et al., 1986).
appears as en echelon folds. The strain of the rubber sheet in the This stress-strain relation occurs instantaneously under
en echelon folds is transferred to the overlying sediment. (4) The simple shear. The en echelon folds in the rubber sheet produce
strain in the form of fractures on the powder surface during dis- a stress field in the overlying sediment that is different in each
location is recorded by camera. Because the Taisho-Kanto earth- domain of the folded rubber. As a result, the stress field makes a
quake occurred on a right-lateral (dextral) oblique-slip fault in conjugate set of shears, R1 and R2.
the Sagami Basin area, our experimental box was designed to The P-shears occur at the left-overstep part of R1, or
efficiently mimic this type of horizontal deformation effect with around the R2 shear zones. This is explained as the result of
minor vertical motions. compressional stress around these areas that gives the shears
Our experiment revealed several different results, described a thrust component. P-shear occurs slightly after R1 and R2.
as follows. First, R1 (synthetic Riedel shear) with right-lateral This is because the compressional stress increases according
Oblique subduction in an island arc collision setting 165

10 km (V. E. = 5)

Figure 10. Three-dimensional (3D) topography at the triple junction (left), showing a 10-km-scale large submarine slide body that has collapsed
onto the trench floor at 9000 m depth (KR99–11 Shipboard Scientific Party, 2000). The dive site of the remotely operated vehicle (ROV) Kaiko
is shown by an arrow, with drawings (right) by Y. Ogawa from video views from the ROV of the collapsed slope. Sediments and rocks of the
slope are composed of inclined strata of Pliocene age with minor blocks of middle Miocene age. V.E.—vertical exaggeration. Depths below sea
level shown in meters.

to the displacement of the basement. The vertical displacement indicates that the fault had some irregular concentrations of stress
of R1 and R2 is also due to the compressional stress driven by and strain.
the left-lateral (sinistral) overstep of R1. Sediment moved in Thus the main fault of the Taisho-Kanto earthquake had
the direction of both R1 and R2, after which sediment accumu- diverse subfaults. Independent preexisting faults or newly formed
lated within the wedge between the two. This uplifts the sedi- faults both moved simultaneously. The subducted-landward, over-
ment, and small-scale pressure ridges form. As displacement riding plate is thought to have deformed by ductile processes out-
increases, R1 and R2 change their form into sigmoidal geom- side the asperity region. This model leads us to think that the duc-
etries because R1 and R2 are connected. tile rubber sheet, overlain by quartz powder, effectively mimics
The 85-km-long, 45-km-wide fault plane for the Taisho- the system, with local folds being similar to the asperity regions.
Kanto earthquake fault in the Sagami Basin area (Ando, 1974;
Aita, 1993) suggests that many subordinate fault planes are Comparison with the Topography of Sagami Bay
responsible for strain. In fact, at the time of that earthquake, the
Sagami Tectonic Line corresponded to the main fault. In locali- In relation to plate rheology, Shimamoto (1989) showed that
ties far from the epicenter at the southernmost tip of the Boso in a thin plate, as in an island arc, where dislocation is dispersed
Peninsula and south of the Miura Peninsula, motion was recorded over a wide area, the plate boundary comprises a wide zone, and
on exposed active faults. These facts indicate that plate motion the island arc is deformed internally. In our study region the Sag-
has accommodated not only one fault plane but associated fault ami Basin area corresponds to the Izu Arc collision zone, and the
systems as well. Furthermore, two asperity parts have recently northern part acts as the hanging wall of the subduction zone in
been discovered for the Taisho-Kanto earthquake (Sato et al., which the North American plate undergoes complicated defor-
2005) that reveal the diversity of the slip plane. This realization mation and fracturing (dislocation) (Lallemant et al., 1996). In
166 Ogawa et al.

Figure 11. 2D Riedel shear box-test equipment with a 10-cm-wide free zone of rubber with right-lateral displacement (left), and one of the rep-
resentative results (right), where R1, R2, and P show synthetic Riedel shear, antithetic Riedel shear, and thrust shear, respectively. After Takami
(1999). D.T.—Dislocating Table; M.R.S.—Movable Rubber Sheet; U.M.R.S.—Unmovable Rubber Sheet.

addition, some convex topography reflects asperities in strain or Our experiments indicate that R2 should also occur, and the
stress concentrations (Sato et al., 2005). If such factors support direction corresponds well to the actual submarine topography.
these facts, then Riedel shears (R1 and R2) must develop as con- In order to corroborate the possibility of these fault systems,
jugate shear fractures in the superficial sediment layers above the we plotted rose diagrams to graphically represent the directions
hanging wall. We now compare examples of topography from the of canyons and faults on the landward slope of Sagami Bay, as
Sagami Bay area with our experimental results. well as all the fractures in our experiments with <3 cm disloca-
Based on the experiments, the overriding plate was predicted tion with all rubber widths (Fig. 12). It is clear that the directions
to deform by ductile shear, resulting in conjugate Riedel shears, in the two diagrams are similar. The internal angle between R1
which we discuss below. The relative plate-movement direction and R2 in the experiments is ~10° smaller than the actual topog-
along the Sagami Trough indicates that the Sagami Tectonic Line raphy, but this may be due to the rotation of topography, as shown
has oblique dextral dislocation (Fig. 4). If we assume that the in experiments discussed below.
strike of the Taisho-Kanto earthquake fault was N59°W (Aita, It was not easy to infer the horizontal displacement sense of
1993), and this is considered to be the Y-shear (main shear) direc- the topography just from bathymetric maps, but the faults around
tion, then the Sagami Tectonic Line must be comparable to R1. the Okinoyama Bank Chain are straightly aligned, suggesting
Oblique subduction in an island arc collision setting 167

a tectonic origin owing to high-angle, horizontal-displacement


faults (Fig. 6). Seismic profiles indicate that the western bound-
ary of the Miura Knoll (part of the Okinoyama Bank Chain) has
a very steep dip (Fig. 5), suggesting horizontal displacement on
a fault. As for the vertical displacement of the Sagami Tectonic
Line, the evidence of the Taisho-Kanto earthquake shows that the
Okinoyama Bank Chain was uplifted by a thrust component.
The experiments are associated with small displacement, but
the actual topography indicates uniform positive relief only on the
eastern side of the Sagami Tectonic Line. The difference between
the model and the observation may be due to our neglecting
the vertical displacement of the Taisho-Kanto earthquake. The
effect where the basement plate has a thrust-component–bearing
oblique fault that affects the overlying sediment was modeled in
3D by Naylor et al. (1986)’s sandbox experiments (Mandl, 1988).
This result shows that the basement dislocation may undergo R1
strain, with a thrust-faulting effect within an en echelon arrange-
ment. Naylor et al.’s experiment predicts uplift along the eastern
side of the Sagami Tectonic Line as a “flower structure.”
The comparison of experiment and topography suggests
that, although some of the submarine fault displacement is not
known in terms of the sense of motion, the present topography of
the Sagami Bay correlates well with the experiment in terms of
the dominant fault and fracture directions. In addition, our study
indicates that the submarine topography is mostly influenced by
a regional dextral shear field.

3D Oblique Subduction Test

As 3D experimental tests are rare, and oblique subduction


models in particular had not previously been performed, we made
a simple but effective test using a steel frame and stiff paper board
(Takazawa, 2003). Wheat powder with colored-chalk layers and
chocolate-powder mesh on the surface was prepared over the sub-
duction boundary in order to mimic the trench wedge sediments.
The inclined board (sloping 30° downward) was moved steadily
by hand at variable degrees of obliqueness. Fractures on the sur-
face and faults and folds inside were shown by cutting the model
with a knife after the movement. For each different degree of
obliqueness (0°, parallel to the trench, to 90° in 10° steps) the sur-
face fracture development and internal structures were recorded Figure 12. Rose diagrams of Riedel shear fractures of sandbox test
by photograph. Representative results are shown in Figure 13. with right-lateral displacement (above; see Fig. 11) and topographic
lineament of the Sagami Basin (below; see Fig. 6). Each dominant di-
Results of 3D Tests rection indicates synthetic Riedel shears and antithetic Riedel shears of
right-lateral displacement, respectively. After Takami (1999).

Figure 13 shows the most representative case, in which the


obliqueness angle was 30°. Surface fractures developed system-
atically. First, Riedel R1 shears formed, then locally R2 shears the ocean side formed at a later stage. In most cases, one or two
were seen, but these were relatively rare. Another direction of back thrusts developed as the deformation proceeded, as part of
shear, interpreted as thrust shears (P-shears) formed at a later a flower structure.
stage. Internally within the wedge, thrusts and their associated This simple experiment clearly indicates that in cases of
folds systematically developed (Fig. 13, left). In some cases greater obliqueness (especially <40°), stronger Riedel shears
the internal thrusts developed during the initial stage, which is develop, but the effect of thrust and fold development is more
similar to the accretionary prism model, because the thrust on or less the same as in the case of smaller obliqueness (close to
168 Ogawa et al.

Figure 13. Results of 3D oblique subduction tests of right-lateral sense with 30° obliquity. Profile (left), and top surface
fractures (right). Notice components of fractures on the surface of the oblique subduction; synthetic Riedel shears and
thrust shears are distinct. Compare the experiment with the real examples from the Middle Sagami Trough Basin in Fig-
ure 8. After Takazawa (2003).

normal convergence). However, in the less oblique experiment, surface deformation driven by a dextral shear zone that is trig-
with convergence at more than 70°, no strong P-shears formed. gered by highly oblique subduction. More detailed constraints
The best fit to the Middle Sagami Trough Basin area in are needed to verify a direct link between the topography, geol-
terms of the surface fracture system is for 20°–30° of oblique- ogy, and tectonics. We suggest that direct observation and sam-
ness (Fig. 13). This corresponds to the natural case in the area, pling at the seabed, supplemented by drilling and dredging,
where measured obliqueness is ~30° (Figs. 4 and 8). In the So-oh are needed to understand the triple junction structure in three
Trough, between the Middle Sagami Trough area and the Sagami dimensions. Only by doing this can our proposed model be rig-
Bay area (NW corner in Fig. 8), just in front of the Boso Pen- orously tested.
insula, obliqueness is almost normal, and no Riedel shears are The different appearance of shear patterns in the Sagami
observed on modern bathymetric maps. Therefore, we conclude Bay area and the Middle Sagami Basin area may be attributed
that even a simple 3D oblique subduction model is successful in to the different relationships between the basement and the
mimicking the natural features of the seafloor (Takazawa, 2003). overlying sediment cover. In the Sagami Basin a wide and thick
accretionary wedge above an irregular basement (modeled with
CONCLUSIONS folds of the rubber sheet) reflects strain-asperity zones and is
deformed by strain partitioning. As a result, both R1 and R2
In this study we verified the origin of the fault system and shears are formed continuously. In contrast, in the case of the
related topography of the Sagami Trough and used experimen- Middle Sagami Trough Basin the strain is transferred in a more
tal modeling to understand the development of Riedel shears in regular fashion to the sediments from the subduction, forming
oblique subduction settings within an island arc collision zone in a regular shear pattern. Further quantitative analysis by more
a TTT-type triple junction. In experiments, well-defined shears sophisticated methods is needed to more fully understand this
associated with horizontal displacement appeared both in 2D and complicated region, but the simple experiments described here
3D tests. The orientation of faults along the Sagami Trough, both offer a plausible solution.
in the northwestern part (Sagami Bay area) and in the middle
part (Middle Sagami Trough Basin area) can be explained as the ACKNOWLEDGMENTS
dominant two directions oblique to the main fault, which cor-
relate well with R1 and R2 shears in sandbox experiments. The We are grateful to Peter Clift and Amy Draut for inviting us to
Sagami Tectonic Line can be understood as a plate-boundary the 2006 Penrose Conference in Price, Utah. Peter Clift, Roland
thrust fault, where an oblique fault has a horizontal displacement von Huene, and one anonymous reviewer kindly read the early
with a thrust component. version of the manuscript, and gave us critical suggestions for
These results indicate that the fault system and topography revision. Some bathymetric charts and figures were prepared
can be attributed to Riedel shears, which generate wide regional through the courtesy of JAMSTEC and Masao Nakanishi.
Oblique subduction in an island arc collision setting 169

Kobayashi, K., 1995, Role of subducted lithospheric slab in uplift and subsid-
REFERENCES CITED ence of the northwestern margins: Marine Geology, v. 127, p. 119–144,
doi: 10.1016/0025-3227(95)00007-L.
Aita, I., 1993, Historical tsunamis in NW Sagami Bay and its numerical model: Kobayashi, K., Kasuga, S., and Okino, K., 1995, Shikoku Basin and its mar-
Journal of Geography, Tokyo Geographical Society, v. 102, p. 427–436. gins, in Taylor, B., ed., Back-Arc Basins in the Western Pacific: New York,
Ando, M., 1974, Seismo-tectonics of the 1923 Kanto Earthquake: Journal of Plenum, p. 381–405.
Physics of the Earth, v. 22, p. 263–277. Kong, Y.S., Kagami, H., Tokuyama, H., Igarashi, C., and Nasu, N., 1984, Sub-
Aoike, K., 1999, Tectonic evolution of the Izu collision zone: Research Report bottom geological structures of Sagami Bay: University of Tokyo, Bul-
of Kanagawa Prefectural Museum of Natural History, v. 3, p. 111–151. letin of Ocean Research Institute, v. 16, p. 1–91.
Bartlett, W.L., Freedman, M., and Logan, J.M., 1981, Experimental folding and KR99–11 Shipboard Scientific Party, 2000, Onboard Report of KR99–11
faulting of rocks under confining pressure: Tectonophysics, v. 79, p. 255– Cruise, JAMSTEC.
277, doi: 10.1016/0040-1951(81)90116-5. Lallemant, S.J., Le Pichon, X., Thoue, F., Henry, P., and Saito, S., 1996, Shear
Fujioka, K., and Sakamoto, I., 1999, Structure of the Izu-Bonin Arc: Research partitioning near the central Japan triple junction: The 1923 great Kanto
Report of Kanagawa Prefectural Museum of Natural History, v. 3, p. 5–20. earthquake revisited—I: Geophysical Journal International, v. 126,
Geological Survey of Japan, 1982, Geological Map of the Northern Ogasawara p. 871–881.
Arc: Marine Geology Map Series 18. Lee, I.T., and Ogawa, Y., 1998, Bottom-current deposits in the Miocene-Pliocene
Hall, R., Ali, J.R., Anderson, C.D., and Baker, S.J., 1995, Origin and motion Misaki Formation, Izu forearc area, Japan: Island Arc, v. 7, p. 315–329.
history of the Philippine Sea plate: Tectonophysics, v. 251, p. 229–250, Lus, W.Y., McDougall, I., and Davies, H.L., 2004, Age of the metamorphic sole
doi: 10.1016/0040-1951(95)00038-0. of the Papuan Ultramafic Belt ophiolite, Papua New Guinea: Tectono-
Hanamura, Y., and Ogawa, Y., 1993, Layer-parallel faults, duplexes, imbricate physics, v. 392, p. 85–101, doi: 10.1016/j.tecto.2004.04.009.
thrust and vein structures of the Miura Group: Key to understanding the Mandl, G., 1988, Mechanics of Tectonic Faulting: New York, Elsevier Science,
Izu fore-arc sediment accretion to the Honshu forearc: Island Arc, v. 3, p. 141–144.
p. 126–141. Maruyama, S., and Seno, T., 1986, Orogeny and relative plate motions: Exam-
Hirano, N., Ogawa, Y., Saito, K., Yoshida, T., Sato, H., and Taniguchi, H., 2003, ple of the Japanese Islands: Tectonophysics, v. 127, p. 305–329, doi:
Multi-stage evolution of the Tertiary Mineoka Ophiolite, Japan: New geo- 10.1016/0040-1951(86)90067-3.
chemical and age constraints, in Dilek, Y., and Robinson, P., eds., Ophio- Matsuda, T., 1978, Collision of the Izu-Bonin arc with central Honshu: Ceno-
lite in Earth History: Geological Society [London] Special Publication zoic tectonics of the fossa magna, Japan: Journal of Physics of the Earth,
218, p. 279–298. v. 26, Supplement, p. S409–S421.
Hirono, T., and Ogawa, Y., 1998, Duplex arrays and thickening of accretionary McKenzie, D.P., and Morgan, W.J., 1969, Evolution of triple junction: Nature,
prisms—An example from Boso Peninsula, Japan: Geology, v. 26, p. 779– v. 224, p. 125–133, doi: 10.1038/224125a0.
782, doi: 10.1130/0091-7613(1998)026<0779:DAATOA>2.3.CO;2. Mohiuddin, M.M., and Ogawa, Y., 1998, Late Paleocene to middle Miocene
Huchon, P., and Labaume, P., 1989, Central Japan triple junction: A three- pelagic sequences in the Boso Peninsula, Japan: New light on NW Pacific
dimensional compression model: Tectonophysics, v. 160, p. 117–133, tectonics: Island Arc, v. 7, p. 301–314.
doi: 10.1016/0040-1951(89)90387-9. Mori, R., and Ogawa, Y., 2005, Transpressional tectonics of the Mineoka Ophiolite
Ishiwatari, A., Yanagida, Y., Li, Y.-B., Ishii, T., Haraguchi, S., Koizumi, K., Ich- Belt in a trench–trench–trench-type triple junction, Boso Peninsula, Japan:
iyama, Y., and Umekawa, M., 2006, Dredge petrology of the boninite- and Island Arc, v. 14, p. 571–581, doi: 10.1111/j.1440-1738.2005.00485.x.
adakite-bearing Hahajima Seamount of the Ogasawara (Bonin) forearc: Nakamura, K., Shimazaki, N., and Yonekura, A., 1984, Subduction, bending
An ophiolite or a serpentinite seamount?: Island Arc, v. 15, p. 102–118, and eduction—Present and Quaternary tectonics of the northern border
doi: 10.1111/j.1440-1738.2006.00512.x. of the Philippine Sea plate: Bulletin of Société Géologie de France, v. 26,
Isozaki, Y., 1996, Anatomy and genesis of subduction-related orogen: A new p. 221–243.
view of subdivision and evolution of the Japanese Islands: Island Arc, Nakamura, K., et al., 1987, Oblique and near collision subduction, Sagami
v. 5, p. 289–320, doi: 10.1111/j.1440-1738.1996.tb00033.x. and Suruga Troughs—Preliminary results of the French-Japanese Kaiko
Iwabuchi, Y., Asada, A., and Kato, Y., 1990, Multichannel seismic profiling of cruise, Leg 2: Earth and Planetary Science Letters, v. 83, p. 229–242, doi:
the Boso triple junction: Marine Survey and Technology, v. 2, p. 29–38. 10.1016/0012-821X(87)90068-9.
KAIKO I Research Group, 1986, Topography and Structure of Trenches around Naylor, M.A., Mandl, G., and Supersteinjn, C.H.K., 1986, Fault geometries
Japan—Data Atlas of Franco-Japanese KAIKO Project, Phase I: Ocean in basement-induced wrench faulting under different initial stress states:
Research Institute, University of Tokyo, Institut Français de Recherches Journal of Structural Geology, v. 8, p. 737–752, doi: 10.1016/0191-
pour l’Exploitation de la Mer (IFREMER) and Centre National de la 8141(86)90022-2.
Recherche Scientifique (CNRS), 305 p. Niitsuma, N., 1988, Neogene tectonic evolution of southwest Japan: Modern
Kanamatsu, T., Herroro-Bervera, E., Taira, A., Saito, S., Ashi, J., and Furumoto, Geology, v. 12, p. 497–532.
A.S., 1996, Magnetic fabric development in the Tertiary accretionary Ogawa, Y., 1982, Tectonics of some forearc fold belts in and around the arc-arc
complex in the Boso and Miura Peninsulas of central Japan: Geophysical crossing area in central Japan, in Legget, J.K., ed., Trench-Forearc Geol-
Research Letters, v. 23, p. 471–474, doi: 10.1029/96GL00147. ogy: Geological Society [London] Special Publication 10, p. 49–61.
Kanie, Y., 1999, Cenozoic System and geological structure of the Miura Penin- Ogawa, Y., 1983, Mineoka ophiolite belt in the Izu forearc area—Neogene
sula and Tokyo Bay to Boso Peninsula, southern-central Japan: Research accretion of oceanic and island arc assemblages in the northeastern cor-
Report of Kanagawa Prefecture Museum of Natural History, v. 9, p. 79– ner of the Philippine Sea plate, in Hashimoto, M., and Uyeda, S., eds.,
94. Accretion Tectonics in the Circum-Pacific Region: Tokyo, Terrapub,
Kato, S., Sato, Y., and Sakurai, M., 1983, Multichannel seismic survey in the p. 245–260.
Nankai: Suruga and Sagami Troughs: Report of Hydrographic Research, Ogawa, Y., 1985, Variety of subduction and accretion processes in Cretaceous
v. 18, p. 1–23. to Recent plate boundaries around southwest and central Japan: Tectono-
Kimura, M., 1976, Marine geology in the Sagami-Nada Sea and its vicinity: physics, v. 112, p. 493–518, doi: 10.1016/0040-1951(85)90192-1.
Geological Survey of Japan, Marine Geological Map Series 3, 9 p. Ogawa, Y., and Horiuchi, K., 1978, Two types of accretionary fold belts in cen-
Kinoshita, H., 1980, Paleomagnetism of sediment cores from Deep Sea Drilling tral Japan: Journal of Physics of the Earth, v. 26, Suppl., S321–336.
Project, Leg 58, Philippine Sea, in Klein, G.DeV., Kobayashi, K., et al., Ogawa, Y., and Sashida, K., 2005, Lower Cretaceous radiolarian bedded chert
Initial Reports of the Deep Sea Drilling Project, v. 58: Washington, D.C., from the Mineoka Belt, Boso Peninsula, Japan: Journal of the Geological
Government Printing Office, p. 765–768. Society of Japan, v. 111, p. 624–627.
Kitazato, H., 1997, Paleogeographic changes in central Honshu, Japan, during Ogawa, Y., and Takahashi, A., 2004, Seafloor spreading, obduction and triple
the late Cenozoic in relation to the collision of the Izu-Ogasawara Arc junction tectonics of the Mineoka ophiolite, Central Japan: Tectonophys-
with the Honshu Arc: Island Arc, v. 6, p. 144–157, doi: 10.1111/j.1440- ics, v. 392, p. 131–141, doi: 10.1016/j.tecto.2004.08.007.
1738.1997.tb00166.x. Ogawa, Y., and Taniguchi, H., 1987, Ophiolitic melange in forearc area and
Kobayashi, K., 1984, Subsidence of the Shikoku back-arc basin: Tectonophys- formation of the Mineoka belt: Science Report, Kyushu University: Geol-
ics, v. 102, p. 105–117, doi: 10.1016/0040-1951(84)90010-6. ogy, v. 15, p. 1–23.
170 Ogawa et al.

Ogawa, Y., and Taniguchi, H., 1988, Geology and tectonics of the Miura-Boso Soh, W., Taira, A., Ogawa, Y., Taniguchi, H., Pickering, K.T., and Stow, D.A.V.,
Peninsulas and the adjacent area: Modern Geology, v. 12, p. 147–168. 1989, Submarine depositional processes for volcaniclastic sediments in
Ogawa, Y., Horiuchi, K., Taniguchi, H., and Naka, J., 1985, Collision of the the Mio–Pliocene Misaki Formation, Miura Group, Central Japan, in
Izu arc with Honshu and the effects of oblique subduction in the Miura- Taira, A., and Masuda, F., eds., Sedimentary Facies in the Active Plate
Boso Peninsulas: Tectonophysics, v. 119, p. 349–379, doi: 10.1016/0040- Margin: Tokyo, Terrapub, p. 619–630.
1951(85)90046-0. Soh, W., Tokuyama, H., Fujioka, K., Kato, S., and Taira, A., 1990, Morphology
Ogawa, Y., Seno, T., Tokuyama, H., Akiyoshi, H., Fujioka, K., and Taniguchi, and development of a deep-sea meandering canyon (Boso Canyon) on
H., 1989, Structure and development of the Sagami Trough and off-Boso an active plate margin, Sagami Trough, Japan: Marine Geology, v. 91,
triple junction: Tectonophysics, v. 160, p. 135–150, doi: 10.1016/0040- p. 227–241, doi: 10.1016/0025-3227(90)90038-L.
1951(89)90388-0. Soh, W., Pickering, K.T., Taira, A., and Tokuyama, H., 1991, Basin evolution
Ogawa, Y., Tamaki, K., Isozaki, Y., and Shikazono, N., 1997, Japan, in Moores, in the arc–arc Izu Collision Zone, Mio–Pliocene Miura Group, Central
E.M., and Fairbridge, R.W., eds., Encyclopedia of European and Asian Japan: Geological Society [London] Journal, v. 148, p. 317–330.
Regional Geology: London, Chapman and Hall, p. 436–449. Stow, D.A.V., Taira, A., Ogawa, Y., Soh, W., Taniguchi, H., and Pickering, K.T.,
Ohkouchi, N., 1990, Active structure and tectonics of the Sagami Bay: Journal 1998, Volcaniclastic sediments, processes interaction and depositional
of Geography, Tokyo Geographical Society, v. 99, p. 458–470. setting of the Mio-Pliocene Miura Group, SE Japan: Sedimentary Geol-
Okino, K., Shimakawa, Y., and Nagaoka, S., 1994, Evolution of the Shikoku ogy, v. 115, p. 351–381, doi: 10.1016/S0037-0738(97)00100-0.
Basin: Journal of Geomagnetics and Geoelectronics, v. 46, p. 463–479. Stow, D.A.V., Ogawa, Y., Lee, In Tae, and Mitsuzawa, K., 2002, Neogene
Pautot, G., et al., 1987, Deep-sea submersible survey in the Suruga, Sagami contourites, Miura-Boso forearc basin, SE Japan, in Stow, D.A.V., et al.,
and Japan Trenches: Preliminary results of the 1985 Kaiko cruise, Leg 2: eds., Deep-Water Contourite Systems: Modern Drifts and Ancient Series,
Earth and Planetary Science Letters, v. 83, p. 300–312, doi: 10.1016/0012- Seismic and Sedimentary Characteristics: Geological Society [London]
821X(87)90073-2. Memoirs, v. 22, p. 409–419.
Renard, V., et al., 1987, Trench triple junction off central Japan—Preliminary Taira, A., Tokuyama, H., and Soh, W., 1989, Accretion tectonics and evolution
results of French–Japanese 1984 Kaiko cruise, Leg 2: Earth and Planetary of Japan, in Ben-Avraham, Z., ed., The Evolution of the Pacific Ocean
Science Letters, v. 83, p. 243–256, doi: 10.1016/0012-821X(87)90069-0. Margins: Oxford, UK, Oxford University Press, p. 100–132.
Saito, S., 1992, Stratigraphy of Cenozoic strata in the southern terminus area of Takahashi, M., and Saito, K., 1997, Miocene intra-arc bending at an arc–arc
Boso Peninsula, Central Japan: Contributions of the Institute of Geology collision zone, central Japan: Island Arc, v. 6, p. 168–182, doi: 10.1111/
and Paleontology, Tohoku University, v. 93, p. 1–37. j.1440-1738.1997.tb00168.x.
Sato, H., and Ogawa, Y., 2000, Sulfide minerals in peridotites as tectonic indi- Takahashi, A., Ogawa, Y., Ohta, Y., and Hirano, N., 2003, The nature of faulting
cators for genesis of ophiolitic rocks: Example from peridotite in the and deformation in the Mineoka Ophiolite, NW Pacific Rim, in Dilek, Y.,
Hayama-Mineoka Belt, central Japan, in Dilek, Y., Moores, E., Ethon, D., and Robinson, P., eds., Ophiolite in Earth History: Geological Society
and Nichols, A., eds., Ophiolites and Oceanic Crust: Geological Society [London] Special Publication 218, p. 299–314.
of America Special Paper 349, p. 427–437. Takami, Y., 1999, Tectonic interpretation of submarine topography in the Sag-
Sato, H., Taniguchi, H., Takahashi, N., Mohiuddin, M.M., Hirano, N., and ami Bay based on Riedel shear experiments [B.S. thesis]: Tsukuba, Japan,
Ogawa, Y., 1999, Origin of the Mineoka Ophiolite: Journal of Geography, University of Tsukuba, 51 p.
Tokyo Geographical Society, v. 108, p. 203–215. Takazawa, S., 2003, The submarine topography and tectonics in the middle of
Sato, H., Hirata, N., Koketsu, K., Okaya, D., Abe, S., Kobayashi, R., Matsub- Sagami Trough based on the subduction model experiment [B.S. thesis]:
ara, M., Iwasaki, T., Ito, T., Ikawa, T., Kawanaka, T., Kasahara, K., and Tsukuba, Japan, University of Tsukuba, 60 p.
Harder, S., 2005, Earthquake source fault beneath Tokyo: Science, v. 309, Tanahashi, M., 1986, Structures of sediments and tectonics of Sagami Trough:
p. 462–464, doi: 10.1126/science.1110489. Earth Monthly, v. 8, p. 238–245.
Seno, T., 1984, Was there a North New Guinea plate?: Geological Survey of Tokuyama, H., et al., 1988, Multi-channel seismic profile in the south of Izu-
Japan Report, v. 263, p. 29–42. Oshima Island: Kazan (Volcano), 2nd Ser., v. 33, p. 67–77.
Seno, T., and Maruyama, S., 1984, Paleogeographic reconstruction and origin Yamamoto, Y., 2006, Systematic variation of shear-induced physical properties
of Philippine Sea: Tectonophysics, v. 102, p. 53–84, doi: 10.1016/0040- and fabrics in the Miura–Boso accretionary prism: The earliest processes
1951(84)90008-8. during off-scraping: Earth and Planetary Science Letters, v. 244, p. 270–
Seno, T., Ogawa, Y., Tokuyama, H., Nishiyama, E., and Taira, A., 1989, Tec- 284, doi: 10.1016/j.epsl.2006.01.049.
tonic evolution of the triple junction off central Honshu for the past 1 Yamamoto, Y., and Kawakami, S., 2005, Rapid tectonics of the Late Miocene
million years: Tectonophysics, v. 160, p. 91–116, doi: 10.1016/0040- Boso accretionary prism related to the Izu–Bonin arc collision: Island
1951(89)90386-7. Arc, v. 14, p. 178–198, doi: 10.1111/j.1440-1738.2005.00463.x.
Seno, T., Stein, S., and Gripp, A.E., 1993, A model for the motion of the Philip- Yamamoto, Y., Ohta, Y., and Ogawa, Y., 2000, Implication for the two-stage
pine Sea plate consistence with NUVEL-1 and geological data: Journal of layer-parallel faults in the context of Izu forearc collision zone—Exam-
Geophysical Research, v. 89, p. 17941–17948. ples from the Miura accretionary prism, central Japan: Tectonophysics, v.
Shimamoto, T., 1989, Rheology of rocks and plate tectonics: Kagaku, v. 59, 325, p. 133–144, doi: 10.1016/S0040-1951(00)00134-7.
p. 170–181. Yamamoto, Y., Mukoyoshi, H., and Ogawa, Y., 2005, Large and small struc-
Shishikura, M., 2003, Cycle of interpolate earthquake along the Sagami Trough, tural characteristics in shallow burial accretionary prism on land: Rapidly
deduced from tectonic geomorphology: Bulletin of Earthquake Research uplifted Neogene accreted sediments in the Miura-Boso Peninsula: Tec-
Institute: University of Tokyo, v. 78, p. 245–254. tonics, v. 24, p. TC5008, doi: 10.1029/2005TC001823.
Soh, W., Taira, A., and Tokuyama, H., 1988, A trench fan in the Izu-Ogasawara
Trench on the Boso Trench triple junction, Japan: Marine Geology, v. 82,
p. 235–249, doi: 10.1016/0025-3227(88)90143-0. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

The West Crocker formation of northwest Borneo: A Paleogene accretionary prism

Joseph J. Lambiase*
Tan Yaw Tzong*
Amelia G. William*
Department of Petroleum Geoscience, Universiti Brunei Darussalam, Tungku, Brunei BE 1410

Michael D. Bidgood
Patrice Brenac
GSS International, Aberdeenshire, UK

Andrew B. Cullen
Sabah Shell Sendirian Berhad, Miri, Malaysia

ABSTRACT

An integrated structural, stratigraphic, and sedimentological analysis of the


West Crocker formation in northwest Borneo suggests that it is best interpreted as an
accretionary prism. The structural geology provides clear evidence of at least two epi-
sodes of syndepositional folding and thrust faulting. A probable Eocene age, indicated
by foraminiferal and palynological assemblages, differs from the generally accepted
Oligocene to early Miocene age and is consistent with deposition of the West Crocker
formation during a phase of tectonism at the northwest Borneo margin. Sandstones
within the West Crocker formation were deposited by high-density turbidity currents
that constructed relatively small, progradational lobes in a slope apron environment,
and trace fossil assemblages confirm bathyal water depths of ~1000 m or more. The
composition of the sandstones, which contain abundant feldspars and lithic fragments,
suggests that their provenance was the first-cycle product of an eroded orogenic belt,
whereas immature textures indicate a short distance of transport.

Keywords: turbidites, syntectonic, accretionary prism, Borneo.

INTRODUCTION Formation (e.g., Hutchison et al., 2000; Fig. 1A). The Crocker-
Temburong Formation has long been recognized as a thick suc-
The West Crocker formation is an informal name for distin- cession of turbidites with northward-directed paleocurrents that
guishing the western and northwestern outcrops of a Paleogene was deposited in response to collisional tectonics associated with
succession in northwest Borneo that constitutes the sand-rich Late Cretaceous to early Miocene rifting in the South China Sea
Crocker Formation and its coeval shaly equivalent, the Temburong (Stauffer, 1967; Wilson, 1964; Tan and Lamy, 1990; Hutchison,
*Present address, Lambiase: Lambiase Geoscience Pte. Ltd., Singapore; e-mail: joe_lambiase@yahoo.com. Present address, Tzong: Sabah Shell Sendirian Berhad,
Miri, Malaysia. Present address, William: ExxonMobil Exploration and Production Malaysia Inc., Kuala Lumpur, Malaysia.

Lambiase, J.J., Tzong, T.Y., William, A.G., Bidgood, M.D., Brenac, P., and Cullen, A.B., 2008, The West Crocker formation of northwest Borneo: A Paleogene
accretionary prism, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological
Society of America Special Paper 436, p. 171–184, doi: 10.1130/2008.2436(08). For permission to copy, contact editing@geosociety.org. ©2008 The Geological
Society of America. All rights reserved.

171
172 Lambiase et al.

the Crocker-Temburong Formation, and unconformities are sus-


pected if not proven. Also, the structural style and burial history
of the various parts of the formation differ significantly, suggest-
ing that the unit comprises a number of successive fan systems,
of which the West Crocker formation is the youngest and least
deeply buried (Hutchison, 2005).
The West Crocker formation lies directly below the Deep
Regional Unconformity of Levell (1987), a profound angular
unconformity that was generated during an orogenic event that
uplifted Paleogene and older rocks, and marks the transition from
that uplift to the rapid subsidence that accompanied deposition of
the thick (up to 14 km), overlying Neogene shallow-marine clas-
tic succession (Fig. 2). The Deep Regional Unconformity is gen-
erally regarded as being of late early Miocene age (e.g., Sandal,
1996), although it is diachronous, and some workers prefer a mid-
dle Miocene age (e.g., Hutchison et al., 2000). In the Luconia and
Balingian provinces of Sarawak a regional unconformity of simi-
lar age separates cycle II and cycle III (Madon, 1999; Madon and
Abolins, 1999) and generally marks the transition from extension
to regional subsidence. There has been no direct age dating of the
West Crocker formation, because the few microfossils that have
been recovered are not age-diagnostic within the Eocene to Mio-
cene (Hutchison, 2005). However, the West Crocker formation is
regarded as a sandy facies that is stratigraphically equivalent to the
shaly Temburong Formation (Fig. 2). The Temburong Formation
is the youngest part of the Crocker Formation-Temburong Forma-
tion complex and was dated as early Oligocene to early Miocene
by Wilson (1964), using foraminifera. Therefore, the West Crocker
formation is presumed to lie directly below the Deep Regional
Unconformity that marks the boundary with the overlying Neo-
gene shallow-marine clastic sedimentary rocks (Fig. 2).
It is generally agreed that the turbidites were deposited in a
foreland basin that developed in response to uplift of the Rajang
Group accretionary prism during the Sabah orogeny that began
in the late Eocene (ca. 36 Ma; e.g., Tongkul, 1990; Hutchison,
1996), although the paleogeography of northwest Borneo and
the exact timing of events during the early Tertiary are poorly
constrained. The thickness of the West Crocker succession is
uncertain partly because postdepositional thrust faulting makes
stratigraphic reconstruction between outcrops nearly impossible;
Wilson (1964) suggested that at least 6000 m of section is pres-
ent, whereas Tongkul (1987) argues for ~1000 m.
Figure 1. (A) Location of Borneo within southeast Asia and the distri-
By the late Oligocene the West Crocker formation was
bution of the West Crocker and Temburong formations, and the Rajang
Group (undifferentiated Crocker and Trusmadi Formations). (B) Loca- being thrust over an area of buried continental shelf and attenu-
tion of the Bukit Melinsung (BM), Jalan Salaiman (JS), and Maju (MJ) ated continental crust called Dangerous Grounds (Hutchison,
outcrops; all are along main roads that are shown as dashed lines. 2004; Hutchison et al., 2000). The SW-directed underthrusting of
Dangerous Grounds during subduction generated NW-SE com-
pression that was responsible for the complex structures of the
Eocene to middle Miocene sedimentary rocks in western Sabah,
1996). Various authors have interpreted the unit as one major including the West Crocker strata (Tongkul, 1990, 1991). Apa-
turbidite fan system derived from a relatively distant southerly tite fission-track data indicate that the West Crocker sedimentary
source (e.g., Crevello, 2002; Tongkul, 1987). However, Hutchison rocks were buried to 4–8 km and were exhumed during develop-
(2005) notes that it is difficult to envision a single fan system per- ment of the middle Miocene regional unconformity (Hutchison
sisting through the entire Paleogene to early Miocene age span of et al., 2000).
West Crocker formation of northwest Borneo 173

Figure 2. Early Tertiary stratigraphy of


northwest Borneo, with suggested revisions
to the scheme of Hutchison et al. (2000).

Present Study consist of very coarse to coarse grained, 1–6-m-thick sandstones


interbedded with thinner shales in a 130-m-thick, continuously
This study investigated the sedimentology, stratigraphic exposed vertical succession.
architecture, and structural style of the West Crocker formation The Maju outcrop was selected because it covers a large area
turbidites in order to develop a tectono-stratigraphic model for and has a wide variety of structures in >800 m of exposed sec-
their deposition. Three outcrops were selected for detailed study: tion, representing >400 m of stratigraphic succession (Fig. 1B).
Bukit Melinsung, Jalan Salaiman, and Maju (Fig. 1B). The first The succession has been deformed by multiple folding and fault-
two were chosen because they are much less deformed than ing events that have generated thrust faults, normal faults, and a
most. One of them has the thickest continuously exposed verti- wide range of fold types.
cal succession, whereas the other represents the most continu-
ous lateral exposure. Approximately 350 m of vertical section is SEDIMENTOLOGY AND STRATIGRAPHY
exposed continuously in the Bukit Melinsung outcrop (Fig. 1B),
although lateral continuity is limited to ~50 m. About 70% of the The outcrops were logged to define grain size trends, sedi-
succession is very fine to granular sandstone, and the remaining mentary structures, the nature of bed contacts, trace fossil dis-
30% is shale. Individual beds in the Jalan Salaiman outcrop near tribution, and stacking patterns. Pollen and foraminifer assem-
Kota Kinabalu (Fig. 1B) can be traced laterally for ~500 m. They blages were determined from the interbedded shales for the
174 Lambiase et al.

interpretation of age and depositional environment, respectively, individual beds generally thicken and coarsen upward. Their
and the sandstones were analyzed petrographically to evaluate basal contacts are straight and sharp, and relatively few are
gross composition. scoured. Flute casts, although not common, indicate a northerly
paleoflow direction when corrected for structural dip (Fig. 3B).
Sedimentary Facies Load structures are moderately common.
Within each unit the lower sandstone beds are relatively thin
The outcrops comprise three facies, which are designated (0.3–1.0 m) and generally grade upward from coarse or very
as thick sandstones, thin interbedded sandstones and shale, coarse sand to fine sand, laminated silt, and finally shale. Most
and shale. beds are massive at the base and pass upward into parallel lami-
nation (Fig. 3C), although some have convoluted bedding that
Thick Sandstones passes upward into parallel lamination. Sandstone beds 1–6 m
Thick sandstones generally occur as amalgamated beds, thick, rarely up to 10 m thick, occur near the top of the amalgam-
although a few are isolated between thin shales. Amalgam- ated units, although Hutchison (2005) reports that 10-m-thick
ated units generally consist of 4–25 sandstone beds with a total beds are more common in a few other outcrops. The thick sand-
thickness of 6–33 m (Fig. 3A). Within the amalgamated units, stones generally are coarse to very coarse sand, massive, and

Figure 3. (A) Amalgamated sandstones in the Jalan Salaiman outcrop. (B) Flute casts at the base of a bed in the Maju outcrop. (C) Typical West
Crocker formation turbidite sandstone, with a scoured base and a sharp upper contact with the overlying shale. The sandstone is massive, with
very coarse sand and granules near the base, and grades into parallel laminated, medium sand. The unit comprises Bouma divisions Ta, Tb, Td,
and Te. (D) Climbing ripples in the siltstone cap of a turbidite sandstone.
West Crocker formation of northwest Borneo 175

structureless. Grain size distribution is uniform throughout but identified are Ophiomorpha, Thalassinoides, Terebellina, Plano-
grades into medium sand a few centimeters from the top of the lites, Paleophycus that are 0.3–0.5 cm in diameter, Asterosoma,
bed. Parallel lamination and, less commonly, contorted lamina- and Rhizocorallium. Ophiomorpha range from 1.0 to 2.5 cm in
tion occur locally. The uppermost parts of the massive sand bod- diameter and are up to 1.3 m long. Tajul Anuar Jamaluddin (1989)
ies locally have thin laminations of carbonaceous material and identified a similar trace fossil assemblage, plus Crossopodia, in
are rippled or wavy-bedded. other West Crocker formation outcrops.
Siltstones and shales that cap the sandstones are usually lam- Trace fossils are rare within most of the shales and are mostly
inated and contain abundant carbonaceous material; isolated rip- Helminthoida and Megagrapton of the deep-water Nereites ich-
ple cross-stratification and climbing ripples occur locally in thin nofacies (Pemberton et al., 1992) as well as Planolites (?) that are
sandy interbeds (Fig. 3D). Generally, the sandstones in the amal- 0.3–0.5 cm in diameter. However, the apparent low abundance
gamated units have flat, straight, and sharp upper contacts; the of trace fossils in the finer grained facies is clearly a product of
few that appear to have gradational upper contacts and scoured intense, tropical surface weathering. A recently excavated area of
basal contacts generally are in the lower part of the unit. the Jalan Salaiman outcrop revealed several bedding plane sur-
faces covered with abundant, high relief Megagrapton, Nereites,
Thin Interbedded Sandstones and Shales and Ophimorpha rudis traces that are indicative of bathyal water
The interbedded sandstone and shale facies comprises depths (Fig. 4; Pemberton et al., 1992). When the site was revis-
monotonous alternations of 4–8-cm-thick sandstones and shale ited 6 weeks later, the traces were barely visible.
within a 0.5–1.5-m-thick unit. The sandstone layers generally are
fine- to medium-grained sand. However, some beds grade upward Petrography
from coarse- to medium-grained sand to parallel-laminated sand
and silt; some layers have discrete and climbing ripples or paral- The modal composition of 36 sandstones was analyzed in
lel lamination, commonly with a high content of carbonaceous thin section. They generally contain 40%–60% quartz, 15%–25%
material. Most of the sandstones have sharp basal and top con- feldspar, and 10%–15% clay matrix (Fig. 5A). The feldspar frac-
tacts and are capped by thin, homogeneous and laminated shale tion is mostly microcline and orthoclase, with a few occurrences
layers, although a few have gradational contacts. of plagioclase. Lithic rock fragments account for 5%–20% of
most samples but reach nearly 30% in a few examples. Thus, the
Shale sandstones are mostly subarkoses with a few sublitharenites and
Shale units are homogeneous and range from 0.5 to 5.0 m in litharenites (Fig. 5B; Pettijohn et al., 1987). Twenty of the samples
thickness. The lower parts of the shale units usually contain thin plot within the field that defines a continental block as the source
interbeds of fine to silty sand that exhibit straight, sharp contacts of sediment, whereas the other 16 appear to have been derived
and ripple cross-lamination. from a recycled orogen (Fig. 5B; Dickinson and Suczek, 1979).
The overall distribution is consistent with sandstones derived
Trace Fossils and Microfossils from older sediment that was the first-cycle product of an eroded
orogenic belt, in this case almost certainly the Rajang Group.
Microfossils The sandstones are texturally immature, as all samples are
Microfossil analysis yielded minimal information about poorly sorted and most grains are angular to subangular, although
the age and depositional setting of the West Crocker forma- subrounded grains are moderately common. The matrix is made
tion because the shale samples contain only a few agglutinating of very fine grains and clay minerals; much of the clay fraction
foraminifera (1–6 individuals per sample). All are either Haplo- comprises authigenic clays formed during diagenesis or surface
phragmoides spp. or Recurvoides spp., except for one possible weathering. Similarly, most of the feldspars in the framework con-
Trochammina sp., that are normally indicative of a dysaerobic, stituents have been replaced by sericite, kaolinite, muscovite, and
deep-water setting and an undifferentiated Tertiary age. a few unidentifiable authigenic clay minerals. The petrographic
Pollen recovery generally was poor, with many of the paly- analysis indicates that the sandstones were originally very imma-
nomorphs consisting of reworked Cretaceous forms or those with ture and contained a considerable amount of feldspar and deposi-
an undifferentiated Tertiary age such as Dicolpopollis spp. How- tional matrix before diagenesis and surface weathering.
ever, the presence of Inaperturopollenites spp., Liquidambar
spp., Tiliaepollenites spp., and several Palmae taxa suggests an Stratigraphy
early Eocene to early Oligocene age.
Generally, the three sedimentary facies stack into coarsening-
Trace Fossils upward parasequences that range in thickness from 25 to 65 m.
A relatively diverse assemblage of trace fossils occurs in the The basal shale in each parasequence is overlain by interbedded
sandstones of the three outcrops, but the number of beds with sand and shale, followed by an amalgamated sandstone unit at the
visible traces is low, and each locality tends to have a low diver- top (Fig. 6). The sandstone beds within the amalgamated units
sity, and some are even monospecific. The traces that have been thicken and coarsen upward, which is a typical progradational
176 Lambiase et al.

Figure 5. (A) Thin section of a typical poorly sorted, texturally imma-


ture West Crocker formation sandstone with some component grains
labeled as quartz (Q), polycrystalline quartz (PQ), microcline (M), and
feldspar (F). (B) Gross composition of selected West Crocker forma-
tion sandstones, based on the classification scheme of Pettijohn et al.
(1987). The provenance fields for recycled orogens and continental
blocks are from Dickinson and Suczek (1979).

stacking pattern (Ricci-Lucchi and Valmori, 1980; Normark,


1978; Macdonald, 1986; Walker, 1992).
Bukit Melinsung is the only outcrop with a vertical succes-
sion that is relatively undeformed and thick enough to discern
stratigraphic trends among parasequences. There, the lower four
parasequences thicken upward, with a corresponding increase in
the proportion of sandstone, indicating that they comprise a sin-
gle progradational parasequence set (Fig. 7). The parasequence
set is incomplete, because a relatively large fault forms its upper
Figure 4. Trace fossils in the shale beds of the West Crocker formation boundary. The upper four parasequences are much sandier than
belong to the bathyal Nereites ichnofacies (Pemberton et al., 1992). They
the lower four. However, they also generally thicken and coarsen
include (A) Megagrapton, (B) Nereites, and (C) Ophiomorpha rudis.
upward and make up part of another progradational parasequence
set (Fig. 7).
West Crocker formation of northwest Borneo 177

Figure 6. Facies stacking pattern of a parasequence in the Jalan Salaiman


outcrop. Relative grain size is schematic and not to scale.

STRUCTURAL STYLE

All of the West Crocker formation outcrops show some


degree of structural deformation. Most are cut by at least sev-
eral, mostly thrust, faults and are folded to some degree but
are not intensely deformed. However, the Maju outcrop cov- Figure 7. Generalized sequence stratigraphy
ers a large area and exposes the temporal relationships among of the Bukit Melinsung outcrop, revised from
the various structures and thus offers a unique opportunity to Mohd. Khalid Jamiran (1999). Note that the
analyze the structural development of the West Crocker forma- thicker sandstones are units composed of stacked
tion. The outcrop consists of three stratigraphic successions; two beds that thicken and coarsen upward within the
unit, whereas individual beds fine upward. Sev-
are separated by a thrust fault across which the offset cannot be eral faults and folds within the succession are not
determined because it is more than the exposed length of the shown; a relatively large thrust fault forms the
fault plane, and the other boundary is stratigraphic (Fig. 8). The upper boundary of the lower parasequence set.
oldest of the three successions forms the hanging-wall of the Relative grain size is schematic and not to scale.
Figure 8. Schematic structural cross-section of the Maju outcrop, showing the major stratigraphic subdivisions. Note that the hanging-wall
of the major thrust fault is more highly deformed than the footwall. Locations of the photographs in Figures 9 and 10 are as indicated.
After Tan (2005).
West Crocker formation of northwest Borneo 179

large thrust fault, and the younger two successions are in the
footwall. The hanging-wall succession has undergone at least
three phases of deformation, of which the older two are com-
pressional and the third extensional, whereas the footwall suc-
cessions were affected only by the second compressional event
and the extensional phase.

Hanging-Wall Succession

Approximately 70 m of the hanging-wall succession is


exposed laterally in near-dip and near-strike surfaces that are a
total of 175 m long and range from 8 to 12 m high. The succes-
sion includes all three of the sedimentary facies described previ-
ously. The first phase of deformation was generated by NE-SW
compression that created two thrust faults that verge southwest
and a relatively large fold that was later refolded during a sub-
sequent NW-SE compressional event. The refolded fold is a key
structure for determining the relative timing of the deformational
events and is bounded by faults to the west and south. It is a
tight recumbent, refolded fold with two fold axes, the older of
which reflects the first, NE-SW compressional event (Fig. 9A).
The rock mass in the core of the tight fold comprises thick, dis-
rupted, overturned sandstone beds, and the remainder is brec-
cia. Structures formed during the first compressional event are
restricted to the hanging-wall succession, suggesting that the
deformation took place before deposition of the footwall suc-
cessions and that these structures were synsedimentary at the
formation time scale.
The second, NW-SE phase of compression generated the
second fold axis in the refolded fold and a series of high-angle,
NW-verging thrust faults that cut across the folded beds, plus
a number of other thrust faults and folds. Parasitic folds are
commonly associated with the larger folds. Generally, thicker,
competent sandstone beds are gently folded, whereas shale-
rich intervals have well-developed detachment folds, indicat- Figure 9. Structural features in the hanging-wall and lower footwall
ing a close relationship between lithology and fold geometry successions of the Maju outcrop. See Figure 8 for locations. (A) Part
that is pervasive in the outcrop. Angular, 5–30-cm-diameter of the large refolded fold in the hanging-wall. The apparent anticline
sandstone clasts within a sheared shale matrix form an inco- on the left side of the figure is actually an overturned syncline; the
hesive, foliated fault breccia on some of the faults. Part of the younging direction is toward the bottom of the photo. (B) Highly de-
formed succession in the hanging-wall that is interpreted as a slump;
hanging-wall succession is characterized by complex folds and the younging direction is variable and locally indeterminable. Some of
sandstone blocks within sheared shale that are interpreted as a the sandstones are outlined with a white line for easier visualization
slump (Fig. 9B). Several small-scale (1–5 m), isolated folds are of the structure. (C) Parasitic folds in a shale bed above gently folded
within the slump. sandstones that form an anticline in the lower footwall succession. The
A few minor normal faults with minimal displacement were younging direction is toward the top of the photo.
generated by the third and final phase of deformation. All the
young normal faults across the entire outcrop have small dis-
placements and are viewed as the product of postorogenic gravi- surface. All three sedimentary facies are present within the succes-
tational collapse after compression rather than the reflection of a sion that was deformed by the second, NW-SE–oriented compres-
significant, extensional tectonic event (Tan, 2005). sional event and later extensional faulting. Structures generated
by NW-SE compression dominate the outcrop and especially the
Lower Footwall Succession footwall successions. Indeed, the entire outcrop can be interpreted
as part of the southeast limb of a large, disharmonic anticline with
The lower footwall succession is 75 m thick and covers an three other folds representing relatively large-scale parasitic folds
area 8 to 10 m high and ~300 m laterally on a near-strike exposure on the anticline (Tan, 2005).
180 Lambiase et al.

The lower footwall succession is a gently folded anticline


within a thick sandstone-dominated succession. There are a num-
ber of parasitic folds that are restricted to shale-rich intervals up
to 7 m thick, and a series of thrust faults that cut the succession
(Fig. 9C). The thrust faults are more abundant in the NW limb
than in the SE limb; this may indicate that the compressional
stress was NW to SE directed, which coincides with the South
China Sea spreading direction (Tan and Lamy, 1990). This sug-
gests that at least some of the deformation in the West Crocker
formation may be the product of regional tectonics, and further
implies that the West Crocker formation may be age-equivalent
to the early Oligocene to early Miocene (32–16 Ma) South China
Sea spreading (Briais et al., 1993). Some of the thrust faults
exhibit drag folding, and others preferentially follow shale-rich
bedding planes. The extensional deformation phase is repre-
sented by several minor normal faults that cut the succession.

Upper Footwall Succession

The ~275 m of upper footwall succession strata also includes


all three sedimentary facies and is exposed in near-strike and
near-dip surfaces that total ~550 m in length and are 8 to 35 m
high. The basal unit is a thick sandstone that directly overlies
the gently folded anticline of the lower footwall succession. The
basal unit also has been folded, but the sandstone beds pinch
out against the flank of the underlying fold, creating an onlap
geometry with a dip difference of 11° across the onlap surface
(Fig. 10A). The onlap angle far exceeds depositional dip in simi-
lar turbidite systems (Kumar and Slatt, 1984), which indicates
that the underlying fold had developed, or was growing, when
the overlying sandstones were deposited, and is clear evidence
of synsedimentary deformation. Dips of up to 12° or 13° are
common on growing structures in unconsolidated sediments in
modern turbidite systems (Morley, 2007), so there is no need to
invoke burial and/or consolidation before folding of the underly-
ing strata, or a significant amount of time, between folding and
deposition of the onlapping sandstones.
The lower part of the upper footwall succession dips con-
sistently at 52°–60° and is cut by numerous thrust faults that
were generated by the NW-SE compressional event, some of
which form a complex fault zone. Near the top of the succession
is a large-scale fold with a NW-SE–striking fold axis in which
dips increase upsection to the point at which the beds eventu-
ally become overturned. The folding caused a distinct thickness
change in a shale unit that is detached from an overlying sandstone Figure 10. Structural features in the upper footwall
bed in which there is no thickness change; this demonstrates the succession of the Maju outcrop. See Figure 8 for lo-
competency of the sandstones relative to the shales (Fig. 10B). cations. (A) Onlap relationship between the upper and
There are also associated small-scale detachment folds within an lower footwall successions. A white line highlights the
overturned thin interbedded sandstone and shale unit (Fig. 10C). onlap surface. (B) Abrupt lateral change in shale thick-
ness adjacent to a fold detachment, as indicated by the
The fold is cut by several thrust faults, some of which are east white arrows. The younging direction is to the right of
verging and some of which have associated breccia. A number of the photograph. (C) Small-scale detachment fold in the
normal faults that represent the final, extensional phase of defor- overturned beds associated with a larger scale anticline.
mation cut the succession and the older structures. The arrow indicates the younging direction.
West Crocker formation of northwest Borneo 181

TECTONO-STRATIGRAPHIC MODEL Age and Tectonic Setting

Depositional Model Age of the West Crocker Formation


The age of the West Crocker formation is poorly constrained
The sedimentary structures found in the West Crocker for- because few age-diagnostic microfossils have been recovered.
mation indicate that the sandstones were deposited by turbid- Many workers assign an Oligocene to early Miocene age to the
ity currents, as noted in previous studies (e.g., Stauffer, 1967; formation (e.g., Hutchison et al., 2000; Hutchison, 2004; Tong-
Tongkul, 1990). Most beds contain partial Bouma sequences kul, 1990, 1991), although some acknowledge that it may be as
with divisions Ta, Tb, and Td; Ta and Td; Ta and Tc; Ta and old as Eocene (e.g., Rangin et al., 1990; van Hattum et al., 2003).
Te; or just Ta. These Ta–Tc dominant, massive, structureless Pollen data from this study suggest that the three outcrops are
but graded amalgamated sandstone beds with sharp and straight Eocene deposits, although the age of the entire West Crocker for-
upper and lower contacts are typical of high-density turbidites mation could be variable.
(e.g., Stow et al., 1996; Lowe, 1982; Walker, 1992; Stow and A preliminary palynological analysis, using identical ana-
Johansson, 2000). The generally structureless and massive tur- lytical techniques, also indicates an Ecoene age for the uncon-
bidite sandstones indicate rapid deposition, probably by col- formably overlying Melingan Formation in Sabah (Fig. 2). It is
lapse fallout from high-density turbidity currents (Stow and important to note that in Sabah the Melingan Formation consists
Johansson, 2000) caused by an abrupt decrease in slope angle of steeply dipping, high-energy storm beds that are markedly
along the transport path. Parallel lamination and climbing rip- different from the gently dipping, low-energy, wave-dominated
ples near the top of the sandstones and in the shale caps are the sandstones with minor limestones in Sarawak and Brunei that
consequence of rapid and continuous fallout of finer grains dur- were dated as early Miocene by Wilson (1964). This suggests
ing the final stages of deposition. The sharp bed contacts are the that either the Melingan Formation has a broader age range than
product of sediment bypassing and sequential sedimentation of previously recognized (Fig. 2) or that the Sabah and Sarawak-
different particle populations from turbulent high-density flows Brunei outcrops belong to different stratigraphic units.
transporting very coarse to fine grains in suspension (Lowe, Assuming that the stratigraphic relationship between the
1982; Stow et al., 1996). West Crocker and Melingan formations as defined by previous
The very immature texture, angularity, and poor sorting of workers (e.g., Liechti et al., 1960; Wilson, 1964; Tate, 1974) and
the sandstones suggest a nearby sediment source and a short illustrated in Figure 2 is accurate, then the West Crocker forma-
transport distance, as does the presence of angular to subangular tion probably is Eocene to Oligocene rather than Oligocene to
feldspars that are easily destroyed by extensive transport (Caro- lower Miocene, and the stratigraphy should be revised as sug-
zzi, 1993; Harwood, 1988). Many of the trace fossils belong to gested in Figure 2. This view indicates that most of the West
the Cruziana ichnofacies that is indicative of a shelfal environ- Crocker formation is not a sandy facies equivalent of the shaly
ment with water depths normally <200 m (Pemberton et al., Temburong Formation, which resolves the problematic paleo-
1992; Frey and Pemberton, 1984). However, those in the shale, geographic facies distribution whereby the sandy fans of the West
Ophiomorpha rudis, Nereites, Helminthoida, and Megagrapton, Crocker formation lie N-NE of the Temburong Formation shales
are deep-water traces of the Nereites ichnofacies (Pemberton et in a more distal position.
al., 1992). In such cases, the assemblage in the shales is the most
reliable water depth indicator, confirming that the West Crocker Tectonic Setting
turbidites probably were deposited in >1000 m of water (G. Pem- Several aspects of the structural evolution of the Maju out-
berton, 2004, personal commun.). crop indicate that at least part of the West Crocker formation
All the sedimentary characteristics are most consistent underwent synsedimentary deformation, which differs from the
with deposition on a slope apron (Macdonald, 1993; Reading traditional view that all the deformation is postdepositional (e.g.,
and Richards, 1994). Sands were deposited rapidly at the toe of Tongkul, 1990, 1991; Hutchison, 1996, 2005). The foremost
the slope, where they formed small-scale lobes; the relatively aspect is that two different phases of deformation affected the
thin sandstone beds (nearly all are <6 m, and most are <2 m hanging-wall succession, whereas only one of them deformed
thick) suggest that the depositional lobes were not large but the footwall. The onlap relationship between the lower and upper
had radii in the range of 1–5 km or less (Stow and Johansson, footwall successions (Fig. 8) not only confirms synsedimentary
2000). This results in a surface slope on the lobes of <1° even deformation but suggests that compressional tectonics was ongo-
for the thickest parasequence of 65 m, which corresponds well ing while most, if not all, of the units were deposited. Therefore,
with other sand-rich slope aprons, e.g., the middle Tonkawa the compressional deformation of the deep-water turbidites is
Sandstone of the Anadarko Basin in the U.S. Midcontinent older than the uplift of the West Crocker formation.
(Kumar and Slatt, 1984). The coarsening and thickening It is generally accepted that early Tertiary compression
upward nature of most of the parasequences suggests that the was related to subduction along the northwest Borneo margin,
lobes were progradational. coupled with seafloor spreading related to the opening of the
182 Lambiase et al.

South China Sea (e.g., Tongkul, 1990; Hutchison, 2005). In lithic fragments than van Hattum et al. (2003) report. This sup-
Tongkul’s (1990) model, NW-SE compression was generated ports their observation that the West Crocker formation is com-
by SE-directed underthrusting of the Dangerous Grounds dur- positionally less mature than the rest of the Crocker Formation
ing late Oligocene (ca. 25 Ma) subduction. Both phases of com- and also suggests that an easterly provenance area was relatively
pressional deformation must have been completed by the time more important. Similarly, William et al. (2003) conclude that
that subduction ceased; many workers think that this occurred the West Crocker formation was derived from nearby, uplifted
in the middle Oligocene (ca. 30 Ma; e.g., Tongkul, 1990), Rajang Group sediments because of the composition and imma-
although some believe that it persisted to the early middle Mio- ture texture, although they prefer an easterly provenance area
cene (Hutchison, 1996). Structures generated by NW-SE com- based on petrographic analysis, regional sedimentary systems,
pression are abundant throughout the West Cocker formation and stratigraphic relationships.
(Stauffer, 1967; Tongkul, 1990, 1991; Hutchison, 1996, 2005), Synsedimentary deformation is more compatible with an
but the older, NE-SW compressional phase has not been docu- Eocene rather than with an Oligocene to early Miocene age for
mented previously. the West Crocker formation. The younger age would require that
The compressional tectonic setting, synsedimentary defor- strong compression, and probably subduction, continued into the
mation, and slope apron depositional model suggest that the West early Miocene (ca. 20 Ma), whereas many workers (e.g., Tong-
Crocker formation is an accretionary prism. Sediments were sup- kul, 1990) contend that subduction ended in the middle Oligo-
plied episodically from uplifted Rajang Group strata in the active cene (ca. 30 Ma). An Eocene age also provides ample time for
orogen and transported down a relatively steep slope to the adja- the estimated 4–8 km of burial that the West Crocker under-
cent basin. They accumulated at the base of the slope in a series went before it was uplifted and exposed in the middle Miocene
of small fans that were deformed by at least two phases of synde- (Hutchison, 2005); the younger age necessitates that both the
formational, compressional folding and thrust faulting. burial and uplift occurred in the early middle Miocene and were,
therefore, extremely rapid.
DISCUSSION The West Crocker formation lies below a regional middle
Miocene unconformity beneath which the Oligocene to early
The probable Eocene accretionary prism interpretation for Miocene geology is characterized by extension in the Natuna Sea
the West Crocker formation not only provides the best reconcili- (Darman and Sidi, 2000), Sarawak (Mat-Zin, 1997; Madon, 1999),
ation of the structural data and sedimentological data collected and the Dangerous Grounds off the shore of western Sabah (Mil-
in the present study, but it overcomes many of the problems and som et al., 1997). In offshore western Sabah there is pre–middle
inconsistencies of previous interpretations. One of the impor- Miocene extension, and many late Miocene anticlines cored with
tant sedimentological issues is the size and type of fan system middle Miocene sedimentary rocks are interpreted as an inversion
that the West Crocker formation represents. Many workers inter- of older structures (Bol and van Hoorn, 1980; Madon et al., 1999).
pret the West Crocker formation as a large submarine fan that The problem posed by development of a major thrust belt syn-
was fed from a point source to the SW, relatively far from the chronous with regional extension has been answered previously
West Crocker depocenter (e.g., Stauffer, 1967; Tongkul, 1987; by having extension in the South China Sea induce subduction
Hutchison, 1996). This argument is partially based on paleogeo- beneath Borneo, thereby causing the Sabah orogeny (Hutchison,
graphic reconstructions that suggest that marine waters covered 1996; Hall, 2002). An Eocene age for the West Crocker forma-
present-day Sabah prior to the uplift that produced the Deep tion eliminates this problem and calls into question models for the
Regional Unconformity (e.g., Hall, 2002), thereby precluding region’s early Neogene tectonic evolution.
a sediment source to the east. However, the thin sandstone beds
and small lobes of the West Crocker formation are too small CONCLUSIONS
for a fan associated with a relatively large river (Reading and
Richards, 1994). An integrated analysis of the structural geology and sedimen-
Another important issue is the provenance area for the West tology of the West Crocker formation concludes that the abun-
Crocker formation sediment. Van Hattum et al. (2003) concluded dant sandstone beds were deposited from high-density turbidity
that the textural immaturity of the West Crocker formation reflects currents and formed small, prograding lobes on a deep-water
a sediment source from nearby, first-cycle orogenic sandstones, slope apron. The composition and immature texture of the sands
possibly the uplifted Rajang Group in Sarawak. These precursor necessitates a short transport distance, suggesting that uplifted
sandstones probably were the erosional products of granites in Rajang Group sediment was the source of the West Crocker
the Schwaner Mountains in southern Borneo (Fig. 1A). Van Hat- formation. Micropaleontological data and stratigraphic relation-
tum et al. (2003) also attribute the subordinate, ophiolite-derived ships indicate that at least this part of the West Crocker formation
material in the West Crocker formation (Stauffer, 1967) to a sec- is Eocene rather than Oligocene to early Miocene. Nearly all of
ondary provenance area to the east. the structural deformation is compressional and synsedimentary;
The samples analyzed in the present study contain consider- the West Crocker formation is a syntectonic deposit, probably an
ably more feldspar (15%–25% as opposed to 0%–13%) and more accretionary prism.
West Crocker formation of northwest Borneo 183

ACKNOWLEDGMENTS Macdonald, D.I.M., 1993, Controls on sedimentation at convergent plate mar-


gins: International Association of Sedimentologists Special Publication
20, p. 225–257.
The authors wish to thank George Pemberton for assistance with Madon, M., 1999, Geological setting of Sarawak, in Petroleum Geology and
trace fossil identifications and Nurhayati and Rhodora Tiamsing Resources of Malaysia: Kuala Lumpur, Petronas, p. 273–290.
for help with drafting. Mohd. Khalid Jamiran contributed to the Madon, M., and Abolins, P., 1999, Balingian Province, in Petroleum Geology
and Resources of Malaysia: Kuala Lumpur, Petronas, p. 343–367.
interpretation of stratigraphic architecture, and the manuscript Madon, M., Leong, K.M., and Anuar, A., 1999, Sabah Basin, in Petroleum Geol-
benefited greatly from discussions with Stefan Back and com- ogy and Resources of Malaysia: Kuala Lumpur, Petronas, p. 501–542.
ments by reviewers Charles Hutchison, Peter Clift, and an anony- Mat-Zin, I.C., 1997, Tectonics, evolution and sedimentation of the Sarawak
Basin: Geological Society of Malaysia Bulletin, v. 41, p. 41–52.
mous reviewer. Amelia William and Tan Yaw Tzong wish to thank Milsom, J., Holt, R., Ayub, D.B., and Smail, R., 1997, Gravity anomalies and
ExxonMobil Exploration and Production Malaysia and Universiti deep structural controls at the Sabah-Palawan margin, South China Sea,
Brunei Darussalam, respectively, for financial support. in Petroleum Geology of Southeast Asia: Geological Society [London]
Special Publication 126, p. 417–427.
Mohd, Khalid Jamiran, 1999, Sedimentology and sequence stratigraphy of
REFERENCES CITED Bukit Melinsung, Papar, Sabah [M.S. thesis]: Universiti Brunei Darus-
salam, 50 p.
Morley, C.K., 2007, Growth of folds in a deepwater setting: Basin Research
Bol, A.J., and van Hoorn, B., 1980, Structural styles in western Sabah offshore: (in press).
Geological Society of Malaysia Bulletin, v. 12, p. 1–16. Normark, W.R., 1978, Fan valleys, channels, and depositional lobes on mod-
Briais, A., Patriat, P., and Tapponnier, P., 1993, Updated interpretation of mag- ern submarine fans; characters for recognition of sandy turbidite envi-
netic anomalies and seafloor spreading stages in the South China Sea: ronments: American Association of Petroleum Geologists Bulletin, v. 62,
Implications for the Tertiary tectonics of Southeast Asia: Journal of Geo- p. 912–931.
physical Research, v. 98, p. 6299–6328. Pemberton, S.G., MacEachern, J.A., and Frey, R.W., 1992, Trace fossil facies
Carozzi, A.V., 1993, Sedimentary Petrography: Englewood Cliffs, NJ, Prentice models: Environmental and allostratigraphic significance, in Walker, R.G.,
Hall, 263 p. and James, N.P., eds., Facies Models: Response to Sea Level Change,
Crevello, P.D., 2002, The great Crocker submarine fan: A world class foredeep Geotext 1: Geological Association of Canada, p. 47–72.
turbidite system: Jakarta, 28th Indonesia Petroleum Association Proceed- Pettijohn, F.J., Potter, P.E., and Siever, R., 1987, Sand and Sandstone (2nd edi-
ings, v. 1, p. 377–407. tion): New York, Springer-Verlag, 571 p.
Darman, H., and Sidi, F.H., 2000, An Outline of the Geology of Indonesia: Rangin, C., Bellon, H., Bernard, F., Letouzey, J., Muller, C., and Sanudin, T., 1990,
Jakarta, Indonesian Association of Geologists, 192 p. Neogene arc-continent collision in Sabah, Northern Borneo, Malaysia:
Dickinson, W.R., and Suczek, C.A., 1979, Plate setting and sandstone com- Tectonophysics, v. 183, p. 305–319, doi: 10.1016/0040-1951(90)90423-6.
position: American Association of Petroleum Geologists Bulletin, v. 63, Reading, H.G., and Richards, M., 1994, Turbidite systems in deep-water basin
p. 2164–2182. margins classified by grain size and feeder system: American Association
Frey, R.W., and Pemberton, S.G., 1984, Trace fossil facies models, in Walker, of Petroleum Geologists Bulletin, v. 78, p. 792–822.
R.G., ed., Facies Models (2nd edition): Geological Association of Can- Ricci Lucchi, F., and Valmori, E., 1980, Basin-wide turbidites in a Miocene,
ada, p. 189–207. over-supplied deep-sea plain: A geometrical analysis: Sedimentology,
Hall, R., 2002, Cenozoic geological and plate tectonic evolution of SE Asia and v. 27, p. 241–270, doi: 10.1111/j.1365-3091.1980.tb01177.x.
Sandal, S.T., ed., 1996, The Geology and Hydrocarbon Resources of Negara
the SW Pacific: Computer-based reconstructions, model and animations:
Brunei Darussalam: Bandar Seri Begawan, Syabas, 243 p.
Journal of Asian Earth Sciences, v. 20, p. 353–431, doi: 10.1016/S1367-
Stauffer, P.H., 1967, Studies in the Crocker Formation, Sabah, in Stauffer, P.H.,
9120(01)00069-4.
ed., Geological Papers 1966: Geological Survey of the Borneo Region of
Harwood, G., 1988, Microscopic techniques: II. Principles of sedimentary
Malaysia, Bulletin 8, p. 1–13.
petrography, in Tucker, M., ed., Techniques in Sedimentology: Oxford,
Stow, D.A.V., and Johansson, M., 2000, Deep-water massive sands: Nature, ori-
UK, Blackwell Science, p. 108–173.
gin and hydrocarbon implications: Marine and Petroleum Geology, v. 17,
Hutchison, C.S., 1996, Geological evolution of South East Asia: Kuala Lumpur,
p. 145–174, doi: 10.1016/S0264-8172(99)00051-3.
Geological Society of Malaysia, 350 p. Stow, D.A.V., Reading, H.G., and Collinson, J.D., 1996, Deep seas, in Reading,
Hutchison, C.S., 2004, Marginal basin evolution: The southern South China H.G., ed., Sedimentary Environments and Facies (3rd edition): Oxford,
Sea: Marine and Petroleum Geology, v. 21, p. 1129–1148, doi: 10.1016/ UK, Blackwell Science, p. 395–453.
j.marpetgeo.2004.07.002. Tajul Anuar Jamaluddin, 1989, Struktur sedimen dalam Formasi Crocker di
Hutchison, C.S., 2005, Geology of Northwest Borneo: Amsterdam, Elsevier, kawasan Tamparuli, Sabah: Geological Society of Malaysia Bulletin,
421 p. v. 24, p. 135–157.
Hutchison, C.S., Bergman, S.C., Swauger, D.A., and Graves, J.E., 2000, A Tan, D.N.K., and Lamy, J.M., 1990, Tectonic evolution of the NW Sabah con-
Miocene collisional belt in north Borneo: Uplift mechanism and isostatic tinental margin since the Late Eocene: Geological Society of Malaysia
adjustment quantified by thermochronology: Geological Society [Lon- Bulletin, v. 27, p. 241–260.
don] Journal, v. 157, p. 783–793. Tan, Y.T., 2005, Structural geology of the Maju outcrop, West Crocker For-
Kumar, N., and Slatt, R.M., 1984, Submarine-fan and slope facies of Tonkawa mation, Kota Kinabalu, Sabah, Malaysia [M.S. thesis]: Universiti Brunei
(Missourian–Virgilian) sandstone in deep Anadarko Basin: American Darussalam, 72 p.
Association of Petroleum Geologists Bulletin, v. 68, p. 1839–1856. Tate, R.B., 1974, Palaeo-environmental studies in Brunei: Brunei Museum
Levell, B.K., 1987, The nature and significance of regional unconformities in Journal, v. 3, p. 285–305.
the hydrocarbon-bearing Neogene sequences offshore West Sabah: Geo- Tongkul, F., 1987, The sedimentology and structure of the West Crocker For-
logical Society of Malaysia Bulletin, v. 21, p. 55–90. mation in the Kota Kinabalu area, Sabah, in GEOSEA IV Proceedings:
Liechti, P., Roe, F.N., Haile, N.S., and Kirk, H.J.C., 1960, The Geology of Sar- Jakarta, Indonesia Association of Geologists, p. 135–156.
awak, Brunei and the Western Part of North Borneo: British Borneo Geo- Tongkul, F., 1990, Structural style and tectonics of Western and Northern Sabah:
logical Survey Bulletin, v. 3, 360 p. Geological Society of Malaysia Bulletin, v. 27, p. 227–239.
Lowe, D.R., 1982, Sediment gravity flows II. Depositional models with spe- Tongkul, F., 1991, Tectonic evolution of Sabah, Malaysia: Journal of South-
cial reference to the deposits of high-density turbidity currents: Journal of east Asian Earth Sciences, v. 6, p. 395–405, doi: 10.1016/0743-
Sedimentary Petrology, v. 52, p. 279–297. 9547(91)90084-B.
Macdonald, D.I.M., 1986, Proximal to distal sedimentological variation in a van Hattum, M.W.A., Hall, R., and Nichols, G.J., 2003, Provenance of northern
linear turbidite trough: Implications for the fan model: Sedimentology, Borneo sediments: Jakarta, 29th Indonesia Petroleum Association Pro-
v. 33, p. 243–259, doi: 10.1111/j.1365-3091.1986.tb00534.x. ceedings, v. 1, paper IPA03-G-016.
184 Lambiase et al.

Walker, R.G., 1992, Turbidites and submarine fans, in Walker, R.G., and James, Wilson, R.A.M., 1964, The geology and mineral resources of the Labuan and
N.P., eds., Facies Models: Response to Sea Level Change, Geotext 1: Padas Valley area, Sabah, Malaysia: Geological Survey of the Borneo
Geological Association of Canada, p. 239–263. Region of Malaysia, Memoir 17.
William, A.G., Lambiase, J.J., Back, S., and Mohd, Khalid Jamiran, 2003, Sedi-
mentology of the Jalan Salaiman and Bukit Melinsung outcrops, western
Sabah: Is the Crocker Formation an analogue for Neogene turbidites off-
shore?: Geological Society of Malaysia Bulletin, v. 47, p. 63–75. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Temporal changes in the composition of Miocene sandstone related to collision between


the Honshu and Izu Arcs, central Japan

Koichi Okuzawa*
Research Center for Deep Geological Environments, National Institute of Advanced Industrial Science and Technology,
Tsukuba 305-8567, Japan

Ken-ichiro Hisada
Graduate School of Life and Environmental Sciences, University of Tsukuba, Tsukuba 305-8572, Japan

ABSTRACT

The Izu Arc has been colliding with the Honshu Arc in central Japan since
ca. 15 Ma. In order to understand the provenance changes related to this collision,
we studied lower to middle Miocene sandstones in and around the collision zone by
analyzing their framework composition and the chemistries of detrital clinopyroxene,
garnet, and chromian spinel.
Sandstone deposited in the trench and forearc basin of the Honshu Arc prior
to collision includes grains of detrital garnet and chromian spinel, which originated
mainly from granites and low pressure-temperature (P-T) metamorphic rocks, and
forearc peridotite, respectively, parts of the Honshu Arc. The forearc and trench-fill
sandstones differ in terms of their framework composition; sedimentary lithics are
more abundant in the forearc sandstone than in the trench. The two groups of sedi-
ments were supplied from different parts of the Honshu Arc.
The lower part of the clastic sequence deposited within the Izu Arc is composed
mainly of volcaniclastic rocks and yields detrital clinopyroxenes that originated from
the Izu Arc. In contrast, the upper part is similar to the lower Miocene trench-fill
deposits in terms of its framework composition and the chemistry of detrital garnet
and chromian spinel. This reflects a change in provenance triggered by the initial con-
tact of the Izu Arc and the trench between the Eurasian and Philippine Sea plates.
The lower part of the middle Miocene trench-fill that was deposited following
initial contact is also similar to the lower Miocene trench-fill. The upper part, how-
ever, resembles lower Miocene sedimentary rocks of the forearc basin. This suggests
that the transport path was changed by collision. During the initial stages of collision
between the Honshu and Izu Arcs, the Honshu Arc was preferentially uplifted, and
therefore supplied most of the detritus to the collision zone.

Keywords: sandstone, provenance, framework composition, detrital, clinopyroxene,


garnet, chromian spinel, Honshu Arc, Izu Arc, collision.

*Present address: Institute for Geo-Resources and Environment, National Institute of Advanced Industrial Science and Technology, Tsukuba 305-8567, Japan;
koichi-okuzawa@aist.go.jp

Okuzawa, K., and Hisada, K., 2008, Temporal changes in the composition of Miocene sandstone related to collision between the Honshu and Izu Arcs, central
Japan, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of
America Special Paper 436, p. 185–198, doi: 10.1130/2008.2436(09). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of
America. All rights reserved.

185
186 Okuzawa and Hisada

INTRODUCTION Although the Izu Collision Zone is a suitable area for studies
that seek to understand the nature of sediments generated in arc-
The Izu Arc, an oceanic island arc, has been colliding with the arc collision settings, there are few petrographical studies of sedi-
Honshu Arc, a continental arc, in central Japan since ca. 15 Ma ments from the collision zone. According to Soh (1986), sand-
(Fig. 1); this zone of collision is termed the Izu Collision Zone stone and conglomerate from upper Miocene to Pliocene trench-
(Taira et al., 1989). There are four main blocks of accreted Izu fill sedimentary rocks are composed mainly of sedimentary lithics
Arc crust: the Koma, Misaka, Tanzawa, and Izu Blocks (Fig. 2). rather than volcanic lithics. In contrast, sandstone from the lower
Each block was accreted to the Honshu Arc at different times Miocene accretionary complex adjacent to the collision zone is
since ca. 15 Ma (Niitsuma and Akiba, 1985; Amano, 1986, 1991; conspicuously rich in quartz and feldspar (Watanabe and Iijima,
Aoike, 1999). The accreted blocks are associated with thick clas- 1990). It appears that a change in provenance occurred during
tic sequences showing coarsening-upward trends (e.g., Ito, 1985, the early to middle Miocene (ca. 13–17 Ma) at approximately the
1987) and regression-type facies (e.g., Huchon and Kitazato, same time that the collision zone formed.
1984). These sequences are interpreted to have been deposited Several types of detrital heavy minerals have been examined
in the trench between the Honshu Arc and the accreted blocks in various provenance studies to determine the source rocks of
(Amano, 1986, 1991; Koyama, 1991, 1993). The collision his- the sediment. Chemical analysis of detrital heavy minerals has
tory of each block was examined in previous studies on the proved to be a powerful tool in provenance studies (e.g., Mor-
basis of the lithology and geochemistry of the accreted blocks, ton, 1991). Chromian spinel, garnet, and clinopyroxene have
sequence stratigraphy, and the geochronology and deformation also been used to infer the nature of the sediment source rocks
history of trench-fill, as summarized by Amano et al. (1999) and (e.g., Hisada and Arai, 1993; Takeuchi, 1994; Cawood, 1983).
Aoike (1999). Chromian spinel generally originates from mafic to ultramafic

138°E 140°E
A B
150°
130°

140°
120°

38°N
50°
North 100 km

Eurasia American
Plate Plate

40°
Japan Sea
HG
r c
h uA Hd
ns
Ho 36°N
Pacific Sb
Izu

Plate 30°
Ch
Arc

Mn Sh Hy
Philippine Sea Ry
Inner zone

Plate
Boso Pen.
L
MT
140°

Mn
130°

Fig. 2
Izu a

Cretaceous to Tertiary accretionary complex


Outer zone

Jurassic accretionary complex


34°N
rc

Permian accretionary complex


High P/T type metamorphic rocks
Low P/T type metamorphic rocks
Circum-Izu massif serpentinite Forearc peridotite (Kurosegawa Belt)
Ophiolite
Continental block

Figure 1. Geologic outline of the Honshu Arc. (A) Plate tectonic setting of Japan. (B) Geologic map of basement rocks of the central Honshu
Arc. Modified from Isozaki (1996). MTL—Median Tectonic Line; Hd—Hida Belt; HG—Hida Gaien Belt; Mn—Mino Belt; Ry—Ryoke Belt;
Sb—Sambagawa Belt; Ch—Chichibu Belt (including Northern Chichibu, Kurosegawa, and Southern Chichibu Belts); Sh—Shimanto Belt (in-
cluding Setogawa, Kobotoke, Hayama, and Mineoka Belts); Hy—Hayama; Mn—Mineoka.
Collision between the Honshu and Izu Arcs 187

138°E 139°E

N
Koma Group
Ko
bo
tok
eB
Study area elt

ISTL

Mt. Fuji
Belt

Mt. Hakone
Figure 2. Geologic map of the Izu Col-
anto

lision Zone. Compiled from Tsuchi


(1986), Koyama (1993), Sugiyama

Sa
Shim

(1995), and Martin and Amano (1999).

ga
35°N

mi
ISTL—Itoigawa-Shizuoka Tectonic Line.
The study area is indicated by the three

Tro
rectangular boxes.

ugh
Suruga Trough

20km

Ryuso Group Trench fill sediments


Kurami Group Setogawa Izu Block Izu
Oigawa Group Belt Tanzawa Block Collision
Setogawa Group Misaka Block Zone
+ Intrusive rocks Koma Block
Cretaceous-Oligocene

rocks; accordingly, the occurrence of detrital chromian spinel This paper describes the framework composition, mineral-
can be used to constrain the timing of emplacement of mafic ogy, and mineral chemistry of sandstone from pre- to syncolli-
or ultramafic rocks. The chemistry of spinel can also provide sional successions within the Izu Collision Zone. The chemistry
clues to the tectonic setting in the source region (e.g., Arai, of detrital minerals was compared with the mineral chemistry of
1992; Barnes and Roeder, 2001). Garnet is commonly found potential source rocks proximal to the study area in order to iden-
within metamorphic rocks and granite. Garnet has a compli- tify provenance changes caused by arc-arc collision.
cated chemistry, and the chemical composition of metamorphic
garnet depends upon both the parent rock and the metamor- OUTLINE OF REGIONAL GEOLOGY
phic grade (Miyashiro, 1953). Clinopyroxene occurs as a stable
phase in almost all types of igneous rocks, as well as various The geology of the central Honshu Arc is characterized by
rocks that formed under conditions of both regional and contact a zonation of geological units (Fig. 1B). The central part of the
metamorphism (Deer et al., 1982). Although chemical analy- Honshu Arc is divided into the Inner and Outer Zones, which
sis of detrital heavy minerals is a useful tool for understanding are separated by the Median Tectonic Line (MTL, Fig. 1B). The
provenance, there has previously been no example applying it Inner Zone is composed of four belts that constitute a continen-
to the sedimentary rocks in the Izu Collision Zone. tal block (Hida Belt), an ophiolite (Hida Gaien Belt), a Jurassic
188 Okuzawa and Hisada

accretionary complex (Mino Belt), and low pressure-temperature Izu Collision


Setogawa Belt Zone
(P-T) metamorphic rocks (Ryoke Belt; Figure 1B; Isozaki, 1996). 10
Cretaceous to early Tertiary granites (e.g., Ryoke Granite) and western central &

Middle
felsic volcanic rocks (e.g., Nohi Rhyolite) are widely distributed part eastern
parts Mm
throughout the Inner Zone. The Outer Zone consists of a high P-T 15 Kurami Group

Miocene
metamorphic accretionary complex (Sambagawa Belt), a Juras- Ks
sic accretionary complex (Northern Chichibu Belt), an ophiolite Koma

Early
(Kurosegawa Belt), a Jurassic accretionary complex (Southern Group
20
Chichibu Belt), and a Cretaceous to Tertiary accretionary com-
plex (Shimanto, Setogawa, Kobotoke, Hayama, and Mineoka
Oigawa
Belts; Figure 1B; Isozaki, 1996). The Tertiary accretionary com- 25 Group
plex (Setogawa, Kobotoke, Hayama, and Mineoka Belts) includes

Late
Age (Ma)
a number of serpentinite bodies, which are called the Circum–Izu

Oligocene
Massif serpentinites (Arai, 1991); these four accrectionary com- 30
plex belts are also termed the Circum–Izu Massif Serpentine Belt
(Arai, 1991). These serpentinite bodies are accompanied by gab-

Early
bro and basaltic rocks of the mid-ocean-ridge basalt (MORB) 35
type and intraplate type (e.g., Ogawa and Taniguchi, 1988; Hirano

Late
et al., 2003), and are considered to have formed as crust gener-
ated by seafloor spreading in a backarc environment (Arai, 1991).

Eocene
40
Alternatively, Arai and Okada (1991) and Arai (1994) proposed

E. Middle
that the Circum–Izu Massif serpentinites were initially heteroge-
neous in their petrological characteristics. Arai (1994) proposed 50 Setogawa Group
that the incipient Circum–Izu Massif serpentinites were formed in
an arc environment. The serpentinites were later amalgmated with clastic rocks tuffaceous mudstone
gabbro and basaltic rocks (e.g., Ogawa and Taniguchi). chert limestone volcanic rocks
The study area covers parts of the Setogawa Belt and the
Figure 3. Correlation diagram of the stratigraphic units targeted in this
Izu Collision Zone (Fig. 2). An early Miocene to early middle study. Ks—Kushigatayama Subgroup; Mm—Momonoki Subgroup.
Miocene accretionary complex is distributed in the Setogawa Compiled from Koyama (1993), Sugiyama (1995), and Hiroki and
Belt and is composed of three stratigraphic units, the Setogawa, Matsumoto (1999).
Oigawa, and Ryuso Groups (Fig. 2). In the western part of the
Setogawa Belt, the Setogawa Group is covered by the lower Mio-
cene Kurami Group and younger sedimentary rocks, which were
deposited in a forearc basin (Fig. 2). The Izu Collision Zone com- Miocene (ca. 22–15 Ma), respectively (Fig. 3; Sugiyama, 1995).
prises accreted Izu Arc crust and trench-fill material (Osozawa et A total of seven sandstone samples were collected from the lower
al., 1990; Hiroki and Matsumoto, 1999; Amano, 1991). The four Miocene accretionary complex: three from the Setogawa Group,
lower to middle Miocene groups were targeted in this study in and four from the Oigawa Group.
order to reconstruct the provenance changes associated with the The Kurami Group crops out in the SW part of the Setogawa
early stage of Honshu-Izu collision: the Setogawa, Oigawa, and Belt (Fig. 2), where it unconformably overlies the Setogawa Group
Kurami Groups of the Setogawa Belt and the Koma Group of the (Fig. 3). The Kurami Group consists of basal conglomerate, sand-
Izu Collision Zone. The Ryuso Group in the Setogawa Belt was stone, alternating sandstone and mudstone, tuffaceous mudstone,
excluded from this study because the group is composed mostly and mudstone in ascending stratigraphic order (Ibaraki, 1986;
of volcanic rocks (Sugiyama, 1995) and is inadequate for the Hiroki and Matsumoto, 1999). The group was deposited in the
framework composition work. late early Miocene (18–15.5 Ma; Hiroki and Matsumoto, 1999)
The Setogawa Group comprises basalt, limestone, chert, and represents forearc basin sediments (Hiroki, 1995; Hiroki and
tuffaceous mudstone, and alternating sandstone and mudstone Matsumoto, 1999). Three sandstone samples were collected from
(Fig. 3; Sugiyama, 1995). The tuffaceous mudstone contains the lower part of the Kurami Group.
tectonic blocks of serpentinite, which is a component of the The Koma Group consists of two subgroups: the lower part
Circum-Izu Massif serpentinites (Arai, 1991), intercalated with is the Kushigatayama Subgroup, and the upper is the Momo-
conglomerate containing clasts of serpentinite. The Oigawa noki Subgroup (Fig. 3; Kosaka and Tsunoda, 1969), with the
Group consists mainly of alternating sandstone and mudstone Kushigatayama Subgroup being subdivided into lower and
with minor amounts of conglomerate and sandstone (Sugiyama, upper parts (Koyama, 1991, 1993). The lower part of the Kushi-
1995). The ages of the clastic rocks in the Setogawa and Oigawa gatayama Subgroup is composed mainly of tholeiitic intermedi-
Groups are early Miocene (24–19 Ma) and early to early middle ate to mafic volcaniclastic rocks and lava flows (Shimazu and
Collision between the Honshu and Izu Arcs 189

Ishimaru, 1987). The upper part consists mainly of intermediate A Qm Setogawa Group
to mafic tuff intercalated with mudstone, sandstone, and con-
glomerate. The Momonoki Subgroup consists of alternating Oigawa Group
sandstone and mudstone, conglomerate, mudstone, siliceous CI Kurami Group
mudstone, and felsic tuff (Koyama, 1991, 1993). The sequence Kushigatayama Subgp.
shows a coarsening-upward trend (Aoike, 1999). Conglomer- TC QR Momonoki Subgp.
ate clasts within the Momonoki Subgroup are derived from
both the Honshu and Izu Arcs (Koyama, 1993). The Kushigata- Mixed
TR
yama Subgroup was deposited within the Izu Arc, whereas the Type A
Momonoki Subgroup represents trench-fill sediment (Amano, DA Type B
1991; Aoike, 1999). The Kushigatayama and Momonoki Sub- BU
groups were deposited at 16–15 Ma and 15–13.5 Ma, respec-
TA LR
tively, and are in faulted contact (Aoike, 1999). We collected
a total of nine sandstone samples from the Koma Group: three Type C UA

from the lower part of the upper Kushigatayama Subgroup, two F Lt


from the uppermost Kushigatayama Subgroup, and two each
from the lower and upper Momonoki Subgroup. B Ls Qm: monocrystalline quartz
F: total feldspar
METHODS Lt: total lithics
Ls: sedimentary lithics
Lvf: felsic volcanic lithics
We mostly collected medium-grained sandstone samples Lvm: intermediate to mafic
from the Setogawa, Oigawa, Kurami, and Koma Groups. Thin volcanic lithics
sections were stained for K-feldspar and point-counted for 500
points per sample, following the Gazzi-Dickinson method (Inger-
soll et al., 1984). Type B
Heavy liquid separation was performed to collect clinopy-
roxene, garnet, and chromian spinel from the selected 13 samples. Type A
A total of 100 g of each specimen was crushed for heavy liquid Type C
separation. Separated heavy minerals were then mounted and thin
sectioned. The chemistry of these grains was analyzed using elec-
tron probe microanalysis (JXA8621 Super Microprobe; JEOL, Lvf Lvm
Tokyo) at the Research Facility Center for Science and Technol- Figure 4. Framework composition data for sandstone analyzed in the
ogy, University of Tsukuba, Japan. Operating conditions were 20 present study. (A) Qm-F-Lt diagram. (B) Ls-Lvf-Lvm diagram. Com-
kV accelerating voltage, 10 nA specimen current, and ~10-μm positional fields in Qm-F-Lt diagram are after Dickinson et al. (1983).
beam diameter. As no zoning was detected microscopically, we CI—craton interior; TC—transitional continent; BU—basement uplift;
analyzed the center of each grain. The results of all the micro- QR—quartzose recycled; TR—transitional recycled; LR—lithic recy-
cled; DA—dissected arc; TA—transitional arc; UA—undissected arc.
probe analyses were published previously by Okuzawa (2004).

RESULTS

Framework Composition of Sandstone Sandstone of the Kurami Group, the uppermost Kushiga-
tayama Subgroup (sample 02102506, Table 1), and the upper
Sandstone in the study area can be divided into types A, Momonoki Subgroup (samples 02102208 and 02102303,
B, and C on the basis of framework composition, as indicated Table 1) are classified as type B. This type of sandstone is com-
by the fields shown in Figure 4. Sandstone of the Setogawa posed mainly of felsic volcanic lithics (13%–36%), monocrystal-
and Oigawa Groups, the uppermost part of the Kushigatayama line quartz (12%–22%), sedimentary lithics (mainly mudstone:
Subgroup (sample 02102504, Table 1), and the lower part of 10%–19%), and plagioclase (7%–17%), with minor K-feldspar
the Momonoki Subgroup (samples 02102101 and 02102207, (0%–5%) and polycrystalline quartz (0%–6%; Table 1). Type
Table 1) belongs to type A. This type of sandstone consists B sandstone is richer in sedimentary lithics than that of type
mainly of monocrystalline quartz (14%–41%), felsic volcanic A (Fig. 4B) and plots in the field of transitional magmatic arc
lithics (11%–31%), and plagioclase (10%–25%), with minor (Dickinson et al., 1983; Fig. 4A).
amounts of K-feldspar (1%–11%; Table 1). Type A sandstone Sandstone from the lower part of the upper Kushigata-
plots in the field of dissected to transitional magmatic arc (Dick- yama Subgroup (samples 02102501, 02102502, and 02102503,
inson et al., 1983; Fig. 4A). Table 1) belongs to type C. This sandstone consists mainly of
190 Okuzawa and Hisada

TABLE 1. FRAMEWORK COMPOSITIONS OF SANDSTONE SAMPLES


If Mtx
Sample Other Total
Group Qm Qp K-fel. Pl. Lvf Lvi Ls H.m. S.m. and and
no. lithics (%)
Aut. Cem.
Setogawa 02092607 36.6 0.4 11.2 15.4 15.6 0.0 0.2 0.0 0.8 2.8 0.4 16.7 100.0
02093005 38.0 0.0 8.6 18.6 13.8 0.0 0.0 0.0 0.4 0.8 0.0 19.8 100.0
02093008 14.8 0.0 1.4 23.8 29.8 0.0 0.4 0.0 0.4 2.0 1.4 26.0 100.0

Oigawa 02122603 22.6 0.6 2.0 15.0 26.6 0.0 0.6 0.0 0.4 1.0 0.0 31.2 100.0
02122604 35.2 0.8 3.4 10.0 30.8 0.0 0.0 0.0 0.4 0.4 0.0 19.0 100.0
02122606 41.2 0.0 7.0 15.4 10.8 0.0 0.0 0.0 2.6 1.2 0.0 21.8 100.0
02122609 36.4 0.2 4.2 16.2 20.2 0.0 0.4 0.0 0.2 1.8 0.0 20.4 100.0

Kurami 02122803 22.4 4.6 4.0 11.2 22.4 0.0 19.2 0.0 0.8 0.0 0.0 15.4 100.0
02122807 11.6 2.2 5.0 12.0 20.8 0.8 9.6 0.0 1.2 0.0 0.0 36.8 100.0
02122810 21.0 3.6 5.0 7.0 26.8 0.0 18.4 0.0 0.4 0.2 0.0 17.6 100.0

Koma 02102501 1.6 0.8 0.2 8.8 27.0 17.4 2.6 0.0 5.6 4.4 0.0 31.6 100.0
(Kushigata-yama 02102502 0.0 0.0 0.0 34.2 15.2 23.6 0.2 0.0 1.2 9.8 0.0 15.8 100.0
Subgroup) 02102503 9.6 0.6 2.8 25.2 27.8 4.0 5.8 0.0 0.0 0.0 9.2 15.0 100.0
02102504 14.2 0.6 6.8 24.6 23.6 1.4 4.8 0.0 1.2 1.0 0.0 21.8 100.0
02102506 17.6 0.4 0.4 17.2 13.2 0.6 13.0 0.0 0.6 0.2 0.0 36.8 100.0

(Momonoki 02102101 27.6 0.0 1.6 15.4 23.4 0.0 0.6 0.0 0.2 0.4 0.0 30.8 100.0
Subgroup) 02102207 36.6 0.2 5.2 12.4 23.6 0.6 3.8 0.0 0.2 0.2 0.0 17.2 100.0
02102208 13.8 0.2 1.8 12.6 21.4 5.8 13.2 0.0 0.8 1.0 0.2 29.2 100.0
02102303 19.4 5.8 4.8 9.4 35.8 0.2 10.0 0.0 0.4 0.4 0.2 13.6 100.0
Note: Qm—monocrystalline quartz; Qp—polycrystalline quartz; K-fel.—K-feldspar; Pl.—plagioclase; Lvf—felsic volcanics; Lvi—intermediate to mafic
volcanics; Ls—sedimentary lithics; H.m.—heavy minerals; S.m.—secondary minerals; If and Aut.—intrabasinal fragments and authigenic minerals; Mtx
and Cem.—matrix and cement.

felsic to mafic volcanic fragments (32%–44%) together with pla- Eighty-eight detrital garnets were analyzed from the Oigawa
gioclase (9%–34%; Table 1). Type C sandstone plots in the field Group, of which 70 are classified as pyrope-rich almandines (Pyr
of undissected and transitional magmatic arc (Dickinson et al., = 5–45 mol%), 10 are spessartine-rich almandines (Sps = 9–40
1983; Fig. 4A). mol%), six are Ca-rich almandines (Grs + And = 14–23 mol%),
and two are almandine-rich spessartines (Sps = 52 and 62 mol%;
Chemistry of Detrital Clinopyroxene Fig. 6A).
Forty-nine detrital garnets were obtained from the Kurami
Sandstone from the lower part of the upper Kushigatayama Group, of which 24 are spessartine-rich almandines (Sps = 7–44
Subgroup (type C sandstone) yields abundant detrital clinopy- mol%), 20 are pyrope-rich almandines (Pyr = 8–41 mol%), three
roxene grains. Almost all of these grains are colorless to light are Ca-rich almandines (Grs + And = 14–17 mol%), one is an
green and are angular to subrounded. We obtained chemical anal- almandine-rich spessartine (Sps = 45 mol%), and one is an Mg-
yses of 30 detrital clinopyroxenes. Twenty-eight of the analyzed rich spessartine (Sps = 53 mol%; Fig. 6B).
grains are augite (Fig. 5A). They are Ca-rich and Fe-poor, with Sixty-one detrital garnets were analyzed from the Koma
low contents of TiO2 (0.1–0.4 wt%), Cr2O3 (0.0–0.1 wt%), Na2O Group. Their compositional range is similar to that of garnets
(0.2–0.4 wt%), and total Al (total Al as Al2O3: 0.6–1.5 wt%). The from the Kushigatayama and Momonoki Subgroups (Fig. 6C).
other two grains are augite and diopside, and are characterized These grains are mostly pyrope-rich almandines (Pyr = 10–45
by a higher Cr content (Cr2O3 >0.5 wt%) and total Al (3.1–4.2 mol%) and spessartine-rich almandines (Sps = 6–45 mol%),
wt%), with TiO2 and Na2O contents of 0.3 wt% and 0.1–0.2 wt%, with just seven grains of Ca-rich almandines (Grs + And = 6–37
respectively. mol%) and one of andradite-rich grossular (Grs = 89 mol%).

Chemistry of Detrital Garnet Chemistry of Detrital Chromian Spinel

A total of 269 chemical analyses of detrital garnets were All the sandstone samples that contain detrital garnets also
obtained from the Setogawa, Oigawa, Kurami, and Koma Groups, yield reddish-brown to black detrital chromian spinels. We obtained
with the exception of the lower part of the upper Kushigatayama a total of 324 chemical analyses of chromian spinel grains.
Subgroup. Seventy-one grains of detrital garnet were analyzed The chemical characteristics of detrital chromian spinels
from sandstones of the Setogawa Group: 49 grains are classified from the Setogawa, Oigawa, Kurami, and Koma Groups are simi-
as pyrope-rich almandine (Pyr = 7–45 mol%), 15 grains are spes- lar. The spinels from the four groups are characterized by a wide
sartine-rich almandines (Sps = 9–47 mol%), 3 grains are Ca-rich range in Cr# [= Cr/(Cr+Al) atomic ratio], from 0.2 to 1.0 (Fig. 7).
almandines (Grs + And = 11–27 mol%), 3 grains are almandine- These spinels show a negative trend in terms of the relationship
rich spessartines (Sps = 43–63 mol%), and one is a Ca-rich spes- between Mg# [= Mg/(Mg+Fe2+) atomic ratio] and Cr# (Fig. 7).
sartine (Sps = 40 mol%; Fig. 6A). Values of Fe3+# [= Fe3+/(Cr+Al+Fe3+) atomic ratio] are mostly
Collision between the Honshu and Izu Arcs 191

A Diopside DIOPSIDE Hedenbergite Detrital garnets from the Setogawa and Oigawa Groups
HEDENBERGITE have similar chemical characteristics (Fig. 6A), and are
L. Kushigatayama
mostly spessartine-rich almandine and pyrope-rich alman-
Subgroup dine (Fig. 6A). Spessartine-rich almandine is found in granitic
rocks and low-grade to amphibolite facies metamorphic rocks
AUGITE (Fig. 6A; Nanayama, 1997). Low and high P-T metamorphic
PIGEONITE rocks and granites from the Honshu Arc (Fig. 1) in the area
proximal to the Setogawa Belt are candidates as source rocks
for the spessartine-rich almandine. Intrusion of granitic rocks in
CLINOENSTATITE CLINOFERROSILITE
the Izu Collision Zone postdates sedimentation of the Setogawa
Enstatite Ferrosilite and Oigawa Groups, and they cannot be source rocks for the
detrital garnet. Most of the detrital spessartine-rich almandines
plot in the same field as garnets from the granites and low P-T
B metamorphic rocks (Fig. 6A). Accordingly, we conclude that the
detrital spessartine-rich almandines were probably derived from
0.04 these rocks.
Orogenic calcalkali basalts Detrital pyrope-rich almandines from the Setogawa and
Oigawa Groups plot in the fields of amphibolite, pegmatite, low-
grade metamorphic rocks, and granulite facies metamorphic rocks
Ti

(Fig. 6A; Nanayama, 1997). Thirty-five grains of the almandine


garnet analyzed from the lower Miocene Setogawa and Oigawa
Groups have relatively high Mg contents (MgO >10 wt%). The
Orogenic tholeiitic basalts Mg/Fe ratio of Ca-poor garnet is known to increase with increas-
ing metamorphic grade (e.g., Coleman et al., 1965). Granulites
0 that contain garnets with Mg contents >10 wt% have not been
0 0.1 0.2 0.3 found in the Honshu Arc or in the Izu Arc (e.g., Teraoka et al.,
Al total 1998), although pyrope-rich almandines have been reported from
Jurassic to Cretaceous sandstones of the Honshu Arc (Adachi
Figure 5. Chemical composition data for detrital clinopyroxenes and Kojima, 1983; Teraoka et al., 1999). Such garnets have also
analyzed in the present study. The subdivision scheme (A) is after
been described from the Sino-Korean Massif of mainland Asia
Morimoto (1988). The range of volcanic rocks of the lower Kushigata-
yama Subgroup is surrounded by broken lines (Shimazu and Ishimaru, (e.g., Jiang, 1988), which is likely to have been the source of the
1987). The fields of orogenic calc-alkali and tholeiitic basalts (B) are detrital pyrope-rich garnets in the lower Miocene Setogawa and
after Leterrier et al. (1982). Oigawa Groups. Because detrital garnets are rare in the sand-
stones of the Setogawa and Oigawa Groups, they may have been
reworked from pre-Miocene sandstones.
The chemistry of detrital garnets from the middle Miocene
<0.1, whereas TiO2 wt% is dominantly in the range 0.0–0.5, trench-fill sedimentary rocks of the Koma Group is similar to that
although some grains record Fe3+# values >1.0. of detrital garnets from the lower Miocene Setogawa and Oigawa
Groups (Figs. 6A, C). However, the Sea of Japan had begun to
PROVENANCE OF PRE-COLLISION SEDIMENTARY open in the middle Miocene (>15 Ma; e.g., Otofuji et al., 1985),
ROCKS and it is therefore difficult to envisage that the Sino-Korean Mas-
sif supplied detrital garnet to the study area at this time. Pyrope-
Trench Sedimentary Rocks rich garnets of the Koma Group were probably reworked from
pre–middle Miocene sandstones.
Sandstone of the Setogawa and Oigawa Groups (labeled Detrital chromian spinels obtained from trench sedimentary
Type A in Fig. 4) that was deposited in the trench prior to col- rocks of the Setogawa and Oigawa Groups are characterized by
lision (Aoike, 1999) contains fragments of granite, felsic vol- a wide range in Cr#, low Fe3+#, and low TiO2 content, although
canic rock, and sedimentary rocks such as chert and mudstone. some have a higher TiO2 content (>1.0 wt%; Fig. 7A). A high
Pre–early Miocene granitic rock, felsic volcanic rock, and chert Cr#, coupled with low Ti and Fe3+ concentrations, is character-
have not been found in the Izu Arc. Possible sources for these istic of chromian spinels from forearc peridotite (Bloomer and
rock fragments include Cretaceous to early Tertiary granites Fisher, 1987; Ishii et al., 1992; Parkinson and Pearce, 1998) or
and felsic volcanic rocks from the Inner Zone of the Honshu boninites (Arai, 1992; Barnes and Roeder, 2001). Chromian
Arc, as well as a Jurassic to Cretaceous accretionary complex spinels with low TiO2 content plot mainly within the field of
in the Honshu Arc. forearc peridotite (Fig. 7A[i]). Consequently, forearc peridotite
192 Okuzawa and Hisada

A Lower Miocene trench sedimentary rocks B Lower Miocene forearc basin sedimentary
Setogawa Group (N=71)
rocks
Kurami Group (N=49)
Oigawa Group (N=88)
Alm GNF Pyr Alm GNF
Pyr
Sps Sps
ECF ECF
PG+Met.
PG+Met.

PG+Low Met. PG+Low Met.

Skarn Skarn
GNF ECF GNF ECF
APF Gro+And APF Gro+And
Alm Pyr Alm Pyr

C Middle Miocene collision zone sedimentary rocks Granites and metamorphic rocks
in the Ryoke Belt
Alm GNF Pyr
Sps Metamorphic rocks in the
Sambagawa Belt
ECF
Sps: spessartine
PG+Met. Alm: almandine
Pyr: pyrope
Gro: grossular
And: andradite

PG+Low Met. PG: pegmatite


Skarn Low Met.: low metamorphic rock
Koma Group (N=61) APF: amphibolite facies
Momonoki Subgroup ECF: eclogite facies
GNF
Kushigatayama GNF: granulite facies.
ECF
Subgroup
APF
Alm Pyr Gro+And
+And
Figure 6. Chemical composition data for detrital garnets. Garnets from (A) lower Miocene trench sedimentary rocks; (B) lower Miocene forearc
basin sedimentary rocks; (C) middle Miocene collision zone sedimentary rocks. Pegmatite and four metamorphic fields with solid lines are after
Nanayama (1997). The fields of Sambagawa Belt metamorphic rocks are compiled from Higashino et al. (1981, 1984), Enami (1983), Sonobe
and Takasu (2000), and Sakurai (2000). The field of granites and metamorphic rocks in the Ryoke belt is compiled from Ono (1969, 1975a, b,
1976, 1977, 1981), Asami and Hoshino (1980), Asami et al. (1982), Nureki et al. (1982), Seo et al. (1982), Shiba (1982, 1989), Wang (1985),
Takagi and Nagahama (1987), Enami (1988), and Miyazaki (1999).

is a possible source rock for most of the detrital chromian spinels Several bodies of mafic and ultramafic rocks occur close to
that have low TiO2 concentrations. the study area: forearc peridotites of the Kurosegawa Belt, peri-
TiO2-rich (>1.0 wt%) detrital chromian spinels are derived dotites and basaltic rocks of the high P-T Sambagawa Metamor-
from a different source rock than that yielding the low-Ti spi- phic Belt, and backarc basin peridotites and basaltic rocks of the
nels. High-Ti spinels with intermediate Cr# are found in intra- Circum–Izu Massif Serpentine Belt (Fig. 1). All of these mafic
plate magmas (Arai, 1992; Barnes and Roeder, 2001). Intra- and ultramafic rocks are exposed within the Honshu Arc. There
plate basalt is therefore a suitable source rock for the high-Ti is no possible source rock for detrital chromian spinel known
detrital spinels. in the Izu Arc. Basalts from the Circum–Izu Massif Serpentine
A Lower Miocene trench sedimentary rocks
1
Cr
i) ii) Setogawa Group (N=125)
iii)
Oigawa Group (N=111)
5
Cr# [Cr/(Cr+Al)]

TiO2wt%
0 Mg# [Mg/(Mg+Fe2+)] 1 0
Cr# [Cr/(Cr+Al)] 1 Al Fe3+

B Lower Miocene forearc basin sedimentary rocks


1 Cr
i) ii) iii)
Kurami Group (N=43)
Forearc
5
peridotite
Cr# [Cr/(Cr+Al)]

Intraplate basalts of Circum-Izu


Massif Serpentine Belt
TiO2wt%

Kurosegawa
serpentinite

Circum-Izu
Massif Serpentinites

0 Mg# [Mg/(Mg+Fe2+)] 1 0
Cr# [Cr/(Cr+Al)]
1 Al Fe3+

C Middle Miocene collision zone sedimentary rocks


1 Cr
i) ii) Koma Group (N=86) iii)
Momonoki Subgroup
5
Kushigatayama Subgroup
Cr# [Cr/(Cr+Al)]

TiO2wt%

Sambagawa
greenstone and peridotite

0 1
Mg# [Mg/(Mg+Fe2+)] 1 0
Cr# [Cr/(Cr+Al)] Al Fe3+

Figure 7. Chemical composition data for detrital chromian spinels analyzed in the present study. The range of the Circum–Izu Massif serpentinites is
enclosed by thin lines (Arai et al., 1990). The range of the intraplate basalts in the Circum–Izu Massif Serpentine Belt is surrounded by a thick line
(Okuzawa and Hisada, 2004). The range of the forearc peridotite is surrounded by a dotted line (Bloomer and Fisher, 1987; Parkinson and Pearce,
1998). The range of the Kurosegawa serpentinite is circled by a broken line (Okuzawa et al., 2004). The range of the peridotite and greenstone in the
Sambagawa Belt is surrounded by a dotted thin line (Uesugi and Arai, 1999).
194 Okuzawa and Hisada

Belt have intraplate and MORB-type ocean floor signatures to be derived from chromitite-bearing ophiolite. It is known that
(e.g., Ogawa and Taniguchi, 1988; Hirano et al., 2003) and are the Circum–Izu Massif serpentinites include a small amount of
interpreted as having been formed within a backarc setting or a chromitite (Kitahara, 1954).
mid-ocean-ridge setting (Hirano et al., 2003). However, Arai and In summary, the framework composition of sandstones and
Okada (1991) and Arai (1994) reported an occurrence of high Cr# the chemistry of detrital chromian spinels and garnets indicate
detrital spinels from tectonic blocks of serpentine-rich sandstone that prior to collision the trench sediments were supplied from
within tectonic blocks of the Circum–Izu Massif serpentinites. the Honshu Arc.
Although the sedimentary age of the serpentine-rich sandstone
is unsettled, Arai (1994) concluded that those sandstones were Forearc Basin Sedimentary Rocks
deposited immediately after emplacement of the Circum–Izu
Massif serpentinites, and that these incipient serpentinites had Sandstone of the upper lower Miocene Kurami Group
petrological characteristics of peridotite formed in an arc setting. (labeled Type B in Fig. 4) was deposited in the forearc basin
Many low-TiO2 detrital chromian spinels from the Setogawa of the Honshu Arc and has a different framework composi-
and Oigawa Groups do not plot within the field of spinels from tion from that of lower Miocene trench sedimentary rocks:
the Kurosegawa forearc peridotite, but instead within the field The forearc basin sandstone is richer in sedimentary lithics in
of spinels from the recent Circum–Izu Massif serpentinites comparison with the trench sandstone, although the assemblage
(Fig. 7A[i], [ii]). There are three possible combinations of source of detrital grains (Table 1) and the chemical characteristics of
rocks for the low-TiO2 detrital spinels: (1) ancient Circum–Izu detrital chromian spinels and garnets (Figs. 6 and 7) are simi-
Massif serpentinites that supplied all the low-TiO2 detrital spi- lar. Marsaglia and Ingersoll (1992) concluded that very little
nels found in the Setogawa and Oigawa Groups, (2) ancient compositional variation is related to forearc basin geometry
Circum–Izu Massif serpentinites and other forearc peridotites or to submarine transport from the forearc into the trench. The
and/or boninites that supplied the low-TiO2 detrital spinels, and compositional differences recorded between lower Miocene
(3) other forearc peridotites and boninites that supplied all the trench and forearc basin sedimentary rocks indicate that detri-
low-TiO2 detrital spinels. In the Boso Peninsula (Fig. 1) the lower tus from the two groups of rocks was derived from different
Miocene Mineoka Group and middle Miocene Sakuma Group provenances. The forearc basin sedimentary rocks appear to
contain detrital spinels with low Ti content and a wide range of have had their source from a different part of the Honshu Arc
Cr# (Cr# = 0.3–0.9: Okuzawa and Hisada, 2004). These authors in comparison with the trench-fill sediments. The detritus of the
concluded that the Circum–Izu Massif serpentinites had supplied Setogawa Group may have been axial currents along the trench
detrital chromian spinels to both groups, because both groups axis, because the framework composition of the Setogawa
include conglomerate containing clasts of serpentinite, gabbro, Group sandstone resembles the trench sandstone of the lower
and basalts, which probably originated from the ophiolitic rocks Miocene Mineoka Group in the Boso Peninsula (Watanabe and
in the Circum–Izu Massif Serpentine Belt. On this basis, the first Iijima, 1990; Okuzawa, 2004).
scenario above appears to be the most plausible.
Most of the high-TiO2 chromian spinels found in the Setogawa CHANGES IN PROVENANCE FOLLOWING
and Oigawa Groups plot within the field of alkaline basalt from the COLLISION
Circum–Izu Massif Serpentine Belt (Fig. 7A[ii]). Peridotites and
basalts in the high P-T Sambagawa Metamorphic Belt also contain The lower part of the upper Kushigatayama Subgroup sand-
high-TiO2 spinels (Fig. 7A[ii]; Uesugi and Arai, 1999), however, stone comprises sandstone of type C (see above; Fig. 4). This
most of the detrital spinels have a lower Fe3+# than spinels in the rock contains intermediate to mafic volcanic lithics (Fig. 4B)
peridotite and basalt of the Sambagawa Belt (Fig. 7A[iii]). The and abundant detrital clinopyroxene. A volcanic arc is therefore a
high-TiO2 detrital chromian spinels were probably derived from likely provenance for this rock type. We used the discrimination
intraplate basalts of the Circum–Izu Massif Serpentine Belt. diagram of Leterrier et al. (1982) to infer the tectonic setting of
The detrital chromian spinels with higher values of Mg# the source rock on the basis of detrital clinopyroxene geochemis-
and Cr# also plot outside the fields of the forearc peridotites try (Fig. 5B). Two detrital clinopyroxenes with Cr2O3 >0.5 wt%
and Circum–Izu Massif serpentinites in an Mg#-Cr# diagram are excluded from analysis because they might have been derived
(Fig. 7A[i]). Intraplate basalt from the Circum–Izu Massif Ser- from different source rocks. Plots of detrital clinopyroxenes from
pentine Belt contains some spinels with higher values of Mg# the lower part of the upper Kushigatayama Subgroup form a tight
and Cr# than those of the forearc peridotites and Circum–Izu cluster on the discrimination diagram, with most grains plotting
Massif serpentinites (Ishida et al., 1990); however, most of the in the field of orogenic tholeiitic basalts (Fig. 5B). On this basis,
detrital chromian spinels have lower TiO2 contents than spinels we conclude that detrital clinopyroxenes from the Kushigata-
in the intraplate basalts. According to Barnes and Roeder (2001), yama Subgroup were derived from orogenic tholeiitic basalts.
chromitites within ophiolite contain chromian spinels with high Lower Miocene volcanic rocks from the Honshu Arc con-
Mg# and Cr# and low Fe3+# and TiO2 contents. On this basis, sist of both tholeiitic and calc-alkaline rocks (e.g., Shuto et al.,
some of the detrital spinels analyzed in the present study appear 1988). Basalts from the Circum–Izu Massif Serpentine Belt
Collision between the Honshu and Izu Arcs 195

have intraplate and MORB-type ocean floor geochemical sig- A Lower part of the upper Kushigatayama Subgroup
natures (e.g., Ogawa and Taniguchi, 1988; Hirano et al., 2003). (eve of the collision)
The chemistry of clinopyroxene in these rocks is different Type A
from that of orogenic tholeiitic basalts. Clinopyroxenes within rc Type B
ua
high P-T metamorphic rocks from the Sambagawa Belt show Ho nsh
Type C
omphacite and aegirine-augite compositions (Higashino et al.,
.
1981; Sakurai, 2000) and are therefore different from detrital i Gp
am
Kur
clinopyroxene grains in the upper Kushigatayama Subgroup. . arc
ted Gp aG
p. Izu
Volcanic rocks from the lower Kushigatayama Subgroup are cre awa w
c
A tog Oiga
island arc tholeiites and contain clinopyroxenes that are mainly Se ma
taya
higa
diopside and augite, with minor chrome diopside (Fig. 5A; Phi
lipp Kus gp.
ine S u b
Shimazu and Ishimaru, 1987). The chemical characteristics of the Sea
Pla
te
detrital clinopyroxenes from the lower part of the upper Kushiga-
tayama Subgroup correspond with those of clinopyroxene-bearing
volcanic rocks from the lower Kushigatayama Subgroup. Fujioka B Upper part of the upper Kushigatayama Subgroup
and lower Momonoki Subgroup (begining of collision)
and Saito (1992) reported chemical analyses of detrital clinopy-
roxenes from Oligocene to Pleistocene volcaniclastic sediments
and sedimentary rocks sampled during Ocean Drilling Program arc
shu
Hon
(ODP) Leg 126 at the Izu Arc. Detrital clinopyroxenes from the
p.
upper Kushigatayama Subgroup analyzed in the present study are ubg
no ki S
similar to those from Oligocene to Miocene sedimentary rocks mo
Mo
sampled from ODP Leg 126 sites. In summary, detritus from the . arc
ted a Gp p. Izu
cre wa G
lower part of the upper Kushigatayama Subgroup appears to have Ac togaw Oiga
Se ma
been derived from the Izu Arc (Fig. 8A). taya
Phi u s higa
K gp.
Sandstone from the uppermost Kushigatayama Subgroup lipp
ine Sub
Sea
and the lower Momonoki Subgroup is classified within our type Pla
te
A category (see above; Fig. 4) and is similar to sandstone within
the lower Miocene trench-fill (Fig. 8B) in terms of the chemis- C Upper Momonoki Subgroup (late stage of collision of
try of detrital heavy minerals (Figs. 6 and 7). The change from Koma Block)
type C to type A that is observed stratigraphically upward within
the Kushigatayama Subgroup is interpreted as a change in prov-
arc
enance related to the arrival of the Izu Arc at the Honshu Trench. shu
Hon
This indicates that the basin within the Izu Arc in which the p.
ubg
Kushigatayama Subgroup was deposited had moved close to the no ki S
mo
trench by 16–15 Ma (Aoike, 1999). Mo
. arc
Sandstone of the upper Momonoki Subgroup is classified as ted a Gp p Izu
cre G
Ac togaw wa
type B and contains detrital heavy minerals of similar chemistry Se Oig
a ma
& taya
to those within the lower Miocene Kurami Group, deposited in Phi u s higa
lip K gp.
the forearc basin (Figs. 6 and 7). As discussed above, sandstone pine
Sea Sub
Pla
of type B was supplied from a different source area of the Hon- te
shu Arc from that of type A (Fig. 8C). The change in provenance
recorded within the Momonoki Subgroup appears to have been Figure 8. Temporal changes to framework composition of the Koma
caused by a change in the passage of detritus to the trench owing Group sandstone at three stages: (A) Lower part of the upper Kushi-
to progressive collision between the two arcs. gatayama Subgroup. (B) Upper part of the upper Kushigatayama Sub-
Conglomerate within the upper Momonoki Subgroup contains group and the lower Momonoki Subgroup. (C) Upper Momonoki Sub-
group. See the text for explanation of types A, B, and C sandstones.
volcanic clasts derived from both the Kushigatayama Subgroup
and the Honshu Arc (Koyama, 1993); however, judging from the
dominance of sedimentary lithics over intermediate to mafic vol-
canic lithics in sandstone (Fig. 4A) and the lack of detrital clino- CONCLUSIONS
pyroxene, sediments of the Momonoki Subgroup were probably
derived mainly from the Honshu Arc. During the early stages of To understand the change in provenance resulting from col-
collision within the Izu Collision Zone, the Honshu Arc appears to lision between the Izu and Honshu Arcs, we studied the frame-
have been preferentially uplifted and thereby supplied more detri- work composition of sandstone and the chemical composition of
tus to the trench in comparison with sources in the Izu Arc. detrital clinopyroxene, garnet, and chromian spinel from lower to
196 Okuzawa and Hisada

middle Miocene coarse-grained sedimentary rocks deposited in Arai, S., and Okada, H., 1991, Petrology of serpentine sandstone as a key to tec-
and around the Izu Collision Zone. Sediments that were depos- tonic development of serpentine belts: Tectonophysics, v. 195, p. 65–81,
doi: 10.1016/0040-1951(91)90144-H.
ited prior to collision were derived from the Honshu Arc, includ- Arai, S., Ito, M., Nakayama, N., and Masuda, F., 1990, A suspect serpentinite
ing Cretaceous to early Tertiary granites, low P-T metamorphic mass in the Tokyo Bay area: Petrology and provenance of serpentinite
rocks, Jurassic to Cretaceous accretionary complexes, and forearc pebbles in the upper Cenozoic system in the Boso Peninsula, central
Japan: Journal of the Geological Society of Japan, v. 96, p. 171–179 (in
peridotite and/or backarc basin ophiolite. Sedimentary rocks of Japanese with English abstract).
the pre-collisional trench and forearc basin of the Honshu Arc Asami, M., and Hoshino, M., 1980, Staurolite-bearing schists from the Hongu-
differ in terms of the framework composition of sandstone. san area in the Ryoke metamorphic belt, central Japan: Journal of the
Geological Society of Japan, v. 86, p. 581–591.
Sediments of the lower part of the upper Kushigatayama Sub- Asami, M., Hoshino, M., Miyakawa, K., and Suwa, K., 1982, Metamorphic
group were derived from volcanic rocks of the Izu Arc, whereas conditions of staurolite schists of the Ryoke metamorphic belt in the
those of the uppermost Kushigatayama Subgroup and the lower Hazu-Hongusan area, central Japan: Journal of the Geological Society of
Japan, v. 88, p. 437–450 (in Japanese with English abstract).
Momonoki Subgroup resemble lower Miocene trench fill. This Barnes, J.S., and Roeder, P.L., 2001, The range of spinel compositions in ter-
change in provenance was caused by the arrival of the Izu Arc at restrial mafic and ultramafic rocks: Journal of Petrology, v. 42, p. 2279–
the Honshu Trench. Sandstone of the upper part of the Momo- 2302, doi: 10.1093/petrology/42.12.2279.
Bloomer, S.H., and Fisher, R.L., 1987, Petrology and geochemistry of igneous
noki Subgroup is similar to lower Miocene sedimentary rocks rocks from the Tonga Trench–A non-accreting plate boundary: Journal of
deposited in the forearc of the Honshu Arc. This suggests that Geology, v. 95, p. 469–495.
the passage of detritus to the trench changed following collision. Cawood, P.A., 1983, Modal composition and detrital clinopyroxene geo-
chemistry of lithic sandstones from the New England Fold Belt (east
In the early stages of collision between the Izu and Honshu Arcs Australia): A Paleozoic forearc terrane: Geological Society of America
the Honshu Arc was preferentially uplifted and consequently pro- Bulletin, v. 94, p. 1199–1214, doi: 10.1130/0016-7606(1983)94<1199:
vided the bulk of the detritus supplied to the trench. MCADCG>2.0.CO;2.
Coleman, R.G., Lee, D.E., Beatty, L.B., and Brannock, W.W., 1965, Eclog-
ites and eclogites: Their differences and similarities: Geological
ACKNOWLEDGMENTS Society of America Bulletin, v. 76, p. 483–508, doi: 10.1130/0016-
7606(1965)76[483:EAETDA]2.0.CO;2.
Deer, W.A., Howie, R.A., and Zussman, J., 1982, An Introduction to the Rock-
We appreciate the informative and helpful discussions with Forming Minerals (2nd edition): London, Longman, 696 p.
Shoji Arai on the chemistry of chromian spinel and with Yujiro Dickinson, W.R., Bread, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C.,
Ogawa on tectonics in the Boso Peninsula. Comments on the Inman, K.F., Kenepp, R.A., Lindberg, F.A., and Ryberg, P.T., 1983, Prov-
enance of North American Phanerozoic sandstones in relation to tectonic
manuscript by Peter D. Clift, Andy Morton, Kantaro Fujioka, setting: Geological Society of America Bulletin, v. 94, p. 222–235, doi:
and an anonymous reviewer are acknowledged. We extend our 10.1130/0016-7606(1983)94<222:PONAPS>2.0.CO;2.
thanks to Teruo Ohtsubo, Naoto Hirano, Jun-ichiro Kuroda, and Enami, M., 1983, Petrology of pelitic schists in the oligoclase-biotite zone of
the Sanbagawa metamorphic terrain, Japan: Phase equilibria in the high-
Shoichiro Tokumine for their help in the field work. Thanks are est grade zone of a high-pressure intermediate type of metamorphic belt:
due to Norimasa Nishida for his help with microprobe analysis. Journal of Metamorphic Geology, v. 1, p. 141–161, doi: 10.1111/j.1525-
1314.1983.tb00269.x.
Enami, M., 1988, Chemistry of garnet of the Ryoke metamorphic belt: Journal
REFERENCES CITED of Geological Society, Nagoya, no. 50, p. 71–80 (in Japanese).
Fujioka, K., and Saito, S., 1992, Composition of heavy minerals from sands and
sandstones of the Izu-Bonin arc, Leg 126, in Taylor, B., Fujioka, K., et al.,
Adachi, M., and Kojima, S., 1983, Geology of the Mt. Hikagedaira area, east eds., Proceedings of the Ocean Drilling Program, Scientific Results, Vol-
of Takayama, Gifu Prefecture, central Japan: Journal of Earth Sciences, ume 126: College Station, Texas, Ocean Drilling Program, p. 155–169.
Nagoya University, v. 31, p. 37–67. Higashino, T., Sakai, C., Otsuki, M., Itaya, T., and Banno, S., 1981, Electron
Amano, K., 1986, Southern Fossa Magna as multiple collision belt: Earth microprobe analyses of rock-forming minerals from the Sanbagawa met-
Monthly (Chikyu), v. 8, p. 581–585 (in Japanese). amorphic rocks, Shikoku, Part I. Asemi River area: Science Report of
Amano, K., 1991, Multiple collision tectonics of the South Fossa Magna in Kanazawa University, v. 26, p. 73–122.
Central Japan: Modern Geology, v. 15, p. 315–329. Higashino, T., Sakai, C., Kurata, H., Enami, M., Hosotani, H., Enami, M., and
Amano, K., Martin, A.J., Tanakadate, H., Kanaguri, S., Yoda, N., and Aizu, T., Banno, S., 1984, Electron microprobe analyses of rock-forming minerals
1999, Tectonics and basin formation in the arc-arc collision zone: The from the Sanbagawa metamorphic rocks, Shikoku, Part II. Sazare, Kotu
example of the South Fossa Magna in central Japan: Structural Geology and Bessi Areas: Science Report of Kanazawa University, v. 29, p. 37–64.
(Journal of the Tectonic Research Group of Japan), no. 43, p. 11–20 (in Hirano, N., Ogawa, Y., Saito, K., Yoshida, T., Sato, H., and Taniguchi, H., 2003,
Japanese with English abstract). Multi-stage evolution of the Tertiary Mineoka ophiolite, Japan: New
Aoike, K., 1999, Tectonic evolution of the Izu Collision Zone: Research Report geochemical and age constraints, in Dilek, Y., and Robinson, P.T., eds.,
of the Kanagawa Prefectural Museum of Natural History, v. 9, p. 111–151 Ophiolites in Earth History: Geological Society [London] Special Publi-
(in Japanese with English abstract). cation 218, p. 279–298.
Arai, S., 1991, The Circum–Izu Massif peridotite, central Japan, as back-arc Hiroki, Y., 1995, Sea-level changes in the early to early middle Miocene series,
mantle fragments of the Izu-Bonin arc system, in Peters, Tj., et al., eds., central Honshu, Japan: Journal of the Geological Society of the Faculty of
Ophiolite Genesis and Evolution of the Oceanic Lithosphere: Dordrecht, Science, University of Tokyo, Section II, v. 22, p. 251–284.
Netherlands, Kluwer Academic Publishers, p. 801–816. Hiroki, Y., and Matsumoto, R., 1999, Magnetostratigraphic correlation of Mio-
Arai, S., 1992, Chemistry of chromian spinel in volcanic rocks as a potential cene regression-and-transgression boundaries in central Honshu, Japan:
guide to magma chemistry: Mineralogical Magazine, v. 56, p. 173–184, Journal of the Geological Society of Japan, v. 105, p. 87–107.
doi: 10.1180/minmag.1992.056.383.04. Hisada, K., and Arai, S., 1993, Detrital chrome spinels in the Cretaceous San-
Arai, S., 1994, The Circum–Izu Massif Serpentine Belt: Geoscience Reports chu sandstone, central Japan: Indicator of serpentinite protrusion into a
of Shizuoka University, v. 20, p. 175–185, in Japanese with English fore-arc region: Palaeogeography, Palaeoclimatology, Palaeoecology,
abstract. v. 105, p. 95–109, doi: 10.1016/0031-0182(93)90109-V.
Collision between the Honshu and Izu Arcs 197

Huchon, P., and Kitazato, H., 1984, Collision of the Izu block with central Japan Program, Scientific Results, Volume 155: College Station, Texas, Ocean
during the Quaternary and geological evolution of the Ashigara area: Tec- Drilling Program, p. 147–168.
tonophysics, v. 110, p. 201–210, doi: 10.1016/0040-1951(84)90261-0. Niitsuma, N., and Akiba, T., 1985, Neogene tectonic evolution and plate sub-
Ibaraki, M., 1986, Planktonic foraminiferal datum levels recognized in the duction in the Japanese island arcs, in Nasu, N., et al., eds., Formation of
Neogene sequence of the Kakegawa area, and their relationship with Active Ocean Margins: Tokyo, Terra Scientific Publishing (TERRAPUB),
the lithostratigraphy: Journal of the Geological Society of Japan, v. 92, p. 75–108.
p. 119–134 (in Japanese with English abstract). Nureki, T., Asami, M., Shibata, T., and Ohira, K., 1982, The Ryoke Belt of
Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., and Pickle, J.D., 1984, the southwestern part of Shiaku-shoto area in the Seto-naikai Inland Sea:
The effect of grain size on detrital modes: A test of the Gazzi-Dickinson Journal of the Geological Society of Japan, v. 88, p. 499–510 (in Japanese
point-counting method: Journal of Sedimentary Petrology, v. 54, p. 103– with English abstract).
116. Ogawa, Y., and Taniguchi, H., 1988, Geology and tectonics of the Miura-Boso
Ishida, T., Arai, S., and Takahashi, N., 1990, Metamorphosed picrite basalts Peninsulas and the adjacent area: Modern Geology, v. 12, p. 147–168.
in the northern part of the Setogawa belt, central Japan: Journal of the Okuzawa, K., 2004, Reconstruction of paleogeography and unraveling of uplift
Geological Society of Japan, v. 96, p. 181–191 (in Japanese with English history in collisional zones using detrital heavy minerals—Setogawa-
abstract). Mineoka Belt and Himalaya-Bengal system [Ph.D. thesis]: Tsukuba,
Ishii, T., Robinson, P.T., Maekawa, H., and Fiske, R., 1992, Petrological studies Japan, University of Tsukuba, 324 p.
of peridotites from diapiric serpentinite seamounts in the Izu-Ogasawara- Okuzawa, K., and Hisada, K., 2004, Detrital chromian spinels from the Mio-
Mariana forearc, Leg 125, in Fryer, P., Pearce, J.A., et al., eds., Proceed- cene Haccho Formation of the Mineoka Group and the Sakuma Group,
ings of the Ocean Drilling Program, Scientific Results, Volume 125: Col- Boso Peninsula, central Japan: Journal of the Geological Society of Japan,
lege Station, Texas, Ocean Drilling Program, p. 445–485. v. 110, p. 237–243 (in Japanese with English abstract).
Isozaki, Y., 1996, Anatomy and genesis of a subduction-related orogen: A new Okuzawa, K., Yagi, N., Hisada, K., and Arai, S., 2004, Detrital chromian spinels
view of geotectonic subdivision and evolution of the Japanese Islands: from the Minokuchi and Uzuhiki Formations in the Itsukaichi area of the
Island Arc, v. 5, p. 289–320, doi: 10.1111/j.1440-1738.1996.tb00033.x. Kanto Mountains, central Japan: Tsukuba, Japan, Annual Report of the
Ito, M., 1985, The Ashigara Group: A regressive submarine fan–fan delta Institute of Geoscience, University of Tsukuba, no. 30, p. 45–51.
sequence in a Quaternary collision boundary, north of Izu Peninsula, Ono, A., 1969, Zoning of the metamorphic rocks in the Takato-Sioziri area,
central Honshu, Japan: Sedimentary Geology, v. 45, p. 261–292, doi: Nagano Prefecture: Journal of the Geological Society of Japan, v. 75,
10.1016/0037-0738(85)90005-3. p. 521–536 (in Japanese with English abstract).
Ito, M., 1987, Middle to Late Miocene foredeep basin successions in an arc-arc Ono, A., 1975a, Spatial distribution, size and chemistry of garnet in a hornfels
collision zone, northern Tanzawa Mountains, central Honshu, Japan: Sed- from the Takato area, Ryoke metamorphic belt, central Japan: Journal of
imentary Geology, v. 54, p. 61–91, doi: 10.1016/0037-0738(87)90004-2. the Japanese Association of Mineralogists, Petrologists, and Economic
Jiang, Y., 1988, Genetic mineralogy and minerogenetic characteristics of Early Geologists, v. 70, p. 169–173.
Precambrian iron formation in Nei Monggol Zizhiqu: Bulletin of the Ono, A., 1975b, Compositions and sizes of garnet crystals from a hornfels, Ten-
Tianjin Institute of Geology and Mineral Resources, v. 19, p. 1–56 (in ryukyo region, central Japan: Journal of the Japanese Association of Min-
Chinese with English abstract). eralogists, Petrologists, and Economic Geologists, v. 70, p. 194–199.
Kitahara, J., 1954, On the chromite from the Hiroosa mine, Shizuoka Prefec- Ono, A., 1976, Crystal growth and zoning of garnet from the Ryoke metamor-
ture: Journal of Mineralogy: Petrology and Economic Geology, v. 38, phic rocks of central Japan: Journal of the Japanese Association of Miner-
p. 176–186 (in Japanese with English abstract). alogists, Petrologists, and Economic Geologists, v. 71, p. 308–325.
Kosaka, T., and Tsunoda, F., 1969, Geology of the Koma Massif in the western Ono, A., 1977, Chemical reactions at the boundary between gneiss and amphib-
part of Yamanashi Prefecture: Journal of the Geological Society of Japan, olite in the Ryoke metamorphic terrain at Takato, Japan: Journal of the
v. 75, p. 127–140 (in Japanese with English abstract). Geological Society of Japan, v. 83, p. 33–40.
Koyama, A., 1991, Collision of the Kushigatayama Block with the Honshu Arc Ono, A., 1981, Geology of the Takato-Kashio area, Ryoke belt, central Japan:
during the Middle Miocene: Modern Geology, p. 15, p. 331–345. Journal of the Geological Society of Japan, v. 87, p. 249–257 (in Japanese
Koyama, A., 1993, Collision of the Kushigatayama block during the Middle with English abstract).
Miocene, northwestern part of South Fossa-Magna, central Japan: Mem- Osozawa, S., Sakai, T., and Naito, T., 1990, Miocene subduction of an active
oirs of Geological Society of Japan, no. 42, p. 245–254 (in Japanese with mid-ocean ridge and origin of the Setogawa Ophiolite, central Japan:
English abstract). Journal of Geology, v. 98, p. 763–771.
Leterrier, J., Maury, R.C., Thonon, P., Girard, D., and Marchal, M., 1982, Otofuji, Y., Matsuda, T., and Nohda, S., 1985, Opening of the Japan Sea inferred
Clinopyroxene composition as a method of identification of the magmatic from the paleomagnetism of the Japan arc: Nature, v. 317, p. 603–604,
affinities of paleo-volcanic series: Earth and Planetary Science Letters, doi: 10.1038/317603a0.
v. 59, p. 139–154, doi: 10.1016/0012-821X(82)90122-4. Parkinson, I.J., and Pearce, J.A., 1998, Peridotites from the Izu-Bonin-Mariana
Marsaglia, K.M., and Ingersoll, R.V., 1992, Compositional trends in arc-related, forearc (ODP Leg 125): Evidence for mantle melting and melt-mantle
deep-marine sand and sandstone: A reassessment of magmatic-arc prov- interaction in a supra-subduction zone setting: Journal of Petrology, v. 39,
enance: Geological Society of America Bulletin, v. 104, p. 1637–1649, p. 1577–1618, doi: 10.1093/petrology/39.9.1577.
doi: 10.1130/0016-7606(1992)104<1637:CTIARD>2.3.CO;2. Sakurai, T., 2000, Chemical compositions of the constituent minerals of the
Martin, A.J., and Amano, K., 1999, Facies analyses of Miocene subaqueous vol- Gazo mass, a tectonic block in the Sambagawa metamorphic belt, Besshi
caniclastics in the Koma Mountains, South Fossa Magna, central Japan: district, central Shikoku, Japan: Geoscience Report of Shimane Univer-
Journal of the Geological Society of Japan, v. 105, p. 552–572. sity, v. 19, p. 167–185.
Miyashiro, A., 1953, Calcium-poor garnet in relation to metamorphism: Geo- Seo, T., Suzuki, M., and Yokoyama, S., 1982, Ryoke metamorphic rocks in
chimica et Cosmochimica Acta, v. 4, p. 179–208, doi: 10.1016/0016- the Uoshima area, Southwest Japan, in Yamamoto, H., ed., Petrological
7037(53)90014-3. Study of the Ryoke Belt, Report of Grant-in-Aid for Scientific Research
Miyazaki, K., 1999, Thermobarometry of the Tsukuba Metamorphic Rocks: (A)(1981): Munakata, Japan, Fukuoka University of Education, p. 7–18
Bulletin of the Geological Society of Japan, v. 50, p. 515–525 (in Japa- (in Japanese).
nese with English abstract). Shiba, M., 1982, Metamorphic conditions of the Tsukuba metamorphic rocks:
Morimoto, N., 1988, Nomenclature of pyroxenes: Mineralogical Magazine, Journal of the Japanese Association of Mineralogists, Petrologists, and
v. 52, p. 535–550, doi: 10.1180/minmag.1988.052.367.15. Economic Geologists, v. 77, p. 345–455 (in Japanese with English
Morton, A.C., 1991, Geochemical studies of detrital heavy minerals and their abstract).
application to provenance research, in Morton, A.C., et al., eds., Develop- Shiba, M., 1989, Grandite-pyralspite garnets from the Tsukuba and Hidaka
ment in Sedimentary Provenance Studies: Geological Society [London] metamorphic rocks: Science Reports of the Hirosaki University, v. 36,
Special Publication 57, p. 31–45. p. 175–183 (in Japanese with English abstract).
Nanayama, F., 1997, An electron microprobe study of the Amazon fan, in Shimazu, M., and Ishimaru, I., 1987, Neogene volcanic rocks of the eastern part
Flood, R.D., Piper, D.J.W., et al., eds., Proceedings of the Ocean Drilling of the Koma Mountains, Yamanashi Prefecture: Journal of the Japanese
198 Okuzawa and Hisada

Association of Mineralogists, Petrologists, and Economic Geologists, Teraoka, Y., Suzuki, M., and Kawakami, K., 1998, Provenance of Cretaceous
v. 82, p. 382–394 (in Japanese with English abstract). and Paleogene sediments in the Median Zone of Southwest Japan: Bul-
Shuto, K., Takimoto, T., Sakai, A., Yamazaki, T., and Takahashi, T., 1988, Geo- letin of the Geological Society of Japan, v. 49, p. 395–411 (in Japanese
chemical variation with time of the Miocene volcanic rocks in northern with English abstract).
part of the Northeast Japan arc: Journal of the Geological Society of Teraoka, Y., Okumura, K., Suzuki, M., and Kawakami, K., 1999, Clastic sedi-
Japan, v. 94, p. 155–172 (in Japanese with English abstract). ments of the Shimanto Supergroup in Southwest Japan: Bulletin of the
Soh, W., 1986, Reconstruction of Fujikawa trough in Mio-Pliocene age and its Geological Society of Japan, v. 50, p. 559–590 (in Japanese with English
geotectonic implication: Kyoto, Japan, Memoirs of the Faculty of Science, abstract).
Kyoto University, Series of Geology and Mineralogy, v. 52, p. 1–68. Tsuchi, R., 1986, 1:200,000 Geologic Map of Shizuoka Prefecture: Shizuoka,
Sonobe, M., and Takasu, A., 2000, Major and trace element chemistry of the Japan, Shizuoka Prefecture.
garnets within the Sambagawa pelitic schists in the Asemigawa area, Uesugi, J., and Arai, S., 1999, The Shiokawa peridotite mass in the Mikabu
central Shikoku, Japan: Geoscience Report of Shimane University, v. 19, belt, central Japan, as a cumulate from intra-plate tholeiite: Memoirs of
p. 151–166. Geological Society of Japan, no. 52, p. 229–242 (in Japanese with English
Sugiyama, Y., 1995, Geology of the northern Setogawa Belt in the Akaishi abstract).
Mountains and the formation of the Setogawa accretionary complex: Bul- Wang, G., 1985, A Ca-Mn-Fe garnet found in the Ryoke metamorphic rocks at
letin of the Geological Society of Japan, v. 46, p. 177–214 (in Japanese the Wazuka area, Kyoto Prefecture: Journal of the Japanese Association of
with English abstract). Mineralogists, Petrologists, and Economic Geologists, v. 80, p. 459–462
Taira, A., Tokuyama, H., and Soh, W., 1989, Accretion tectonics and evolution (in Japanese with English abstract).
of Japan, in Ben-Avraham, Z., ed., Evolution of the Pacific Ocean Margin: Watanabe, Y., and Iijima, A., 1990, Evolution of the Tertiary Setogawa-
New York, Oxford University Press, p. 100–123. Kobotoke-Mineoka turbidite fills: Tokyo, Journal of the Faculty of Sci-
Takagi, H., and Nagahama, H., 1987, The Ryoke Belt in the Hiki Hills, north- ence, University of Tokyo, Section II, v. 20, p. 425–441.
eastern marginal area of the Kanto Mountains: Journal of the Geological
Society of Japan, v. 93, p. 210–215 (in Japanese with English abstract).
Takeuchi, M., 1994, Changes in garnet chemistry show a progressive denu-
dation of the source areas for Permian-Jurassic sandstones, Southern
Kitakami Terrane, Japan: Sedimentary Geology, v. 93, p. 85–105, doi:
10.1016/0037-0738(94)90030-2. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Cenozoic volcanic arc history of East Java, Indonesia: The stratigraphic record of
eruptions on an active continental margin

Helen R. Smyth*
Robert Hall
Gary J. Nichols
SE Asia Research Group, Geology Department, Royal Holloway University of London, Egham TW20 0EX, UK

ABSTRACT

The stratigraphic record of volcanic arcs provides insights into their eruptive
history, the formation of associated basins, and the character of the deep crust
beneath them. Indian Ocean lithosphere was subducted continuously beneath Java
from ca. 45 Ma, resulting in formation of a volcanic arc, although volcanic activity
was not continuous for all of this period. The lower Cenozoic stratigraphic record
on land in East Java provides an excellent opportunity to examine the complete
eruptive history of a young, well-preserved volcanic arc from initiation to termi-
nation. The Southern Mountains Arc in Java was active from the middle Eocene
(ca. 45 Ma) to the early Miocene (ca. 20 Ma), and its activity included significant
acidic volcanism that was overlooked in previous studies of the area. In particu-
lar, quartz sandstones, previously considered to be terrigenous clastic sedimentary
rocks derived from continental crust, are now known to be of volcanic origin. These
deposits form part of the fill of the Kendeng Basin, a deep flexural basin that formed
in the backarc area, north of the arc. Dating of zircons in the arc rocks indicates
that the acidic character of the volcanism can be related to contamination of mag-
mas by a fragment of Archean to Cambrian continental crust that lay beneath the
arc. Activity in the Southern Mountains Arc ended in the early Miocene (ca. 20 Ma)
with a phase of intense eruptions, including the Semilir event, which distributed ash
over a wide area. Following the cessation of the early Cenozoic arc volcanism, there
followed a period of volcanic quiescence. Subsequently arc volcanism resumed in
the late Miocene (ca. 12–10 Ma) in the modern Sunda Arc, the axis of which lies
50 km north of the older arc.

Keywords: East Java, Indonesia, stratigraphic record, Cenozoic arc volcanism.

*Present address: Cambridge Arctic Shelf Programme (CASP), 181A Huntingdon Road, Cambridge CB3 0DH, UK

Smyth, H.R., Hall R., and Nichols, G.J., 2008, Cenozoic volcanic arc history of East Java, Indonesia: The stratigraphic record of eruptions on an active continental
margin, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of
America Special Paper 436, p. 199–222, doi: 10.1130/2008.2436(10). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of
America. All rights reserved.

199
200 Smyth et al.

INTRODUCTION creates the central spine of the island of Java (Fig. 2). An arc of
older, Eocene to Miocene volcanoes is parallel to and south of the
The island of Java lies within the Indonesian archipelago Sunda Arc and is known as the Southern Mountains Arc (Smyth
between the landmasses of Eurasia and Australia, on the margin et al., 2005).
of the Eurasian plate. The southeastern part of this plate is known Exposures of basement rocks are rare in East Java, and
as Sundaland (Fig. 1), which is the Mesozoic continental core before this study the basement was thought to have formed dur-
of SE Asia (e.g., van Bemmelen, 1949; Hamilton, 1979). Java is ing the Cretaceous, when fragments of arc and ophiolitic mate-
located on Sundaland’s southern margin. rial were accreted to the southern margin of Sundaland (Wakita,
Many studies of volcanic arcs are based on well-exposed 2000). There are no continental basement rocks at the surface
areas of arcs active long ago, such as the Cretaceous Allistos or reported from drilling in East Java. The basement is often
Arc in Baja California (Busby et al., 2006) and the Cambrian– referred to as “transitional,” as to the north and west in Sunda-
Ordovician Lough Nafooey Arc of the Irish Caledonides (e.g., land the basement is continental, and in the east near to Flores, it
Draut and Clift, 2001). There are few studies of younger arcs is oceanic (e.g., Hamilton, 1979).
in areas where active volcanism continues today. One reason for The Cenozoic stratigraphic record preserved on land in East
this is that younger arc products cover older arc sequences. Fur- Java provides an excellent opportunity to examine the eruptive
thermore, many modern arcs are in tropical areas, such as Indo- history of a young, well-preserved volcanic arc. The present-
nesia, the Philippines, and the Western Pacific, that are difficult day arc, active since the middle to late Miocene (ca. 12–10 Ma;
to work in, commonly are not well exposed, and are remote. The Soeria-Atmadja et al., 1994), is immediately obvious, but the
modern volcanoes on Java form part of the Sunda Arc, which is importance of the older Southern Mountains Arc has previously
well known for volcanic activity, including the famous nineteenth been underestimated. The relatively basic volcanic products of
century eruptions of Krakatau and Tambora, and the Pleistocene this arc, the “Old Andesites” (van Bemmelen, 1949), are widely
eruption of Toba. However, despite a long volcanic record in known because they are well exposed and form prominent topo-
Java, owing to subduction of Indian Ocean lithosphere during the graphic features. However, the acidic products of the arc, which
Cenozoic, little is known of its pre-Pleistocene arc history. are widespread, have been largely overlooked. As a result of the
Since the middle Eocene (ca. 45 Ma) there has been north- work reported here, abundant quartz-rich sedimentary rocks in
ward subduction of the Indo-Australian plate at the Java Trench East Java are now known to be the primary and epiclastic prod-
to the south of Java (Hall, 2002). Consequently, Java is essen- ucts of acidic volcanism but were previously interpreted as conti-
tially a volcanic island, and a long history of arc volcanism is nental sediments (e.g., Harahap et al., 2003). The acidic nature of
recorded in its Cenozoic stratigraphy. Products of both active and the erupted material is a reflection of the character of the underly-
ancient volcanic arcs can be observed. An east-west–trending ing crust, which as a result of new zircon dating is now thought to
chain of >30 modern volcanoes, forming part of the Sunda Arc, include a fragment of Archean continental crust.
The full cycle of the Paleogene Southern Mountains Arc,
from initiation to termination, can be documented in the stratigra-
phy of East Java. This paper reports the stages of arc development
Malaya and the sequence and timing of events within the arc, determined
SOUTH by detailed field investigations accompanied by provenance anal-
CHINA
SEA ysis and isotopic dating. Most of the middle Miocene was marked
by a period of volcanic quiescence that followed the termination
0° Borneo of Paleogene arc volcanism in the early Miocene (ca. 20 Ma).
Sumatra Near the end of the middle Miocene (ca. 12–10 Ma) the modern
Sunda Arc began activity in its present position, 50 km north of
Ja

the Southern Mountains Arc. Possible explanations for the shift


va

SUNDALAND
in position of arc volcanism are discussed later in this paper.
Tre

JAVA
n
ch

SEA
INDIAN Java PRESENT STRUCTURE OF EAST JAVA
OCEAN

East Java can be subdivided into three parts, broadly par-


10°S
allel to the elongation of the island, representing (1) the early
100°E 110°E Cenozoic Southern Mountains Arc, (2) a deep basin north of the
arc, and (3) a marine shelf north of the basin (Figs. 3 and 4). The
Figure 1. Current plate tectonic setting of western Indonesia. The mar- modern arc is built mainly on top of the basin north of the early
gin of Sundaland (Hamilton, 1979) is defined by a black dashed line.
The black arrows show relative plate motion of boundaries—upper
Cenozoic arc. We first summarize the principal features of these
plate movement relative to lower (McCaffrey, 1996). The study area is components of Java and then go on to discuss the stratigraphy of
enclosed by a rectangle. the Southern Mountains Arc.
Cenozoic volcanic arc history of East Java, Indonesia 201

110o E o
112 E EAST JAVA SEA
o
114 E A
 Muria
igh
ang H
Remb
o
7 S
MADURA
Dieng Plateau
Kendeng Hills Surabaya

Karangsambung Madura Straits


Nanggulan EAST JAVA
Jiwo Hills

o
8 S
Yogyakarta Southern Mountains

0 50 100
INDIAN OCEAN
Kilometers

o
110 E
o
112 E
o
114 E B

o
7 S

Merapi

o
8 S

0 50 100
Kilometers

Figure 2. East Java location map. (A) Important locations and topographic features. The black star marks the epicenter of the 27 May 2006
earthquake, and the currently active volcano Merapi is marked by the black triangle. (B) Digital elevation model from SRTM data (Shuttle Radar
Transect Mission, NASA) of East Java, with the centers of Oligocene–Miocene volcanism marked.

Southern Mountains Arc interpreted to be arc and ophiolitic rocks of Cretaceous age. A
volcanic arc was built upon basement rocks from the middle
There are few exposures of basement rocks in Java, and in Eocene to the Miocene in southern Java (Smyth, 2005), which
East Java they are known only in the western part of the study is known to extend from East Java into West Java (Soeria-
area. Based upon this limited evidence, the basement has been Atmadja et al., 1994). The stratigraphic thickness of the arc
202 Smyth et al.

110o E 112o E 114o E


A
Java Sea


Section line
o
7 S SUNDA SHELF
Madura Island

KENDENG BASIN

Merapi
CENTRAL JAVA
PROVINCE

MODERN SUNDA ARC


8o S
SOUTHERN MOUNTAINS
ARC
Indian Ocean

0 50 100 B
EAST JAVA
Kilometers FORE-ARC BASIN

EOCENE TO MIOCENE
SOUTHERN
FORE-ARC
SUNDA SHELF KENDENG BASIN MOUNTAINS
BASIN
ARC
A B
?? ??
0 50 100
Km

Figure 3. Structure of East Java in map-and-sketch Eocene to Miocene profile, showing the three structural provinces—Southern Mountains Arc,
Kendeng Basin, and the edge of the Sunda Shelf—and the modern Sunda Arc building on top.

products is >2500 m, and within this sequence andesites are well the Dieng Plateau (Fig. 5). To the west, in West Java, the anomaly
known (van Bemmelen, 1949; Soeria-Atmadja et al., 1994), but is not well defined, as it becomes positive (40 μm–2).
acidic volcanic rocks have not been previously reported. The No basement rocks are exposed or known from drilling in this
zone containing the arc products is 50 km wide. The Southern region. The basin fill has an age range of middle Eocene to Mio-
Mountains Arc is now uplifted and partially eroded. The strata cene, similar to the Southern Mountains Arc (de Genevraye and
typically dip uniformly toward the south between 20° and 30°. Samuel, 1972). The basin is east-west oriented, at least 400 km
long, parallel to the Southern Mountains Arc, and filled with a
Kendeng Basin succession of volcaniclastic turbidites and pelagic mudstones that
are reported to be at least 6 km thick (Untung and Sato, 1978).
The Kendeng Basin lies directly behind and to the north of Gravity calculations indicate that the basin may contain as much
the Southern Mountains Arc (Fig. 2). The deposits of the basin are as 10 km of sediment (C.J. Ebinger, 2005, personal commun.).
poorly exposed. The depocenter is marked by a strong negative Based on field observations, seismic sections, and regional gravity
Bouguer gravity anomaly of more than −580 μm–2 (Fig. 5). The interpretation (Smyth, 2005), the basin is estimated to have been
negative anomaly can be traced eastward into a negative marine ~100–120 km wide during the early Cenozoic. It is now partially
free air anomaly (Sandwell and Smith, 1997), which extends from exposed at the surface in the Kendeng Fold-Thrust Belt, where
the Straits of Madura eastward to the north of Bali. At the west there is an estimated 10–30 km of shortening (de Genevraye and
end of the Kendeng Basin a relatively abrupt change in the charac- Samuel, 1972; Smyth, 2005); de Genevraye and Samuel (1972)
ter of the anomaly is around the modern volcanoes of Merapi and report that deformation commenced in the Pliocene.
Cenozoic volcanic arc history of East Java, Indonesia 203

110o E 112o E 114o E

Rembang Java Sea


Muria 0 50 100
Kilometers

7 S
o SUNDA SHELF Madura Island
Semarang
(geology not shown)
Dieng
Plateau
Surabaya
KENDENG BASIN

Merapi

Karangsambung
8o S
SOUTHERN MOUNTAINS
Yogyakarta
ARC
Indian Ocean
Jiwo Hills Pacitan
Southern Mountains Arc

Intrusive rocks Kendeng Basin


Kendeng

Modern volcanic and alluvial


Basin

Middle Miocene Carbonates Approximate southern deposits


Sunda Shelf

Middle Miocene reworked limit of the Kendeng


Basin (not exposed) Pliocene clastic and
volcaniclastic deposits
carbonate rocks


Oligo-Miocene volcanics
Modern Arc

Post-peneplane carbonate Fault


Eocene Sedimentary rocks Sunda Arc rocks
Basement rocks - Modern volcanic Fault (inferred)
Shelf edge clastic and
Cretaceous and older centers carbonate rocks

Figure 4. Simplified geological map of East Java, showing the main geological provinces and stratigraphic units.

Edge of Sunda Shelf of the Southern Mountains Arc. Arc activity commenced in
its present position at ca. 10 Ma in the late Miocene (Soeria-
To the north of the Kendeng Basin (Figs. 2 and 4) the hills of Atmadja et al., 1994). The volcanoes exceed 3000 m in elevation
North Java and the offshore East Java Sea constitute an area inter- and form the central spine of the island (Fig. 2). The average com-
preted to be the edge of the early Cenozoic Sunda Shelf (Hamil- position of the present-day volcanic products is basaltic andesite
ton, 1979). This region has been the focus of most hydrocarbon (Nicholls et al., 1980), which is much more basic than the average
exploration. Pre-Cenozoic basement rocks sampled by drilling composition of the Southern Mountains Arc products. Most of
are known to be ophiolitic and arc rocks, which include chert and the active volcanoes are situated ~100 km above the subducting
basic volcanic and metasedimentary rocks (e.g., Hamilton, 1979). slab (England et al., 2004). There are also a number of unusual
Basin development began in the Eocene. There are between 2000 K-rich backarc volcanoes (Edwards et al., 1991), which occur to
and 6000 m of Eocene to Pliocene shallow marine clastic and the north of the axis of the arc, including Muria (Fig. 2). The erup-
extensive carbonate sedimentary rocks within fault-controlled tive and epiclastic products of the Sunda Arc cover a significant
basins (e.g. Ardhana, 1993; Ebanks and Cook, 1993). proportion of the island, contributing to its fertile soils.
The sedimentary sequences have been deformed since the
late Miocene by numerous open, east-west–oriented folds; north- STRATIGRAPHIC RECORD OF THE EARLY
ward-verging, east-west–oriented thrusts interpreted to be Plio- CENOZOIC SOUTHERN MOUNTAINS ARC
cene in age; and ENE-WSW normal faults (Chotin et al., 1984;
Hoffmann-Rothe et al., 2001). The sedimentary rocks in the Southern Mountains Arc in
East Java were deposited on the basement above a poorly dated
Modern-Day Volcanic Arc regional unconformity (Fig. 6). The unconformity separates
Upper Cretaceous basement rocks and the Cenozoic succession,
The active volcanoes of the Sunda Arc are built mainly on suggesting a long period, potentially up to 30 m.y., during which
the Kendeng Basin (Figs. 3 and 4) but locally overlap the edge uplift and erosion occurred. The stratigraphic succession above
204 Smyth et al.

-750 to -500

 -500 to -250
-250 to 0
0 to 250
250 to 500
500 to 750
7o S SUNDA SHELF
750 to 1000
Dieng
1000 to 1250
Plateau 1250 to 1500
75
0m

1500 to 1750
KENDENG BASIN 1750 to 2000
500

µms-2
m

250m
o
8 S Merapi
SOUTHERN MOUNTAINS ARC

0 50 100

Kilometers

o o o
110 E 112 E 114 E

Figure 5. Bouguer gravity anomaly map of East Java. The map also shows topographic contours at 250 m intervals.

the unconformity in the Southern Mountains Arc and the Ken- Karangsambung, Nanggulan, and the Jiwo Hills (Fig. 2). Syn-
deng Basin has been subdivided into three “synthems” (Smyth them One is not exposed in the Kendeng Basin but was sampled
et al., 2005), each representing a different period of arc activity in blocks brought to the surface by mud volcanoes.
(Fig. 6). The term synthem has been used because much of the
work undertaken and published in East Java has been hydrocar- Southern Mountains Arc
bon oriented, and as a result the term sequence is often used with Basement exposures in East Java are rare and are found only
seismic and sequence stratigraphic implications (e.g., Catuneanu, in the western part of the area at Karangsambung and the Jiwo
2006). We have therefore chosen to avoid the term sequence for Hills (Fig. 2). The exposed basement rocks are Cretaceous in age
the major stratigraphic subdivisions and in this study used the (Parkinson et al., 1998; Wakita and Munasri, 1994) and include
term synthem, defined as an unconformity-bounded stratigraphic basaltic pillow lavas, radiolarian cherts, various metasedimen-
package (Rawson et al., 2001). The three synthems are: tary lithologies, quartz-mica schist, and high-grade metamor-
• Synthem One: records the initiation of arc volcanism and phic rocks including jadeite-quartz-glaucophane–bearing rocks
the early stages of arc development during the middle and eclogites. Quartz veins are commonly observed within the
Eocene (ca. 45 Ma) to early Oligocene (ca. 34–28 Ma). basement exposures and have not been identified in the overly-
• Synthem Two: records the growth and termination of arc vol- ing Cenozoic rocks. The high-grade metamorphic rocks have
canism in the Southern Mountains Arc during the late Oli- been interpreted to indicate subduction zone metamorphism
gocene (ca. 28–23 Ma) to the early Miocene (ca. 20 Ma). (Miyazaki et al., 1998). The basement rocks are interpreted to
• Synthem Three: records widespread carbonate growth, represent fragments of arc and ophiolitic material accreted to the
accompanied by the erosion and redeposition of rocks margin of Sundaland during the Late Cretaceous. Similar rocks
from earlier synthems during the middle Miocene (ca. 20– have previously been assumed to extend beneath the rest of East
10 Ma), with no significant volcanic activity. Java, where there is no information from surface exposures or
In the account of the stratigraphy that follows, the strati- drilling (e.g., Hamilton, 1979; Parkinson et al., 1998).
graphic ages have been converted to numerical ages, using the Synthem One is ~1000 m thick but is exposed only in the
Gradstein et al. (2004) time scale. western part of the Southern Mountains Arc in East Java. The
oldest sedimentary rocks rest directly on the basement above a
Synthem One: Initiation of the Southern Mountains Arc regional angular unconformity. They are poorly dated fluvial con-
glomerates and interbedded sandstones (Figs. 7A and 8) and are
Most parts of the Southern Mountains rocks of Synthem at least 50 m thick, but their total thickness is difficult to assess
One are covered by younger deposits and are exposed only at owing to limited and patchy exposures. These are the only rocks
TECTONIC ENVIRONMENT SEDIMENT
Age (Ma) SETTING OF DEPOSITION CHARACTER
0
KEY

VOLCANIC ACITIVITY
PLIOCENE Active arc volcanism Volcaniclastic turbidites

RESURGENCE OF
UPLIFT Erosion or non deposition Crystal-rich volcaniclastics
AND
Subsidence Proximal volcanic rocks
5 EROSION
Super-eruption Siliciclastic sediments
Reefal limestones Basal conglomerates
Foraminiferal packstones Unconformity
10
CARBONATE ROCKS
SYNTHEM THREE

PLATFORMS AND
TURBIDITE FANS
EPICLASTIC

CARBONATE
REWORKING OF
EXTINCT VOLCANIC
VOLCANICS +
REWORKED

ARC AND
15 WIDESPREAD
MIOCENE

CARBONATE
DEPOSITION

Fig.11A
SEMILIR ERUPTION
20
EXPLOSIVE VOLCANISM

TERRESTRIAL-MARINE
VOLCANIC DEPOSITS

VOLCANIC ISLANDS
SYNTHEM TWO

IN TERRESTRIAL-
SHALLOW MARINE-
OLIGO-MIOCENE

EXPLOSIVE SLOPE. SOME Fig.10


VOLUMINOUS EPICLASTIC
25 REWORKING.
VOLCANISM
OLIGOCENE

DEPOSITION IN THE
ARC, INTRA-ARC,
FORE-ARC AND
BACK-ARC.

30
EUSTATIC EVENT APPARENT GAP IN SECTION
NO EVIDENCE OF
DEFORMATION

Fig.7C
TERRESTRIAL-SHELF-BASIN

35
EARLY BASIN SEQUENCE

OPEN - DEEP
TRANSGRESSION

MARINE
EOCENE
SYNTHEM ONE

40 SLUMPING
VOLCANIC
ACTIVITY SHALLOW MARINE

Fig.7B
SUBSIDENCE DELTAIC
45
Fig.7A
NO VOLCANIC TERRESTRIAL
ACTIVITY

~30 m.y. time gap


CRETACEOUS

CRETACEOUS
+ OLDER
BASEMENT

Figure 6. Simplified stratigraphic column of the Southern Mountains Arc. The column shows the three synthems men-
tioned in the text and indicates the phases of arc development recorded by the sedimentary succession.
206 Smyth et al.

A Basal part of Synthem One


Clay Silt Sst Cong
B Lower part of Synthem One
Clay Silt Sst Cong
C Upper
Clay
part of Synthem One
Silt Sst Cong
1.7 TOP

10 1.6

3
1.5

9
1.4
Thickness (m)

Lithology Logs A and B


Thickness (m)

Thickness (m)
Organic-rich muds
8 2 Quartz-rich sandstones
+ siltstones

1.2 Conglomerates
Lithology Logs C

Increasing volcanic material


7 Crystal-rich
1.1 volcaniclastic sandstone
Tuffaceous mudstones Figure 7. Stratigraphic logs illustrating
2 + siltstones features of the type sections of Syn-
1m
Sedimentary structures them One. (A) Basal conglomerates
6
and sandstones of the Lukulo Member,
Rip up clasts
0.9 Karangsambung Formation. (B) Quartz-
Parallel lamination rich sandstones and interbedded organ-
5 0.8
Cross lamination ic-rich muds of the middle Eocene Kali
Trough cross Songo Member, Nanggulan Formation.
lamination Volcanic components increase upsec-
Convolute lamination
0.7 tion. (C) Volcanogenic sandstones and
Channels tuffaceous mudstones of the Jetis Mem-
4
0.6
Irregular erosive ber, Nanggulan Formation.
bed base
Planar sharp bed base
1
0.5 Conglomerate
3

0.4

2
0.3

0.2
1

0.1

0 0 0
BASAL TERRESTRIAL SANDSTONES + TERRESTRIAL - SHALLOW MARINE MARINE VOLCANOGENIC
CONGLOMERATES SANDSTONES + ORGANIC MUDSTONES SANDSTONES + MUDSTONES
o, o
Loc. 7.5477 S, 109.6694 E Loc. 7.7323 S, 110.1972 E Loc. 7.7395 S, 110.1887 E

exposed onshore that contain no fresh volcanic material (Fig. 9). and mudstones (Fig. 8). The middle Eocene age of the lower
They are dominated by quartz grains and metamorphic and igne- part of the succession is based on palynomorphs in the coals
ous lithic clasts (Figs. 8 and 9), including vein and metamorphic and organic-rich mudstones (Lelono, 2000), and the occurrence
quartz, chert, phyllite, schist, metasedimentary rock, and basalt. of Nummulites and Discocyclina higher in the section. The
The lithologies identified as clasts are typical of those exposed in lower part is at least 200 m thick and is a noncalcareous unit of
the Cretaceous basement (described above) and are interpreted to well-bedded coals, conglomerates, quartz-rich sandstones, and
be the product of erosion and reworking of these basement rocks. organic-rich muds. The coals and conglomerates are restricted
The terrestrial conglomerates and sandstones lack palynomorphs to the lower 50 m. The conglomerates are commonly channel-
and so cannot be directly dated, but they are overlain by a succes- ized, are ~10–75 cm thick, and contain a range of lithic clasts
sion of well-dated middle Eocene strata (Lelono, 2000). similar to those identified in the basal section directly overly-
The middle Eocene to lower Oligocene of the Southern ing the basement. The coals vary laterally in thickness from 5
Mountains Arc is represented by a transgressive succession, from to >50 cm and have an average vitrinite reflectance of 0.4% Ro
base to top, of coals and conglomerates, sandstones, siltstones, (Smyth, 2005). For normal and high geothermal gradients typical
Cenozoic volcanic arc history of East Java, Indonesia 207

A WEST EAST B
Foraminiferal packstone
30 750 Slump - schist and
foraminifer mixture
OLIGOCENE

Volcaniclastic sandstones
Conglomerate clast
Quartz-rich sandstones lithology types
Quartz
Basal Conglomerate Chert
Phyllite
35
Basalt
500
Metasediments

C D
40
EOCENE

250
Thickness (m)

45

Cretaceous and older basement lithologies

E F

Figure 8. Field observations of Synthem One. (A) Sketch stratigraphy of the middle Eocene–early Oligocene deposits. Volcanic material has
increased upsection from ca. 42 Ma. (B) Field photograph of the basal conglomerate. The pie chart shows the abundance of quartz and lithic
clasts. These conglomerates contain no acidic volcanic material. (C) Quartz-rich sandstones and coals from the lower part of Synthem One. (D)
Interbedded quartz-rich sandstones and organic-rich and tuffaceous mudstones, upsection from photo C. The sandstones contain a significant
percentage of volcanic quartz, plagioclase feldspar laths, and pyroclastic zircons. (E) Volcanogenic mudstone rich in planktonic foraminifers. (F)
Volcanogenic sandstones with abundant volcanic lithic fragments.

of arc regions this would imply very shallow burial depths of <1 pumice and andesite. In addition, a significant proportion of
or 2 km (Madon et al., 1997; Watts, 1997). the quartz has a volcanic origin (the identification of volcanic
Above this transgressive succession lie thinly laminated quartz is discussed further below). At the base of the quartz-
quartz-rich sandstones and organic-rich mudstones that are at rich sandstones the quartz is dominated by metamorphic grains,
least 150 m thick. These sediments were deposited in a ter- but this content gradually decreases in its relative contribution
restrial setting with some marine incursions indicated by the upsection, and at the top, volcanic quartz is dominant (Fig. 9B).
presence of calcareous zones rich in Nummulites and Discocy- The heavy mineral assemblage contains zircons that yielded
clina. In addition to the organic-rich mudstones there are sev- spot sensitive high-resolution ion microprobe (SHRIMP) U-
eral gray mudstones, which are rich in smectite. The quartz-rich Pb ages of 41.8 ± 1.6 and 42.7 ± 1.5 Ma (Smyth, 2005). These
sandstones contain metamorphic lithic clasts, like those of the ages are the same as the biostratigraphically determined ages
conglomerates, but also contain volcanic lithic clasts such as for the host rocks, indicating that volcanic activity occurred
208 Smyth et al.

Basement
A B TOP

80 20

60 40 Up-section
increase in volcanic
quartz within the
quartz-rich
sandstones of
40 60 Synthem One (Fig
7B), accompanied
by a decrease in
metamorphic
quartz.
20 80

80 60 40 20 Other
Volcanic
sources

BOTTOM
Pre-Middle Eocene basal Late Eocene-Oligocene
conglomerates and sandstones volcaniclastic deposits
(Fig. 7A) (Fig. 7C)
Thin section analysis of quartz type %
Middle Eocene quartz-rich Up-section increase in volcanic Metamorphic Sedimentary
sandstones (Fig. 7B) material Volcanic Unknown
Plutonic

C D

Figure 9. Petrographic details of Synthem One. Scale bars are 1 mm. (A) Basement, volcanic, and other sources are de-
picted in a triangular diagram, showing the upsection trend identified in East Java, with a reduction in basement-derived
material and an increase in volcanic contribution. (B) Pie charts showing the quartz types identified within quartz-rich
sandstones of East Java; example from the Nanggulan Formation. Volcanic quartz increases upsection. (C) Photomi-
crograph of the basal sandstone, dominated by metamorphic quartz. (D) Volcaniclastic turbidite dominated by laths of
plagioclase feldspar.

as the sediments were deposited. Higher in the section, in the zeolites (Fig. 9A). These sandstones are interbedded with tuff-
sandstones, a number of Nummulites-rich zones mark the onset aceous mudstones. Water depth is uncertain but is interpreted
of fully marine conditions. to be deeper than that for the nummulitic zones. There are no
Above the nummulitic units the sandstones become diagnostic sedimentary structures, but the sandstones contain
increasingly arkosic and rich in fresh laths of plagioclase feld- pelagic foraminifers.
spar (Fig. 9D), volcanic quartz, volcanic lithic fragments, elon- Higher in the section the presence of volcaniclastic turbi-
gate volcanic zircons, and volcanic clays such as smectite and dites indicates an increase in water depth. Some turbidites are
Cenozoic volcanic arc history of East Java, Indonesia 209

characterized by Bouma divisions A, C, and E, but more com- (Fig. 4) and within the fold-thrust belt of the Kendeng Basin. The
monly C, D, and E. They are >150 m thick, with individual oldest rocks of this synthem in the Southern Mountains Arc are
bed thicknesses ranging from 5 to 50 cm. The turbidites have a reworked bioclastic tuffaceous mudstones dated biostratigraphi-
diverse planktonic foraminiferal assemblage (Fig. 8E), includ- cally as Oligocene, NP24 (27.3–30 Ma; M. Fadel, 2002, personal
ing Helicosphaera euphratis, H. reticulate, and H. wolcoxonii commun.).
(P. Lunt, 2002, personal commun.), which provides an age of Throughout the late Oligocene and early Miocene the vol-
NP18 (36.8–36.2 Ma) for the lower parts of the succession. The canic activity in the Southern Mountains Arc was extensive,
turbidites extend into the lower Oligocene, based on biostratig- explosive, and intermediate to acidic in composition. The depos-
raphy (P. Lunt, 2002, personal commun.). They are dominated its range from andesite to rhyolite, with an average SiO2 con-
by volcanic debris (Fig. 9) such as laths of plagioclase feldspar, tent of 67 wt% (Smyth, 2005), and include thick mantling tuffs
volcanic lithic clasts, and volcanogenic clays, and there are no (Fig. 10A), crystal-rich tuffs, block and ash flows, pumice-lithic
metamorphic lithic clasts within the coarser beds. The heavy breccias, andesitic breccias (Fig. 10B), silicic lava domes, and
mineral assemblage is dominated by volcanic zircons, which lava flows. The Oligocene–Miocene volcanic centers can be
yielded a weighted mean SHRIMP U-Pb age from 17 grains of mapped (Smyth, 2005) by the occurrence of vent proximal facies,
41 ± 1.4 Ma, similar to the biostratigraphic age of the lower part and at least 13 centers are presently exposed (Fig. 2), which show
of the section, suggesting some reworking. a similar spacing to the volcanoes of the present-day arc. The
The upper boundary of Synthem One is an intra-Oligo- thickness of the proximal volcanic deposits ranges from 250 to
cene unconformity, which is interpreted to be the result of sea >2000 m.
level change, as the sedimentary rocks directly above and below Thick (>700 m) successions of volcaniclastic sedimentary
the gap have similar bedding orientations with no indication of rocks surround the volcanic centers. These reworked deposits are
deformation. This could be a local sea level change but could commonly interbedded with beds of unreworked mantling tuffs
also be a global change. Unconformities of this age that record a and volcanic breccias >1 m in thickness. The reworked volcani-
global intra-Oligocene sea level fall are widely known, from the clastic beds vary in thickness from 5 to >100 cm and are crystal-
Haq et al. (1987) curve (30 Ma), the Marshall Paraconformity and volcanic lithic-rich sandstones and tuffaceous mudstones.
(32–29 Ma) in the Canterbury Basin, New Zealand (Fulthorpe et They commonly contain abundant fragments of charcoal, indi-
al., 1996), and the carbonates of Baldwin County, Alabama, USA cating the presence of vegetated slopes on the terrestrial volcanic
(Baum et al., 1994). centers. Both terrestrial and shallow marine deposits have been
Within Synthem One the contribution of volcanic mate- identified; the latter do not contain abundant bioclastic mate-
rial increases upsection as the proportion of basement-derived rial but are weakly calcareous, and their upper bedding surfaces
material (metamorphic quartz, metamorphic lithic fragments, are commonly intensely bioturbated with Cruziana-type facies
and illite, chlorite, and serpentinite clays) decreases, and the traces. There are slump folds and mass wasting deposits on the
upper Eocene sedimentary rocks are almost entirely dominated flanks of the volcanic centers.
by volcanic debris (Fig. 9A, B). The volcanic centers supplying In the Southern Mountains Arc is a record of a major erup-
this material cannot be separately mapped owing to their limited tion, the Semilir Eruption, toward the end of the period of arc
exposure, but they are interpreted to follow the same trend and activity. Extensive deposits of this eruption are widespread to
occupy the same positions as the volcanoes of the Oligocene– the east of Yogyakarta (Fig. 4). The Semilir and Nglanggran
Miocene arc (see below). Formations (Fig. 10A, B) are the products of this event, and
they were deposited in a short period, possibly during one
Kendeng Basin eruptive phase between 21 and 19 Ma (Smyth, 2005). Based on
In the Kendeng Basin there are limited exposures, and Syn- measured sections, the combined thickness of the two forma-
them One is not seen at the surface, but blocks of the older litholo- tions varies between 250 and 1100 m. They are well exposed
gies have been brought to the surface by Pleistocene and modern over an area of 800 km2, and the total volume of volcanic mate-
mud volcanoes. These blocks are terrestrial and are composed rial is estimated to be at least 480 km3. The Semilir Formation
of shallow marine sandstones and conglomerates (de Genevraye (Fig. 6) is a thick accumulation of dacitic air-fall, pyroclastic
and Samuel, 1972) similar in character to the middle Eocene sed- surge and flow deposits produced by an explosive eruption.
imentary rocks in the Southern Mountains Arc. Both terrestrial and shallow marine deposits have been identi-
fied, indicating that the erupted material entered the sea from
Synthem Two: Growth and Catastrophic Termination of the flanks of the volcanic center. The Nglanggran Formation
Arc Volcanism (Fig. 6) is a series of monomict andesitic volcanic breccias.
The individual beds are up to 10 m thick, have flat bases and
Southern Mountains Arc tops, and can be mapped for tens of kilometers. The breccias
The upper Oligocene to lower Miocene deposits of Synthem contain some blocks >3 m across. They are interpreted as vent
Two are primary volcanic rocks and epiclastic rocks. These rocks proximal facies such as flow breccias or block-and-ash–flow
are exposed extensively throughout the Southern Mountains Arc deposits. These deposits mark the end of volcanism in the
210 Smyth et al.

A B

Basement
C
D

80 20

60 40
Quartz Type %
Volcanic Unknown

40 60
SOUTHERN MOUNTAINS ARC
Kebobutak
Jaten
20 80

Volcanic 80 60 40 20 Other
Sources

Figure 10. Details of Synthem Two. (A) Field photograph of thick, well-bedded ashes (bar is 1m). (B) Field photograph
of monomict andesitic breccia; the breccia clasts commonly exceed 1 m in size (bar is 1m). (C) Basement, volcanic, and
other sources are depicted in a triangular diagram, showing a selection of formations from Synthem Two. These deposits
lack material of basement origin. (D) Pie chart showing the quartz types identified petrographically in a quartz-rich sand-
stone in the upper part of Synthem Two. The quartz is dominated by volcanic quartz; the image is a bipyramidal grain
from the Jaten Formation, Pacitan (bar is 1m).

Southern Mountains Arc. There is no significant break within (Smyth, 2005). Volcanic zircons from the sandstones form a
or between the Semilir and Nglanggran Formations, based on single population with ages that are the same as that of the
zircon dating and biostratigraphic dating of the overlying for- Semilir Formation (Smyth, 2005) and are contemporane-
mations (Smyth, 2005). ous with the Semilir Eruption. The quartz within these sand-
A number of quartz-rich sandstones (Fig. 10D) are in the stones is entirely of volcanic origin (Fig. 10D). Volcanic quartz
upper part of Synthem Two in the Southern Mountains Arc grains commonly appear clear and bright in thin section, have
Cenozoic volcanic arc history of East Java, Indonesia 211

nonundulose extinction, and are monocrystalline (Leeder, Synthem Three: Reworking of the Southern Mountains Arc
1982). These features and other characteristics, including
well-developed crystal faces, bipyramidal shape, melt embay- Southern Mountains Arc
ments, and melt inclusions, can be used to distinguish volcanic In the Southern Mountains, to the south of the Oligocene–
quartz from other types such as metamorphic, vein, plutonic, Miocene volcanic centers, extensive calcareous volcanogenic
and sedimentary quartz. Volcanic crystal-rich sandstones can turbidites (Fig. 11) occur with Bouma divisions A, C, D, and E.
be produced by primary eruptive mechanisms and/or second- The turbidites are at least 500 m thick and have been dated within
ary epiclastic processes. The sandstones and other quartz-rich nannofossil subzones NN2–NN8 (Kadar, 1986) that correspond
volcaniclastic rocks in East Java previously were interpreted to dates between 19 and 10 Ma. Beds vary in thickness from 20
as continental siliciclastic deposits (Harahap et al., 2003), and to 75 cm, and upper bedding surfaces are commonly bioturbated
it is for this reason that many of the acidic products of the with traces of Cruziana facies. Flute casts indicate that the flow
Southern Mountains Arc have been overlooked (as discussed direction was toward the southeast, and there are slump folds
below). Here they are interpreted to have a volcanic origin and with southeasterly vergence. Thin section examination shows
to be largely the product of the Semilir Eruption with some that volcanic crystals and lithic fragments are well rounded, and
subsequent reworking. there is no fresh or unreworked volcanic debris present. Dating of
zircons (see below) supports the field and petrographic evidence
Kendeng Basin that these are reworked volcaniclastic deposits, as the zircon ages
Synthem Two exposures in the Kendeng Basin are very lim- are similar to the age range of rocks in Synthem Two. Synthem
ited, occurring only in a small thrust-bound sliver. They comprise Three is therefore interpreted to be the product of reworking of
poorly lithified Globigerina-rich tuffaceous mudstones ranging older arc rocks of Synthem Two rather than the product of con-
in age from late Oligocene to early Miocene (de Genevraye temporaneous volcanism.
and Samuel, 1972). The mudstones are at least 85 m thick (de Several tuff beds have been identified at the top of the tur-
Genevraye and Samuel, 1972) but have limited surface exposure. bidite sequence in the Southern Mountains but are not the result
The mudstones are volcanogenic and apparently lack continen- of volcanic activity in the Southern Mountains Arc. The tuffs are
tal terrigenous material. The volcanic material is fine grained, distal air-fall deposits and yield zircons with ages between 12
reworked, and distal in character, very different from that identi- and 10 Ma (P.J. Hamilton, 2003, personal commun.). These tuffs
fied within the arc at this time. The abundance of Globigerina are the product of the modern Sunda Arc (Soeria-Atmadja et al.,
and other planktonic foraminifers is indicative of pelagic sedi- 1994), and the ages mark the initiation of volcanic activity some
mentation in an open marine setting at water depths of a few hun- 50 km to the north of the Southern Mountains Arc.
dred meters or more. Thick volcanogenic turbidites are reported In addition to the turbidites, there are the first widespread
from wells in this area (P. Lunt, 2002, personal commun.), but no carbonates in the Southern Mountains (Fig. 11D). The carbon-
descriptions of the stratigraphy or well log interpretations have ates range in age from late early Miocene to middle Miocene
been published. (Lokier, 2000; Smyth, 2005) and formed isolated reefs and exten-
sive carbonate platforms, which can be observed overstepping
Edge of the Sunda Shelf the deposits of Synthem Two. The limestones are at least 200 m
The offshore Eocene to Pliocene successions in the East Java thick and are the source of the carbonate within the volcanogenic
Sea that have been investigated during hydrocarbon exploration turbidites.
are not reported to contain any volcanic material (e.g., Matthews
and Bransden, 1995). Field investigations during this study in Kendeng Basin
northeast Java have identified distal volcanic material on the Volcanogenic sedimentary rocks identified in the Kendeng
edge of the Sunda Shelf within carbonates, including thin tuff Basin include channelized volcanic lithic conglomerates, crystal-
layers, zones rich in smectite clays, and concentrations of volca- rich and volcanic lithic-rich sandstones, tuffaceous mudstones,
nic quartz and zircons. This suggests that there may be more vol- and Globigerina mudstones. A thickness of up to 400 m of rocks
canic debris offshore than recognized up to now, and that some of is exposed (Smyth, 2005), and hydrocarbon exploration shows
the clay layers may be fine air-fall ash deposits. there is up to 3000 m in the basin (de Genevraye and Samuel,
The stratigraphic record of Synthem Two within the South- 1972). The sandstones contain Bouma divisions B, C, and E and
ern Mountains Arc and Kendeng Basin accounts for only vol- are locally cut by the channelized conglomerates. These conglom-
canic and volcaniclastic rocks at this interval. There is no evi- erates can cut as much as 20 cm into the underlying beds and are
dence of input of material from basement or other sources. This up to 1 m thick. Slump folds are common, and most verge toward
indicates that the volcanic arc was the only source of sediment the northeast. Measurements of scours, grooves, channel struc-
at this time. There are no significant exposures of carbonates tures, flutes, and slumping indicate that sediment was transported
within Synthem Two in the Southern Mountains Arc or Kend- from the south toward the north. The Globigerina mudstone is
eng Basin, but carbonates do occur farther to the north on the similar in character to those of Synthem Two. A sample from
edge of the Sunda Shelf. the top of the exposed section yielded a biostratigraphic age
212 Smyth et al.

A Clay Silt Sand Loc. 7.8727oS B


F M C 110.5496oE

20
Thickness (m)

Bioturbation
Asymmetrical ripples
Parallel lamination
Convolute parallel
lamination
Breccia
C
Pebbles
15 Figure 11. Stratigraphic details of Syn-
them Three. (A) Logged section typi-
cal of the volcanogenic turbidites of
Synthem Three, Sambipitu Formation.
Synthem Three

(B) Field photograph showing asym-


metrical ripples on the upper bedding
surfaces of the volcanogenic turbidites,
Sambipitu Formation. (C) Field photo-
graph of volcanogenic turbidites within
10 the Kendeng Basin, Kerek Formation.
(D) Field photograph of the reefal car-
bonates typical of Synthem Three in the
Southern Mountains Arc, Campurdarat
Formation.

D
5

Upper part of
0 Synthem Two

of foraminifer subzones N8–N9 (M. Fadel, 2003, personal com- in the East Java Sea (Bishop, 1980; van Bemmelen, 1949), and
mun.) corresponding to ages between 17.2 and 14.4 Ma. redeposited on the edge of the Sunda Shelf.

Edge of the Sunda Shelf INTERPRETATION OF THE EOCENE TO MIOCENE


On the edge of the Sunda Shelf are a number of lower to mid- STRATIGRAPHY
dle Miocene quartz-rich sandstones. They were previously inter-
preted to have a continental provenance and to have been derived Initiation of the Southern Mountains Arc
from reworking of basement rocks of Sundaland (e.g., Ardhana,
1993; Sharaf et al., 2005), but new studies (Smyth et al., 2007) The first evidence of arc volcanism in East Java is identified
have shown that these sandstones contain a significant proportion as middle Eocene at ca. 42 Ma with the first occurrence of tuff
of volcanic quartz, volcanic zircons, and reworked volcanogenic layers, volcanic lithic fragments, volcanic quartz, laths of plagio-
clays. These volcanic particles are interpreted to have fallen as clase feldspar, and volcanic zircons. Prior to this, the sedimentary
ash onto the Sunda Shelf, to have been subsequently reworked rocks, which are the oldest rocks exposed on East Java, lack evi-
and mixed with material derived from uplifted basement blocks dence of contemporaneous arc volcanism and are the only rocks
Cenozoic volcanic arc history of East Java, Indonesia 213

exposed on East Java that contain no volcanic debris. This sug- reef growth. Volcanogenic turbidites built out as thick aprons
gests that there was no arc volcanism on Java prior to 42 Ma, northward from the inactive arc into the Kendeng Basin. Sedi-
which also implies no subduction of the Indian-Australian plate mentation was rapid and on an unstable northward-dipping slope,
beneath Java at this time. and the deposits began to fill the accommodation space within
the basin. Material also traveled southward from the eroding arc
Growth and Development of the Southern Mountains Arc into the Java forearc. On fault-bounded highs or in the shelter of
the extinct volcanoes, reefs and carbonate platforms developed,
Following the initiation of volcanism in the Southern Moun- marking the first period of extensive carbonate growth within the
tains the contribution of volcanic debris increased rapidly, and Southern Mountains during the middle Miocene (ca. 20–10 Ma).
the contribution of basement-derived material decreased. The Arc volcanism resumed in the late Miocene, after a lull of
volcanic centers probably formed an east-west chain of volcanic ~8 m.y., 50 km to the north of the extinct Southern Mountains
islands separated by interarc basins, much like the present Izu- Arc at the position of the modern Sunda Arc.
Bonin-Mariana and Aleutian Islands Arcs. A narrow volcaniclas-
tic shelf built up around the isolated volcanic islands, where thick CHARACTER OF THE CRUST BENEATH THE
sequences of volcanic and epiclastic material were deposited. To SOUTHERN MOUNTAINS ARC
the north and directly behind the arc, the Kendeng Basin began
to subside, and material was transported from the volcaniclastic Little is known of the crust beneath many young arcs, espe-
shelf into the basin. cially in Indonesia, and East Java is no exception. However, study
No reef carbonates are preserved within Synthem Two near of the stratigraphy of the Southern Mountains Arc has provided
the arc, suggesting that environmental conditions, such as the some new and surprising insights into the character of the deep
influx of volcanic detritus, prevented reef growth or, alternatively, crust beneath the arc.
that these carbonates were not preserved or exposed. However, As discussed above, exposures of basement rocks in East
some distance to the north, on the edge of the Sunda Shelf, a Java are limited and are restricted to the west of the study area.
large area of carbonate platform was developing at this time; These rocks have been interpreted to be fragments of arc and
the shelf received volcanic debris as air fall, but much less than ophiolitic material accreted to the continental margin of Sunda-
that deposited close to the arc, and the volcanic material did not land during the Late Cretaceous (Hamilton, 1979; Wakita, 2000).
substantially inhibit carbonate growth. Wilson and Lokier (2002) It has been generally considered that these rocks are typical of the
showed that volcanic input can influence carbonate development basement beneath the whole of East Java. This is supported by oil
close to active arcs without killing carbonate-producing organ- company drilling on land in East Java and farther to the north in
isms and causing carbonate deposition to cease. the East Java Sea, where deep wells penetrate a varied basement,
During the Oligocene to early Miocene, volcanic activity including basic and acidic volcanic rocks, slaty metasedimentary
along the Southern Mountains Arc was at its most voluminous rocks, quartzites, and cherts (e.g., Hamilton, 1979; Matthews and
and explosive. The volcaniclastic shelf close to the active volcanic Bransden, 1995). However, the abundance of acidic volcanism
centers increased in width and thickness. Material was fed into the in the Southern Mountains Arc and the dating of zircons indicate
Kendeng Basin, and only in the deepest part of the basin was there that the crust beneath much of the southern part of East Java is
pelagic sedimentation without a volcanogenic component. very different, and is much older.

Termination of Volcanism in the Southern Mountains Arc Volcanism

Following a long period of arc activity between 42 and The intermediate to acidic volcanic rocks of the Southern
18 Ma, volcanism in the Southern Mountains Arc ceased. The Mountains Arc range in composition from andesite to rhyolite
final stages of activity were marked by a phase of explosive vol- (60–77 wt% SiO2), with an estimated average SiO2 content of
canism, which included a major event, the Semilir Eruption. The 67 wt%. In addition, many of the high-level intrusive bodies
age range, thicknesses, area, and estimated volumes of volcanic exposed within the Southern Mountains Arc are granodiorites.
deposits in the vicinity of the Toba volcanic center (Chesner These rocks are considerably more evolved than the basic to
and Rose, 1991) are similar to those of the Semilir center, and intermediate eruptive products of the present-day volcanoes. The
the deposits of the Semilir Eruption may be distributed, like the chemistry of the modern arc volcanic rocks in East Java (Hand-
Youngest Toba Tuffs (Song et al., 2000), over large parts of SE ley, 2006; Nicholls et al., 1980; Wheller et al., 1987) indicates
Asia. Work is in progress to assess their distribution. that they are the products of relatively primitive subduction melts,
which reveal no significant interaction with underlying crust. The
Lull and Resurgence of Volcanic Activity difference in chemistry of the products of the Southern Moun-
tains Arc and the Sunda Arc could be explained in a number of
Following the termination of volcanism, there was wide- ways: fractional crystallization of more basic magma, interaction
spread erosion of the deposits of Synthem Two and extensive of more basic magma with felsic crust, and partial melting of
214 Smyth et al.

felsic crust. Evidence from zircons (discussed below) suggests If all the zircon ages are grouped together, five age peaks
that involvement of continental basement in petrogenesis is the can be identified (Fig. 13): a Cenozoic peak (5–42 Ma) recording
most likely explanation. volcanism in the Southern Mountains and Sunda Arcs, a Creta-
ceous peak (65–135 Ma) consistent with the interpreted age of
Ancient Zircons basement, a Cambrian to Neoproterozoic peak (500–1000 Ma), a
Mesoproterozoic to Paleoproterozoic peak (1000–2500 Ma), and
At the beginning of this study there were very few good dates an Archean peak (2500–3200 Ma). The Cambrian and older ages
for volcanic rocks of the older arc on Java (50 analyses), and were not expected, based on knowledge of the regional geology.
all published ages were K-Ar dates (Ben-Avraham and Emery,
1973; Soeria-Atmadja et al., 1994). No acidic rocks had been Source of the Ancient Zircons
dated. During this study, zircons from 16 Cenozoic samples of
igneous (9), volcaniclastic (3), and sedimentary rocks (4) were Sundaland, the continental core of SE Asia, is the closest
dated by the SHRIMP U-Pb method at Curtin University of Tech- and most obvious source of the very old zircons (Figs. 1 and 13).
nology, Australia, using the methods described in Smyth et al. Rocks that could have provided abundant old zircons include gran-
(2005, 2007). More than 453 spot ages were measured, and 270 ites of SW Borneo (Hamilton, 1979), the Malay Peninsula (Liew
of these were Cretaceous and older. and Page, 1985), the Thai-Malay Tin Belt (Cobbing et al., 1986),
Prior to SHRIMP dating, the expected age range of the zir- and Sumatra (Imtihanah, 2000; McCourt et al., 1996) (Fig. 13).
con samples was Cenozoic to Cretaceous. This reflects the age of However, in these areas there are no known Archean rocks, nor
arc activity known from Java, and a contribution from the Cre- is there any indication that they are underlain by Archean crust.
taceous arc and ophiolitic terranes interpreted to form the base- Geochemical and isotopic studies suggest a basement no older
ment of East Java. However, in addition to the expected ages, an than Proterozoic in areas of Sundaland, such as the Thai-Malay
unexpected range of Archean–Cambrian grains was identified in peninsula, that have been studied (e.g., Liew and Page, 1985). The
a large number of samples (Smyth, 2005, 2007). The samples closest area with extensive granites is the Schwaner Mountains of
containing Cretaceous ages and those containing Cambrian and SW Borneo. Paleogene sedimentary rocks of north Borneo contain
older zircons occur in distinct areas of East Java. debris eroded mainly from the Schwaner granites and the Malay
Tin Belt, including detrital zircons (van Hattum et al., 2006). The
Lithology and Zircon Age Range zircon populations are dominated by Cretaceous zircons with age
The five intrusive and four extrusive igneous rocks analyzed peaks different from those of East Java; there are abundant Perm-
contain a range of Cambrian and older zircons (n = 155) but lack ian–Triassic zircons, rare in East Java; and Precambrian zircons
Cretaceous zircons. In contrast, the three volcaniclastic rocks are mainly Paleoproterozoic with very rare Archean grains. The
analyzed yielded a significant number of Cretaceous zircons (n differences in zircon ages rule out a Schwaner or Thai-Malay
= 42) but only one Cambrian or older zircon. The four quartz- provenance, and other granite sources in Sundaland are even more
rich sandstones analyzed yielded varied zircon ages. All of these distant and paleogeographically unlikely.
sandstones contain Cretaceous grains (n = 33). Three sandstones The largest and closest area of continental crust of Archean
from the Southern Mountains and Kendeng Basin contain a range age is Australia, which formed part of Gondwana until the Cre-
of Cambrian and older grains (n = 22). One sandstone from the taceous. It has been suggested that small Gondwana continental
edge of the Sunda Shelf contains a single Jurassic–Triassic grain fragments collided with the east Sundaland margin in Sulawesi
but no older grains. during the Cretaceous (e.g., Wakita et al., 1996; van Leeuwen and
Muhardjo, 2005), although the source of the fragments was not
Age Ranges and Spatial Distribution identified. There is evidence of Late Jurassic–Early Cretaceous
The samples containing different zircon ages do have a clear rifting of continental fragments from the northwest and west Aus-
geographical pattern. The Archean grains are found only in rocks tralian margins (Müller et al., 2000) during continental breakup
of the Southern Mountains (Fig. 12B), and the igneous rocks of preceding the separation of India from Gondwana. One of these
this area contain no Cretaceous zircons (Fig. 12A). Cretaceous continental fragments could have collided with the Sundaland
zircons are found only in the north and west of East Java, and margin. Bergman et al. (1996) speculated that there could be
these zircons have been identified only in sedimentary and vol- old continental crust beneath west Sulawesi, on the basis of lead
caniclastic rocks. This is interpreted to indicate that the South- isotopic compositions of Neogene plutonic rocks, which were
ern Mountains are underlain by material providing Cambrian to suggested to have had an Australian origin. Basement rocks in
Archean zircons but no Cretaceous zircons. To the north and west western Australia are dated as Proterozoic and Archean. Detrital
of the Southern Mountains the underlying rocks do not include zircons from young sediments of the Perth Basin, derived from
Cambrian to Archean material but do include Cretaceous mate- the erosion of western Australian basement, have been well stud-
rial. Figure 12C shows the distribution of basement rocks inter- ied and dated (Cawood and Nemchin, 2000; Pell et al., 1997;
preted from the zircon ages. Sircombe and Freeman, 1999). The zircon populations in Perth
Cenozoic volcanic arc history of East Java, Indonesia 215

A o
110 E
o
112 E Inherited population includes
Cretaceous zircons
 NO Cretaceous zircons

o
7S

8oS

0 50 100

Kilometers

B o
110 E
o
112 E
Inherited population contains

 NO Archean zircons
Archean zircons

7oS
Figure 12. Distribution of samples that
contained zircon populations of (A)
Cretaceous ages, and (B) Archean ages
(Smyth, 2005). (C) Interpreted character
of the basement terranes in East Java.
o
8S

0 50 100

Kilometers

C 110oE 112oE
Edge of Sunda shelf: Nature of the basement
 uncertain; potentially arc and ophiolitic fragments
accreted during the Cretaceous.

o
7S
CENTRAL JAVA:
Accreted
Cretaceous KENDENG BASIN: Nature of the basement
metamorphosed uncertain; potentially arc and ophiolitic
arc and ophiolitic fragments accreted during the Cretaceous
crust.

o
8S SOUTHERN MOUNTAINS ARC:
Underlain by a fragment of continental
crust of Gondwana origin

0 50 100
Kilometers

Basin sedimentary rocks are remarkably similar in age to those Mesoproterozoic Capricorn and Albany-Fraser orogenic belts, and
of the Southern Mountains in East Java. Both areas yield zircons 800–500 Ma ages similar to the Neoproterozoic Pinjara orogenic
with 4200–2500 Ma ages (predominantly 3200–2500 Ma) simi- belt (Fig. 14). A west Australian origin is a geographically simple
lar to the Archean Yilgarn and Pilbara Blocks, 2000–1600 Ma and plausible explanation for the Precambrian zircons of East
and 1300–1000 Ma ages similar to the Paleoproterozoic and Java, but how did they become incorporated in igneous rocks?
216 Smyth et al.

Malay
Peninsula
Sabah
Brunei CELEBES
SEA
Sarawak Malino Metamorphic
SUNDA Borneo Complex
SHELF
Kalimantan

it
Stra

r
Schwaner

assa
INDIAN
Mountains

ins
OCEAN

Mounta s

Mak
Meratu
Sulawesi
Sumatra
Su

JAVA
ch
nd

r
West Java SEA aA
aT

Basin jn aw
ren

u East Java
rim
ch

Extent of continental Ka Sea Basin FLORES


crust of Sundaland SEA
(Hamilton, 1979) Java Bali Lombok
Granites of Sundaland
(Cobbing, 2005;
10°S van Hattum, 2005)
Java Trench
100°E 110°E 120°E

Figure 13. Extent of area interpreted to be underlain by continental crust, and locations of granitic bodies that may have
been available for erosion during the Cenozoic. The NE-SW–trending Karimunjawa Arch, which was a significant topo-
graphic barrier throughout the Cenozoic, is also denoted.

Incorporation of Zircons into East Java Igneous Rocks via the Bengal Fan, or from the Bird’s Head microcontinent,
New Guinea. If sediment was produced by the erosion of the
Zircons are highly refractory and are well known for sur- Himalayas, it would require transport via the Bengal Fan to the
viving repeated episodes of melting (e.g., Hanchar and Hoskin, Java Trench and total transport distances >4000 km. Although
2003). Ancient zircons could have been introduced into Paleo- mud-rich material is considered to have reached northern
gene magmas of the Southern Mountains Arc either by subduc- Sumatra in the distal parts of the Bengal Fan (Curray et al.,
tion of Gondwana-derived sediments or crust, or by passage of 2002), there were topographic barriers at the Investigator Frac-
magmas through a previously emplaced Gondwana-origin conti- ture Zone and the Ninety East Ridge, and even then a further
nental fragment at depth beneath the Southern Mountains. 2000 km of along-trench transport. Today there is no material
from the Bengal Fan reported in the Java Trench off the shore
Subduction of Gondwana Sediment of East Java (e.g., Masson et al., 1990) and no significant sedi-
Sediment from western Australia deposited on the Indo-Aus- ment cover on the Indian plate south of Java (Kopp, 2002). It
tralian plate could have been carried north to the Java Trench as is equally unlikely that sediment from the Himalayan region
Australia moved north. However, even today a fan similar in size could have traveled southeastward from Indochina across Sun-
to the modern Bengal Fan would be required to bring material daland, bypassing numerous basins and associated structural
from Australia to the site of subduction at the Java Trench, and highs (Hall and Morley, 2004) to enter East Java. This would
there is no evidence of a sediment supply and distribution system require transport distances >3500 km.
like this in the past. The huge Bengal Fan is fed by erosion from The basement rocks of the Bird’s Head microcontinent were
the Himalayas, but there is no evidence for a comparable orogenic not available for erosion during most of the Cenozoic, and the
belt in Australia during Cenozoic times. During the early Ceno- Bird’s Head was the site of deposition of thick marine carbon-
zoic, Australia lay much farther south of its current position (Hall, ates of the New Guinea limestones (Fraser et al., 1993; Pieters
2002), and there is little sedimentary cover today on the Indian- et al., 1983). Even had the Bird’s Head been supplying large vol-
Australian plate south of Java (Masson et al., 1990; Kopp, 2002) umes of sediment, along-trench transport distances >2000 km
in what would have been more proximal parts of such a fan. would have been required for material to enter the subduction
It is unlikely that material could have been introduced zone beneath Java. This is equivalent to material from southern
into the Java Trench by axial transport from the Himalayas Mexico or northern Washington State entering the Allistos Arc
Cenozoic volcanic arc history of East Java, Indonesia 217

40
CENOZOIC
VOLCANISM
Range of U-Pb SHRIMP zircon ages from the A
Southern Mountains Arc, East Java (Smyth,
2005). N = 453

35

30
Numbers of grains

CRETACEOUS
BASEMENT
25

20

15

10

0
500 1000 1500 2000 2500 3000
Figure 14. Comparison of the U-Pb
PHANEROZOIC NEOPROT MESOPROT PALEOPROTEROZOIC ARCHEAN
SHRIMP ages of (A) East Java (Smyth,
2005), and (B) Western Australia (Bruguier
40 Range of U-Pb SHRIMP ages from placer B et al., 1999; Sircombe and Freeman, 1999).
deposits in Western Australia (Bruguier et al, 1999; Revised from Smyth et al. (2005).
Sircombe and Freeman, 1999) and some of the
probable source regions. N = 260
35

30 PINJARA ALBANY-FRASER CAPRICORN


YILGARN CRATON
OROGEN OROGEN OROGEN
500-800 Ma 1000-1300 Ma 1600-2000 Ma 2500-4200 Ma
Numbers of grains

25

20

15

10

500 1000 1500 2000 2500 3000


Age (Ma)

(Busby et al., 2006) of Baja California, or for northern Norway or even if these were overcome, the fragment would have been
Alabama to have been sources for the Lough Nafooey Arc (Draut required to supply material to the arc from the middle Eocene
and Clift, 2001) of the Irish Caledonides. to the early Miocene, a period of ~20 m.y. The fragment of crust
either remained stationary beneath the Southern Mountains or
Subduction of a Gondwana Fragment was 1500 km in length, based on the present subduction rate of
Subduction of continental crust is even less likely. Buoy- 75 mm/year (McCaffrey, 1996), or on Cenozoic plate tectonic
ancy forces make subduction of continental crust difficult, and reconstructions (Hall, 2002).
218 Smyth et al.

Accretion of a Gondwana Fragment and the processes of eruption and tropical weathering combined
The most likely explanation is that continental crust was to remove the unstable volcanic constituents, leaving well-sorted
already present beneath the Southern Mountains Arc when sub- quartz-rich sandstones. These were previously interpreted as hav-
duction began in the middle Eocene. We suggest that this was a ing been eroded from a continental Sundaland source, but careful
western Australia–rifted continental fragment that collided with examination of the sandstones reveals abundant evidence of their
the margin of Sundaland during the Late Cretaceous and may volcanic origin on the basis of textures, light mineral constitu-
have terminated the Cretaceous phase of subduction. Subduction ents, quartz character, clay mineralogy, and zircon character and
was renewed in the middle Eocene, and Cenozoic melts were ages (Smyth, 2005).
contaminated by interaction with continental crust beneath the
Southern Mountains. The distribution of pre-Cenozoic zircons Age of Subduction
indicates that north of the Southern Mountains there is no conti-
nental basement, and the deep crust is arc or ophiolitic, perhaps The character of the oldest parts of the East Java stratigraphic
representing material accreted during Cretaceous subduction or successions indicates that Cenozoic volcanic activity began in the
underplated during collision. The continental fragment could middle Eocene (ca. 45 Ma). There is little evidence to support the
have been quite large and may be traceable into Sulawesi. Zircon common assumption (e.g., Hall, 2002; Heine et al., 2004; Met-
dating and geochemical evidence (van Leeuwen and Muhardjo, calfe, 1996; Scotese et al., 1988) that subduction continued from
2005) show that the Malino Metamorphic Complex of NW the Mesozoic into the Cenozoic without significant interruption.
Sulawesi (Fig. 13) contains zircons with ages up to 3500 Ma. The stratigraphy of East Java indicates that older subduction
probably terminated in the Late Cretaceous with the collision of
DISCUSSION a continental fragment rifted from western Australia. We suggest
that the continental fragment extended from East Java to North
The stratigraphic record in East Java provides insights into Sulawesi. Ophiolites were emplaced from Java to North Borneo
the nature of the deep crust beneath the arc, the history of Indian- during this collision. The oldest sedimentary rocks above the
Australian plate subduction, and the contribution of the volcanic basement in East Java are of uncertain age but are middle Eocene
arc to basin formation. It also poses some questions about the or older and lack volcanic debris. There is little evidence for lat-
continuity and location of volcanic activity, which have relevance est Cretaceous to early Eocene volcanic activity in most of the
for other arcs. Sundaland margin between Sumatra and Sulawesi. Subduction
resumed in the middle Eocene when Australia began to move
Deep Character of the East Java Volcanic Arc northward rapidly after 45 Ma (Hall, 2002; Müller et al., 2000;
Schellart et al., 2006), and has continued to the present day.
Little is known about the deep crust beneath the young arcs
of the Western Pacific and Indonesia. There have been few seis- Formation of the Kendeng Basin
mic refraction studies and few xenolith studies, and the deep crust
is almost invariably covered by younger volcanic and sedimen- When volcanic activity resumed in the middle Eocene a basin
tary rocks. In East Java the old zircons in the Eocene to Miocene began to form directly north of the arc. Most of the investigations
arc products provide evidence of the character of the deep crust. of sedimentary basins on land in East Java and in the East Java
The insight gained was completely unexpected. When this study Sea (Fig. 1) have been carried out as part of hydrocarbon explora-
began, it was thought that the entire region was underlain by Cre- tion activity, and little attention has been given to areas close to
taceous arc and ophiolitic fragments, as suggested by Hamilton the modern arc. The Sunda Shelf basins are typically >100 km
(1979). The distribution of ages shows that north of the South- from the arc and are characterized by many features typical of
ern Mountains there is indeed crust of this character, as indicated fault-controlled basins; explanations for their origin have there-
by the small exposures of basement, but beneath the Southern fore interpreted them as types of backarc basins resulting from
Mountains themselves there is a fragment of continental crust of subduction rollback, rift and thermal sag, or strike-slip faulting
Gondwana character and western Australian origin. (e.g., Hamilton, 1979; Matthews and Bransden, 1995).
The presence of a continental fragment beneath the arc The Kendeng Basin is not a typical backarc basin. Backarc
accounts for the unusually acidic character of arc volcanism in basins (Taylor and Karner, 1983) may form by trench rollback,
the Eocene to early Miocene arc. The acidic products of this arc causing generation of oceanic crust (Dewey, 1980; Karig, 1971),
have been overlooked partly because the resistant “Old Andes- rifting of continental crust (Kobayashi, 1985), gravitational col-
ites” are topographically so much more obvious, although acidic lapse in the wake of arc-continent collision (Clift et al., 2003),
volcanic and minor intrusive rocks are well exposed in many parts or by the trapping of old oceanic crust behind an arc (Scholl et
of the Southern Mountains. However, the main reason why the al., 1986). Retroarc basins (Dickinson, 1974) form as compres-
acidic products of the Eocene to early Miocene arc were missed sive foreland basins behind a volcanic arc. Busby and Ingersoll
by earlier studies is because they were erupted explosively and (1995) defined backarc basins as oceanic basins behind intraoce-
dispersed as volcanic ash, the ash was reworked into sediments, anic magmatic arcs, and continental basins behind continental
Cenozoic volcanic arc history of East Java, Indonesia 219

margin arcs that lack foreland fold and thrust belts. Marsaglia melting was inhibited as the wedge became depleted. The cause
(1995) identified similar categories of backarc basins with the of hinge advance was the result of the collision of Australia with
addition of boundary basins resulting from extension along plate eastern Indonesia, and the counterclockwise rotation of the Bor-
boundaries with translational components. Busby and Ingersoll neo-Java region this collision induced (Hall, 2002). Arc volcanism
(1995) claim that the Sunda Shelf basins of Indonesia are non- resumed when hinge advance ceased and the mantle wedge was
extensional backarc basins formed in a neutral strain regime, replenished by fertile mantle (Macpherson and Hall, 2002).
although the basis for this assertion is not clear, because there is Why the volcanic arc moved to its new position 50 km north
abundant evidence for rifting of many of these basins (e.g., Cole of the Southern Mountains Arc is uncertain. There is evidence of
and Crittenden, 1997; Hall and Morley, 2004). thrusting and contraction in Java in the late Neogene, but its tim-
There is no evidence for either newly formed oceanic crust ing is not precisely known; this is part of our current research. It
under the Kendeng Basin or for strike-slip faulting. The basin is is possible that contraction was linked to large-scale arc dynam-
much closer to the arc than other backarc basins and most of the ics, such as coupling of the subduction and overriding slab. There
“backarc basins” of the Sunda Shelf. The late Cenozoic deforma- is evidence of an old consolidated accreted material in the Java
tion in the region led to ~10–30 km of contraction at the outer forearc that acts as a backstop to the active accretionary prism
edge of the basin, so throughout the early Cenozoic this basin was (Kopp et al., 2001). This suggests two phases of forearc develop-
very close to the arc. There was extension in the Sunda Shelf from ment, and these may be linked to the two phases of arc volcanism
the Eocene, but nowhere was oceanic crust formed. The Kend- on land in East Java. It is also possible that when the subduction
eng Basin is an asymmetrical depression, the deepest part of the hinge advanced and the mantle wedge was not being replenished,
basin lies directly behind the arc, and the Eocene to lower Mio- one result was a rigid, stronger overriding lithospheric plate that
cene volcanic and sedimentary sequence thins toward the edge of later failed when hinge advance ceased and weaker, warmer man-
the Sunda Shelf (Waltham et al., this volume). The basin fill was tle replenished the wedge. Another control could have been the
derived mainly from the Southern Mountains Arc on the south nature of the crust beneath the arc. The zircon ages indicate that
side of the basin. The basin began to form at the same time as there was a change in basement type north of the Southern Moun-
arc activity began, and its subsidence history is closely linked to tains and that the boundary between continental and ophiolitic
activity in the volcanic arc. There are no seismic lines crossing the crust at depth may have been a preexisting structural weakness
basin, and it is not well exposed at the surface, so we cannot assess that influenced the position at which melts could rise when arc
the role of faulting in the basin development, but it does not have activity resumed.
the typical synrift-postrift stratigraphy of many other Sunda Shelf Other possibilities for explaining arc migration include sub-
basins. There is no evidence of compressional loading having duction erosion and a change in dip of the subducting slab. Sub-
contributed to basin formation. Thrusting in East Java occurred in duction erosion seems unlikely, as the width of the arc-trench gap
the late Neogene long after the Kendeng Basin had formed. The today is >300 km, and because marine data support active accre-
close relationship between volcanic activity and basin subsidence tion at that time (e.g., Kopp et al., 2001). A change in subduction
suggests that the two are linked, and we suggest that the load of angle cannot be ruled out, but it is notable that today the slab
the volcanic arc was the major cause of basin subsidence. The dip has increased to a very high angle after the slab descended
contribution of volcanic arc loading to basin formation in Indone- to 100 km, as seen in seismicity data (England et al., 2004). A
sia is discussed in greater detail by Waltham et al. (this volume). lower, not a higher, angle would be expected if the arc had moved
north owing to the change in dip of the slab.
Movement of the Volcanic Arc The jump in position of the volcanic arc is a feature of other
Indonesian arcs (e.g., Halmahera Arc; Nichols and Hall, 1991)
Activity in the Southern Mountains Arc terminated during and reflects the changing stresses at the plate boundary over time.
the early Miocene (ca. 20 Ma). After a period with little mag- We prefer an explanation that links the change in position of the
matism, a new episode of arc volcanism began in East Java at arc to plate reorganization in the region. However, it is difficult to
ca. 12–10 Ma in a new location. The Miocene to Holocene arc is test different hypotheses. Comparison of arc development to the
parallel to the Southern Mountains Arc but ~50 km north of it. A east and west of East Java, from Bali eastward, and toward West
significant reduction in volcanic activity in the middle Miocene Java and Sumatra, might provide insights, but unfortunately even
is a well-known feature of the Sunda Arc from Java eastward, less is known of arc history in these regions than in East Java.
although subduction was continuous during this period (Hall, This East Java study shows that our understanding of arc tec-
2002). Vigorous volcanic activity resumed at the end of the mid- tonics is still incomplete and that studies of arc stratigraphy are
dle Miocene in the Sunda and Banda Arcs at ca. 10 Ma (Hall, essential if our knowledge of arc dynamics is to be improved.
2002; Macpherson and Hall, 2002).
Macpherson and Hall (1999, 2002) suggest that the decline in CONCLUSIONS
volcanic activity resulted from northward advance of the subduc-
tion hinge. Hinge advance prevented replenishment of the mantle The case study of early Cenozoic arc volcanism in East Java
wedge by fertile mantle, and consequently subduction-induced shows the importance of examining arc stratigraphy. Insights
220 Smyth et al.

have been gained into the eruptive history, formation of a deep REFERENCES CITED
basin behind the arc, and the character of the deep crust.
The early Cenozoic stratigraphy of East Java provides Ardhana, W., 1993, A depositional model for the early middle Miocene Ngray-
ong Formation and implications for exploration in the East Java basin:
a record of a cycle of arc activity from initiation in the mid- Jakarta, Proceedings, Indonesian Petroleum Association Annual Conven-
dle Eocene (ca. 42 Ma) to termination in the early Miocene tion, 22nd, p. 395–443.
(ca. 20 Ma). The Kendeng Basin, directly behind the Southern Baum, J.S., Baum, G.R., Thompson, P.R., and Humphery, J.D., 1994, Stable
isotope evidence for relative and eustatic sea-level changes in Eocene
Mountains Arc, contains >6 km of volcaniclastic and sedimen- to Oligocene carbonates, Baldwin County, Alabama: Geological
tary rocks. The basin is not a typical backarc basin, and its sub- Society of America Bulletin, v. 106, p. 824–839, doi: 10.1130/0016-
sidence history is linked to volcanic activity within the South- 7606(1994)106<0824:SIEFRA>2.3.CO;2.
Ben-Avraham, Z., and Emery, K.O., 1973, Structural framework of Sunda
ern Mountains Arc. Shelf: American Association of Petroleum Geologists Bulletin, v. 57,
The final stage of volcanic activity in the Southern Moun- p. 2323–2366.
tains Arc is marked by the Semilir Eruption (ca. 20 Ma), which Bergman, S.C., Coffield, D.Q., Talbot, J.P., and Garrard, R.J., 1996, Tertiary
tectonic and magmatic evolution of Western Sulawesi and the Makassar
distributed ash over a wide area and may be comparable to the Strait, Indonesia: Evidence for a Miocene continent-continent collision,
Pleistocene eruption of Toba in Sumatra. Following this phase of in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE Asia: Geo-
major eruptions, there was a lull in volcanic activity during the logical Society [London] Special Publication 106, p. 391–430.
Bishop, W.P., 1980, Structure, stratigraphy and hydrocarbons offshore southern
middle Miocene, followed by resumption in arc activity to the Kalimantan, Indonesia: American Association of Petroleum Geologists
north of the Southern Mountains Arc, along the axis of the mod- Bulletin, v. 64, p. 37–58.
ern Sunda Arc during the late Miocene. The mechanisms that Bruguier, O., Bosch, D., Pidgeon, R.T., Byrne, D.I., and Harris, L.B., 1999,
U-Pb chronology of the Northampton Complex, Western Australia—
resulted in the decline in volcanism, and northward movement in Evidence for Grenvillian sedimentation, metamorphism and deformation
arc axis, are not yet understood and show that our understanding and geodynamic implications: Contributions to Mineralogy and Petrol-
of arcs is still incomplete. ogy, v. 136, p. 258–272, doi: 10.1007/s004100050537.
Busby, C.J., and Ingersoll, R.V., 1995, Tectonics of Sedimentary Basins: Cam-
The stratigraphic record of volcanic arcs can provide insights bridge, Massachusetts, Blackwell Science, 579 p.
into the character of the deep crust. The entire East Java region Busby, C.J., Adams, B.F., Mattinson, J., and Deoreo, S., 2006, View of an intact
was previously thought to be underlain by Cretaceous arc and oceanic arc, from surficial to mesozonal levels: Cretaceous Alisitos arc,
Baja California: Journal of Volcanology and Geothermal Research, v. 149,
ophiolitic fragments, but Archean to Cambrian zircons within p. 1–46, doi: 10.1016/j.jvolgeores.2005.06.009.
acidic products of the Southern Mountains Arc point to the Catuneanu, O., 2006, Principles of Sequence Stratigraphy: Amsterdam, Else-
occurrence of a continental crust of Gondwanan character and vier, 375 p.
Cawood, P.A., and Nemchin, A.A., 2000, Provenance record of a rift basin: U/Pb
western Australian origin beneath the old arc. This continental ages of detrital zircons from the Perth Basin, Western Australia: Sedimen-
fragment is thought to have collided with Sundaland during the tary Geology, v. 134, p. 209–234, doi: 10.1016/S0037-0738(00)00044-0.
Cretaceous and is interpreted to have terminated the Cretaceous Chesner, C.A., and Rose, W.I., 1991, Stratigraphy of the Toba tuffs and the
evolution of the Toba caldera complex, Sumatra, Indonesia: Bulletin of
phase of subduction. The extent of this fragment is not known Volcanology, v. 53, p. 343–356, doi: 10.1007/BF00280226.
but may be traceable into Sulawesi. The use of inherited zircons Chotin, P., Giret, A., Rampnoux, J.P., Rasplus, L., and Suminta, S., 1984,
to determine the character of the deep crust may be applicable in Etude de la fracturation de l’ile de Java, Indonesie: Bulletin de la Société
Géologique de France, v. 7, p. 1259–1262.
other arcs. Clift, P.D., Schouten, H., and Draut, A.E., 2003, A general model of arc-
continent collision and subduction polarity reversal from Taiwan and the
ACKNOWLEDGMENTS Irish Caledonides, in Larter, R.D., and Leat, P.T., eds., Intra-Oceanic Sub-
duction Systems; Tectonic and Magmatic Processes: Geological Society
[London] Special Publication 219, p. 81–98.
This project was funded by the SE Asia Research Group at Cobbing, E.J., 2005, Granites, in Barber, A.J., Crow, M.J., and Milsom, J.S.,
Royal Holloway University, supported by an oil company con- eds., Sumatra: Geology, Resources, and Tectonic Evolution: Geological
Society [London] Memoirs 31, p. 267–282.
sortium. Zircons were analyzed using the SHRIMP facility at Cobbing, E.J., Mallick, D.I.J., Pitfield, P.E.J., and Teoh, L.H., 1986, The gran-
Curtin University of Technology, Perth, Australia. Financial ites of the Southeast Asian tin belt: Geological Society [London] Journal,
assistance for SHRIMP dating was provided by the University v. 143, p. 537–550.
Cole, J.M., and Crittenden, S., 1997, Early Tertiary basin formation and the devel-
of London Central Research Fund and CSIRO, Australia. The opment of lacustrine and quasi-lacustrine/marine source rocks on the Sunda
analyses were undertaken by Joseph Hamilton (CSIRO and Shelf of SE Asia, in Fraser, A.J., et al., eds., Petroleum Geology of SE Asia:
University of the West Indies) and Pete Kinny (Curtin Univer- Geological Society [London] Special Publication 126, p. 147–183.
Curray, J.R., Emmel, F.J., and Moore, D.G., 2002, The Bengal Fan: Morphol-
sity of Technology). We are grateful to LIPI (Lembaga Ilmu ogy, geometry, stratigraphy, history and processes: Marine and Petroleum
Pengetahuan Indonesia) for field-work permissions. We wish to Geology, v. 19, p. 1191–1223, doi: 10.1016/S0264-8172(03)00035-7.
thank Eko Budi Lelono, Peter Lunt, Theo van Leeuwen, Moyra de Genevraye, P., and Samuel, L., 1972, The geology of Kendeng Zone (East
Java): Jakarta, Proceedings, Indonesian Petroleum Association Annual
Wilson, Colin Macpherson, Heather Handley, Cindy Ebinger, Convention, 1st, p. 17–30.
and Dave Waltham. Marcelle BouDagher-Fadel of University Dewey, J.F., 1980, Episodicity, sequence and style at convergent plate boundar-
College London provided biostratigraphic analyses. Peter Clift, ies: Geological Association of Canada Special Paper 20, p. 553–573.
Dickinson, W.R., 1974, Plate tectonics and sedimentation, in Dickinson, W.R.,
Jason Ali, and an anonymous reviewer are thanked for their ed., Tectonics and Sedimentation: Tulsa, Society of Economic Paleontolo-
extensive reviews, which greatly improved this manuscript. gists and Mineralogists, p. 204.
Cenozoic volcanic arc history of East Java, Indonesia 221

Draut, A.E., and Clift, P., 2001, Geochemical evolution of arc magmatism dur- Leeder, M.R., 1982, Sedimentology: Process and Product: London, George
ing arc-continental collision, South Mayo, Ireland: Geology, v. 29, p. 543– Allen and Unwin, 344 p.
546, doi: 10.1130/0091-7613(2001)029<0543:GEOAMD>2.0.CO;2. Lelono, E.B., 2000, Palynological study of the Eocene Nanggulan Formation,
Ebanks, W.J.J., and Cook, C.B.P., 1993, Sedimentology and reservoir proper- central Java, Indonesia [Ph.D. thesis]: University of London, 453 p.
ties of Eocene Ngimbang clastics sandstones in cores of the Pagerungan-5 Liew, T.C., and Page, R.W., 1985, U-Pb zircon dating of granitoid plutons from
well, Pagerungan Field, East Java Sea, Indonesia, in Atkinson, C.D., et al., the west coast of peninsular Malaysia: Geological Society [London] Jour-
eds., IPA Core Workshop Notes, Clastic Rocks and Reservoirs of Indone- nal, v. 142, p. 515–526.
sia: Jakarta, Indonesian Petroleum Association, p. 9–36. Lokier, S.W., 2000, The Miocene Wonosari Formation, Java, Indonesia: Vol-
Edwards, C., Menzies, M., and Thirlwall, M., 1991, Evidence from Muriah, caniclastic influences on carbonate platform development [Ph.D. thesis]:
Indonesia, for the interplay of supra-subduction zone and intraplate pro- University of London, 648 p.
cesses in the genesis of potassic alkaline magmas: Journal of Petrology, Macpherson, C.G., and Hall, R., 1999, Tectonic controls of geochemical evolu-
v. 32, p. 555–592. tion in arc magmatism of SE Asia, in Proceedings, PACRIM Congress,
England, P., Engdahl, R., and Thatcher, W., 2004, Systematic variation in the 4th, Australian Institute of Mining and Metallurgy, p. 359–368.
depths of slabs beneath arc volcanoes: Geophysical Journal International, Macpherson, C.G., and Hall, R., 2002, Timing and tectonic controls on magma-
v. 156, p. 377–408, doi: 10.1111/j.1365-246X.2003.02132.x. tism and ore generation in an evolving orogen: Evidence from Southeast
Fraser, T.H., Bon, J., and Samuel, L., 1993, A new dynamic Mesozoic stratig- Asia and the western Pacific, in Blundell, D.J., et al., eds., The Timing
raphy for the West Irian micro-continent Indonesia and its implications: and Location of Major Ore Deposits in an Evolving Orogen: Geological
Jakarta, Proceedings, Indonesian Petroleum Association Annual Conven- Society [London] Special Publication 204, p. 49–67.
tion, 22nd, p. 707–761. Madon, M.B., Azlina, A., and Wong, R., 1997, Structural evolution, maturation
Fulthorpe, C.S., Carter, R.M., Miller, K.G., and Wilson, J., 1996, Marshall history and hydrocarbon potential of the Penyu basin, offshore peninsula
Paraconformity: A mid-Oligocene record of inception of the Antarctic Malaysia, in Howes, J.V.C., and Noble, R.A., eds., Proceedings of the
Circumpolar and coeval glacio-eustatic lowstand?: Marine and Petroleum International Conference on Petroleum Systems of SE Asia and Australia:
Geology, v. 13, p. 61–77, doi: 10.1016/0264-8172(95)00033-X. Jakarta, Indonesian Petroleum Association, p. 403–424.
Gradstein, F., Ogg, J., and Smith, A., 2004, A Geological Time Scale 2004: Marsaglia, K.M., 1995, Interarc and backarc basins, in Busby, C.J., and Inger-
Cambridge, UK, Cambridge University Press, 589 p. soll, R.V., eds., Tectonics of Sedimentary Basins: Cambridge, Massachu-
Hall, R., 2002, Cenozoic geological and plate tectonic evolution of SE Asia and setts, Blackwell Science, p. 299–330.
the SW Pacific: Computer-based reconstructions, model and animations: Masson, D.G., Parson, L.M., Milsom, J., Nichols, G., Sikumbang, N., Dwi-
Journal of Asian Earth Sciences, v. 20, p. 353–434, doi: 10.1016/S1367- yanto, B., and Kallagher, H., 1990, Subduction of seamounts at the Java
9120(01)00069-4. Trench: A view with long-range sidescan sonar: Tectonophysics, v. 185,
Hall, R., and Morley, C.K., 2004, Sundaland basins, in Clift, P., et al., eds., Con- p. 51–65, doi: 10.1016/0040-1951(90)90404-V.
tinent-Ocean Interactions within the East Asian Marginal Seas: American Matthews, S.J., and Bransden, P.J.E., 1995, Late Cretaceous and Cenozoic
Geophysical Union Geophysical Monograph 149, p. 55–85. tectono-stratigraphic development of the East Java Sea Basin, Indonesia:
Hamilton, W., 1979, Tectonics of the Indonesian Region: U.S. Geological Sur- Marine & Petroleum Geology, v. 12, p. 499–510, doi: 10.1016/0264-
vey Professional Paper 1078, 345 p. 8172(95)91505-J.
Hanchar, J.M., and Hoskin, P.W.O., eds., 2003, Zircon: Reviews in Mineralogy McCaffrey, R., 1996, Slip partitioning at convergent plate boundaries of SE
and Geochemistry, v. 53, 500 p. Asia, in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE Asia:
Handley, H., 2006, Geochemical and Sr-Nd-Hf-O isotopic constraints on volca- Geological Society [London] Special Publication 106, p. 3–18.
nic petrogenesis at the Sunda arc, Indonesia [Ph.D. thesis]: Durham, UK, McCourt, W.J., Crow, M.J., Cobbing, E.J., and Amin, T.C., 1996, Mesozoic and
University of Durham, 268 p. Cenozoic plutonic evolution of SE Asia: Evidence from Sumatra, Indone-
Haq, B.U., Hardenbol, J., and Vail, P.R., 1987, Chronology of fluctuating sea sia, in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE Asia:
levels since the Triassic: Science, v. 235, p. 1156–1167, doi: 10.1126/sci- Geological Society [London] Special Publication 106, p. 321–335.
ence.235.4793.1156. Metcalfe, I., 1996, Gondwanaland dispersion, Asian accretion and evolution of
Harahap, B.H., Bachri, S., Baharuddin, Surwarna, N., Panggabean, H., and eastern Tethys: Australian Journal of Earth Sciences, v. 43, p. 605–624.
Simanjuntak, T.O., 2003, Stratigraphic Lexicon of Indonesia: Bandung, Miyazaki, K., Sopaheluwakan, J., Zulkarnain, I., and Wakita, K., 1998, A jade-
Geological Research and Development Centre, 729 p. ite-quartz-glaucophane rock from Karangsambung, central Java, Indone-
Heine, C.R., Müller, D., and Gaina, C., 2004, Reconstructing the lost East- sia: Island Arc, v. 7, p. 223–230, doi: 10.1046/j.1440-1738.1998.00164.x.
ern Tethys ocean basin: Convergence history of the SE Asian margin and Müller, R.D., Gaina, C., and Clarke, S., 2000, Seafloor spreading around Aus-
marine gateways, in Clift, P., et al., eds., Continent-Ocean Interactions tralia, in Veevers, J.J., ed., Billion-Year Earth History of Australia and
within the East Asian Marginal Seas: American Geophysical Union Geo- Neighbours in Gondwanaland: Sydney, GEMOC Press, p. 18–28.
physical Monograph 149, p. 37–54. Nicholls, I.A., Whitford, D.J., Harris, K.L., and Taylor, B., 1980, Variation in
Hoffmann-Rothe, A., Ritter, O., and Hanna, W.F., 2001, Magnetotelluric and the geochemistry of mantle sources for tholeiitic and calc-alkaline mafic
geomagnetic modelling reveals zones of very high electrical conductivity magmas, western Sunda volcanic arc, Indonesia: Chemical Geology,
in the upper crust of Central Java: Physics of the Earth and Planetary Inte- v. 30, p. 177–199, doi: 10.1016/0009-2541(80)90105-9.
riors, v. 124, p. 131–151, doi: 10.1016/S0031-9201(01)00196-0. Nichols, G.J., and Hall, R., 1991, Basin formation and Neogene sedimentation
Imtihanah, 2000, Isotopic dating of the Sumatran Fault System [MPhil. thesis]: in a backarc setting, Halmahera, eastern Indonesia: Marine and Petroleum
University of London, 150 p. Geology, v. 8, p. 50–61, doi: 10.1016/0264-8172(91)90044-2.
Kadar, D., 1986, Neogene planktonic foraminiferal biostratigraphy of the south Parkinson, C.D., Miyazaki, K., Wakita, K., Barber, A.J., and Carswell, D.A.,
central Java area, Indonesia: Geological Research and Development Cen- 1998, An overview and tectonic synthesis of the pre-Tertiary very-high-
tre Special Publication 5, 83 p. pressure metamorphic and associated rocks of Java, Sulawesi and Kali-
Karig, D.E., 1971, Origin and development of marginal basins in the Western mantan, Indonesia: Island Arc, v. 7, p. 184–200, doi: 10.1046/j.1440-
Pacific: Journal of Geophysical Research, v. 76, p. 2542–2561. 1738.1998.00184.x.
Kobayashi, K., 1985, Sea of Japan and Okinawa Trough, in Naira, A.E.M., et Pell, S.D., Williams, I.S., and Chivas, A.R., 1997, The use of protolith zircon-
al., eds., The Pacific Ocean, Volume 7: The Ocean Basins and Margins: age fingerprints in determining the protosource areas for some Australian
New York, Plenum Press, p. 419–458. dune sands: Sedimentary Geology, v. 109, p. 233–260, doi: 10.1016/
Kopp, H., 2002, BSR occurrence along the Sunda margin: Evidence from seis- S0037-0738(96)00061-9.
mic data: Earth and Planetary Science Letters, v. 197, p. 225–235, doi: Pieters, P.E., Pigram, C.J., Trail, D.S., Dow, D.B., Ratman, N., and Sukamto,
10.1016/S0012-821X(02)00484-3. R., 1983, The stratigraphy of western Irian Jaya: Bandung, Bulletin of
Kopp, H., Flueh, E.R., Klaeschen, D., Bialas, J., and Reichert, C., 2001, Crustal Geological Research and Development Centre, v. 8, p. 14–48.
structure of the central Sunda margin at the onset of oblique subduction: Rawson, R.F., Allen, P.M., Brenchley, P.J., Cope, J.C.W., Gale, A.S., Evans,
Geophysical Journal International, v. 147, p. 449–474, doi: 10.1046/ J.A., Gibbard, P.L., Gregory, F.J., Hailwood, E.A., Hesselbo, S.P., Knox,
j.0956-540x.2001.01547.x. R.W.O.B., Marshall, J.E.A., Oates, M., Riley, N.J., Smith, A.G., Trewin,
222 Smyth et al.

N., and Zalasiewicz, J.A., 2001, Stratigraphical Procedure: Bath, Geo- van Bemmelen, R.W., 1949, The Geology of Indonesia: The Hague, Nijhoff,
logical Society [London], 64 p. Government Printing Office, 732 p.
Sandwell, D.T., and Smith, W.H.F., 1997, Marine gravity anomaly from Geosat van Hattum, M.W.A., 2005, Provenance of Cenozoic rocks of northern Borneo
and ERS 1 satellite altimetry: Journal of Geophysical Research, v. 102, [Ph.D. thesis]: University of London, 457 p.
p. 10,039–10,054, doi: 10.1029/96JB03223. van Hattum, M.W.A., Hall, R., Pickard, A.J., and Nichols, G.J., 2006, Southeast
Schellart, W.P., Lister, G.S., and Toy, V.G., 2006, A Late Cretaceous and Ceno- Asian sediments not from Asia: Provenance and geochronology of north
zoic reconstruction of the Southwest Pacific region: Tectonics controlled Borneo sandstones: Geology, v. 34, p. 589–592, doi: 10.1130/G21939.1.
by subduction and slab rollback processes: Earth-Science Reviews, v. 76, van Leeuwen, T.M., and Muhardjo, 2005, Stratigraphy and tectonic setting of
p. 191–233, doi: 10.1016/j.earscirev.2006.01.002. the Cretaceous and Paleogene volcanic-sedimentary successions in north-
Scholl, D., Vallier, T.L., and Stevenson, A.J., 1986, Terrane accretion, produc- west Sulawesi, Indonesia: Implications for the Cenozoic evolution of
tion and continental growth: A perspective based on the origin and tec- Western and Northern Sulawesi: Journal of Asian Earth Sciences, v. 25,
tonic gate of the Aleutian–Bering Sea region: Geology, v. 14, p. 43–47, p. 481–511, doi: 10.1016/j.jseaes.2004.05.004.
doi: 10.1130/0091-7613(1986)14<43:TAPACG>2.0.CO;2. Wakita, K., 2000, Cretaceous accretionary-collision complexes in central Indo-
Scotese, C.R., Gahagan, L.M., and Larson, R.L., 1988, Plate tectonic recon- nesia: Journal of Asian Earth Sciences, v. 18, p. 739–749, doi: 10.1016/
structions of the Cretaceous and Cenozoic ocean basins: Tectonophysics, S1367-9120(00)00020-1.
v. 155, p. 27–40, doi: 10.1016/0040-1951(88)90259-4. Wakita, K., and Munasri, B.W., 1994, Cretaceous radiolarians from the Luk-
Sharaf, E., Simo, J.A., Carrol, A.R., and Shields, M., 2005, Stratigraphic evolu- Ulo Melange Complex in the Karangsambung area, Central Java, Indo-
tion of Oligocene–Miocene carbonates and siliciclastics, East Java basin, nesia: Journal of Southeast Asian Earth Sciences, v. 9, p. 29–43, doi:
Indonesia: American Association of Petroleum Geologists Bulletin, v. 89, 10.1016/0743-9547(94)90063-9.
p. 799–819. Wakita, K., Sopaheluwakan, J., Miyazaki, K., and Zulkarnain, I., and Munasri,
Sircombe, K.N., and Freeman, M.J., 1999, Provenance of detrital zircons 1996, Tectonic evolution of the Bantimala Complex, South Sulawesi,
on the Western Australian coastline—Implications for the geological Indonesia, in Hall, R., and Blundell, D.J., eds., Tectonic Evolution of SE
history of the Perth basin and denudation of the Yilgarn craton: Geol- Asia: Geological Society [London] Special Publication 106, p. 353–364.
ogy, v. 27, p. 879–882, doi: 10.1130/0091-7613(1999)027<0879: Waltham, D., Hall, R., Smyth, H.R., and Ebinger, C., 2008, this volume, Basin
PODZOT>2.3.CO;2. formation by volcanic loading, in Draut, A.E., et al., eds., Lessons from
Smyth, H., 2005, Eocene to Miocene basin history and volcanic activity in East the Stratigraphic Record in Arc Collision Zones: Geological Society of
Java, Indonesia [Ph.D. thesis]: University of London, 476 p. America Special Publication 436, doi:10.1130/2008. 2436(02).
Smyth, H., Hall, R., Hamilton, J.P., and Kinny, P., 2005, East Java: Cenozoic Watts, K.J., 1997, The Northern Nam Con Son Basin petroleum system, based
basins, volcanoes and ancient basement: Jakarta, Proceedings, Indonesian on exploration data from block 04D2, Vietnam, in Howes, J.V.C., and
Petroleum Association Annual Convention, 30th, p. 251–266. Noble, R.A., eds., Proceedings of the International Conference on Petro-
Smyth, H.R., Hamilton, P.J., Hall, R., and Kinny, P.D, 2007, The deep crust leum Systems of SE Asia & Australia: Jakarta, Indonesian Petroleum
beneath island arcs: Inherited yircons reveal a gondwana continental frag- Association, p. 481–498.
ment beneath East Java, Indonesia: Earth and Planetary Science Letters Wheller, G.E., Varne, R., Foden, J.D., and Abbot, M.J., 1987, Geochemistry
(EPSL), v. 258, p. 269–282. of Quaternary volcanism in the Sunda-Banda arc, Indonesia, and three-
Soeria-Atmadja, R., Maury, R.C., Bellon, H., Pringgoprawiro, H., Polve, M., and component genesis of island-arc basalt magmas: Journal of Volcanol-
Priadi, B., 1994, Tertiary magmatic belts in Java: Journal of Southeast Asian ogy and Geothermal Research, v. 32, p. 137–160, doi: 10.1016/0377-
Earth Sciences, v. 9, p. 13–17, doi: 10.1016/0743-9547(94)90062-0. 0273(87)90041-2.
Song, S.R., Chen, C.H., Lee, M.Y., Yang, T.F., Iizuka, Y., and Wei, K.Y., 2000, Wilson, M.E.J., and Lokier, S.W., 2002, Siliciclastic and volcaniclastic influences
Newly discovered eastern dispersal of the youngest Toba Tuff: Marine on equatorial carbonates: Insights from the Neogene of Indonesia: Sedi-
Geology, v. 167, p. 303–312, doi: 10.1016/S0025-3227(00)00034-7. mentology, v. 49, p. 583–601, doi: 10.1046/j.1365-3091.2002.00463.x.
Taylor, B., and Karner, G.D., 1983, On the evolution of marginal basins:
Reviews of Geophysics and Space Physics, v. 21, p. 1727–1741.
Untung, M., and Sato, Y., 1978, Gravity and Geological Studies in Java, Indo-
nesia: Geological Survey of Indonesia and Geological Survey of Japan
Special Publication 6, 207 p. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

New constraints on the sedimentation and uplift history of the Andaman-Nicobar


accretionary prism, South Andaman Island

R. Allen
Department of Environmental Science, Lancaster University LA1 4YQ, UK

A. Carter
Research School of Earth Sciences, Birkbeck and University College London, Gower St., London WC1E 6BT, UK

Y. Najman
Department of Environmental Science, Lancaster University LA1 4YQ, UK

P.C. Bandopadhyay
Geological Survey of India, Geodata Division, Salt Lake, Kolkata, 91, India

H.J. Chapman
M.J. Bickle
Department of Earth Sciences, Cambridge University, Downing St., Cambridge CB2 3EQ, UK

E. Garzanti
G. Vezzoli
S. Andò
Dipartimento di Scienze Geologiche e Geotecnologie, Universita Milano-Bicocca, Piazza della Scienza 4, 20126 Milano, Italy

G.L. Foster
Department of Earth Sciences, Bristol University, Queens Rd., Bristol BS8 1RJ, UK

C. Gerring
Department of Earth Sciences, The Open University, Walton Hall, Milton Keynes MK7 6AA, UK

ABSTRACT

The Andaman Islands are part of the Andaman-Nicobar Ridge, an accretionary


complex that forms part of the outer-arc ridge of the Sunda subduction zone. The
Tertiary rocks exposed on the Andaman Islands preserve a record of the tectonic
evolution of the surrounding region, including the evolution and closure of the Tethys
Ocean. Some of the Paleogene sediments on Andaman may represent an offscraped
part of the early Bengal Fan. Through field and petrographic observations, and use of
a number of isotopic tracers, new age and provenance constraints are placed on the

Allen, R., Carter, A., Najman, Y., Bandopadhyay, P.C., Chapman, H.J., Bickle, M.J., Garzanti, E., Vezzoli, G., Andò, S., Foster, G.L., and Gerring, C., 2008, New
constraints on the sedimentation and uplift history of the Andaman-Nicobar accretionary prism, South Andaman Island, in Draut, A.E., Clift, P.D., and Scholl,
D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special Paper 436, p. 223–255, doi:
10.1130/2008.2436(11). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All rights reserved.

223
224 Allen et al.

key Paleogene formations exposed on South Andaman. A paucity of biostratigraphic


data poorly define sediment depositional ages. Constraints on timing of deposition
obtained by dating detrital minerals for the Mithakhari Group indicate sedimenta-
tion after 60 Ma, possibly younger than 40 Ma. A better constraint is obtained for the
Andaman Flysch Formation, which was deposited between 30 and 20 Ma, based on
Ar-Ar ages of the youngest detrital muscovites at ca. 30 Ma and thermal history mod-
eling of apatite fission-track and U-Th/He data. The latter record sediment burial and
inversion (uplift) at ca. 20 Ma. In terms of sediment sources the Mithakhari Group
shows a predominantly arc-derived composition, with a very subordinate contribution
from the continental margin to the east of the arc. The Oligocene Andaman Flysch at
Corbyn’s Cove is dominated by recycled orogenic sources, but it also contains a sub-
ordinate arc-derived contribution. It is likely that the sources of the Andaman Flysch
included rocks from Myanmar affected by India-Asia collision. Any contribution of
material from the nascent Himalayas must have been minor. Nd isotope data discount
any major input from cratonic Greater India sources.

Keywords: Andaman, accretionary wedge, arc, subduction, thermochronology, prov-


enance, uplift.

INTRODUCTION aim of this paper is to build on previous field-based studies and


apply petrologic, isotopic, and thermochronometric techniques
The Andaman-Nicobar Islands are part of an accretionary to better understand the provenance, sediment deposition, and
complex that forms the outer arc ridge of the northern Sunda uplift history.
subduction zone (Fig. 1). The Andaman Islands are in the south- The origin of the Andaman Flysch has been debated for
eastern part of the Bay of Bengal and make up part of a 3000– >20 yr. It has been variously proposed that the Andaman Flysch
5000 km chain that runs from the Myanmar Arakan-Yoma down was derived from the Irrawaddy Delta (Karunakaran et al., 1968;
to Sumatra and Java in the south. The Indian plate is subducting Pal et al., 2003) or, alternatively, from Bengal Fan material shed
northward below the Euarasian plate and obliquely below the from the nascent Himalaya sourced either directly or by emplace-
Sino-Burman plate along the Burma-Andaman-Java Trench. The ment as an allochthon into the accretionary prism by oblique sub-
structure of the Andaman Islands comprises an accretionary prism duction (Curray et al., 1979; Curray, 2005). We are particularly
formed by an imbricate stack of east-dipping fault slices and folds interested to determine whether the Himalayan-Tibetan orogen
that young to the west (Fig. 2), linked to a westward-shifting sub- contributed sediment to the Andaman Islands, because this might
duction zone (e.g., Roy, 1992; Pal et al., 2003). The geology of an reveal information on the early evolution of the orogen not pre-
accretionary wedge is complex, reflecting its dynamic environ- served elsewhere.
ments and involving subduction, folding, and thrusting. Deposi-
tional ages and environments can change abruptly over relatively LITHOLOGIES
short distances, and uplift leads to recycling of sediment from the
eroding wedge. Throughout subduction, new material introduced Previous Work
at the bottom of the accretionary wedge is accreted, uplifted, or
subducted. Some of the accreted material may be uplifted and The current stratigraphy (Table 1) of the Andaman Islands
brought to the seafloor. Slope basins may develop behind folds in is based on lithological mapping and can be traced back to
the accreted sedimentary rocks and trap sediment in a deep-water the pioneering work of Oldham (1885), who first divided the
environment. Simultaneously, shallow-water sediments such as Andaman geology into an older Port Blair Series and a younger
reefs can form on the prism top and be eroded and transported Archipelago Series, separated by volcanic rocks and serpenti-
down the slope. Given this inherent complexity, many accretion- nites later recognized as an ophiolite. Over the past 50 yr the
ary complex rocks are referred to as mélange; thus unraveling stratigraphy has been modified and formation names changed,
the sedimentation history in a subduction-accretionary setting is but it was not until the 1960s that paleontological constraints
a major task. were used to place the Paleogene–Neogene lithostratigraphic
Interpretation of the geology of the Andaman Islands is units within a temporal framework (Guha and Mohan, 1965;
hindered by the lack of isotopic age constraints, limited bio- Karunakaran et al., 1968). The stratigraphy now comprises four
stratigraphy, and poor outcrop exposure (Bandopadhyay and units, which, in ascending order, are Cretaceous sedimentary
Ghosh, 1999, Bandopadhyay, 2005; Pal et al., 2003, 2005). The rocks and ophiolite, the Eocene Baratang-Mithakhari Group,
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 225

90 100 drilled to the east and west of the islands (Fig. 2) that helped
Himalaya g es
n place the exposed geology within the context of the accretionary
Ra Mogok setting. Below we briefly describe the main rock units exposed
Belt

a
Shillong

rm
on the Andaman Islands and report key observations from previ-

obu
Bengal Basin Myanmar Shan
ous field and petrographic studies.

I nd
Plateau
Central
India Basin 20 Ophiolite
Bengal Fan Thailand The Andaman ophiolite contains the main components of an
ophiolite sequence that includes upper mantle–depleted harzbur-
Andam gites and dunites, lower crustal cumulate gabbros and peridotites,
and upper crustal sheeted dikes, pillow lavas, and marine pelagic
sediments (Halder, 1985; Ray et al., 1988; Roy, 1992). However,
Andaman
the sequence is tectonically disturbed, and much of the crustal
an Nicobar R

Sea
10
section is deformed and difficult to identify in the field. Pillow
Mergui lavas are abundant, but sheeted dikes have been identified only
Ma

basin in disrupted small-scale faulting and folding. Both massive and


lay

layered gabbros are recognized in the preserved ophiolite. South


idge

Pe

Andaman has the best preserved and most complete sequence of


nin

ophiolite, which extends for ~30 km from Corbyn’s Cove in the


su
la

north to Chiriyatapu in the south (Fig. 3).


Su

0
Pelagic Sedimentary Rocks
ma
tra

The topmost part of the ophiolite complex contains thin and


discontinuous lenses and streaks, and laterally continuous (at
95 E outcrop scale) bedded sequences of pelagic sedimentary rocks
consisting of jasper, chert, cherty limestone, and shales (Bando-
Myanmar padhyay and Ghosh, 1999). Outcrops commonly show evidence
Bengal Fan of significant deformation and folding (Fig. 4).

Bedded Chert
Irrawaddy delta Rhythmic alternations of centimeter-thick, milky white chert
15 N and millimeter- to centimeter-thick reddish-brown and purple
shale-mudstone beds constitute the bedded chert facies. Chert and
shale normally show uniform (0.5–4.0-cm-thick) beds that have
sharp bases and tops and planar contacts. Some 10–15-cm-thick
nch

North beds, and massive beds, of chert are present. Soft-sediment defor-
Andaman v
e

vv mation is evident in some localities. Radiolarians are preserved


obar Tr

v Andaman Sea to varying degrees in most cherts and indicate a Late Cretaceous
v
Richie’s to Paleocene depositional age, which constrains the underlying
an - Nic

arc

Archipelago ophiolite sequence to a Late Cretaceous age.


& Havelock
nic

South v Shale
a
Andam

Andaman
Volc

v At outcrop the shale facies form interbedded sequences


vv of extremely variable thickness and lateral continuity. Basaltic
vv
vv volcanic rocks occur as thin intercalations, conformable lenses,
and, at places, small crosscutting dikes. Thin beds of fine-grained
Figure 1. General location map of Andaman Islands and location of
potential source regions. Lower figure corresponds to boxed area in sandstone, siltstone, and cherty limestone are also present. There
top figure. is evidence of soft-sediment deformation and slump folds as well
as cutting by normal and thrust faults. Some shales are clearly
the Eocene to upper Oligocene Port Blair–Andaman Flysch tuffaceous with plagioclase phenocrysts, vitric fragments, pum-
Group, and the lower to upper Miocene Archipelago Group. In ice clasts, and diagenetically altered volcanic lithic fragments.
the mid-1970s the Indian state Oil and Natural Gas Commission Chlorite is abundant in the matrix of the altered tuff and tuffa-
(ONGC) carried out a detailed seismic reflection study across ceous shale. Occasionally sharp edged, cuspate or platy, fresh
the Andaman Islands, calibrated against offshore boreholes glass shards can be found (Bandopadhyay and Ghosh, 1999).
226 Allen et al.

A Andaman - Nicobar A'


0-
two-way time (sec) West East
1- 50 km

2-

3- Middle to
Andaman - Nicobar
nch

Ridge
South Andaman
4- B B'
a Tre

0-
N
Sund

1000-

depth (m)
2000-

3000-
B
B'
25 Km
Melange/ Neogene Archipelago Group
Cretaceous

A
A' deformed
Upper

pelagic cover Andaman Flysch


Paleogene
Ophiolite Mithakhari Group

Figure 2. Cross section through Andaman Ridge, based on seismic sections from Roy (1992) and Curray (2005) that show
the underlying structure of the accretionary wedge, comprising a series of folded east-dipping thrust slices.

TABLE 1. SIMPLIFIED STRATIGRAPHY OF THE ANDAMAN ISLANDS


Approximate depositional Group Formation Lithology
age range
Cross-stratified and graded sandstones, silty
mudstones and limestones marls, and chalky
Miocene to Pliocene Archipelago Group
limestones

Andaman Flysch Group Bouma sequences, sandstone-shale and


(formerly Port Blair mudstones
Oligocene–late Eocene(?)
Group)

Pebbly and coarse to fine-grained volcaniclastic


Namunagarh Grit
sandstones and grits
Mithakhari Group
Interstratified massive and graded polymict
Early to middle Eocene(?) (formerly Baratang and
Hope Town Conglomerate conglomerates, massive cross-stratified and
Port Meadow Groups)
graded sandstones, shales, and thin coals
Lipa Black Shale Pyritiferous black shale

Pillow lava, basalt, gabbro, pyroxinite,


harzburgite, serpentinite, andesite, diorite,
Late Cretaceous to
Ophiolite Group plagiogranite, rhyolite, serpentinized
Paleocene(?)
harzburgite, pyroxinite, and pelagic sediments;
radiolarian chert and hematitic mudstones

Mithakhari Group a paucity of good exposures and poor access make it difficult to
obtain continuous sections, and exposures are limited to isolated
The Mithakhari Group consists of immature gravels and stone quarries, coastal areas, and road cuts. In South Andaman
coarse- to fine-grained sandstones, pebbly to fine-grained pyro- the Mithakhari Group occurs as a north-south–trending outcrop
clastic sandstones, and minor thin beds of mudstones and coal. that extends for ~50 km, but the best sections, exposing the least
Whereas the Mithakhari Group dominates the outcrop geology of weathered outcrops, are found only near Hope Town, Mungleton,
the Andaman Islands, particularly in North and Middle Andaman, Namunagarh, and Chiriyatapu (Fig. 3). Karunakaran et al. (1968)
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 227

92.30' 92.38' 92.43' A)


Mithakhari Hope
South Group Town

Andaman
Namunagarh
Mungleton
Port Blair 11.39'
Hobdaypur

Andaman Corbyn's
12 Flysch Cove

Ophiolite Figure 3. General location map of An-


daman Islands and local geology (A, B)
for the areas on South Andaman Island
3km
sampled for this study.

B)
Andaman
Andaman
Flysch
Flysch Ophiolite

A) Mithakhari
Group

Chiriyatapu 11.29'

B) 92.42'

Hope Town Conglomerates


This unit is best seen near Hope Town on South Andaman
Island, where ~6 m of conglomerates and pebbly sandstones are
well exposed (Fig. 5), interbedded with thin beds of greenish
gray coarse- and fine-grained sandstones. The sequence shows
fining- and thinning-upward sequences with evidence of slump-
ing and soft sediment deformation. Bed contacts are generally
sharp and planar, and some evidence of fluvial channels can be
found. Conglomerates are polymict, of mainly basic-ultrabasic
sources, and with subordinate to minor amounts of andesite, sedi-
mentary limestone, and cherts plus sporadic mudstone clasts and
metamorphic quartz.

Namunagarh Grit Formation


Figure 4. Folded cherts and shales that represent the pelagic cover to This unit is characterized by coarse- to fine-grained sand-
the Cretaceous ophiolite exposed on the shore at Chiriyatapu, South
Andaman Island. stones and siltstone, with minor conglomerate at the base. On
South Andaman the type section and best exposures are found in
quarry sections near Namunagarh village (Fig. 3). These display
first introduced the term Mithakhari Group, dividing the group 3–5-m-thick, green, matrix-supported sandstones. The sandstones
into a lower Lipa Black Shale, a middle Hope Town Conglom- are well bedded and laterally persistent along the quarry sections,
erate, and an upper Namunagarh Grit Formation (Table 1). The and consist of coarse- and fine-grained beds. The coarse-grained
Lipa Black Shale is a minor unit and not well exposed, and so beds, at the base of the section, are ≥1 m thick, with sharp non-
will not be considered further. erosive contacts. The finer grained beds consist of 4–8-cm-thick
228 Allen et al.

A)

1m

B) Figure 6. Namunagarh Quarry. Thin sections show that the sediments


are largely tuffaceous with clear evidence of fresh arc volcanic mate-
rial including devitrified glass.

been derived from the ophiolite and its pelagic–shallow marine


cover. The succession, which includes thin coals and gypsum, was
deposited in a delta-slope setting with facies associations ranging
from subaerial alluvial plain to prodelta slope (Chakraborty et al.,
1999). Its depositional age is not well defined owing to the lack
of distinct biostratigraphic evidence, but shallow benthic fora-
minifers in the Hope Town Conglomerate, including Nummulites
atacicus, constrain the age from late Ypresian to early Lutetian
(Karunakaran et al., 1968). Many of the foraminifers, however,
are broken and abraded (i.e., reworked). The relationship of the
10cm Namunagarh grits to the Hope Town conglomerates is not clear,
although the Namunagarh grits are presumed to be younger.
Figure 5. Massive polymict conglomerates exposed at Hope Town
Quarry (A), South Andaman Island. These polymict conglomerates Andaman Flysch
are dominated by basic and ophiolitic clasts (B) whereas the matrix
contains abundant volcanic glass and felsitic volcanic grains, suggest-
The Andaman Flysch is a siliciclastic turbidite sequence
ing a dominant volcanic arc source, although zircon fission-track data
also show evidence of continental Mesozoic sources. deposited on a submarine fan. It is bounded between the
Mithakhari Group below and the Archipelago Group above.
The (misleading) term flysch is derived from the resemblance
beds of fine- to medium-grained volcaniclastic sandstones inter- of the turbidites to the classic Bouma turbidites described in the
bedded with thin mudstones. Some large fragments of volcanic Swiss Alps. Similar looking beds are seen throughout the Anda-
rock fragments, including elongated pumice lapilli, are present, man Islands; hence the Andaman Flysch is described as crop-
which resemble floating clasts in turbidites. For a long time the ping out over a N-S strike length of 200 km from the southern
sandstones exposed at Namunagarh stone quarries had been part of South Andaman to the northern tip of North Andaman.
described as graywacke formed from weathering and erosion of The overall thickness is not well defined, with estimates varying
accreted ophiolite (Acharyya et al., 1989). Recently Bandopad- from 750 m (Roy, 1983) to 3000 m (Pal et al., 2003). The best
hyay (2005) identified beds with abundant pyroclasts, including and most completely documented exposures are found on South
vesiculated glass fragments, pumice clasts and shards, euhedral Andaman at Corbyn’s Cove (Fig. 3), where outcrops of steep,
feldspars, and angular lithic fragments diagnostic of tuff, indicat- westerly dipping beds are seen adjacent to the pillow basalt of
ing that some of the Namunagarh Grit beds were derived from the ophiolite sequence (Fig. 7), although the nature of the con-
direct volcanic arc sources (Fig. 6). tact is uncertain. Individual sandstone beds can be traced along
The polymict conglomerates (Hope Town Conglomerate) strike for distances of several kilometers, but the total thickness is
and grits (Namunagarh Grit Formation) are interpreted as having only 250–300 m. Current directional structures in sandstone beds
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 229

METHODOLOGICAL DETAILS

The aim of this work is to provide improved understanding


of the burial and uplift history of the accretionary wedge and
provenance of the constituent sedimentary rocks, now exposed
on the Andaman Islands, through the application of thermochro-
nometric, petrographic, and geochemical analyses of these rocks.
Multiple proxies are used to get the best insight to the source
provenance and to avoid any potential bias that might arise by
relying on a single type of mineral. For example, mica is gener-
ally not present in arc-derived volcanic rocks, whereas apatite
and zircon are. Sampling is confined to South Andaman, where
Figure 7. Outcrop of the Andaman Flysch, looking south across there are accessible exposures of the main type localities, stud-
Corbyn’s Cove (see Fig. 3) toward outcrops of pillow basalts and ied by Bandopadhyay and Ghosh, (1999), Chakraborty and Pal
ophiolite. (2001), Pal et al. (2003, 2005), and Bandopadhyay (2005). Both
Middle and North Andaman are less developed and thus are less
accessible, with extensive areas of jungle that include large tribal
include flute casts, groove casts, and current bedding. The orien- reserves that are restricted to aboriginal peoples.
tation of flute casts at the base of overturned sandstone beds near Samples from South Andaman were collected for detrital
Corbyn’s Cove reveals southward-directed paleocurrents (Pal et thermochronometric, heavy-mineral, biostratigraphic, and Sm-Nd
al., 2003). The relationship between the turbidites and underlying whole rock and single grain analyses.
lithostratigraphic units is unclear. Onlap of the Andaman Flysch
with the Mithakhari Group has been reported (Chakraborty and Zircon and Apatite Fission-Track Analysis
Pal, 2001), but no supporting evidence was found in this study.
However, there is a marked change in lithology and provenance Apatite and zircon fission-track (FT) analyses are used to
with up to 50% quartz in the Andaman Flysch in contrast to define provenance and low-temperature histories, and for places
the relatively quartz-free Mithakhari Group (Pal et al., 2003). where sediments have been subjected to significant (>1.5 km)
Lithic fragments in the Andaman Flysch range from micaceous burial to determine their postdepositional burial and uplift his-
metamorphic clasts diagnostic of continental sources to cherts, tory. Fission tracks in apatite are sensitive to relatively low tem-
basalts, and weathered volcanic glass consistent with derivation peratures (typically <60–110 °C) and are ideally suited to con-
from volcanic arc and ophiolite sources. Biostratigraphic evi- strain levels of postdepositional burial and timing of subsequent
dence is vague and spans the Oligocene to the early Miocene, rock uplift and exhumation (e.g., Green et al., 1995). Zircon
ca. 36–21 Ma (Pal et al., 2003). FT data record higher temperature cooling histories (typically
~200–310 °C) and for sedimentary rocks are more suited to
Archipelago Group provenance studies that record either volcanic formation ages or
postmetamorphic cooling-exhumation ages (Carter, 1999). Sam-
The Archipelago Group represents the topmost stratigraphic ples for FT analysis were irradiated in the well-thermalized (Cd
unit of the Tertiary succession. The lower units comprise basal ratio for Au >100) Hifar Reactor at Lucas Heights in Australia.
conglomerates and sandstones, overlain by calcareous arenites Apatite grain compositions were monitored using etch pits and
of the Strait Formation. This is followed by chalk and limestone direct measurement on a JEOL microprobe using a defocused 15
with some argillaceous limestones and shale, described as the KeV Beam to prevent F and Cl migration. Durango apatite and
Melville Limestone (Shell Limestone) Formation. Deposition rock salt were used as standards. Samples with mixed ages, indi-
was mostly in a slope environment (Pal et al., 2005; Roy, 1983). cated by χ2 values <5% and large age-dispersion values (>20%),
These sedimentary rocks most likely covered most of Andaman, were deconvolved into their principal age components using the
but recent uplift and erosion means that today only small patches approach of Sambridge and Compston (1994) and incorporating
can be found on the main islands, with most exposures confined the method of Galbraith and Green (1990).
to Havelock Island and associated smaller islands to the east of
South Andaman in Richie’s Archipelago (Fig. 1). Radiolarians, Apatite Helium Analysis
planktonic foraminifers, and calcareous nannofossils from John
Lawrence Island (Singh et al., 2000) indicate a maximum depo- Apatite helium dating was used to provide additional con-
sitional age for the calcareous chalk in the Archipelago Group straint on rock uplift histories. This method complements FT
of 18.3–15.6 Ma (Burdigalian to Serravallian), but elsewhere the dating, as it is sensitive to closure at ~60 °C, the temperature at
Archipelago Group may span any age from Miocene to Pliocene which FT begins to lose sensitivity to changes in cooling rate.
(Pal et al., 2005). Helium ages were based on replicate analyses of apatite grains
230 Allen et al.

handpicked to avoid mineral and fluid inclusions. Each selected 337.1 ± 0.7 Ma) to correct for instrumental mass bias. The results
grain was first photographed, and then its dimensions were have not been corrected for common lead or ranked according to
recorded for later alpha-ejection correction. Samples were loaded degree of discordance, as the latter involves choosing an arbitrary
into platinum microtubes for helium outgassing and U/Th deter- value and is therefore open to analyzer bias.
mination. Outgassing was achieved using an induction furnace
at a temperature of 950 °C. The abundance of 4He was measured Sm-Nd Isotope Analysis
relative to a 99.9% pure 3He spike in a Pfeiffer Prisma 200 quad-
rupole mass spectrometer. The quantification of U/Th was per- Whole rock Sm and Nd isotope data in sedimentary rocks
formed on an Agilent 7500 quadrupole mass spectrometer using are widely used to fingerprint sediment source. 143Nd/144Nd
spiked solutions of the dissolved apatite. Repeated analysis of ratios are generally normalized and expressed in epsilon units as
the California Institute of Technology (CIT) laboratory Durango deviation from a chondritic uniform reservoir (CHUR), where
apatite standard gives an age of 31.3 ± 1.2 Ma (2σ), based on 39 εNd = 0. A single epsilon unit is equivalent to a difference in the
143
analyses. This error of the mean (6.7%), combined with the U/Th Nd/144Nd ratio at the 4th digit. For clastic sedimentary rocks,
and He analytical uncertainties, is used as a measure of the total εNd will in part represent the weighted average of the time when
uncertainty in sample age. the sediment sources were extracted from the mantle. When melt
is extracted from the mantle, it has a lower Sm/Nd ratio than its
Ar-Ar Age Dating of Detrital White Micas parent and therefore evolves over time to have a lower εNd than
CHUR; the residual has a higher Sm/Nd than CHUR, evolving to
The 40Ar/39Ar age of a detrital muscovite records the timing of a higher εNd over time.
cooling and exhumation (or crystallization) of rocks in the source For this study, sandstone and mudstone samples were collected
region, aiding the discrimination of potential source regions for from type localities from the Andaman Flysch and the Mithakhari
clastic sequences (e.g., Sherlock and Kelley, 2002; Haines et al., Group. Whole rock samples were ignited overnight at 900 °C to
2004). For this study, single grains were totally fused using an remove any organic material. Dissolution and analytical methods
infrared laser ablation microprobe (IRLAMP). Samples were follow Ahmad et al. (2000), with the exception that the samples
monitored using the GA1550 biotite standard with an age of 98.8 were spiked with a mixed 150Nd-149Sm spike and the 143Nd-144Nd
± 0.5 Ma (Renne et al., 1998). The calculated J value for the sam- ratios were measured on the spiked fraction. εNd is calculated
ples was 0.0138 ± 0.000069. Blanks were measured both before relative to the present day (i.e., at t = 0) using CHUR 143Nd/144Nd
and after each pair of sample analyses, and the mean of the two = 0.512638. Sm and Nd blanks were <10−3 of the sample, and the
blanks was used to correct the sample analyses for the measured laboratory Johnson Mathey Nd internal standard gave 143Nd/144Nd
isotopes. Overall mean blank levels for 40Ar, 39Ar, and 36Ar were = 0.511119 ± 5 (1σ = 24) over the period of the analyses. As whole
(378, 6, and 11) × 10−12 cm3 at a standard temperature and pres- rock εNd typically represents the weighted average of sediment
sure. The resulting analyses were also corrected for mass spec- sources, further insight into the origin of the Andaman Flysch was
trometer discrimination, 37Ar decay, and neutron induced interfer- achieved by analyzing the Nd isotopic character of single apatite
ences. The correction factors used were (39Ar/37Ar)Ca = 0.00065, grains. This was achieved using a 193 nm homogenized ArF New
(36Ar/39Ar)Ca = 0.000264, and (40Ar/39Ar)K = 0.0085; these were Wave/Merchantek laser ablation system linked to a ThermoFini-
based on analyses of Ca and K salts. Samples were irradiated for gan Neptune multicollector mass spectrometer at the University of
33 h in the McMaster University reactor (Canada). Cambridge (UK). All ablation was carried out in a He environment
and mixed with Ar and N after the ablation cell. Laser spot sizes
U-Pb Dating of Detrital Zircon were 65–90 µm. During the analytical period, standards repro-
duced to better than 0.5 epsilon units, while samples typically gave
Zircon U-Pb dating reflects the time of zircon growth, which internal precisions of 1–2 epsilon units. The full methodology of
in most cases is the igneous rock’s crystallization age. The U-Pb this in situ approach is detailed in Foster and Vance (2006).
system is mostly unaffected by high-grade metamorphism and
is effectively stable up to ~750 °C (Cherniak and Watson, 2001; Petrography and Heavy Minerals
Carter and Bristow, 2000). Zircon U-Pb ages from detrital grains
in a sedimentary rock are therefore expected to be representa- A total of 400 points were counted in six selected samples
tive of the range of crustal ages within the contributing drainage according to the Gazzi-Dickinson method (Dickinson, 1985). A
basin. Samples for this study were analyzed at University College classification scheme of grain types allowed for the collection of
London by laser ablation-inductively-coupled-plasma mass spec- fully quantitative information on the sampled sandstones. Trans-
trometry (LA-ICP-MS) using a New Wave 213 aperture imaged parent dense minerals were counted on grain mounts according
frequency quintupled laser ablation system (213 nm) coupled to to the “ribbon counting” method, and 200 minerals were counted
an Agilent 750 quadrupole ICP-MS. Real-time data were pro- also to assess the percentage of etched and corroded grains. Dense
cessed using GLITTER and repeated measurements of the exter- minerals were concentrated with sodium metatungstate (density,
nal zircon standard PL (Svojtka et al., 2001; TIMS reference age 2.9 g/cm3) using the 63–250 µm fraction treated with hydrogen
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 231

peroxide, oxalic acid, and sodium ditionite to eliminate organic reason, FT data sets from the Mithakhari Group were smaller in
matter, iron oxides, and carbonates, respectively. comparison with those from the Andaman Flysch. Nevertheless
the results adequately provide a measure of the underlying detri-
RESULTS AND INTERPRETATION tal FT signatures for each key lithostratigraphic unit. Data are
summarized as radial plots in Figures 9 and 10, and tabulated in
Biostratigraphy Table A1 (Appendix).
The interpretation of detrital apatite FT data requires com-
Attempts to identify new, more robust biostratigraphic con- paring youngest ages with sediment depositional age. If all single
trol for the Paleogene sedimentary rocks, based on nannofossils, grain ages are less than depositional age, total resetting took place
failed owing to the barren nature of the mudstones. Samples of (generally indicating burial heating to >100–120 °C; e.g., Green
mudstone were taken from the Archipelago Group, Andaman et al., 1995, their Fig. 12), whereas if the population of measured
Flysch Group, and the Mithakhari Group. Whereas nannofossils single grain ages ranges from younger to older than depositional
are present in the calcareous Neogene sedimentary rocks (Archi- age, partial resetting has taken place (generally indicating burial
pelago Group) (Singh et al., 2000), we can only conclude that heating to <100 °C). The main issue with interpreting the Anda-
nannofossils either were never present in the Paleogene sedimen- man FT data is that suitable depositional ages are missing, pre-
tary rocks or have since been dissolved by weathering or dis- venting robust use of FT analysis to determine exhumation rates.
solution below the carbonate compensation depth. Similarly, the Samples from the Hope Town Conglomerate collected at the
sampled rocks yield few (as yet undated) foraminifers. Those Hope Town quarry (Fig. 5) yield a single population of apatites
found were either broken or abraded, consistent with rework- with an age of 57 ± 9 Ma. The large uncertainty in age is due
ing. At Chiriyatapu, clasts of limestone were present within the to low uranium concentrations. The zircon ages comprise three
coarse-grained Mithakhari sedimentary rocks (Fig. 8A). These age modes, with the youngest (majority of analyzed grains) at
were found to have shallow marine reef assemblages, including 61 ± 2 Ma, within error of the apatite age. The other zircon ages
Nummulites spp., small miliolids and rare Morozovella spp., frag- indicate sources with Late Cretaceous and Permian cooling sig-
ments of rhodophyte algae, and dasycladaceans (Belzungia spp.) natures. Given that the zircon grain ages from the Hope Town
of Thanetian–Ypresian age (ca. 58–49 Ma) (Fig. 8B). Conglomerate are at or older than the Eocene biostratigraphic age
(maximum age owing to the reworked nature of the fossils), the
Fission-Track and (U/Th)-He Thermochronometry FT ages must reflect different sources. Furthermore, given that
the apatite grains have the same age as the youngest zircon ages,
Samples from the Mithakhari Group contained relatively the apatite data must also reflect provenance.
low concentrations of heavy minerals and yielded fewer apa- The youngest apatite and zircon ages at ca. 60 Ma constrain
tites and zircons than those from the Andaman Flysch. For this depositional age to being at or after this time for the Hopetown
Conglomerate Formation at this location. The apatites are notice-
ably euhedral, contain variable chlorine, and are low in uranium
B) A) (Fig. 11), typical of volcanic apatites, consistent with sample
petrography that records a dominant flux of volcanic detritus
(see below). Similar FT ages and volcanic affinities are seen in
the apatite and zircon data from the Mungleton quarry (20 km
inland from the Hope Town quarry; Fig. 5), exposing a 6-m-thick
sequence of interbedded greenish-gray, fine- to medium-grained
sandstones, siltstones, and mudstones, underlain by conglom-
erates. The succession here was described as the Namunagarh
Grit Formation. In contrast, the apatite FT data from a quarry at
Namunagarh village (also mapped as Namunagarh Grit Forma-
tion) gave a single population age of 40 ± 4 Ma, based on 24
C) grains. The zircon content of this sample was too low for FT
analysis. This quarry, studied by Bandopadhyay (2005), com-
prises tuffaceous beds with well-preserved pumice fragments and
glass shards. The apatites are euhedral and have variable chlorine
and low uranium contents typical of volcanic sources. Detrital
Figure 8. A and B show clasts of limestone from Mithakhari sedi- assemblages in these rocks contain pyroxenes, epidote, sphene,
ments. These contain shallow-marine reef assemblages that in thin sec- green-brown hornblende, and chrome spinel that are consistent
tion (C) are seen to include Nummulites spp., small miliolids and rare
Morozovella spp., fragments of rhodophyte algae, and dasycladaceans with provenance from a volcanic arc. Major diagenetic dissolu-
(Belzungia spp.) that indicate a Thanetian–Ypresian depositional age. tion is also evident, so it is questionable whether the apatite age
No in situ outcrops of these beds have been found on South Andaman. reflects source or resetting. No track lengths were measured as
Apatite Andaman Flysch Zircon
Partially reset by burial Unreset by burial
AND-1a 100 AND-1a
400
Central Age: 31±2 Ma Central Age: 61±5 Ma 300
P(X^2): 0.0% P(X^2): 0.0% 200
Relative Error: 47% 147±10
Relative Error: 29%
50 No. of grains: 43
No. of grains: 38 100
40 +2
+2
0 30 0
67±4
-2 50
-2
20
44±2
% relative error % relative error
10 25
0 0
10 20 30 40 10 20 30 40
Precision (1/sigma) Precision (1/sigma)

AND-1B 400 342±83


AND-1B 300
Central Age: 32±2 Ma 100 Central Age: 54±4 Ma 200
P(X^2): 0.7% P(X^2): 0.0%
Relative Error: 22% 150
Relative Error: 41%
No. of grains: 60 No. of grains: 32 101±14
+2 50 100
40
0 +2
30
-2 0 ?50
20 -2 41±2
% relative error % relative error
10
0 0
10 20 10 20 30 40
Precision (1/sigma) Precision (1/sigma)

AND-3 100 AND-3 616±102


600
Central Age: 32±2 Ma 80 Central Age: 56±1 Ma 450
P(X^2): 0.7% P(X^2): 73% 300
Relative Error: 39% 60 Relative Error: 0%
No. of grains: 50 No. of grains: 15 150

+2 40 +2
0 30 0
-2 20 -2 60

36±2
?30
% relative error % relative error
0 10 0
10 20 30 40 10 20 30 40
Precision (1/sigma) Precision (1/sigma)
100 AND-4
AND-4
Central Age: 68±8 Ma 600
Central Age: 31±2 Ma
P(X^2): 0.09% P(X^2): 0.0% 270±16
300
Relative Error: 20% Relative Error: 70%
50 No. of grains: 42 200
No.of grains: 47
40
+2
+2 100
0 30
0
-2 -2
20 58±4
50
% relative error % relative error
0 10 0 38±1
10 20 30 40 10 20 30 40
Precision (1/sigma) 25
Precision (1/sigma)

Figure 9. Radial plots of fission-track data from the Andaman Flysch, showing the distribution of single grain
ages. Where the sample data comprise mixed grain ages the deconvolved age modes are also shown.
Apatite Zircon
Unreset by burial Unreset by burial
Hopetown Conglomerate

Central Age: 57±9 Ma Central Age: 92±11 Ma 600


P(X^2): 69% P(X^2): 0.0% 400
Relative Error: 5% 200 Relative Error: 56% 280±32
No. of grains: 19 No. of grains: 25 200
+2
80
60 +2 129±10
0 100
40 0
-2 -2
20 61±2
10
% relative error % relative error
0 0 30
10 20 10 20 30
Precision (1/sigma) Precision (1/sigma)

Namunagarh Grit
Mangluton Mangluton
Central Age: 69±9 Ma Central Age: 75±8 Ma 400
P(X^2): 0.0% 400 314±51 P(X^2): 0.0% 226±21
Relative Error: 65% 300 Relative Error: 42% 200
No. of grains: 21 No. of grains: 18
200
+2 151±68 100
0 +2
100 0
-2
-2
50 55±8 59±2
% relative error 50
% relative error
0 0
10 20 10 10 20 30 40 30
Precision (1/sigma) Precision (1/sigma)

Namunagarh Chiriyatapu
Central Age: 40±4 Ma 100 Central Age: 143±22 Ma
P(X^2): 96.0% 80 P(X^2): 0.0%
Relative Error: 0% Relative Error: 93% 800
600
No. of grains: 24 60 No. of grains: 38 400
+2 329±17
+2
0 40 0 200
-2
-2
100
20 91±7

% relative error 10 % relative error


0 0 40±3
10 20 30 30
10 20 30
Precision (1/sigma) Precision (1/sigma)

Figure 10. Radial plots of fission-track data from the Mithakhari Group, showing the distribution of single grain
ages. Where the sample data comprise mixed grain ages the deconvolved age modes are also shown.
234 Allen et al.

1.5

Apatite grain composition


Hopetown Conglomerate

Andaman Flysch
1
Chlorine (wt%)

Figure 11. Plot comparing uranium and


chlorine for apatite grains from the An-
daman Flysch and Hopetown Conglom-
erate, Mithakhari Group. The plot clearly
shows that the apatite came from differ-
ent sources. The low uranium Hopetown
Conglomerate apatites, which also have
variable amounts of chlorine, are typical
0.5 of volcanic apatites. The volcaniclastic
petrography and euhedral form of the
apatites support this and suggest that
they came from arc related sources.

0
0 20 40 60 80 100 120
Uranium (ppm)

a result of the low spontaneous track density. A sample of the is not robust (Pal et al., 2003). While both zircon and apatite FT
Namunagarh Grit Formation from Chiriyatapu on the southern ages are older than the Oligocene, some partial resetting (burial-
coast of South Andaman yielded only zircon, which produced related heating) may have taken place.
three FT provenance age modes at 40 ± 3 Ma, 91 ± 7 Ma, and
329 ± Ma. The youngest zircon FT age mode, at 40 ± 3 Ma, indi- Thermal Modeling
cates that the Thanetian–Ypresian limestone clasts found in these
rocks were derived from erosion of significantly older material To constrain postdepositional thermal history, apatites from
(by ca. 16 Ma). Thus we conclude that the age of the Mithakhari one of the samples were analyzed by the (U/Th)-He method
Group is constrained to be younger than 60 Ma at two locations, (Table A2, Appendix). Replicates yielded an FT corrected age
and younger than 40 Ma at a third location. Nevertheless, the of 16 ± 1 Ma, crudely representing the time at which the sam-
volcanic origin of the material, with little sign of reworking, sug- ple cooled to <50 ± 10 °C. To define more robustly the sample
gests that the rocks were deposited only shortly after the recorded postdepositional thermal history, the FT and helium data can be
mineral ages. This is consistent with diachronous deposition in jointly modeled. Ideally this requires incorporating a sample
local basins, as expected in a subduction zone setting (e.g., Draut depositional age, but, as discussed, this is not well defined, and
and Clift, 2006). so we resorted to using the youngest detrital ages as an upper
Apatite and zircon FT data from four samples collected limit for the time of deposition. For the Andaman Flysch, deposi-
through the section of Andaman Flysch at Corbyn’s Cove are tion must have taken place at or after 35–30 Ma on the basis of
closely similar in age (most are 35–40 Ma) to the youngest age the youngest argon mica ages. In addition the Andaman Flysch
modes in the Namunagarh Grit Formation beds. The youngest sedimentary rocks must have been at or near surface tempera-
ages comprise most analyzed grains for zircon and all analyzed tures by the middle Miocene (ca. 16 Ma) because in the Hobday-
grains for the apatites. In addition, two of the samples show zir- pur area of South Andaman (Fig. 2) these rocks are seen to be
con age modes at 58–67 Ma, similar to the age modes detected in conformably overlain by sedimentary rocks of the Archipelago
the Mithakhari Group. Three of the samples also show Mesozoic Group, although in the eastern part of the island Archipelago
and Paleozoic FT ages. The Andaman Flysch is widely consid- rocks are juxtaposed with rocks of the Mithakhari Group along a
ered to be Oligocene in age, although biostratigraphic evidence faulted contact (Pal et al., 2005). A phase of reburial up to ~60 °C
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 235

is also required on the basis of recent diagenetic evidence from Detrital Argon Data
the Archipelago Group sedimentary rocks (Pal et al., 2005).
With these constraints the combined apatite FT and helium data Four samples were analyzed from the Andaman Flysch,
were modeled, using the data-driven modeling approach of Gal- Corbyn’s Cove section, which yielded 114 grain ages. One addi-
lagher (1995) that combines multicompositional FT annealing tional sample from Mithakhari Group sedimentary rock at Chiri-
and helium diffusion models (Ketcham et al., 1999; Meesters yatapu was also analyzed (33 grain ages). The results are dis-
and Dunai, 2002). The best-fit solution (Fig. 12) highlights played as probability plots in Figure 13 and tabulated in Table A3
three important stages: (1) a requirement for deposition and (Appendix). Ages for the Mithakhari Group sample extend from
burial to peak temperatures of ~80–90 °C between ca. 30 and 70 Ma to the Archean, with the bulk of ages spanning the Meso-
25 Ma, (2) uplift to the surface between ca. 25 and 20 Ma, and zoic and Paleozoic. The Andaman Flysch ages are in general
(3) reburial in the Miocene (ca. 25–5 Ma) to temperatures of younger, ranging from 30 Ma to the Proterozoic. Most of the
50–60 °C (broadly equivalent to depths of ~1–1.5 km, assuming ages (71 grains) are <200 Ma, with 20 grains falling between
geothermal gradients of 30 °C/km). 30 and 60 Ma, and 32 grains between 60 and 80 Ma. Given that
postdepositional heating was modest, as constrained by the FT
data, which measure lower closure temperatures than argon, the
young mica ages, some of which are younger than the FT data,
Oldest point to an Oligocene depositional age: i.e., the Andaman Flysch
depositional age
Depositional of Archipelago was deposited at or after 30 Ma.
age range Group
0
Modeled results Detrital Zircon U-Pb Data
20 are consistent with
Temperature °C

Archipelago Group
40 (Pal et al., 2005) Preliminary zircon U-Pb data from the Andaman Flysch
60 diagenetic data
Youngest that indicate (Fig. 14; Table A4, Appendix) show a large Proterozoic popula-
80 U-Pb and maximum burial tion, with a few grains showing Archean and Cretaceous–Eocene
argon temperatures of
100 source
range of max burial ages. The youngest zircons (five grains) give an average age of
45-55°C
120 ages 48 ± 5 Ma, which overlaps both zircon FT and mica ages. These
50 40 30 20 10 0 grains are euhedral and have concentric zoning typical of mag-
Age (Ma) matic zircon, suggesting a direct contribution from an early
20 Eocene igneous source rather than reworking of older sedimen-
Obs. Age : 36 Ma
Number of tracks

Pred. Age : 35 Ma tary deposits.


Obs. Mean length : 13.24μm
10 Pred. Mean length 13.66μm
Obs. S.D. : 1.61μm Petrography and Heavy Mineral Data
Pred. S.D. : 1.61μm
Oldest track 39 Ma
The very fine to fine-grained turbidites of the Andaman
1 2 3 4 5 6 7 8 9 10 1112 1314151617181920 Flysch have an intermediate quartz content (Q 49 ± 2%) (Fig. 15)
Track Length (microns) and contain subequal amounts of feldspar (F 22 ± 6%; plagio-
clase feldspar 39 ± 9%) and lithic grains (L 29 ± 6%) as seen
He/P
Helium diffusion model 50 in Table A5A (Appendix). The latter chiefly consist of very low
40
rank to medium rank and subordinately high rank metapelite-
He Obs : 13.0 metapsammite grains (Garzanti and Vezzoli, 2003), indicating
He Pred: 14.7 Ma 30 provenance from a collisional orogen (Dickinson, 1985). Volca-
Grain R: 75.0 microns
20 nic grains are mainly felsitic and are significant (Lv 8 ± 1%), sug-
gesting subordinate contributions from a volcanic arc. Transparent
10
heavy-mineral assemblages in the Andaman Flysch (Table A5B,
0 Appendix) have very low concentrations (0.3 ± 0.2%). Ultra-
50 40 30 20 10 0
Age (Ma)
stable species (zircon, tourmaline, rutile, chromian spinel) rep-
resent 21 ± 2% of the assemblage, and ferromagnesian minerals
Figure 12. Best-fit thermal-history model for apatite fission-track and (pyroxenes, amphiboles) are invariably absent, clearly showing
(U-Th)/He data from the Andaman Flysch. Key features are a require- the strong influence of diagenetic dissolution.
ment for maximum burial temperatures between ca. 20 and 30 Ma
and rapid cooling at ca. 20 Ma. Values given are for observed data
The very coarse grained sandstones of the Namunagarh Grit
(Obs.), model predicted values (Pred.), and standard deviation (S.D.). Formation chiefly consist of plagioclase and microlitic, lath-
For the helium plot, R is the grain radius and He/P is the helium con- work, and felsitic volcanic grains. The quartz content is low (Q
centration profile. 14 ± 10%), and only a few K-feldspar and granitic lithic grains
236 Allen et al.

Andaman Flysch-1A
(90 grains)
Mithakhari Group

Relative probability
Relative probability

Chiriyatapu
(33 grains)

0 500 1000 1500 2000 2500 3000


0 400 800 1200 1600 2000 2400 2800 Age (Ma)
Age (Ma)

Andaman Flysch
Relative probability

Corbyn’s Cove
Relative probability

(114 grains)

0 200 400 600 800 1000 1200 1400 1600 0 50 100 150 200 250 300 350 400
Age (Ma) Age (Ma)

Figure 14. Probability plots showing the distribution of concordant de-


trital zircon U-Pb ages analyzed in this study.

Younger grain ages


Relative probability

(71 grains)
occur, indicating provenance from a largely undissected volcanic
arc (Marsaglia and Ingersoll, 1992). A few terrigenous (shale-
sandstone) and very low rank to medium rank metapelite and
metapsammite grains also occur, suggesting minor reworking
of sedimentary and low-grade metasedimentary sequences. Sig-
nificant chert, along with traces of metabasite and serpentinite
grains, point to minor contributions from oceanic rocks.
In the Namunagarh Grit Formation, opaque grains represent
21 ± 16% of total heavy minerals. Transparent heavy-mineral
assemblages have poor concentrations, pointing to significant
20 60 100 140 180 220
diagenetic dissolution. These mainly include pyroxenes (mostly
Age (Ma) green augite; 35 ± 24%) and epidote (34 ± 19%), with subordinate
Figure 13. Probability plots showing the distribution of detrital argon sphene (11 ± 8%), green-brown hornblende (8 ± 8%), chromian
mica ages analyzed in this study. spinel (8 ± 8%), minor apatite (2 ± 1%), garnet, rutile, and other
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 237

Figure 15. Petrography of Tertiary sand-


Q Lmf stones of South Andaman. Whereas
composition of the Namunagarh Grit
A) Andaman Flysch B) Formation sandstones points to prov-
enance from an undissected-transitional
Namunagarh Grit
+ Bengal Fan magmatic arc, a recycled orogenic prov-
Irrawaddy 125 enance is clearly indicated for the An-
+ Irrawaddy 180 + daman Flysch (Dickinson, 1985). These
O
O
O
Irrawaddy 80 very fine to fine-grained turbidites com-
rec pare closely with modern sands of ho-
ton

yc
led mologous grain size from the Irrawaddy
cra

oro
ge O
Delta. Q—quartz; F—feldspars; L—
n O
O lithic grains; Lv + Lmv—volcanic and
magmatic arc low-rank metavolcanic; Ls + Lms—sed-
O
O O O imentary and low-rank metasedimenta-
F L Lv + Lmv Lms + Ls ry; Lmf—high-rank metamorphic.

titanium oxides, consistent with provenance from a volcanic arc. Sm/Nd ratios of 0.2–0.5, therefore for non-age–corrected apatite
It must be kept in mind, however, that the original composition of the 143Nd/144Nd ratio tends to reflect the Nd isotopic composition
the Namunagarh Grit Formation sandstones (abundant volcanic of their parent whole rock, enabling a direct linkage between ther-
glass and ferromagnesian minerals) was markedly different with mochronometric data and source in terms of Nd isotopic signa-
respect to orogen-derived turbidites, and therefore that dissolu- ture. This approach is particularly effective if the source rocks are
tion reactions of ferromagnesian minerals may have progressed young and/or their potential range of Sm-Nd has been quantified.
at different rates and in the presence of partially buffered or even It is also important to note that this approach yields information
saturated interstitial waters. that is not the same as sediment whole rock values that record an
average value (weighted by Nd concentration) of source rocks.
Sm-Nd Whole Rock and Single Grain Analyses It is also possible that the detrital apatite may be dominated by
a single apatite-rich source. The majority (~60%) of the single
Sandstone and mudstone samples from the Mithakhari grain analysis of apatite from the Andaman Flysch produced εNd
Group yielded whole rock εNd values of –7.2 and 3.1(+), respec- values of –5 to +5, typical of juvenile volcanic whole rock val-
tively. The Andaman Flysch samples gave whole rock values of ues and contrasting significantly with the bulk sediment values
εNd –11.1 and –8.2 for the sandstone sample and mudstone sam- of εNd –11.1 and –8.2 (Fig. 17; Table A6A, B, Appendix). The
ple, respectively. Such values are lower than those of Himalayan restricted range of εNd values for these apatites, which gave par-
sources, which typically range between εNd –12 and –26 (Galy tially reset FT ages between 30 and 40 Ma (Table A1, Appen-
and France-Lanord., 2001) and are inconsistent with Indian dix), imply a single source region for the apatite grains. Little
Shield sources (Peucat et al., 1989). Very little data exist for is known about the apatite Sm-Nd systematics for the source
the relevant Myanmar source regions. Analyses of modern river rocks of the Andaman Flysch. However, as Figure 17 shows, the
sediment from the Irrawaddy River give εNd values between Andaman Flysch apatite Nd isotope data contrast with apatite Nd
−10.7 (Colin et al., 1999) and −8.3 (this paper: Table 2). The data from a Holocene sand dominated by Himalayan sources col-
Western Indo-Burman ranges (Fig. 16), which may have been lected from the Bengal Basin near Joypur, West Bengal. This plot
exhuming by the Oligocene (Mitchell, 1993), have more nega- clearly shows that there is little evidence for material eroded from
tive values (Colin et al., 2006; Singh and France-Lanord, 2002; Himalayan sources in this sample.
this paper, Table 2, and Table A6A, Appendix). Figure 16, which
compares the whole rock data from the Andaman Flysch with DISCUSSION
possible source areas, shows that the most probable source area
lies within Myanmar. Constraints on Sedimentation and Uplift
To understand the source area in more detail, and in particu-
lar the significance of apatite from the Andaman Flysch, which New thermochronometric evidence and field observations
appears to come from more distant sources in comparison with from type locations on South Andaman, combined with existing
local arc and ophiolite sources in the Mithakhari Group sedi- biostratigraphy and petrography, provide an improved chronol-
mentary rocks, Nd isotope composition was measured on single ogy for deposition and uplift. The accretionary setting means
grains of apatite from the Andaman Flysch. Apatite typically has that sedimentation history will be intimately tied to subduction
TABLE 2. TYPICAL SIGNATURES OF POTENTIAL SOURCE REGIONS
40
Source region Rock description, heavy minerals, and petrography Sm–Nd U-Pb ages of zircons Ar-39Ar ages of white Zircon fission-track
Whole rock mica ages
εNd(0)
Himalaya (Southern Flank)
Drained by Ganges and tributaries In the Oligocene, less metamorphosed “Higher Signature today: Signature today: Signature today: Signature today:
Oligocene bedrock characteristics Himalayan protolith” would likely have been Higher Himalayan As determined from As determined from No FT data available
interpolated from Eocene and Miocene exposed. Rocks were low-grade metamorphic, as determined Ganges sediments: Ganges tributaries: for Ganges
foreland basin rocks sub-garnet grade, nonmicaceous (as determined from Ganges: av. mostly >400 Ma– Neogene peak; sediment; He data
Bedrock signal today taken from modern from Miocene foreland basin rocks).1 –17.52 Precambrian. Grains subordinate grains <55 Ma, Pliocene–
river sediments; higher Himalaya <30 Ma rare. Grains 30– spanning to Pleistocene peak9
dominates detritus 400 Ma very rare5 Precambrian7
Oligocene bedrock: Oliogocene bedrock:
Eocene: av. –83 Oligocene bedrock: Oligocene bedrock: Eocene: peaks at 45
Miocene: Eocene and Miocene: Micas absent in Eocene, Ma, 119 Ma, and
–14 to –174 Tertiary grains absent; very rare in late 343 Ma;10
65–400 Ma grains very Oligocene–Miocene8 Miocene: peaks at 30
rare6 Ma, 60–75 Ma,
117 Ma, 300–370
Ma11

Burma
Region drained by Irrawaddy Cretaceous arc rocks and Triassic forearc-backarc Irrawaddy: Irrawaddy: Irrawaddy: Irrawaddy:
Data from modern Irrawaddy River sediments on metamorphic basement. Mogok –10.7, –8.313 Proterozoic, Ordovician, Cretaceous to Miocene Neogene,
sediment. Shan-Thai block lies to east, schists, gneisses, and intrusives, Shan-Thai Jurassic–Cretaceous, grains; peak 30–55 Paleogene,
forearc-backarc of Indo-Burman Proterozoic–Cretaceous sedimentary rocks on Paleogene, and Ma15 Cretaceous
Ranges (IBR) to west. schist basement. Irrawaddy River sediment plots Neogene grains14 grains15
Paleocontinental margin. in “recycled orogenic” province field of QFL plot
(Dickinson, 1985).12

Western Indo-Burman Ranges Dark-gray, very fine grained volcanic arenite of 3 samples give Precambrian: Archean and No mica present in pre- Cretaceous and
Accretionary prism; may have been Eocene (Oligocene?) age. Only limited mica and values of –4.0; 1 Proterozoic; Cretaceous Miocene sedimentary Paleocene to
exhuming and thus a sediment source heavy minerals. Many shale and siltstone lithics.16 sample (farther and Paleogene (55–65 rocks16 Eocene grains16
in late Eocene–Oligocene. Data north) gives a Ma) grains16
determined from Arakan modern river value of –7.416
sands (R. Allen et al., unpublished
data) unless otherwise stated.

Indian Shield
Dominantly Arc hean craton. Subordinate Predominantly gneisses and granites; opaques, Values from south Dominantly Archean19 No data available No data available
Proterozoic mobile belts and orthoproxene, and sillimanite found in rivers and east craton
Gondwanan sedimentary cover. draining east craton.17 only;
–30 to –3618
1
Najman and Garzanti (2000).
2
Galy and France-Lanord (2001).
3,4
Robinson et al. (2001); Najman et al. (2000).
5
Campbell et al. (2005).
6
DeCelles et al. (2004).
7
Brewer et al. (2003); Najman and Pringle (unpublished data).
8
DeCelles et al. (1998); Najman and Garzanti (2000).
9
Campbell et al. (2005).
10
Najman et al. (2005).
11
Najman et al. (2004, 2005).
12
Mitchell (1993); Pivnik et al. (1998); Bertrand et al. (1999); R. Allen et al. (unpublished data).
13
Colin et al. (1999); R. Allen et al. (unpublished data).
14
Bodet and Schärer (2000).
15
R. Allen et al. (unpublished data).
16
R. Allen et al. (unpublished data).
17
Mallik (1976).
18
Peucat et al. (1989); Saha et al. (2004).
19
Mishra and Rajamani (1999); Auge et al. (2003).
Andaman Flysch
Namunagarh Grit
IndoBurman Ranges
Modern Irrawaddy River
-30 to -36
Indian Shield
Himalayan Sources

present-day
Brahmaputra /
Ganges TSS
HHS
LHS

-28 -24 -20 -16 -12 -8 -4 0 4 8


Epsilon Nd
Figure 16. Plot comparing Sm-Nd whole rock values from the Namunagarh Grit Formation and Andaman Flysch For-
mation with data from possible source regions (Colin et al., 2006; DeCelles et al., 2004; Peucat et al., 1989; Singh and
France-Lanord, 2002). The data do not support the Himalayan region as the main source. TSS—Tethys Sedimentary
Sequence; HSS—Higher Himalayan Sequence; LHS—Less Himalayan Sequence.

0.5134

0.5132 Holocene, Bengal delta


Andaman Flysch
0.5130 +8

0.5128 +4

0
143 Nd / 144 Nd

0.5126

Epsilon Nd
0.5124 Field of typical Himalayan -4
apatites
0.5122 -8

0.5120 -12

0.5118 -16

0.5116 -20

0.5114 -24

0.5112
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
147Sm / 144Nd
Figure 17. Single grain Nd measured on apatite from the Andaman Flysch Formation compared with a Holocene sand
from the Bengal Basin known to have its source in the Himalayas. The clear difference between these two samples sup-
ports the Sm-Nd whole rock data plotted in Figure 16.
240 Allen et al.

behavior, offscraping, accretion, and uplift. While in such a set- modeling, which indicates that burial to maximum temperatures
ting uplift may be viewed as a more or less continuous process, of ~90 ± 5 °C was reached ca. 25–20 Ma, ending with a phase
we recognize some distinct uplift events that have affected the of rapid cooling at ca. 20 Ma. This implies a discrete episode
South Andaman geology. of major uplift at ca. 20 Ma consistent with deposition of the
The earliest conglomerates and grits of the Mithakhari Archipelago Group shallow marine volcanic-rich sedimentary
Group record subaerial exposure and erosion of the arc sequence rock from ca. 18 Ma onward (Pal et al., 2005; Singh et al., 2000),
and shallow marine limestone cover, as described above, con- which was deposited on top of the Andaman Flysch. Uplift at
straining the age of ophiolite as pre–Mithakhari Formation (early ca. 20 Ma is also regionally significant, coinciding with major
to middle Eocene) and post–shallow marine sedimentation. The stratigraphic changes in the Irrawaddy Delta, Mergui Basin
petrography of the Mithakhari Group is indicative of an undis- (Fig. 1), in sediments accreted onto the Andaman-Nicobar Ridge
sected volcanic arc with minor reworking of sedimentary and (Curray, 2005), and formation of an unconformity in the Indo-
low-grade metasedimentary sequences and minor contributions Burman Ranges (Acharyya et al., 1989). Uplift of the Andaman
from oceanic igneous rocks. Ridge at 20 Ma is also coincident with the uplift of the Owen and
Our data provide maximum and minimum age constraints Murray Ridges in the Arabian Sea (Mountain and Prell, 1990),
for deposition of the Andaman Flysch. Detrital Ar-Ar mica and suggestive of a wider plate tectonic trigger.
FT data constrain maximum depositional age to ca. 30 Ma. The An outstanding question is, what was the original maximum
minimum depositional age is constrained by the thermal history thickness of the Andaman Flysch? Hydrocarbon-exploration

180-
convergence (mm/yr)
Rate of Indian

150-

100-
Lee and Lawver, 1995

Guillot et al., 2003


convergence
(from N)
Angle of

50-
40-
30-
20-
10-

deposi-
uplift

Archipelago Group Figure 18. A comparison between In-


tion
dian convergence history (after Lee and
unconformity Lawver, 1995, and Guillot et al., 2003)
and the uplift and sedimentation history
deposi-
uplift

Andaman Flysch of the rocks studied from South Anda-


tion man Island.
contact unclear
Namunagarh
Mithakhari

continuous
Group

Grit
Hopetown deposition
Conglom. ?????
unconformity

Ophiolite pelagic uplift deposition


cover sediments

? ? ?
Ophiolite obduction
? ? ?
0 20 40 60 80 100
Age (Ma)
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 241

wells drilled to the east and the west of Middle Andaman indicate TARIM QD AL
thicknesses of <1000 m (Roy, 1992), whereas outcrop estimates KL NORTH CHINA
range from ~750 to 3000 m (Pal et al., 2003). Results from ther- QT QS SPG
mal history modeling indicate burial of the sampled Andaman
Flysch beds beneath at least 2.5 km of cover (assuming geother- LHASA
mal gradients in the range of 35–30 °C). In a growing accretion- SOUTH CHINA
ary wedge, burial is as likely to be due to thrust stacking as it
is to sedimentation. Crucially ~2.5 km of burial is incompatible India WB
with the unannealed apatite data from the stratigraphically lower
Mithakhari Group, highlighting the fact that these different units
must have been deposited in different locations (sub-basins?)

SIBUMA
within the accretionary wedge. Thermal history modeling also

INDOCH
shows a phase of accelerated cooling from ca. 10 to 5 Ma. This
is probably tied to a phase of regional Miocene–Pliocene uplift
linked to spreading in the Andaman Sea driven by increases

SU
in the subduction rate and dip of the subducted slab (Khan

INA
and Chakraborty, 2005). This last phase of uplift, the result of
squeezing between subduction in the west and extension in the
east, is likely responsible for the present topography of the Anda- QD = Quidam terrane
man Islands. QS = Qamdo-Simao terrane SWB
The relationship of these episodes of uplift and erosion to
QT = Qiantang terrane
regional events can be appreciated when these new constraints
KL = Kunlun terrane
are compared against reconstructions of Indian plate convergence
AL = Alao Shan terrane
history. Figure 18 plots the India-Asia convergence rate and con-
SPG = Song Pan Ganzi terrane
vergence angle from the studies of Lee and Lawver (1995) and
WB = West Burma
more recently from that of Guillot et al. (2003). The graph shows
SWB = Southwest Burma
that the studied outcrops of Andaman Flysch were deposited at a
time of more northerly convergence and ended when subduction
Figure 19. Map showing the principal terranes that have amalgamated
shifted to a less oblique angle. Seismic lines taken from across to form SE Asia (after Metcalfe, 1996).
the Andaman Arc show how the angle of subduction can influ-
ence wedge development where the intensity of deformation,
which increases from north to south, changes with obliquity of
subduction (Curray, 2005). This relationship may explain the down to peninsular Thailand, adjacent to the forearc (Metcalfe,
evidence for a discrete episode of uplift, involving the Corbyn’s 1996) (Fig. 19).
Cove Andaman Flysch, at ca. 20 Ma. The data show a clear change in petrographic and isotopic
characteristics between the Mithakhari Group and the overlying
Sediment Provenance Andaman Flysch. Andaman Flysch composition represents the
first sedimentary material to plot in the recycled orogenic prov-
The Mithakhari Group shows clear evidence of a dominant ince (Fig. 15). Initial results from the pilot study of Sm-Nd whole
contribution from an arc, in keeping with its interpreted forearc rock values differ markedly between the Mithakhari Group and
basin depositional environment and in line with previous prov- the Andaman Flysch, particularly when the finer grained facies
enance work on the basis of petrography (Bandopadhyay, 2005). are compared. The positive εNd signal of the Mithakhari Group
The rocks plot in the magmatic arc field on the QFL petrographic contrasts strongly with the more negative εNd signal (−11) of the
plot (Fig. 15); volcanic fragments and glass shards are common, Andaman Flysch mudstone, suggestive of a contribution from
and mineral grains, e.g., apatite, are euhedral. A positive whole older continental crustal sources (Fig. 15). Detrital micas are of
rock εNd signal (+3) is indicative of derivation from a mafic- radically different ages, with the age peaks for the Mithakhari
juvenile source. The source was likely that of the Cretaceous– Group falling within the Precambrian–Paleozoic spectrum,
Eocene arc that stretched from the Himalaya collision zone whereas those from the Andaman Flysch are typically of late
through Myanmar to Sumatra (Mitchell, 1993), as reflected in Mesozoic and Tertiary age (Fig. 13). A provenance contrast is
the FT ages of the euhedral apatites. A subordinate continental not apparent from the zircon FT data (Figs. 9 and 10), as the
source is confirmed by petrography and Precambrian–Paleozoic detrital modes are similar. The apatite FT data cannot be directly
Ar-Ar mineral grain ages. Such a source was probably the con- compared, as the Andaman Flysch apatites have seen some post-
tinental margin of the Shan-Thai Block (Sibumasu) extending depositional partial resetting. However, comparison of apatite
242 Allen et al.

chemistry (chlorine-uranium content, Fig. 11) does highlight a older than 50 Ma, present in both the Mithakhari Group and the
major difference in the apatite source. Andaman Flysch, are consistent with known sources along the
To some extent these differences between the Mithakhari western margin of the Shan-Thai Block (Sibumasu) adjacent to
Group and the Andaman Flysch may reflect different mixtures the forearc. The presence in the Andaman Flysch of a small but
of the same sources. The continental-derived material that domi- recognizable population of mica ages between 30 and 40 Ma
nates the Andaman Flysch represents only a minor contribution is not consistent with these Sibumasu sources, as the youngest
in the arc-dominated Mithakhari Group. Variations may have as granites on the Thai peninsula are ca. 50 Ma (Charusiri et al.,
much to do with depositional setting than source variation. In 1993). Possible sources for the Tertiary mica ages might instead
this regard, sedimentary rocks of the Namunagarh Grit Forma- include a region affected by India-Asia collision at ca. 50 Ma,
tion may represent a perched forearc basin, whereas the Anda- possible candidates being either the nascent Himalaya to the
man Flysch represents deep-water turbidites. north, Transhimalaya, or a more northeastern source—namely,
It was proposed by Karunakaran et al. (1968) and Pal et Myanmar locations of the Burman-Thai Block—where Tertiary
al. (2003) that these “recycled orogenic” Andaman Flysch rocks metamorphism, magmatism, and grain isotopic characteristics
were sourced by the Irrawaddy River of Myanmar and deposited and ages are ascribed to the effects of the India-Asia collision
in a forearc basin. In contrast, Curray et al. (1979) considered (Bertrand et al., 1999; 2001, Bodet and Schärer, 2000; Barley et
the Andaman Flysch to be trench sediments offscraped from al., 2003). Isotopic characteristics of these Myanmar rocks are
the Himalayan-derived Bengal Fan on the downgoing Indian incompletely documented at present. Those that are available are
slab. Significant contribution from the northward-drifting cra- taken from bedrock, and analyses of sands from the Irrawaddy
tonic Greater India can be ruled out by the dissimilar petrog- River, which drains the areas under discussion—i.e., the Central
raphy, Sm-Nd signature, and mineral cooling ages (Fig. 16; Myanmar Basin, to its east the Mogok Belt and Shan Plateau, and
Table 2) as well as by the apatite Nd data (Fig. 17). No apa- to its west the Indo-Burman Ranges (Bodet and Schärer, 2000;
tites were measured with significantly negative εNd values (e.g., Colin et al., 2006; Singh and France-Lanord, 2002; Table 2, this
−30) diagnostic of Indian Shield sources. The mica argon ages study)—as verified by 30 Ma Ar-Ar ages from the Irrawaddy

Paleogene
nas
cen
Eurasia
fore tH Tsangpo River, flows
land imalay
bas a along Himalayan
ins
suture draining east to
South China Sea
rivers draining
southern flanks continental Figure 20. Cartoon illustrating the Paleo-
drainage gene paleogeography and paleodrainage
of SE Asia. Although this sketch simpli-
volc

fies the location of the subduction zone,


Sh

Greater India Early the Asian margin and Andaman crust


anic

an

Bengal would have changed throughout the Pa-


-T

Fan
arc

leogene as Greater India moved north-


ha

ward. The cartoon does serve to illustrate


i-

Paleogene the general location of continental sourc-


Su

location of es and drainages to the Andaman region.


Andaman
nd

crust
ala
nd
vo und
lca a T
S

nic re

Indian Ocean
ar nch
c
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 243

region (Table 2). Discrimination between these two possible similar to data obtained from the Andaman Flysch, showing a
sources, the Himalayan or Myanmar region, requires an under- mica age peak of ca. 30–60 Ma with a population of older Creta-
standing of the different paleogeography and paleo-drainage of ceous grains and zircon FT ages of Miocene and Paleogene (20–
the region in the Oligocene in comparison with the present day, 65 Ma) to Cretaceous (Table 2). Nevertheless, some disparities
as discussed below (Fig. 20). are evident. The lack of zircon grains with Jurassic U-Pb ages
Prior to the India-Asia collision, the northward extension of for the Andaman Flysch is surprising, considering their preva-
the Sunda Arc as far as Pakistan permits correlation between the lence in the Mogok Belt (Barley et al., 2003) and their presence
southern margin of Asia in the Himalayan region with equiva- in modern Irrawaddy River sand (Bodet and Schärer, 2000).
lent rocks in Myanmar (Mitchell, 1993) and farther south. Sub- Recourse to a Himalayan contribution through the Bengal Fan
sequent to continental collision at ca. 50 Ma (Rowley, 1996), the is not required to explain the recycled orogenic component of
Himalayan thrust belt started to develop in the north while the the Andaman Flysch. However, a minor contribution from the
eastern (Myanmar) region remained an active continental margin Himalayan thrust belt cannot be ruled out: for example, grains
with, for example, the Central Basin along which the Irrawaddy aged 100–1500 Ma in the Andaman Flysch are common to both
now drains, the site of the continental margin between the sub- Himalayan and Myanmar rock types. In addition, the intermedi-
duction zone to the west and the arc to the east (Pivnik et al., ate whole rock εNd values may represent a mixture between a
1998). Movement along the Sagaing dextral strike-slip fault (ini- more negative Himalayan source and more positive Myanmar
tiation in pre-Miocene times: Mitchell, 1993) and the opening of sources, as implied by the single grain apatite analyses. The pos-
the Andaman Sea in the late Miocene (Curray, 2005; Khan and sibility of dual provenance, with Himalayan-derived Bengal Fan
Chakraborty, 2005) were responsible for changes in the coastal material and Myanmar-derived material meeting and mixing in
paleogeography of this region. the trench, is a model that would be consistent with the regional
In the Paleogene the Yarlung-Tsangpo River (draining seismic data, which show folding and uplift of Bengal Fan sedi-
the Himalayan Arc and suture zone) probably drained into the ments at the base of the slope (Curray, 2005). Definitive discrim-
South China Sea prior to its capture by the Irrawaddy River, and ination between Myanmar and Himalayan sources, if possible
finally the Brahmaputra River in the Neogene (Clark et al., 2004, at all, awaits more information on the source rock geology of
2005). Thus, Ganges-type rivers—those draining only the south- Myanmar, in particular that of the Shan Plateau and Mogok Belt,
ern slopes of the Himalayan mountain belt, with no arc com- and analysis of as-yet-undrilled Oligocene sediments preserved
ponent—would have contributed the Himalayan signal to the in the Bengal Fan, and coeval sedimentary rocks of the Central
Indian Ocean (Fig. 17). The southern slopes of the Himalayas Myanmar Basin.
consist of Indian crust. Characteristics of these rocks changed
with time as Neogene Himalayan metamorphism subsequently CONCLUSIONS
increased the metamorphic grade and overprinted the metamor-
phic cooling ages of minerals and rocks exposed at the surface The Mithakhari Group was deposited in the late Paleocene–
today. In contrast, the northeastern Myanmar region was devoid Eocene, with a maximum age of ca. 60 Ma. This current study
of Indian crustal rocks, and no thrust belt was present to bar the shows that the sedimentary rocks are predominantly arc derived
Asian and arc sources from draining south. Prior to the open- from a proximal source in keeping with the interpreted forearc
ing of the Andaman Sea, rivers draining the Myanmar region depositional environment. A subordinate contribution from an
presumably would have drained into the trench-forearc system, older continental source was most likely to have been the western
which acted as a “sink,” preventing deposition farther west onto Sibumasu margin, but a Transhimalayan arc unit source cannot
the Indian plate. be ruled out.
The Andaman Flysch records mixed orogenic and arc prov- The Oligocene Andaman Flysch was deposited between 30
enance. The young (<100 Ma) U-Pb–dated detrital zircons from and 20 Ma, and then uplifted by 20 Ma. It shows a clear change
the Andaman Flysch are consistent with derivation from igneous in petrographic and isotopic characteristics from the Mithakhari
sources, most likely an eastern (Myanmar) provenance based on Group and is composed mainly of “recycled orogenic” sources
similar aged grains found in modern Irrawaddy River sediment with a subordinate arc provenance. It represents the earliest
(Bodet and Schärer, 2000), as grains with young U-Pb zircon record in the region of major influx from a continental area. The
ages are extremely rare in the southern flanks of the Himalayas recycled orogenic component is most simply explained by ero-
(Campbell et al., 2005; DeCelles et al., 1998, 2004). sion from the northeastern (Myanmar) continental region. Iso-
Myanmar sources also can adequately account for the topic and petrographic differences between formations may be
majority of the “recycled orogenic” component of the detritus explained to some extent by different mixtures and contributions
in the Andaman Flysch. Both petrography and εNd whole rock from the same source regions. Although dual provenance is a
signatures, especially of the fine-grained material, are nearly favorable model for this region, detailed discrimination between
identical to those of the modern Irrawaddy River (Figs. 15 and Himalayan and Myanmar sources, at present, awaits more data
16). 40Ar-39Ar and FT data from the Irrawaddy River also are from Myanmar source rocks.
TABLE A1. FISSION-TRACK ANALYTICAL DATA FOR ANDAMAN ISLANDS–SOUTH ANDAMAN
Sample Latitude Mineral No. of ρd ρs ρι Central age Age components Mean S.D. No. of
Dispersion
Longitude grains Nd NS Ni ±1σ (Ma) track tracks
length
(Pχ2) RE% (μm)

MITHAKHARI GROUP
Hopetown Conglomerate Formation
Hopetown N11°41.407′ Zircon 25 0.528 12.16 4.32 0.0 56.2 91.8 ± 10.8 61 ± 2 (16) 129 ± 10 (5) 280 ± 32 (4)
Quarry E92°43.501′ (3450) (4942) (1756)
Apatite 19 1.106 0.064 0.209 0.03 69.1 56.5 ± 8.5 57 ± 9 (19)
(6131) (58) (191)
Namunagarh Grit Formation
Mungleton N11°34.223′ Zircon 18 0.535 10.972 4.872 0.0 42.3 75.15 ± 7.9 59 ± 2 (15) 226 ± 21(3)
Quarry E92°39.541′ (3450) (4182) (1857)
Apatite 21 1.106 0.150 0.246 0.0 64.5 68.9 55 ± 8 (19) 151 ± 68 (1) 314 ± 51 (1)
(6131) (322) (946)
Namunagarh N11°41.339′ Apatite 24 1.106 0.060 0.282 99.6 0.0 39.8 ± 4.2 40 ± 4 (24)
Quarry E92°41.106′ (6131) (108) (506)
Chiriyatapu N11°42.005′ Zircon 38 0.525 12.53 2.255 0.0 92.6 143.0 ± 22 40 ± 3 (9) 91 ± 7 (8) 329 ± 17 (22)
E93°32.301′ (3450) (7468) (1344)

ANDAMAN FLYSCH
AND-1A N11°38.430′ Apatite 38 1.106 0.264 1.653 0.02 28.8 30.85 ± 2.2 31 ± 2 (38) 13.25 ± 0.17 1.51 76
E92°45.292′ (6131) (497) (3116)
Zircon 43 0.516 8.341 4.07 0.0 44.5 62.15 ± 4.6 44 ± 2 (24) 67 ± 4 (12) 147 ± 10 (6)
(3450) (6439) (3147)
AND-1B N11°38.457′ Apatite 57 1.106 0.499 2.499 7.3 16.3 37.15 ± 1.8 37 ± 2 (57) 13.06 ± 0.20 1.75 75
E92°45.318′ (6131) (738) (3696)
Zircon 32 0.515 6.115 3.685 0.0 41.5 53.75 ± 4.4 41 ± 2 (26) 101 ± 14 (6)
(3450) (3085) (1859)
AND-3 N11°38.459′ Apatite 50 1.106 0.497 2.556 0.03 22.6 35.85 ± 1.9 36 ± 2 (50) 13.24 ± 0.22 1.61 54
E92°45.368′ (6131) (852) (4739)
Zircon 15 0.108 10.24 4.828 0.0 73.0 58.55 ± 11.3 36 ± 2 (12) 616 ± 102 (3)
(3450) (1930) (910)
AND-4 N11°39.461′ Apatite 47 11.106 0.427 2.652 0.04 20.2 30.75 ± 1.5 31 ± 2 (47) 13.42 ± 0.23 1.80 62
E92°45.417′ (6131) (923) (5733)
Zircon 42 0.505 10.63 4.506 0.0 70.1 67.95 ± 7.6 38 ± 1 (20) 58 ± 4 (10) 270 ± 16 (12)
(3450) (7575) (3210)
6 –2
Note: (i) Track densities are (x10 tr cm ); numbers of tracks counted (N) shown in brackets. (ii) Analyses by external detector method using 0.5 for the 4π/2π geometric correction factor. (iii) Ages calculated
using dosimeter glass CN-5; (apatite) ξCN5 = 338 ± 4; CN-2 (zircon) ξCN2 = 127 ± 4 calibrated by multiple analyses of IUGS apatite and zircon age standards (see Hurford, 1990). (iv) Pχ2 is probability for obtaining χ2
value for v degrees of freedom, where v = no. crystals –1. (v) Central age is a modal age, weighted for different precisions of individual crystals (see Galbraith and Green, 1990). (vi) Age modes deconvolved using
approach of Sambridge and Compston (1994) and Galbraith and Laslett (1993). RE—% relative error about the central age; S.D.—standard deviation; bold type—apatite; bold italic type—zircon ages.
TABLE A2. RAW APATITE-HELIUM DATA
4 4 4 238 232
Sample Location He HB He-HB %S.D. in He Absolute ± U ± Th ± He age* Error Grain FT FT
(ncc) (ncc) (ncc) Q± (atoms) (ng) (%) (ng) (%) (Ma) ± 7% radius corrected corrected
age
(Ma)
AND-A N11°38.430′ E92°45.292′ 0.356 0.023 0.333 0.452 8.95E + 09 4.04E + 07 0.108 1.2 0.537 1.33 11.704 0.819 77 0.81 14.4
AND-B N11°38.457′ E92°45.318′ 0.209 0.021 0.188 0.119 5.05E + 09 6.01E + 06 0.062 2.18 0.215 1.8 13.756 0.962 66 0.76 18.1
AND-C N11°38.459′ E92°45.368′ 0.222 0.018 0.204 0.119 5.48E + 09 6.52E + 06 0.058 1.98 0.315 1.4 12.664 0.886 82 0.81 15.6
Note: All samples are taken from the Andaman Flysch Formation at Corbyn’s Cove. Helium ages are based on replicate analyses of apatite grains. The total uncertainty in sample age
is based on the reproducibility of 39 analyses at the California Institute of Technology laboratory. Durango apatite standard (this error of the mean = 6.7%) combined with the U/Th and He
analytical uncertainties. HB—hot blank; ncc—nanno cc; S.D.—standard deviation; FT—fission track.
*Uncorrected He age.
TABLE A3. RAW ARGON DATA
Andaman Flysch (Corbyn’s Cove section)
40 39 38 37 36 40 39
A1B Ar Ar Ar Ar Ar Ar*/ Ar ± Age (Ma) ±
Ms 1 0.68068 0.213559 0.003536 0.005284 0.000779 2.109962 0.020067 51.8 0.5
Ms 2 0.203759 0.060257 0.000705 0.000626 6.48E-05 3.063559 0.006231 74.7 0.4
Ms 3 0.451179 0.185698 0.002571 0.002597 0.000354 1.865821 0.022963 45.9 0.6
Ms 4 0.329659 0.110714 0.001661 0.002407 0.000334 2.085147 0.038311 51.2 1.0
Ms 5 0.37799 0.046614 0.000787 0.000994 0.00013 7.286482 0.032957 172.9 1.1
Ms 6 0.269821 0.037161 0.000531 0.000573 5.98E-05 6.784923 0.089994 161.5 2.2
Ms 8 2.277688 0.11746 0.001906 0.002831 0.000274 18.7013 0.031643 414.2 2.0
Ms 9 2.143176 0.070945 0.00093 0.000402 2.99E-05 30.08465 0.047492 626.5 2.8
Ms 10 1.384602 0.074126 0.001043 0.001053 0.00018 17.96253 0.023198 399.5 1.9
Ms 11 0.253605 0.015673 0.000291 5.75E-05 0 16.18154 0.035317 363.6 1.8
Ms 12 0.259847 0.082319 0.00117 0.000824 7.98E-05 2.870187 0.003993 70.1 0.4
Ms 13 3.391288 0.385463 0.004783 0.003164 0.000389 8.499624 0.011701 200.1 1.0
Ms 15 0.790418 0.252174 0.003255 0.003477 0.000284 2.801527 0.02637 68.4 0.7
Ms 16 0.531821 0.070046 0.000772 0.000291 5.49E-05 7.360776 0.036057 174.5 1.2
Ms 17 0.323132 0.105295 0.001329 0.001905 0.000189 2.537016 0.031682 62.1 0.8
Ms 18 0.832588 0.098415 0.001318 0.000894 4.98E-05 8.310543 0.035372 195.9 1.2
Ms 19 0.176399 0.061786 0.00071 0.000428 7.99E-05 2.472932 0.067916 60.5 1.7
Ms 20 0.458665 0.086807 0.001068 0.00179 0.00013 4.842794 0.034779 116.7 1.0
Ms 21 0.269899 0.070387 0.000792 0.000117 0.0002 2.994993 0.047513 73.1 1.2
Ms 22 1.721992 0.356929 0.004472 0.009619 0.000637 4.296717 0.010036 103.9 0.6
Ms 23 1.216759 0.459022 0.005575 0.00187 0.00031 2.451514 0.006712 60.0 0.3
Ms 24 0.085958 0.027352 0.000404 0.000195 0 3.142686 0.011197 76.6 0.5
Ms 25 0.130443 0.038038 0.000618 0.002262 0.000284 1.219882 0.023433 30.1 0.6
Ms 26 2.371162 0.106391 0.001354 0.001404 0.000115 21.96896 0.014042 477.7 2.1
Ms 27 0.314057 0.064276 0.000828 0.000507 4.99E-05 4.656836 0.012754 112.4 0.6
Ms 28 0.336883 0.064142 0.000767 0.000156 8E-05 4.883813 0.04699 117.7 1.2
Ms 29 0.999109 0.106174 0.001329 0.00041 0.000125 9.062499 0.029797 212.6 1.2
Ms 30 0.124323 0.075656 0.000961 0.000488 7.49E-05 1.350842 0.039184 33.3 1.0
Ms 31 1.604755 0.163177 0.001993 0.000528 0.000145 9.572115 0.020485 223.8 1.1
Ms 32 0.176028 0.107915 0.001318 0.000411 8.49E-05 1.398714 0.006655 34.5 0.2
Ms 33 0.094113 0.056227 0.000787 0.000782 5.48E-05 1.385834 0.005724 34.2 0.2
40 39 38 37 36 40 39
A3 Ar Ar Ar Ar Ar Ar*/ Ar ± Age (Ma) ±
Ms 1 0.274897 0.083084 0.000981 0.00039 0.000145 2.793324 0.037215 68.2 1.0
Ms 2 0.444523 0.143237 0.00185 0.001424 0.000445 2.186142 0.029916 53.6 0.8
Ms 3 0.297818 0.112368 0.0014 0.000746 0.000115 2.348489 0.038715 57.5 1.0
Ms 4 1.30819 0.466538 0.005703 0.002087 0.000109 2.734713 0.013397 66.8 0.5
Ms 5 0.346651 0.128071 0.001671 0.000476 0.000155 2.349362 0.023267 57.6 0.6
Ms 6 3.632999 0.500069 0.006087 0.001156 0.000305 7.084947 0.012826 168.3 0.9
Ms 8 0.311219 0.100244 0.001252 0.000476 0.000185 2.55964 0.004776 62.6 0.3
Ms 9 0.281996 0.054435 0.00068 0.000511 0.000105 4.611138 0.010622 111.3 0.6
Ms 10 0.180333 0.044863 0.000593 0.000392 2.49E-05 3.855628 0.00766 93.5 0.5
Ms 11 0.518299 0.095269 0.001186 0.000528 0.000125 5.053064 0.006582 121.6 0.6
Ms 12 0.604087 0.186293 0.002397 0.000648 0.000155 2.997079 0.016331 73.1 0.5
Ms 13 0.514859 0.086648 0.001109 0.000392 2.49E-05 5.857044 0.034464 140.2 1.0
Ms 15 0.190225 0.061068 0.000797 0.000672 0.000145 2.414197 0.048452 59.1 1.2
Ms 16 0.138431 0.029423 0.000409 0.000534 0.000115 3.551304 0.003826 86.3 0.4
Ms 17 0.560451 0.085584 0.001022 6.89E-05 6.5E-05 6.324168 0.019494 151.0 0.9
Ms 18 1.927513 0.246058 0.003455 0.002792 0.000814 6.855692 0.06817 163.1 1.7
Ms 19 0.549605 0.188385 0.002402 0.00138 0.000265 2.502356 0.018307 61.2 0.5
Ms 20 0.199519 0.062808 0.000767 0.000897 0.000185 2.307352 0.05385 56.5 1.3
Ms 21 0.792813 0.159513 0.00209 0.002278 0.000214 4.573025 0.019797 110.4 0.7
Ms 22 0.531614 0.155877 0.00209 0.001657 0.000385 2.681443 0.009091 65.6 0.4
Ms 23 0.290937 0.092139 0.001272 0.000881 7.98E-05 2.901768 0.033228 70.8 0.9
Ms 24 0.481724 0.087635 0.000905 0.00038 9.99E-05 5.160096 0.036978 124.1 1.0
Ms 25 1.163152 0.352631 0.004461 0.001418 0.00038 2.980373 0.009667 72.7 0.4
Ms 26 0.272653 0.094083 0.001191 0.005552 8.85E-05 2.619921 0.032203 64.1 0.8
Ms 27 0.422343 0.128727 0.00161 0.000571 0.00011 3.028746 0.023467 73.9 0.7
Ms 28 0.24212 0.086183 0.001145 0.000242 9.94E-06 2.775297 0.03496 67.8 0.9
Ms 29 0.358268 0.11507 0.001441 0.000277 2.99E-05 3.036643 0.005754 74.1 0.4
Ms 30 0.178945 0.044807 0.000537 0.000211 3.49E-05 3.763259 0.033264 91.3 0.9
Ms 31 0.273151 0.093756 0.001232 0.000316 0.000115 2.551233 0.016583 62.4 0.5
Ms 32 0.173772 0.089489 0.00115 1.76E-05 5E-05 1.77673 0.016944 43.7 0.5
(continued)
TABLE A3. RAW ARGON DATA (continued)
Andaman Flysch (Corbyn’s Cove section)
40 39 38 37 36 40 39
A4 Ar Ar Ar Ar Ar Ar*/ Ar ± Age (Ma) ±
Ms 1 1.757257 0.228175 0.003051 0.001952 0.000479 7.080388 0.0316 168.2 1.1
Ms 2 0.350481 0.117885 0.001625 0.000598 0.000135 2.63508 0.030076 64.4 0.8
Ms 3 0.226229 0.086674 0.001165 0.000669 0.000115 2.218651 0.038875 54.4 1.0
Ms 4 0.84056 0.203536 0.00254 0.000722 0.000305 3.687254 0.018742 89.5 0.6
Ms 5 0.387886 0.07517 0.001007 0.000476 3.49E-05 5.023023 0.044232 120.9 1.2
Ms 6 0.827359 0.256024 0.00326 0.001164 0.00039 2.781793 0.00436 68.0 0.3
Ms 8 0.124755 0.077014 0.001012 0.000247 0.00011 1.198084 0.007833 29.6 0.2
Ms 9 3.009747 0.428003 0.005417 0.000953 0.000195 6.897607 0.010471 164.0 0.8
Ms 10 1.209192 0.187151 0.00232 0.000706 0.000105 6.29557 0.006128 150.3 0.7
Ms 11 0.497906 0.161033 0.00184 0.000212 7.49E-05 2.954419 0.018552 72.1 0.6
Ms 12 3.00958 0.046496 0.000608 0.000395 7.99E-05 64.22046 0.067476 1144.9 4.3
Ms 13 0.239454 0.097718 0.001129 0.000897 0.00013 2.058057 0.005784 50.5 0.3
Ms 15 1.727941 0.097393 0.001247 3.59E-05 7.5E-05 17.51444 0.014541 390.6 1.8
Ms 16 0.259758 0.122808 0.001441 0.000395 3.49E-05 2.0312 0.024229 49.9 0.6
Ms 17 0.102435 0.029899 0.000307 0 1.5E-05 3.277664 0.007241 79.8 0.4
Ms 18 0.565702 0.067938 0.000869 0.000198 3.49E-05 8.174691 0.007531 192.8 0.9
Ms 19 0.461503 0.030601 0.000363 0 5E-06 15.03286 0.049761 340.1 1.9
Ms 20 0.084844 0.03198 0.000358 3.6E-05 9.99E-06 2.560674 0.046289 62.7 1.2
Ms 21 0.051231 0.028354 0.000327 0.000144 9.96E-06 1.70302 0.116858 41.9 2.9
Ms 22 0.539888 0.175027 0.002034 0.000648 0.000225 2.705021 0.007126 66.1 0.4
Ms 23 0.224278 0.125809 0.001584 0.000793 0.000115 1.513072 0.003637 37.3 0.2
Ms 24 2.435008 0.11556 0.001303 0.000343 8.99E-05 20.84143 0.030878 456.1 2.1
Ms 25 0.28137 0.043934 0.000537 0.000162 3E-05 6.202931 0.034616 148.2 1.1
Ms 26 1.447467 0.066026 0.00092 0.001368 0.000185 21.09624 0.0327 461.0 2.1
Ms 27 0.131921 0.043514 0.000777 0.001825 0.000255 1.303279 0.01007 32.2 0.3
Ms 28 0.381247 0.061408 0.000899 0.001902 0.000264 4.935659 0.026829 118.9 0.8
Ms 29 0.400761 0.030962 0.000603 0.001484 0.000155 11.46811 0.056235 265.1 1.7
Ms 30 0.124287 0.055912 0.000761 0.001294 0.000205 1.141272 0.059711 28.2 1.5
Ms 31 0.077119 0.024908 0.000368 0.000686 8.98E-05 2.030571 0.013455 49.9 0.4
Ms 32 0.042558 0.024805 0.000409 0.000952 4.97E-05 1.123077 0.119683 27.7 2.9
Ms 33 0.057102 0.020714 0.000225 0.000305 5.99E-05 1.901885 0.160137 46.7 3.9
Mithakhari Group, Chiriyatapu
40 39 38 37 36 40 39
Ar Ar Ar Ar Ar Ar*/ Ar ± Age (Ma) ±
Ms 1 0.8284261 0.0365259 0.0004242 0.0004699 0.0000000 22.6805064 0.0750144 491.3 2.6
Ms 2 0.7446774 0.0323472 0.0003628 0.0000000 0.0000550 22.5188268 0.0957791 488.2 2.8
Ms 3 0.7987535 0.0472551 0.0006388 0.0002743 0.0000249 16.7471471 0.0197808 375.1 1.7
Ms 4 0.3026851 0.0938337 0.0011243 0.0001764 0.0000750 2.9897197 0.0315933 72.9 0.8
Ms 5 0.6288437 0.0690488 0.0008483 0.0003334 0.0000449 8.9150267 0.0157014 209.3 1.0
Ms 6 0.2056404 0.0084045 0.0001022 0.0000000 0.0000000 24.4679238 0.0604556 524.8 2.5
Ms 8 2.0267770 0.0181880 0.0002760 0.0003637 0.0000349 110.8679577 0.1272294 1674.6 5.6
Ms 9 1.3058958 0.0589862 0.0007666 0.0002830 0.0000949 21.6634575 0.0479489 471.9 2.3
Ms 10 1.0895856 0.0467023 0.0006133 0.0004044 0.0000849 22.7933251 0.0380826 493.4 2.3
Ms 11 1.7582128 0.0745294 0.0008279 0.0005462 0.0000899 23.2345815 0.0322423 501.7 2.3
Ms 12 0.7200124 0.0307043 0.0004191 0.0003844 0.0001499 22.0072636 0.1088272 478.5 3.0
Ms 13 0.7633280 0.0327602 0.0004344 0.0003442 0.0000499 22.8502553 0.0157553 494.5 2.2
Ms 15 1.4321915 0.0604223 0.0007308 0.0002229 0.0000499 23.4587880 0.0323211 506.0 2.3
Ms 16 3.1189393 0.0694722 0.0008943 0.0007095 0.0000998 44.4702431 0.0356104 863.3 3.5
Ms 17 0.5782068 0.0518422 0.0006746 0.0002434 0.0000349 10.9540815 0.0577358 254.0 1.7
Ms 18 1.2848422 0.0599626 0.0007052 0.0001623 0.0000750 21.0580001 0.0540885 460.2 2.3
Ms 19 0.5290525 0.0217523 0.0003015 0.0002639 0.0000199 24.0509065 0.1038536 517.1 3.0
Ms 20 5.1561196 0.0246141 0.0003628 0.0002639 0.0000799 208.5188362 0.2883512 2444.9 6.9
Ms 21 0.6433217 0.0339743 0.0003577 0.0001625 0.0000500 18.5010347 0.0472177 410.2 2.1
Ms 22 1.1030049 0.0956311 0.0011039 0.0005283 0.0001499 11.0708902 0.0175933 256.5 1.3
Ms 23 0.3852341 0.0186428 0.0002044 0.0000000 0.0000000 20.6639642 0.0133670 452.6 2.0
Ms 24 0.8201835 0.0345528 0.0004804 0.0002349 0.0000499 23.3100383 0.0867000 503.2 2.7
Ms 25 0.2651809 0.0818851 0.0009965 0.0008358 0.0001048 2.8603315 0.0027919 69.8 0.3
Ms 26 0.4454907 0.0287515 0.0004242 0.0007315 0.0001248 14.2118131 0.0282976 323.1 1.6
Ms 27 0.2823107 0.0124546 0.0001073 0.0000000 0.0000000 22.6671453 0.0457513 491.0 2.3
Ms 28 7.5439887 0.1794691 0.0021310 0.0008891 0.0001048 41.8625264 0.0295576 822.6 3.3
Ms 29 0.4338841 0.0171655 0.0002402 0.0000000 0.0000000 25.2765779 0.0434406 539.8 2.5
Ms 30 0.9228281 0.0225529 0.0003373 0.0004192 0.0000299 40.5267640 0.1467725 801.4 4.0
Ms 31 2.4734857 0.0295677 0.0004191 0.0005765 0.0000398 83.2566784 0.1423053 1380.1 5.1
Ms 32 1.5381052 0.0256109 0.0003475 0.0004719 0.0000000 60.0566236 0.1661772 1089.0 4.7
Ms 33 0.3426377 0.0155174 0.0001942 0.0002623 0.0000299 21.5108920 0.0746433 469.0 2.5
TABLE A4. ZIRCON U-Pb DATA (UNCORRECTED AGES) FOR ANDAMAN FLYSCH SAMPLE 1A
206 238 207 235 207 206 206 238 207 235 207 206
Grain no. Pb/ U ± Pb/ U ± Pb/ Pb ± Pb/ U ±1σ Pb/ U ±1σ Pb/ Pb ±1σ
ratio ratio ratio Age (Ma) Age (Ma) Age (Ma)
8 0.56820 0.01491 18.27837 0.00211 0.23368 0.01172 2900.4 61.3 3004.5 46.4 3077.4 77.9
21 0.52114 0.01876 17.80391 0.00319 0.24751 0.01869 2704.0 79.5 2979.2 72.0 3168.9 114.9
18 0.45548 0.02377 10.17671 0.00390 0.16214 0.01534 2419.6 105.3 2451.0 83.7 2478.1 151.3
28 0.30919 0.01417 5.34072 0.00228 0.12476 0.01304 1736.7 69.8 1875.4 88.2 2025.4 174.4
1 0.25134 0.01476 3.17496 0.00192 0.09164 0.01199 1445.4 76.0 1451.1 95.3 1459.9 230.3
37 2.68537 0.14191 0.23801 0.00394 0.08188 0.00394 1376.4 20.5 1324.4 39.1 1242.7 99.7
7 0.20471 0.00761 2.27006 0.00089 0.08055 0.00685 1200.5 40.7 1203.0 56.6 1210.5 158.8
13 0.19499 0.00490 2.12506 0.00101 0.07915 0.00497 1148.4 26.5 1157.0 42.0 1175.9 119.3
31 2.04475 0.18301 0.19077 0.00536 0.07779 0.00536 1125.6 29.0 1130.5 61.0 1141.5 172.8
16 0.19076 0.00770 2.00881 0.00087 0.07645 0.00752 1125.5 41.7 1118.5 63.8 1106.9 184.9
23 0.19007 0.00612 2.22350 0.00114 0.08469 0.00712 1121.8 33.2 1188.5 58.4 1308.4 154.9
2 0.18013 0.00766 2.41204 0.00081 0.09716 0.00881 1067.7 41.8 1246.2 60.4 1570.4 160.9
9 0.17942 0.00506 1.81366 0.00083 0.07343 0.00503 1063.8 27.7 1050.4 42.7 1025.9 132.6
5 0.17514 0.00495 1.78122 0.00078 0.07384 0.00488 1040.3 27.1 1038.6 40.6 1037.0 128.1
3 0.15851 0.00357 1.66696 0.00100 0.07632 0.00404 948.5 19.8 996.0 31.7 1103.5 102.3
35 1.54569 0.07373 0.15281 0.00239 0.07342 0.00239 916.7 13.4 948.8 29.4 1025.7 94.7
4 0.15069 0.00383 1.56797 0.00082 0.07553 0.00458 904.8 21.5 957.6 35.5 1082.6 117.0
29 0.15001 0.00490 1.70213 0.00107 0.08191 0.00734 901.0 27.5 1009.3 58.2 1243.4 166.2
12 0.13848 0.00336 1.34274 0.00107 0.07044 0.00431 836.1 19.1 864.4 34.2 941.1 120.5
39 1.43337 0.20932 0.12528 0.00491 0.08302 0.00491 760.9 28.2 903.0 87.3 1269.6 266.2
26 0.12184 0.00484 1.12034 0.00076 0.06648 0.00719 741.2 27.8 763.1 57.1 821.5 210.9
6 0.11741 0.00687 0.91179 0.00108 0.05641 0.00919 715.6 39.6 658.0 75.1 467.6 325.9
22 0.09140 0.00258 0.76164 0.00058 0.06035 0.00473 563.8 15.2 575.0 34.3 616.3 160.8
11 0.08908 0.00363 0.96494 0.00048 0.07869 0.00822 550.1 21.5 685.8 49.1 1164.5 194.1
32 0.70554 0.09115 0.08743 0.00262 0.05857 0.00262 540.3 15.6 542.1 54.3 551.3 264.1
25 0.08390 0.00418 0.73573 0.00076 0.06343 0.00948 519.4 24.8 559.9 62.5 722.7 288.8
34 0.54771 0.03128 0.06830 0.00115 0.05821 0.00115 425.9 6.9 443.5 20.5 537.1 122.9
10 0.03673 0.00221 0.25590 0.00024 0.05061 0.01065 232.5 13.7 231.4 41.9 223.1 425.1
40 0.16376 0.10117 0.02074 0.00132 0.05728 0.00132 132.3 8.4 154.0 88.3 501.8 982.4
19 0.01567 0.00179 0.10372 0.00025 0.04802 0.02354 100.2 11.3 100.2 45.6 99.0 874.6
27 0.01427 0.00191 0.10618 0.00022 0.05378 0.03759 91.3 12.1 102.5 67.0 361.9 1092.3
33 0.09256 0.01678 0.01397 0.00047 0.04809 0.00047 89.4 3.0 89.9 15.6 103.5 384.6
17 0.01183 0.00088 0.07776 0.00012 0.04771 0.01416 75.8 5.6 76.0 21.1 84.0 585.7
30 0.09413 0.02893 0.01041 0.00056 0.06564 0.00056 66.7 3.5 91.3 26.9 795.0 545.5
24 0.00927 0.00087 0.06214 0.00010 0.04851 0.01680 59.5 5.5 61.2 19.9 124.2 660.0
20 0.00910 0.00077 0.11849 0.00010 0.09449 0.02208 58.4 4.9 113.7 23.6 1518.0 386.4
38 0.05721 0.00743 0.00799 0.00020 0.05193 0.00020 51.3 1.3 56.5 7.1 282.3 275.3
15 0.00778 0.00052 0.05499 0.00005 0.05135 0.01135 49.9 3.3 54.4 11.2 256.6 441.6
36 0.13338 0.08419 0.00723 0.00081 0.13385 0.00081 46.5 5.2 127.1 75.4 2149.0 828.8
14 0.00705 0.00024 0.04889 0.00004 0.05037 0.00528 45.3 1.6 48.5 4.8 212.2 225.9
TABLE A5A. PETROGRAPHIC DATA
S it e Formation Region Sample Grain size Q KF P Lv Lc Lp Lch Lm M HM Total MI
(μm)
Corbyn’s Cove section N11°38′43.0″ E92°45′29.2″ Andaman Flysch S Andaman FT1B 115 50 13 5 4 0 1 0 25 1 1 100.0 175
Corbyn’s Cove section ~100 m upsection from 1A Andaman Flysch S Andaman FT3 140 48 17 11 3 0 2 0 16 2 1 100.0 160
Corbyn’s Cove section below hotel Andaman Flysch S Andaman FT4 130 47 10 9 5 0 3 1 24 1 0 100.0 150
Namunagarh Quarry Namunagarh Grit S Andaman NAM 3D 1500 2 1 41 52 0 0 0 1 0 2 100.0 2
Hopetown Quarry N11°41′40.7″ E92°43′50.1″ Namunagarh Grit S Andaman NAM 25A 850 21 2 27 40 0 2 2 5 0 1 100.0 15
Mungleton Quarry N11°34′22.3″ E92°39′54.1″ Namunagarh Grit S Andaman NAM 26 1800 19 1 30 45 0 1 1 3 0 0 100.0 11
Irrawaddy River ~ Nyaungdoun Modern river sand Myanmar MY05 23A — 47 11 11 6 0 6 1 15 2 1 100.0 112
Note: The MI (“Metamorphic Index”; Garzanti and Vezzoli, 2003) expresses the average rank of metamorphic rock fragments in the studied samples, and varies from 0 in detritus from
sedimentary and volcanic cover rocks to 500 in detritus from high-grade basement rocks. Q—quartz; KF—potassium feldspar; P—plagioclase; Lv—volcanic; Lc—carbonate lithic fragments
(including marble); Lp—terriginous lithic fragments (shale, siltstone); Lch—chert lithic fragments; Lm—metamorphic lithic fragments; M—micas; HM—heavy minerals.
TABLE A5B. HEAVY-MINERAL DATA

HM% vfs–fs
HM% transparent
% transparent
% opaque
% turbid
Total
Zircon
Dravite
Schorlite
Rutile
Sagenite
Sphene

S it e Formation Sample
Corbyn’s Cove section N11°38′43.0″ E92°45′29.2″ Andaman Flysch FT1B 0.9 0.5 54 9 37 100 3 4 4 5 0 3
Corbyn’s Cove section ~100 m upsection from 1A Andaman Flysch FT3 0.4 0.2 44 7 49 100 4 4 5 5 2 2
Corbyn’s Cove section below hotel Andaman Flysch FT4 0.6 0.2 35 8 57 100 4 3 6 7 1 0
Namunagarh Quarry Namunagarh Grit NAM 3D 4.0 1.5 37 38 25 100 0 0 0 0 0 1
Hopetown Quarry N11°41′40.7″ E92°43′50.1″ Namunagarh Grit NAM 25A 1.3 0.6 42 16 42 100 1 0 0 1 0 15
Mungleton Quarry N11°34′22.3″ E92°39′54.1″ Namunagarh Grit NAM 26 1.0 0.5 53 7 40 100 0 0 0 1 0 16
Irrawaddy River ~ Nyaungdoun River sand MY05 23A 5.0 4.0 79 3 17 100 0 0 0 0 0 3

Anatase
Brookite
Ti aggregates
Apatite
Xenotime
Monazite
Barite
Others
Blue-green
hornblende
Green
hornblende
Green-brown
hornblende
Brown
hornblende

Site Formation Sample


Corbyn’s Cove section N11°38′43.0″ E92°45′29.2″ Andaman Flysch FT1B 0 0 24 10 0 0 0 0 0 0 0 0
Corbyn’s Cove section ~100 m upsection from 1A Andaman Flysch FT3 0 0 30 17 0 1 0 0 0 0 0 0
Corbyn’s Cove section below hotel Andaman Flysch FT4 0 0 24 11 0 0 0 0 0 0 0 0
Namunagarh Quarry Namunagarh Grit NAM 3D 0 0 0 3 0 0 0 0 1 2 20 0
Hopetown Quarry N11°41′40.7″ E92°43′50.1″ Namunagarh Grit NAM 25A 0 0 3 1 0 0 0 0 0 0 0 0
Mungleton Quarry N11°34′22.3″ E92°39′54.1″ Namunagarh Grit NAM 26 0 0 1 3 0 0 0 0 0 0 0 0
Irrawaddy River ~ Nyaungdoun River sand MY05 23A 0 0 0 0 0 0 0 0 44 2 2 1
(continued)
TABLE A5B. HEAVY-MINERAL DATA (continued)

Oxyhornblende
Glaucophane
Sodic amphiboles
Tremolite
Actinolite
Antofillite
Other amphiboles
Hypersthene
Olivine
Spinel
Epidote
Clinozoisite

Site Formation Sample


Corbyn’s Cove section N11°38′43.0″ E92°45′29.2″ Andaman Flysch FT1B 0 0 0 0 0 0 0 0 0 2 18 0
Corbyn’s Cove section ~100 m upsection from 1A Andaman Flysch FT3 0 0 0 0 0 0 0 0 0 3 2 0
Corbyn’s Cove section below hotel Andaman Flysch FT4 0 0 0 0 0 0 0 0 0 2 0 0
Namunagarh Quarry Namunagarh Grit NAM 3D 0 0 0 0 0 0 0 0 0 0 11 0
Hopetown Quarry N11°41′40.7″ E92°43′50.1″ Namunagarh Grit NAM 25A 0 0 0 0 0 0 0 0 0 16 43 0
Mungleton Quarry N11°34′22.3″ E92°39′54.1″ Namunagarh Grit NAM 26 0 0 0 0 0 0 0 0 0 7 44 0
Irrawaddy River ~ Nyaungdoun River sand MY05 23A 0 0 0 0 0 2 0 0 0 0 33 0

Zoisite
Other epidotes
Prehnite
Pumpellyite
Chloritoid
Lawsonite
Carpholite
Garnet
Staurolite
Andalusite
Kyanite
Sillimanite
Total transparent

Site Formation Sample


Corbyn’s Cove section N11°38′43.0″ E92°45′29.2″ Andaman Flysch FT1B 0 0 0 0 12 0 0 14 0 0 0 0 100
Corbyn’s Cove section ~100 m upsection from 1A Andaman Flysch FT3 0 0 0 0 12 0 0 13 0 0 0 0 100
Corbyn’s Cove section below hotel Andaman Flysch FT4 0 0 0 0 25 0 0 15 1 0 0 0 100
Namunagarh Quarry Namunagarh Grit NAM 3D 0 0 0 0 0 0 0 0 0 0 0 0 100
Hopetown Quarry N11°41′40.7″ E92°43′50.1″ Namunagarh Grit NAM 25A 0 2 0 0 0 0 0 0 0 0 0 0 100
Mungleton Quarry N11°34′22.3″ E92°39′54.1″ Namunagarh Grit NAM 26 0 0 0 0 0 0 0 1 0 0 0 0 100
Irrawaddy River ~ Nyaungdoun River sand MY05 23A 0 1 0 0 0 0 0 6 0 0 0 0 100
Note: HM—heavy minerals; vfs—very fine sand; fs—fine sand.
TABLE A6A. WHOLE ROCK Sm-Nd DATA
147 144 143 144
Sample Location Formation / GPS Lithology Sm Nd Sm/Nd Sm/ Nd Nd/ Nd 1σ εNd
(ppm)
Andaman Islands
AN05-31 Namunagarh Quarry Namunagarh Grit Sandstone 4.83 25.3134 0.1909 0.1154 0.512271 16 –7.2
AN05 3E Namunagarh Quarry Namunagarh Grit Mudstone 1.99 7.18860 0.2769 0.1674 0.512799 19 3.1
AF1A Corbyn’s Cove Andaman Flysch Mudstone 11.17 50.6900 0.2204 0.1332 0.512068 10 –11.1
AF2 Corbyn’s Cove Andaman Flysch Sandstone 5.05 25.7129 0.1966 0.1188 0.512220 12 –8.2

Myanmar—Rivers draining Eocene rocks


MY05 8A Lemyu River N20°49.212′, E93°18.576′ MR sand 4.02 20.0489 0.2010 0.1215 0.512300 18 –7.4
MY05 15A Kyeintuli River N17°57.139′, E94°33.087′ MR sand 3.33 16.4129 0.2035 0.1250 0.512434 14 –4.0
MY05 17B Thanlwe River N18°59.097′, E94°15.244′ MR mud 3.48 17.6156 0.1977 0.1195 0.512427 18 –4.1
MY05 22B Thandwe River N18°27.466′, E94°23.563′ MR mud 3.60 18.0331 0.1997 0.1207 0.512424 14 –4.2
MY05 23B Irrawaddy River N18°48.391′, E95°12.218′ MR mud 3.67 17.9317 0.2052 0.124 0.512210 18 –8.3
Note: MR—modern river sample; GPS—Global Positioning System.
TABLE A6B. SINGLE-GRAIN APATITE Sm-Nd DATA
147 144 143 144
Sample Formation / Age Lithology Sm/ Nd 2 S.D. of NIST610 2σ Nd/ Nd 2σ ppm εNd
(internal) (0)
1 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.123272 0.001140 0.000069 0.512855 134 4.24
2 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.070462 0.000408 0.000065 0.512803 127 3.21
3 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.090724 0.000581 0.000058 0.512237 114 –7.83
4 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.073481 0.000171 0.000054 0.512588 105 –0.98
5 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.082561 0.000180 0.000031 0.512485 61 –2.98
6 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.100785 0.002208 0.000062 0.512404 121 –4.57
7 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.106736 0.000552 0.000046 0.512211 90 –8.33
8 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.093898 0.000488 0.000033 0.511800 65 –16.35
9 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.108342 0.001112 0.000047 0.512545 93 –1.81
10 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.099159 0.000388 0.000033 0.512669 64 0.61
11 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.113837 0.000346 0.000051 0.512794 99 3.03
12 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.120108 0.000750 0.000044 0.512621 86 –0.33
13 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.078972 0.000114 0.000067 0.512291 131 –6.77
14 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.067596 0.000236 0.000075 0.512425 147 –4.16
15 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.114968 0.000317 0.000043 0.511822 83 –15.92
16 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.105069 0.001427 0.000067 0.512792 131 3.01
17 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.117803 0.000238 0.000055 0.512607 108 –0.60
18 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.085210 0.000165 0.000056 0.512321 109 –6.19
19 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.102799 0.000424 0.000051 0.512782 99 2.80
20 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.082935 0.000202 0.000136 0.512486 266 –2.96
21 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.136207 0.000550 0.000118 0.512192 230 –8.71
22 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.131323 0.000741 0.000062 0.512791 122 2.99
23 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.244791 0.003438 0.000076 0.512607 148 –0.61
24 Andaman Flysch AF1A Corbyn’s Cove Apatite grain 0.172124 0.000850 0.000079 0.512771 154 2.59
Note: Global Positioning System (GPS) reference for AF1A Corbyn’s Cove: N11°38.430' E92°45.292'; S.D.—standard deviation.
254 Allen et al.

ACKNOWLEDGMENTS
Mithakhari Group, Middle Andaman, India: Journal of the Geological
Society of India, v. 53, p. 271–284.
This paper has benefited from thoughtful reviews by Peter Clift, Charusiri, P., Clark, A.H., Farrar, E., Archibald, D., and Charusiri, B., 1993,
Joseph Curray, and Christophe Colin. The work was funded by Granite belts in Thailand: Evidence from the 40Ar/39Ar geochronological
and geological syntheses: Journal of Southeast Asian Earth Sciences, v. 8,
NERC grants NE/B503192/1 and NER/S/A/2004/12158, with p. 127–136, doi: 10.1016/0743-9547(93)90014-G.
additional support from the Royal Society short-visits scheme. Cherniak, D.J., and Watson, E.B., 2001, Pb diffusion in zircon: Chemical Geol-
ogy, v. 172, p. 5–24, doi: 10.1016/S0009-2541(00)00233-3.
Clark, M.K., Schoenbohm, L.M., Royden, L., Whipple, K.X., Burchfiel, B.C.,
REFERENCES CITED Zhang, X., Tang, W., Wang, E., and Chen, L., 2004, Surface uplift, tec-
tonics, and erosion of eastern Tibet from large-scale drainage patterns:
Acharyya, S.K., Ray, K.K., and Roy, D.K., 1989, Tectonic stratigraphy and Tectonics, v. 23, p. TC1006, doi: 10.1029/2002TC001402.
emplacement history of the ophiolite assemblage from Naga Hills and Clark, M.K., House, M.A., Royden, L.H., Whipple, K.X., Burchfiel, B.C.,
Andaman island arc, India: Journal of the Geological Society of India, Zhang, X., and Tang, W., 2005, Late Cenozoic uplift of southeastern
v. 33, p. 4–18. Tibet: Geology, v. 33, p. 525–528, doi: 10.1130/G21265.1.
Ahmad, T., Harris, N., Bickle, M., Chapman, H., Bunbury, J., and Prince, C., Colin, C., Turpin, L., Bertaux, J., Desprairies, A., and Kissel, C., 1999, Ero-
2000, Isotopic constraints on the structural relationships between the sional history of the Himalayan and Burman ranges during the last two
Lesser Himalayan Series and the High Himalayan Crystalline Series, Gar- glacial-interglacial cycles: Earth and Planetary Science Letters, v. 171, p.
hwal Himalaya: Geological Society of America Bulletin, v. 112, p. 467– 647–660, doi: 10.1016/S0012-821X(99)00184-3.
477, doi: 10.1130/0016-7606(2000)112<0467:ICOTSR>2.3.CO;2. Colin, C., Turpin, L., Bertaux, J., Frank, N., Kissel, C., and Duchamp, S., 2006,
Auge, T., Cocherie, A., Genna, A., Armstrong, R., Guerrot, C., Mukherjee, Evolution of weathering patterns in the Indo-Burman Ranges over the
N.M., and Patra, R.N., 2003, Age of the Baula PGE mineralization (Oris- last 280 kyr: Effects of sediment provenance on 87Sr/86Sr ratios tracer:
sam India) and its implications concerning the Singhbhum Archaean Geochemistry, Geophysics, Geosystems, v. 7, p. 1–16, doi:10.1029/
nucleus: Precambrian Research, v. 121, p. 85–101, doi: 10.1016/S0301- 2005GC000962.
9268(02)00202-4. Curray, J.R., 2005, Tectonics and history of the Andaman Sea region:
Bandopadhyay, P.C., 2005, Discovery of abundant pyroclasts in Namunagarh Journal of Asian Earth Sciences, v. 25, p. 187–232, doi: 10.1016/
Grit, South Andaman; evidence for arc volcanism and active subduction j.jseaes.2004.09.001.
during the Palaeogene in the Andaman area: Journal of Asian Earth Sci- Curray, J.R., Moore, D.G., Lawver, L.A., Emmel, F.J., Raitt, R.W., Henry, M.,
ences, v. 25, p. 95–107, doi: 10.1016/j.jseaes.2004.01.007. and Kieckhefer, R., 1979, Tectonics of the Andaman Sea and Burma, in
Bandopadhyay, P.C., and Ghosh, M., 1999, Facies, petrology and depositional Watkins, J., et al., eds., Geological and Geophysical Investigations of
environment of the Tertiary sedimentary rocks around Port Blair, South Continental Margins: American Association of Petroleum Geologists
Andaman: Journal of the Geological Society of India, v. 52, p. 53–66. Memoir 29, p. 189–198.
Barley, M.E., Pickard, A.L., Zaw, K., Rak, P., and Doyle, M.G., 2003, Jurassic DeCelles, P.G., Gehrels, G.E., Quade, J., and Ojha, T.P., 1998, Eocene to
to Miocene magmatism and metamorphism in the Mogok metamorphic early Miocene foreland basin development and the history of Himalayan
belt and the India-Eurasia collision in Myanmar: Tectonics, v. 22, p. 1019, thrusting, western and central Nepal: Tectonics, v. 17, p. 741–765, doi:
doi: 10.1029/2002TC001398. 10.1029/98TC02598.
Bertrand, G., Rangin, C., Maluski, H., Han, T.A., Thein, M., Myint, O., Maw, DeCelles, P.G., Gehrels, G.E., Najman, Y., Martin, A.J., and Garzanti, E., 2004,
W., and Lwin, S., 1999, Cenozoic metamorphism along the Shan scarp Detrital geochronology and geochemistry of Cretaceous–Early Miocene
(Myanmar): Evidences for ductile shear along the Sagaing fault or the strata of Nepal: Implications for timing and diachroneity of initial Hima-
northward migration of the eastern Himalayan syntaxis?: Geophysical layan orogenesis: Earth and Planetary Science Letters, v. 227, p. 313–330,
Research Letters, v. 26, p. 915–918, doi: 10.1029/1999GL900136. doi: 10.1016/j.epsl.2004.08.019.
Bertrand, G., Rangin, C., Maluski, H., and Bellon, H., 2001, Diachronous cool- Dickinson, W.R., 1985, Interpreting provenance relations from detrital modes
ing along the Mogok Metamorphic Belt (Shan scarp, Myanmar): The trace of sandstones: Provenance of arenites, in Zuffa, G.G., ed., NATO ASI:
of the northward migration of the Indian syntaxis: Journal of Asian Earth Dordrecht, Netherlands, Reidel, v. 148, p. 333–361.
Sciences, v. 19, p. 649–659, doi: 10.1016/S1367-9120(00)00061-4. Draut, A.E., and Clift, P.D., 2006, Sedimentary processes in modern and ancient
Bodet, F., and Schärer, U., 2000, Evolution of the SE-Asian continent from U- oceanic arc settings: Evidence from the Jurassic Talkeetna Formation of
Pb and Hf isotopes in single grains of zircon and baddeleyite from large Alaska and the Mariana and Tonga Arcs, western Pacific: Journal of Sedi-
rivers: Geochimica et Cosmochimica Acta, v. 64, p. 2067–2209, doi: mentary Research, v. 76, p. 493–514, doi: 10.2110/jsr.2006.044.
10.1016/S0016-7037(00)00352-5. Foster, G.L., and Vance, D., 2006, In situ Nd isotopic analysis of geological
Brewer, I.D., Burbank, D.W., and Hodges, K.V., 2003, Modelling detrital cool- materials by laser ablation MC-ICP-MS: Journal of Analytical Atomic
ing-age populations: Insights from two Himalayan catchments: Basin Spectrometry, v. 21, p. 288–296, doi: 10.1039/b513945g.
Research, v. 15, p. 305–320, doi: 10.1046/j.1365-2117.2003.00211.x. Galbraith, R.F., and Green, P.F., 1990, Estimating the component ages in a finite
Campbell, I.H., Reiners, P.W., Allen, C.M., Nicolescu, S., and Upadhyay, R., mixture: Nuclear Tracks and Radiation Measurements, v. 17, p. 197–206,
2005, He-Pb double dating of detrital zircons from the Ganges and Indus doi: 10.1016/1359-0189(90)90035-V.
Rivers: Implication for quantifying sediment recycling and provenance Galbraith, R.F., and Laslett, G.M., 1993, Statistical models for mixed fission
studies: Earth and Planetary Science Letters, v. 237, p. 402–432, doi: track ages: Nuclear Tracks and Radiation Measurements, v. 21, p. 459–
10.1016/j.epsl.2005.06.043. 470, doi: 10.1016/1359-0189(93)90185-C.
Carter, A., 1999, Present status and future avenues of source region discrimina- Gallagher, K., 1995, Evolving temperature histories from apatite FT data: Earth
tion and characterization using fission-track analysis: Sedimentary Geol- and Planetary Science Letters, v. 136, p. 421–435, doi: 10.1016/0012-
ogy, v. 124, p. 31–45, doi: 10.1016/S0037-0738(98)00119-5. 821X(95)00197-K.
Carter, A., and Bristow, C.S., 2000, Detrital zircon geochronology: Enhancing Galy, A., and France-Lanord, C., 2001, Higher erosion rates in the Himalaya:
the quality of sedimentary source information through improved method- Geochemical constraints on riverine fluxes: Geology, v. 29, p. 23–26, doi:
ology and combined U-Pb and fission track techniques: Basin Research, 10.1130/0091-7613(2001)029<0023:HERITH>2.0.CO;2.
v. 12, p. 47–57, doi: 10.1046/j.1365-2117.2000.00112.x. Garzanti, E., and Vezzoli, G., 2003, A classification of metamorphic grade
Chakraborty, P.P., and Pal, T., 2001, Anatomy of a forearc submarine fan: in sands based on their composition and grade: Journal of Sedimentary
Upper Eocene–Oligocene Andaman Flysch Group, Andaman Islands, Research, v. 73, p. 830–837.
India: Gondwana Research, v. 4, p. 477–487, doi: 10.1016/S1342- Green, P.F., Duddy, I.R., and Bray, R.J., 1995, Applications of thermal history
937X(05)70347-6. reconstruction in inverted basins, in Buchanan, J.G., and Buchanan, P.G.,
Chakraborty, P.P., Pal, T., Dutta Gupta, T., and Gupta, K.S., 1999, Facies pat- eds., Basin Inversion: Geological Society [London] Special Publication
tern and depositional motif in an immature trench-slope basin, Eocene 88, p. 149–165.
Sedimentation and uplift history of the Andaman-Nicobar accretionary prism 255

Guha, D.K., and Mohan, M., 1965, On the ostracod from the Neogene of Anda- Najman, Y., Carter, A., Oliver, G., and Garzanti, E., 2005, Provenance of early
man Island: Journal of the Geological Survey of India, v. 9, p. 58–66. foreland basin sediments, Nepal: Constraints to the timing and diachro-
Guillot, S., Garzanti, E., Baratoux, D., Marquer, D., Mahéo, G., and de neity of early Himalayan orogenesis: Geology, v. 33, p. 309–312, doi:
Sigoyer, J., 2003, Reconstructing the total shortening history of the NW 10.1130/G21161.1.
Himalaya: Geochemistry, Geophysics, Geosystems, v. 4, p. 1064, doi: Oldham, R.D., 1885, Notes on the geology of Andaman Islands: Record of the
10.1029/2002GC000484. Geological Survey of India. v. 18, p. 135–145.
Haines, P.W., Turner, S.P., Kelley, S.P., Wartho, J.-A., and Sherlock, S.C., Pal, T., Chakraborty, P.P., Gupta, T.D., and Singh, C.D., 2003, Geodynamic
2004, Ar/Ar dating of detrital muscovite in provenance investiga- evolution of the outer-arc–forearc belt in the Andaman Islands, the cen-
tions: A case study from the Adelaide Rift Complex, South Australia: tral part of the Burma-Java subduction complex: Geological Magazine,
Earth and Planetary Science Letters, v. 227, p. 297–311, doi: 10.1016/ v. 140, p. 289–307, doi: 10.1017/S0016756803007805.
j.epsl.2004.08.020. Pal, T., Gupta, T.D., Chakraborty, P.P., and Gupta, S.C.D., 2005, Pyroclastic
Halder, D., 1985, Some aspects of the Andaman Ophiolite Complex: Geologi- deposits of Mio-Pliocene age in the Arakan Yoma-Andaman-Java subduc-
cal Survey of India, v. 115, p. 1–11. tion complex, Andaman Islands, Bay of Bengal, India: Geochemical Jour-
Hurford, A.J., 1990, Standardization of fission track dating calibration: Recom- nal, v. 39, p. 69–82, doi: 10.2343/geochemj.39.69.
mendation by the Fission Track Working Group of the IUGS Subcommis- Peucat, J.J., Vidal, P., Bernard-Griffiths, J., and Condie, K.C., 1989, Sr, Nd and
sion on Geochronology: Chemical Geology, v. 80, p. 177–178. Pb isotopic systematics in the Archaean low-to-high-grade transition zone
Karunakaran, C., Ray, K.K., and Saha, S.S., 1968, Tertiary sedimentation in of southern India: Syn-accretion granulites: Journal of Geology, v. 97,
Andaman-Nicobar geosyncline: Journal of the Geological Society of p. 537–550.
India, v. 9, p. 32–39. Pivnik, D.A., Nahm, J., Tucker, R.S., Smith, G.O., Nyein, K., Nyunt, M., and
Ketcham, R.A., Donelick, R.A., and Carlson, W.D., 1999, Variability of apa- Maung, P.H., 1998, Polyphase deformation in a fore-arc/back-arc basin,
tite fission-track annealing kinetics: III. Extrapolation to geological time- Salin sub-basin, Myanmar (Burma): American Association of Petroleum
scales: American Mineralogist, v. 84, p. 1235–1255. Geologists Bulletin, v. 82, p. 1837–1856.
Khan, P.K., and Chakraborty, P.P., 2005, Two-phase opening of Andaman Sea: Ray, K.K., Sengupta, S., and Van Den Hul, H.J., 1988, Chemical characters of
A new seismotectonic insight: Earth and Planetary Science Letters, v. volcanic rocks from Andaman ophiolite, India: Geological Society [Lon-
229, p. 259–271. don] Journal, v. 145, p. 392–400.
Lee, T.T., and Lawver, L.A., 1995, Cenozoic plate reconstruction of South- Renne, P.R., Swisher, C.C., Deino, A.L., Karner, D.B., Owens, T.L., and
east Asia: Tectonophysics, v. 251, p. 85–138, doi: 10.1016/0040- DePaolo, D.J., 1998, Intercalibration of standards, absolute ages and
1951(95)00023-2. uncertainties in 40Ar/39Ar dating: Chemical Geology, v. 145, p. 117–152,
Mallik, T.K., 1976, Shelf sediments of the Ganges delta with special empha- doi: 10.1016/S0009-2541(97)00159-9.
sis on the mineralogy of the western part, Bay of Bengal, Indian Ocean: Robinson, D.M., DeCelles, P.G., Patchett, P.J., and Garzione, C.N., 2001,
Marine Geology, v. 22, p. 1–32, doi: 10.1016/0025-3227(76)90007-4. The kinematic evolution of the Nepalese Himalaya interpreted from Nd
Marsaglia, K.M., and Ingersoll, R.V., 1992, Compositional trends in arc-related, isotopes: Earth and Planetary Science Letters, v. 192, p. 507–521, doi:
deep-marine sand and sandstone; a reassessment of magmatic-arc prov- 10.1016/S0012-821X(01)00451-4.
enance: Geological Society of America Bulletin, v. 104, p. 1637–1649, Rowley, D.B., 1996, Age of initiation of collision between India and Asia: A
doi: 10.1130/0016-7606(1992)104<1637:CTIARD>2.3.CO;2. review of stratigraphic data: Earth and Planetary Science Letters, v. 145,
Meesters, A.G.C.A., and Dunai, T.J., 2002, Solving the production diffusion p. 1–13, doi: 10.1016/S0012-821X(96)00201-4.
equation for finite diffusion domains of various shapes; Part II, Appli- Roy, S.K., 1992, Accretionary prism in Andaman forearc: Geological Survey of
cation to cases with alpha-ejection and nonhomogeneous distribution of India Special Publication 29, p. 273–278.
the source: Chemical Geology, v. 186, p. 57–73, doi: 10.1016/S0009- Roy, T.K., 1983, Geology and hydrocarbon prospects of Andaman–Nicobar
2541(01)00423-5. basin, in Bhandari, L.L., ed., Petroliferous Basins of India: Petroleum
Metcalfe, I., 1996, Pre-Cretaceous evolution of SE Asian terranes, in Hall, R., Asia Journal, p. 37–50.
and Blundell, D., eds., Tectonic Evolution of Southeast Asia: Geological Saha, A., Basu, A.R., Garzione, C.N., Bandyopadhyay, P.K., and Chakrabarti,
Society [London] Special Publication 106, p. 97–122. A., 2004, Geochemical and petrological evidence for subduction-accretion
Mishra, M., and Rajamani, V., 1999, Significance of the Archaean bimodal vol- processes in the Archean Eastern Indian Craton: Earth and Planetary Sci-
canics from the Ramagiri schist belt in the formation of the Eastern Dhar- ence Letters, v. 220, p. 91–106, doi: 10.1016/S0012-821X(04)00056-1.
war craton: Journal of the Geological Society of India, v. 54, p. 563–583. Sambridge, M.S., and Compston, W., 1994, Mixture modelling of multi-compo-
Mitchell, A.H.G., 1993, Cretaceous–Cenozoic tectonic events in the western nent data sets with application to ion-probe zircons: Earth and Planetary
Myanmar (Burma)–Assam region: Geological Society [London] Journal, Science Letters, v. 128, p. 373–390, doi: 10.1016/0012-821X(94)90157-0.
v. 150, p. 1089–1102. Sherlock, S.C., and Kelley, S.P., 2002, Excess argon in HP-LT rocks: A UV
Mountain, G.S., and Prell, W.L., 1990, A multiphase plate tectonic history of laserprobe study of phengite and K-free minerals: Chemical Geology,
the southeast continental margin of Oman, in Robertson, A.H.F., et al., v. 182, p. 619–636, doi: 10.1016/S0009-2541(01)00345-X.
eds., The Geology and Tectonics of the Oman Region: Geological Society Singh, O.P., Subramanya, S.M., and Sharma, V., 2000, Early Neogene multiple
[London] Special Publication 49, p. 725–743. microfossil biostratigraphy, John Lawrence Island: Andaman Sea: Micro-
Najman, Y., and Garzanti, E., 2000, Reconstructing early Himalayan tec- palaeontology, v. 46, p. 343–352.
tonic evolution and paleogeography from Tertiary foreland basin sedi- Singh, S.K., and France-Lanord, C., 2002, Tracing the distribution of erosion
mentary rocks, northern India: Geological Society of America Bul- in the Brahmaputra watershed from isotopic compositions of stream
letin, v. 112, p. 435–449, doi: 10.1130/0016-7606(2000)112<0435: sediments: Earth and Planetary Science Letters, v. 202, p. 645–662, doi:
REHTEA>2.3.CO;2. 10.1016/S0012-821X(02)00822-1.
Najman, Y., Bickle, M., and Chapman, H., 2000, Early Himalayan exhumation: Svojtka, M., Ko, J., and Venera, Z., 2001, Dating granulite-facies structures and
Isotopic constraints from the Indian foreland basin: Terra Nova, v. 12, p. the exhumation of lower crust in the Moldanubian Zone of the Bohemian
28–34, doi: 10.1046/j.1365-3121.2000.00268.x. Massif: International Journal of Earth Sciences, v. 91, p. 373–385.
Najman, Y., Johnson, C., White, N.M., and Oliver, G., 2004, Evolution of the
Himalayan foreland basin, NW India: Basin Research, v. 16, p. 1–24, doi:
10.1111/j.1365-2117.2004.00223.x. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Post-collisional collapse in the wake of migrating arc-continent collision in the


Ilan Basin, Taiwan

Peter D. Clift
School of Geosciences, University of Aberdeen, Aberdeen, AB24 3UE, UK, and DFG-Research Centre Ocean Margins (RCOM),
Geowissenschaften, Universität Bremen, Klagenfurter Strasse, 28359 Bremen, Germany

Andrew T.S. Lin


Department of Earth Sciences, National Central University, 300, Jungda Road, Jungli, Taiwan

Andrew Carter
School of Earth Sciences, University and Birkbeck College London, Gower Street, London, WC1E 6BT, UK

Francis Wu
State University of New York at Binghamton, Department of Geological Sciences and Environmental Studies, P.O. Box 600,
Binghamton, New York 13902-6000, USA

Amy E. Draut
U.S. Geological Survey, Pacific Science Center, 400 Natural Bridges Drive, Santa Cruz, California 95060, USA

T.-H. Lai
L.-Y. Fei
Central Geological Survey, 2, Lane 109, Hua-Hsin Street, Chung-Ho, Taipei, Taiwan 235

Hans Schouten
Department of Geology and Geophysics, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts 02543, USA

Louis Teng
Department of Geosciences, National Taiwan University, Roosevelt Road, Taipei, Taiwan

ABSTRACT

The Ilan Basin of northern Taiwan forms the western limit of the Okinawa Trough,
where the trough meets the compressional ranges of central Taiwan. Apatite fission-
track ages of 1.2 ± 0.5 Ma and 3.5 ± 0.5 Ma, measured north and south of the basin,
respectively, indicate faster exhumation rates in the Hsüehshan Range to the north
(>1.6 mm/yr) than in the Backbone Range to the south (0.7 mm/yr). Reconstructed
subsidence rates along the northern basin margin are also faster than in the south
(6–7 compared with 3–5 mm/yr). Global positioning system (GPS) and active seismo-
logical data indicate motion of the southern basin margin to the east and southeast.

Clift, P.D., Lin, A.T.S., Carter, A., Wu, F., Draut, A.E., Lai, T.-H., Fei, L.-Y., Schouten, H., and Teng, L., 2008, Post-collisional collapse in the wake of migrating
arc-continent collision in the Ilan Basin, Taiwan, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc
Collision Zones: Geological Society of America Special Paper 436, p. 257–278, doi: 10.1130/2008.2436(12). For permission to copy, contact editing@geosociety.
org. ©2008 The Geological Society of America. All rights reserved.

257
258 Clift et al.

We propose that the Ilan Basin is being formed as a result of extension of northern
Taiwan, largely controlled by a major southeast-dipping fault, modeled at ~30° dip,
and mapped as a continuation of the Lishan Fault, a major thrust structure in the
Central Ranges. Flexural rigidity of the lithosphere under the basin is low, with elastic
thickness ~3 km. A southwest-migrating collision between the Luzon Arc and south-
ern China, accompanied by subduction polarity reversal in the Ryukyu Trench, has
allowed crustal blocks that were previously held in compression between the Eurasian
and Philippine Sea plates to move trenchward as they reach the northern end of the
collision zone. Subduction polarity reversal permits rapid extension and formation of
the Ilan Basin and presumably, at least, the western Okinawa Trough, as a direct con-
sequence of arc-continent collision, not because of independent trench rollback forces.
This conceptual model suggests that migrating arc-continent collision causes the rapid
formation of deep marginal basins that are then filled by detritus from the adjacent
orogen, and that these should be common features in the geologic record.

Keywords: collision, extension, erosion, subduction, seismology.

INTRODUCTION 1981; Chemenda et al., 1997), which is itself built on the remains
of a Cretaceous continental arc (Davis et al., 1997). The collision
The Earth’s continental crust is generally considered to be between the Luzon oceanic arc and mainland Asia is oblique, so
largely generated by magmatism along active plate margins, a that the collision point must have migrated along the Asian con-
process that balances the long-term loss of continental crust back tinental margin through time. At any given time the north end
into the mantle through subduction zones (e.g., von Huene and of the island will represent arc units and Chinese passive-margin
Scholl, 1991; Rudnick and Fountain, 1995; Clift and Vannucchi, sequences that began to be involved in collision earlier than those
2004). This balance is maintained because arc crust generated in in the south, where collision is just starting. Thus northern Tai-
oceanic subduction settings is not entirely subducted when these wan is the optimal place to examine the processes that occur dur-
features collide with other trench systems or passive continental ing and after the peak of arc-continent collision. These processes
margins. Instead, continental masses appear to be built up as a form the subject of this paper.
result of the progressive accretion of primitive arc blocks to older, Radiometric dating of the metamorphic rocks exposed in
preexisting continental blocks. Understanding the tectonism of the Central Ranges, combined with biostratigraphic dating of
arc-continent collision is important, not only because this pro- orogenic sediment from the Coastal Ranges, indicates that col-
cess is fundamental to the conservation of the total volume of lision of this part of the Luzon Arc with mainland Asia started
continental crust, but also because it is likely the most important ca. 6–9 Ma (e.g., Suppe, 1981; Teng, 1990; Sibuet et al., 2002;
method by which active continental margins are created (e.g., Malavieille et al., 2002; Huang et al., 2006). A wide consensus in
Casey and Dewey, 1984; Konstantinovskaya, 1999). the community agrees that Taiwan represents a simple collision
Accreted oceanic island-arc terranes have been recognized between the Luzon Arc and mainland Asia, an assumption that
in a number of orogenic suture zones (e.g., Kohistan in the Hima- underlies this present study. Nonetheless, we note that alternative
laya, Talkeetna in southern Alaska, South Mayo in the Irish Cale- models have been advanced to explain the mountain building.
donides), and the process of arc-continent collision forms an inev- One theory involves collision of an exotic terrane with the Chi-
itable part of the plate tectonic cycle. Arc-continent and continent- nese passive margin, followed by later collision of the Luzon Arc
continent collisions are unlikely to be identical because of the (Lu and Hsü, 1992). Another model invokes compressional tecto-
different mechanical properties of arc and cratonic lithospheres. nism following collision of the Luzon Arc with the Ryukyu Arc,
In this paper we investigate the later stages of arc-continent colli- which would previously have extended farther southwest along
sion, specifically the collapse of the collisional mountain belts to the southern margin of China (Hsu and Sibuet, 1995; Sibuet and
form sedimentary basins whose fill might be used to understand Hsu, 1997).
the collision process. It is unclear when the collision between Luzon and Eurasia
In order to understand the collapse of arc collisional moun- began. Whereas some models favor collision to have initiated
tains, we examine the Ilan Basin of northern Taiwan. Taiwan, only ca. 6–9 Ma, effectively just to the east of the present col-
located off the coast of southeastern China, comprises some of lision zone (e.g., Suppe, 1984; Sibuet and Hsu, 1997; Huang et
the tallest and most rapidly uplifting and eroding mountains in al., 2000, 2006), other models suggest a more steady-state colli-
East Asia. The orogen is formed by the collision between the oce- sion that may have started earlier (e.g., Suppe, 1984; Teng, 1990,
anic Luzon Arc and the passive margin of southern China (Suppe, 1996; Clift et al., 2003). In the steady-state model the modern
Migrating arc-continent collision in the Ilan Basin, Taiwan 259

collision represents a snapshot of a continuous collision process of the free edge provided by the new trench. Specifically, there is
that is migrating to the southwest along the Chinese margin, no need for slab break-off to allow polarity reversal or as a trigger
accreting at least part of the crust of the Luzon Arc to the edge of of orogenic extension.
Eurasia, while generating a new Ryukyu Arc-Trench system and
associated Okinawa Trough in its wake. Debate continues as to REGIONAL GEOLOGY
whether the Okinawa Trough is an active rift system formed by
trench tectonic forces linked to the Ryukyu subduction zone and The Ilan Basin forms the onshore western limit of the Oki-
propagating to the southwest into the northern tip of the Taiwan nawa Trough and lies in northeast Taiwan (Figs. 1 and 2). The
orogen (e.g., Suppe, 1984; Sibuet et al., 1998; Wang et al., 1999), basin is supplied with sediment from the southwest via the Lan-
or whether it represents the product of gravitational collapse of yang River. This river terminates in a small delta, facing into the
the Taiwan orogen, following subduction polarity reversal (e.g., deeper waters of the Okinawa Trough. To the north the basin is
Teng, 1996; Clift et al., 2003). In the latter case, extension of bounded by exposures of shale and sandstone of the Meichi, Sze-
the Ilan Basin has occurred because of the reversal of subduc- leng, and Chiayang Formations, which are dated as Oligocene,
tion polarity following arc-continent collision, which removes and typically interpreted to be part of the passive margin of China
the stresses that support the high topography in central Taiwan. (Suppe, 1981; Teng, 1990; Lundberg et al., 1997), as well as minor
The earliest phases of extension form the Ilan Plain Basin. The amounts of the underlying Eocene Tachien Sandstone, which
Okinawa Trough represents the continuation of this extension represents a synrift deposit, predating Oligocene seafloor spread-
to form a deep marine basin offshore. Without the tectonic push ing (Fig. 3). The whole sedimentary sequence was deformed and
from the Philippine Sea plate the orogen is able to extend because thrust to the northwest during the Taiwan orogeny, forming the

East China Shelf

Figure 3 h
Tr oug
wa
Figure 2 Okina
rc
kyu A
Ryu
it
ra
St
an Figure 1. Regional bathymetric map of
iw
Ta the Taiwan region, showing the collision
ch between the Luzon Arc and the passive
Ryukyu Tren margin of southern China. The Okinawa
Trough is in active extension to the east
Gagua Ridge

of Taiwan, and its western end comes


onshore in the Ilan Plain. Dashed white
line marks the base of the continental
slope in the South China Sea.
Philippine Sea
h
ug

Arc

Relative plate
Tro
h

motion,7 cm/yr
nc

on
n
e

zo

Luz
Tr

Lu
ila

South China Sea


an

th
r
M

No

Luzon
260 Clift et al.

1.2 ± 0.5 Ma
an
hsh
üe
Hs Well TC

Well TL Figure 2. Shaded topographic map of


the Ilan Plain area, showing the location
Philippine Sea of the drill sites considered in this study,
Ilan the position of apatite fission-track sam-
ples (with average age of the dominant
population), and the orientation of topo-
graphic profiles shown in Figure 5. Map
Lanyang River is from NASA Shuttle Radar data, made
available through www.geomapapp.org.
Well CH
ult
Fa
n
ha
Lis

Suao
Tungao

3.5 ± 0.5 Ma
Backbone Range

modern Hsüehshan Range. Peak metamorphic grades are low Mesozoic and subsequently overprinted by greenschist condi-
(e.g., Chen, 1984). Liu et al. (2001) used zircon fission-track data tions in the Pliocene (Lo and Yu, 1996; Wang et al., 1998).
to show that these units had not been buried beyond the partial Southwest of the Ilan Basin the Lushan and Tatungshan For-
annealing temperature for that mineral (~200 °C; Tagami et al., mations are juxtaposed across the Lishan Fault. In central Taiwan
1998) and inferred a lower greenschist metamorphic facies. the Lishan Fault is manifested as a major, steep, NW-dipping
The southern edge of the basin is marked by outcrops of the thrust structure, having an ESE-vergent geometry, interpreted by
shale-rich Miocene Lushan Formation, which Liu et al. (2001) Clark et al. (1993) as an antithetic backthrust relative to the domi-
considered to have reached prehnite-pumpellyite facies, and nant northwest-vergent thrust stack in the Central Ranges (Wu,
thus is of a slightly lower grade than the Hsüehshan Range. The 1978). Changes in stratigraphic thickness across the fault suggest
Lushan Formation makes up much of the Backbone Range and that it is a reactivated structure that originally formed within an
unconformably overlies the Tananao Schist locally (Suppe et al., extensional rift structure on the South China Sea passive margin
1976), which is formed largely of Paleozoic and Mesozoic sedi- (Teng et al., 1991).
mentary rocks, metamorphosed to greenschist grade. The old-
est sedimentary rocks overlying the Tananao Schist are Eocene NEOTECTONICS OF THE LISHAN FAULT
conglomerates of the Pilushan Formation, but the onlap is time
transgressive between basin and highs on the paleo-Chinese Although the Lishan Fault is a NW-dipping thrust in central
margin. Localized slices of higher grade Tananao metamor- Taiwan, it can be readily traced northward to the SW apex of
phic rocks, including mafic amphibolites, are well exposed on the Ilan Plain. In this region the fault is overturned and has an
the coast at Tungao, south of Suao (Chen and Jahn, 1998). The extensional sense of motion at the western limit of the Ilan Basin,
Tananao Schist is dated to have been metamorphosed during the meaning that the slip sense remains the same. The Lishan Fault is
120 121 122

Pleistocene
volcanic rock Miocene

Taipei Basin
25
Pliocene-
Pleistocene Ilan
h an
e hs Plain
Paleogene
Hs

nge
Ra
Recent

ne
Alluvium

bo
ck
Longitudinal

Ba
Valley 24
(accreted Luzon
Forearc)

Lishan
Fault
Accreted
Chinese Line of section
margin
Coastal
Range 23
(accreted
Luzon Arc)

Paleozoic-
Mesozoic

22

Lishan Tananao Subduction


Fault Complex
Hsüehshan Backbone
Range Longitudinal
Miocene Valley Fault Coastal
Plio-Pleistocene Range

Figure 3. Simplified geological map of Taiwan, showing the map tectonic units that constitute the island and the location
of the Ilan Basin along the strike of the Lishan Fault, which separates the Hsüehshan and Backbone Ranges. The lower
cross section is modified from Lee et al. (2002) and shows the large-scale structure of the thrust belt in central Taiwan and
the geometry of the Lishan Fault, which subsequently has become the dominant structure of the Ilan Plain Basin.
262 Clift et al.

such a major tectonic division that it can be readily traced to pass are relatively fresh and prominent at the northern edge of the
from being a thrust in central Taiwan to an extensional fault in basin along much of its length. Figure 4B shows an example of
the Ilan Plain (Teng and Lee, 1996). In the western Ilan Basin the well-bedded, relatively unweathered sedimentary rocks, exposed
Lishan Fault forms a normal fault of relatively high angle (>60°) close to the trace of the fault near the southwestern end of the
that separates the Hsüehshan Range from the Pleistocene basin basin. In this area the southern boundary of the basin is also sharp
fill (Teng and Lee, 1996). The topographic break along the north- and consistent with neotectonic faulting, as shown in map view
ern edge of the basin is commonly sharp and suggestive of an (Fig. 2) and in topographic profiles drawn across the strike of the
active neotectonic structure (Fig. 4A; Shyu et al., 2005). Expo- basin (Fig. 5). However, the shaded topographic map (Fig. 2) also
sures of the sedimentary rocks that form the Hsüehshan Range shows that the southern basin margin is less linear and poorly

Figure 4. Field photographs. (A) Steep


escarpment along the northeast edge
of the Ilan Plain, testifying to the fault-
bounded nature of the boundary. (B) A
B relatively fresh exposure of sandstone
and shale exhumed by the Lishan Fault
at the western end of the Ilan Plain, tes-
tifying to the recent activity of the fault
along the northern edge of the basin.

2m
Migrating arc-continent collision in the Ilan Basin, Taiwan 263

Eastern transect
N S
800
Elevation (m)

600

400

200 Lishan Fault

0
0 10 20 30 40
1400 Figure 5. Topographic profiles across
Western transect the Ilan Plain Basin, showing the asym-
metry of the basin margins, especially
1200 in the east, where the northern, fault-
N S controlled margin rises rapidly owing to
1000 unloading along the Lishan Fault.
Elevation (m)

800
Lishan Fault
600

400

200

0
0 10 20 30 40
Distance (km)

defined in comparison with the northern boundary. Exposures where motion is transferred laterally. As is often the case in fault
along the southern basin margin are generally not fresh but show relay zones in rifts, this area is marked by rivers that enter the basin
strong deep weathering of the basement, inconsistent with recent laterally, adding to the sediment provided by the axial Lanyang
tectonic exhumation. The eastern part of the plain in particular River (e.g., Leeder and Gawthorpe, 1987).
shows a strong topographic contrast between a gently sloping
southern margin and a steep northern margin (Fig. 5). DRILLING CONSTRAINTS ON NEOTECTONISM
This general contrast between the northern and southern
expressions of the Lishan Fault suggests that the northern bound- Drilling by the Central Geological Survey of Taiwan has
ary is mostly fault controlled. However, the relationship is less allowed the basement contact along the faulted northern edge of
clear in the south and is more consistent with erosion of a tilted the Ilan Basin to be determined at depth as well as at outcrop.
hanging block rather than an uplifted footwall block. Although Cores from three sites were examined in this project, wells TC,
Shyu et al. (2005) proposed a neotectonic fault along the east- TL, and CH, trending east to west close to the northern edge of
ernmost part of the southern boundary, it is noteworthy that this the basin (Fig. 2). These wells intersect the basement at depths of
feature has little topographic expression, suggesting that motion 17, 84, and 38 m, respectively, suggesting an average dip of the
on this structure is rather less than on the Lishan Fault. Sharp basement-cover interface under the northern edge of only ~10°,
increases in topographic gradient within the Backbone Range although this is over a short distance and may not be representa-
south of the basin boundary nonetheless indicate that faulting is tive of the dip over the whole basin width. The overlying sedi-
active in this region. ment comprises sequences of alluvial, coastal, and some shallow
Although the eastern and western ends of the northern margin marine sediment that has been dated as late Pleistocene to Holo-
of the Ilan Basin are well defined, and apparently fault-controlled, cene by AMS 14C methods (Lai and Hsieh, 2003). The sediment
these trends do not align with one another but instead are separated recovered at well CH includes fine-grained muds, silts, and fine
by a section ~10 km wide with less topographic relief (Fig. 2). We sands but is dominated by thick-bedded debris-flow gravels, with
infer a relay section in the Lishan Fault extension in this area, clasts up to 10 cm across (Fig. 6). Well TL recovered the finest
West East
Well CH Well TL Well TC

M. Sand
M. Sand
M. Sand

C. Sand
C. Sand
C. Sand

F. Sand
F. Sand
F. Sand

Cong.
Cong.
Cong.

Mud
Mud
Mud

Silt
Silt
Silt
0 20 0

10 30 10

20 40 20

Basement
30 50 30
Depth (m)

40 60 40
Basement

50 70

60 80
Basement

70 90

80 100
Figure 6. Sedimentary logs showing the cored sequences found along the northern edge of the Ilan Plain, where Pleisto-
cene sediment is deposited over low-grade metamorphosed sedimentary rock, commonly below a zone of fault breccia
that marks a detachment surface.
Migrating arc-continent collision in the Ilan Basin, Taiwan 265

grained sand of the three drilled sequences, though the sediment Because the sediment cored in the Ilan Basin is all close
is still characterized by coarse sands and gravels. Well TC recov- to sea level and preserves evidence of the environment of its
ered a sequence of medium- to thick-bedded sands and muds. sedimentation, these materials can be used to determine rates of
In each location the sediment is underlain by well-lithified, low- tectonic subsidence across the plain. In the Ilan Basin the fast
grade metamorphosed shale, typical of that seen in outcrop in the rate of the sediment supply, largely from the Lanyang River,
Hsüehshan Range. appears generally to match rates of basement subsidence, keep-
All drilled basement samples are distinguished by the pres- ing the basin full while allowing shoreline progradation toward
ence of discrete brecciated layers, generally 2–5 m thick at the the Okinawa Trough. Lai and Hsieh (2003) demonstrated that
basement contact but also present within the basement below that subsidence reaches a maximum rate of ~19 mm/yr around the
level. These breccias are monomict and comprise only fragments mouth of the Lanyang River and decreases toward the edges
of metamorphosed shale (Fig. 7). The lack of material in the of the plain, both north and south, as well as toward the west.
breccias from the shallower basin fill sediment probably reflects However, what is important toward understanding the tectonics
the fact that they are loose, unconsolidated materials and unable of the Ilan Basin is the recognition that the northern part of the
to support the confining stresses needed to be incorporated into plain subsides faster than the southern sector, reaching average
the gouge. Texturally the breccias are generally very angular, Holocene rates of 6–7 mm/yr versus 3–5 mm/yr. This disparity
unsorted, and matrix supported. The breccias are relatively well is consistent with the observations presented above that indicate
lithified in comparison with the overlying soft Pleistocene sedi- a dominant asymmetric character to the basin structure. We thus
ment. We interpret these rocks to represent fault breccias formed favor basin formation above a dominant south-dipping exten-
by cataclasis in the shallow subsurface of the Lishan Fault. The sional detachment (Fig. 9).
brittle, angular, broken character of the fault-breccia clasts is
apparent in thin section. Critically, the fault breccias postdate the MODELING BASIN GEOMETRY
development of cleavage in the metapelites that form the footwall
(Fig. 8). As a result, motion on the fault is constrained to being The large-scale structure of the basin can be considered and
post–peak metamorphism and relatively shallow (<8 km) in the modeled using a theoretical forward of an asymmetric basin, con-
crust. The breccias are preferentially formed close to the basement trolled by a dominant detachment. We employed the flexural can-
contact, because this is the fault plane exposed to the surface by tilever model of Kusznir et al. (1991) to forward model the defor-
the exhumation of the footwall block. Not all the breccia and fault mation and subsidence that would be expected to result from
layers lie exactly along the basement-sediment contact, indicat- the extension measured across the normal fault that bounds the
ing that the Lishan Fault has been active along several fault splays northern edge of the basin, and to test the idea that the regional
within the footwall block during uplift of the Hsüehshan Range. dip is 10°, as inferred from the drilling. Although the flexural
The drill data confirm the presence of a major southeast-dipping cantilever model is not the only, or universally accepted, theo-
structure that controls the basin’s northern margin (Fig. 9). retical model for extensional basins, it does have a number of

A B

Figure 7. Photographs of cored pre-


Pleistocene rocks from well CH in the
western Ilan Basin. (A) Fault-brecciated
zone at 75.78 m depth. (B) Coherent
block of metapelite from 84.50 m depth.

5 cm
266 Clift et al.

A B

Figure 8. Photomicrographs of thin sec-


tions cut from basement rocks in well
CH. (A) Typical fine-grained charac-
ter of the basement metapelites, with a
well-developed cleavage. (B) Fault rock
from within one of the breccia zones,
demonstrating low-temperature shatter-
ing textures.

250 µm 500 µm

Distance (km)
0 10 20 30 40
1000
N S
500
Figure 9. Schematic cross section of the
Elevation (m)

0 regional basin structure through the east-


ern Ilan Basin, showing the dominant
control of a south-dipping detachment
500 in controlling uplift of the Hsüehshan
Range and preferential subsidence of
the northern plain. Shading patterns are
1000 not intended to accurately portray de-
formation structures in the footwall and
1500
hanging-wall blocks.
Lishan Fault

2000
Lushan Formation (Miocene)
Pleistocene to Recent sediments
meta-sedimentary rocks
Paleogene low grade
Mesozoic subduction complex meta-sedimentary rocks

important features that allow the basin tectonics to be understood continental lithosphere using the major detachment fault and
at a first-order level, and it has been used to effect in understand- predicted what sort of basin this would form. Flexural rigidity
ing basins in several parts of the world (Kusznir et al., 1995; Rob- is expected to be low in an arc orogenic setting. Effective elas-
erts et al., 1993). tic thickness (Te) is only 13 km in the Taiwan foreland (Lin and
The model is complicated by our lack of subsurface data and Watts, 2002) and is expected to be rather less in the orogenic
constraint on the degree of extension across the fault, although core that we consider here. The deformation in the flexural can-
this is also limited by the total size of the basin. For modeling tilever model assumes brittle faulting in the upper crust, set at
purposes we chose a default of 5 km horizontal extension and a 10 km thickness here, and ductile deformation distributed in a
crustal thickness of 40 km. In this approach we made no attempt sinusoidal fashion over a wavelength of 100 km. The results of
to replicate the basin morphology but simply extended a model the models are shown in Figure 10.
Migrating arc-continent collision in the Ilan Basin, Taiwan 267

Figure 10. Proposed basin geometries and comparisons with possible model basins in order to limit the possible faulting
geometries and strength of the lithosphere under the Ilan Plain. (A) Proposed basin cross section inferred from outcrop,
coring, and gravity constraints. (B–D) Forward models, made using the flexural cantilever model of Kusznir et al. (1991)
and a 30° detachment fault, with variations in the flexural rigidity of the lithosphere. (E–G) Forward models using a 10°
detachment fault, showing the effects of varying amounts of extension and flexural rigidity. (H) Forward models using a
45° detachment fault, showing a poor match to observed basin shape. Te—elastic thickness.

An important result of our forward models is that basins using a 30° detachment, with a 5 km extension and a Te of 3 km
generated above a 10° detachment surface do not produce real- (Fig. 10A, B). Although this is not a unique solution, it does give
istic basin geometries (Fig. 10A, E–G), mostly because they some guidance as to the type of fault and mechanical state of
do not predict significant footwall uplift of the size seen in the the crust under the Ilan Plain. Changing flexural rigidity to either
Hsüehshan. Also, 10° detachments fail to produce footwall 1 km or 5 km changed the basin width to less close fits (Fig. 10A,
uplift even when Te is increased to 10 km (Fig. 10A, G), or if C, D). The modeling is also effective at eliminating very high
the extension across the fault is increased to 10 km. We conclude angle faults as possible mechanisms, because even a 45° fault
that the 10° dip inferred from coring cannot be representative of produces too much footwall uplift, hanging-wall subsidence, and
the basin. The best first-order fit to basin geometry was acquired a wide basin (Fig. 10A, H).
268 Clift et al.

FISSION-TRACK ANALYSIS Of 12 samples analyzed, only 2 yielded sufficient apatite


to produce meaningful results. Fortunately these samples came
The exhumation history of the ranges around the Ilan Basin from opposite sides of the basin (Fig. 2) and allow the exhuma-
can be examined using low-temperature thermochronometry. tion histories of the margins to be compared. Central ages of 1.2
Although previous studies have used fission-track methods in the ± 0.5 Ma and 3.5 ± 0.5 Ma were recorded at the coast on the
central parts of Taiwan (e.g., Liu et al., 2000, 2001; Willett et north and south basin margins, respectively (Fig. 11). The south-
al., 2003), no similar studies around the Ilan Basin have been ern sample contains a minor number of older grains (older than
undertaken. Liu et al. (2001) used zircon fission-track methods to 150 Ma), reflecting an earlier phase of cooling. However, for this
date the cooling history and included a transect north of the basin, study we focus on most of the population, which records cooling
across the Hsüehshan Range. Fission-track studies in zircon grains linked only to Pliocene–Holocene tectonism. The single, young
record cooling through a partial annealing zone of 200–320 °C grain population seen on the northern margin indicates total reset-
(Tagami et al., 1998), and so this method is sensitive to exhu- ting and recent, rapid cooling. Exactly why the southern margin
mation driven by erosion and has been widely used in orogenic sample contains a few grains that are not reset at 3.5 Ma is not
exhumation studies. However, Liu et al. (2001) recorded zircon clear. However, the vast majority of grains in that sample show a
fission tracks north of the Ilan Basin that are mostly older than well-defined 3.5 Ma cooling trend (Fig. 11). The spread of ages
the Taiwan orogeny. Central ages are as low as 22 Ma close to the would be greater if only partial annealing had occurred. We con-
topographic front of the Hsüehshan Range but become older far- clude that the sample underwent rapid cooling through the AFTA
ther north, consistent with more exhumation at the margin of the annealing temperature zone ca. 3.5 Ma. Our result indicates more
basin (Liu et al., 2001). That the zircon fission tracks are this old recent cooling of the north margin in comparison with the south,
indicates that they have been partially but not fully reset by late as might be expected for an asymmetric basin controlled by a
Miocene–Holocene collision-related burial. A lower temperature south-dipping detachment. What is surprising is that this result is
thermochronometer is thus required to reconstruct the cooling of the opposite of Liu et al.’s (2001) zircon fission-track result that
rock units in shallower parts of the crust. has central ages of only 1.5 ± 0.3 Ma in the south close to Suao
In this study we employed the fission-track method applied and 22.2 ± 2.3 Ma along the northern margin, which would imply
to apatite, which records cooling through ~125–60 °C over time the opposite sense of motion. Such motion is not consistent with
scales of 1–10 m.y. (Green et al., 1989). Apatite fission-track the structural asymmetry presented above. Although the apatite
analysis was performed at University College, London, UK. and zircon results are consistent in the north, it is impossible to
Polished grain mounts were etched with 5N HNO3 at 20 °C for have younger fission-track ages for zircon than for apatite, call-
20 s to reveal the spontaneous fission tracks. Subsequently, the ing into question whether the zircon or apatite data from Suao are
uranium content of each crystal was determined by irradiation, representative of the region. Assuming a geothermal gradient of
which induced fission of 235U. The induced tracks were registered 30 °C/km (Barr and Dahlen, 1988; Willett et al., 2003), the apa-
in external mica detectors. The samples for this study were irradi- tite data imply average exhumation rates of 1.6 ± 0.3 km/m.y. in
ated in the thermal facility of the Hifar Reactor at Lucas Heights, the north and 0.7 ± 0.1 km/m.y. in the south. Because we demon-
Australia. The neutron flux was monitored by including Corning strate the importance of extensional motion on the Lishan Fault
glass dosimeter CN-5, with a known uranium content of 11 ppm, in generating uplift of the Hsüehshan Range, much of the exhu-
at either end of the sample stack. After irradiation, sample and mation on the north margin may be linked to tectonic unroofing
dosimeter mica detectors were etched in 48% HF at 20 °C for by detachment faulting rather than by erosion.
45 min. Only crystals with sections parallel to the c-axis were
counted, as these crystals have the lowest bulk etch rate. To avoid SEISMIC EVIDENCE
biasing results through preferred selection of apatite crystals, the
samples were systematically scanned, and each crystal encoun- Observation of recent earthquakes is important to under-
tered with the correct orientation was analyzed, irrespective of standing the neotectonic evolution of the Ilan Basin, as they
track density. The results of the fission-track analysis are pre- provide a measure of where faulting is active and a snapshot of
sented in Table 1. the current stress regime. The seismicity used here is a subset of

TABLE 1. FISSION-TRACK ANALYTICAL DATA


Sample no./ Mineral No. Dosimeter Spontaneous Induced Age dispersion Central age Age components
ρd ρs ρi
2
field no. of Nd Ns Ni Pχ RE% (Ma) mode 1 mode 2 mode 3
crystals ±1 (number of grains)
TW114-1 Apatite 30 1.387 3845 0.185 141 3.639 2952 0.0 260.4 14.4 ± 7.0 3.5 ± 0.5 22 ± 12 294 ± 78
(26) (2) (2)
TW116-3 Apatite 15 1.387 3845 0.009 13 1.666 2116 19.8 97.0 1.2 ± 0.5 Single pop
6 –2
Note: (i) Track densities are (×10 tr cm ) numbers of tracks counted (N) shown in parentheses; (ii) analyses by external detector method using 0.5 for the 4π/2π geometric
correction factor; (iii) ages calculated using dosimeter glass CN-5; (apatite) ξCN5 = 338 ± 4; CN-2; (zircon) ξCN2 = 127 ± 5 calibrated by multiple analyses of IUGS apatite and zircon
age standards (see Hurford, 1990); (iv) Pχ is probability for obtaining χ value for v degrees of freedom, where v = no. crystals – 1; (v) central age is a modal age, weighted for
2 2

different precisions of individual crystals (see Galbraith, 1990).


Migrating arc-continent collision in the Ilan Basin, Taiwan 269

400 However, a coherent array of deep earthquakes is not seen under


TW114-1 300
Central age: 14.4 ± 7 Ma 200 the western Okinawa Trough in the 0–50 km depth range consid-
P(χ2): 0.0% 150 ered here. Our deeper profile (Fig. 12) indicates that the Philip-
Relative error: 256%
100 pine oceanic slab is deeper in this location and that therefore the
Number of crystals: 30 array of seismic events seen in Figure 14A–C reflects a separate
lithospheric structure. The NE-trending seismic zone in the cen-
tral part of the Ilan Plain is populated by shallow normal-faulting
events with NW-SE directed T-axes (extension). In the southeast
basin, shallow seismicity appears to be arranged into two shallow
+2 10 E-W–trending zones dominated by sinistral strike-slip events and
0 N-S–oriented normal faulting, but with T-axes in the same general
-2
direction as in the basin center. The two areas of intense seismic-
3.5 ± 0.5 Ma ity are not that different in that both show a shallow layer of seis-
% Relative error
63 10 0 micity between a narrow range of 8–13 km. Analysis of the map
0 and sections in Figures 13 and 14 reveals the subtle differences
10 20 30 40 in seismicity associated with different belts. In the first place, the
Precision (1/sigma)
shallow events are concentrated in a fairly narrow depth range.
TW116-3 This is curious; one possible explanation is that the seismic layer
Central age: 1.2 ± 0.5 Ma 4 is limited on its upper surface by the presence of relatively incom-
P(χ2): 19.8% petent and therefore nonbrittle sediment. In contrast, the lower
Relative error: 97%
limit of seismic activity may be bounded by hot middle crust at
Number of crystals: 15
~10 km depth, where ductile deformation precludes seismogen-
+2 3 esis. Sections B and E (Fig. 14) show the depth distribution of the
2 E-W–trending strike-slip events around 24°34′N (southern end
of the profiles). The NE-SW–trending belt does show well in the
0 1 cross section, with only a few events >30 km in Figure 14B, C.
No evidence exists for a shallow-dipping active slip plane bound-
ing the basin to the north on the basis of seismicity. Whatever fault
-2 motion we can confirm is either diffused or concentrated in a nar-
0 row depth range, probably between the bottom of the sedimentary
% Relative error
53 8 layer and the top of the basement, judging from the rheological
0 properties of these materials.
10 20 30 40 50
Precision (1/sigma)
The events >50 km deep under the basin axis are interpreted
as thrust earthquakes, consistent with their association with the
Figure 11. Radial plots of apatite fission-track analyses from Samples Ryukyu subduction zone, which is now starting under northern
TW114-1 and TW116-3, respectively, from the southern and northern Taiwan by lateral motion of the Philippine Sea plate into a tear
margins of the Ilan Basin. Locations shown in Figure 2.
in the passive margin lithosphere of southern China (Lallemand
et al., 2001). However, analysis of earthquakes shallower than
20 km from the nearby Okinawa Trough shows a dominant
extensional character, with extension perpendicular to the strike
events initially reported by the Central Weather Bureau of Tai- of the trough, i.e., NW-SE. These shallow earthquakes form an
wan. The phase data for these events are used for relocation by offshore continuation of the shallow events seen under the Ilan
means of the double-difference method of Waldhauser and Ells- Basin, which are similarly interpreted as being extensional. No
worth (2000). The relocation procedure minimizes the errors in strong evidence exists from earthquakes to support the presence
hypocentral determination because of lateral velocity variations. of a currently active major south-dipping detachment (e.g., the
The resulting events tend to cluster around known structures, as Lishan Fault). However, if the major extensional fault is shal-
shown by Waldhauser (2001). low dipping, then it is too shallow to store elastic stresses and to
The Benioff zone in this area is below 50 km and strikes nearly be seismogenic under the northern part of the Ilan Plain. Deeper
E-W and dips at ~30° to the north (Fig. 12; Wu et al., 1997). In the faulting under the southern edge of the basin is consistent with
Ilan area the deeper seismicity (~50 km) is concentrated along the these asymmetric structural models. Strike-slip faulting in the
southwestern margin of the basin, as well as more generally across southeast basin may be interpreted as part of the strain accom-
the northern half of the basin. The seismicity in the upper 50 km modation related to the bend in the tectonic fabric from a N-S ori-
is shown in Figure 13. Cross sections through the basin show that entation in central Taiwan to an E-W orientation in the Ryukyu
the deeper earthquakes form a NW-dipping array (Fig. 14A–C). Arc. The strain observed indicates motion of the southern margin
270 Clift et al.

S N
Elevation

3
(km)

0
121˚ 122˚
0
25˚ 25˚
Figure 12. Cross section off the east
coast of Taiwan, showing recent seis-
Depth (km)

50 micity down to 150 km depth. The loca-


tion of the section is shown as a black
line on the map (right). Seismicity di-
vides broadly into the shallow seismic-
ity seen in detail in Figures 13 and 14,
and the well-defined north-dipping Be-
100 24˚ 24˚ nioff zone of the Ryukyu Arc.

121˚ 122˚

0 50 100 150
Distance (km)

of the Ilan Basin to the ESE. The sense of motion is the opposite on the seismology alone, implying extrusion of this crustal block
of that expected from the bending of the structural fabric, but it toward the Okinawa Trough. We do recognize the importance of
is consistent with the lateral motion of orogenic crust away from the strike-slip zone resolved by the seismic data and note that
the Taiwan mountains toward the Ryukyu Trench, as might be GPS motions are more south-oriented, directly toward the trench,
predicted for orogenic collapse driven by a change in stresses south of that lineament.
triggered by generation of that trench. The GPS motion data can be understood in the context
of the tectonic setting that is characterized by E-W compres-
GEODYNAMIC EVOLUTION sion in central Taiwan, contrasting with N-S subduction to the
east along the Ryukyu Trench. In practice the Ryukyu Trench
Additional constraints on the nature of current tectonic strain provides a free edge, allowing the southward motion of crustal
are provided by global positioning system (GPS) monitoring of blocks in the Ryukyu forearc, compared with the compression
the region. Here we consider the motions reconstructed by Chang by the Luzon-China collision (Fig. 16). If Taiwan is considered
et al. (2003), as shown in Figure 15. Relative to stable southern an arc-continent collision orogen that is progressively migrat-
China the ranges north of the Ilan Basin are moving slowly to ing along the Chinese margin toward the southwest, then the
the northwest (~10–15 mm/yr), effectively a continuation of the passive-margin units that are compressed and uplifted by col-
west-directed thrusting that characterizes most of Taiwan, albeit lision may then collapse and extend, once the restraining but-
slightly rotated clockwise in this particular area. In contrast, GPS tress of the Philippine Sea plate has been removed and sub-
locations in the Ilan Basin itself are moving almost due east, duction with the opposite polarity is initiated. The load of the
while the basement around Suao on the southern boundary is Taiwan orogen causes flexure of the underlying Chinese conti-
being displaced toward the southeast at 15–20 mm/yr. The net nental crust (Yu and Chou, 2001; Lin and Watts, 2002; Lin et
result indicates motion of the Hsüehshan and Backbone Ranges al., 2003). Where that load slides southeastward away from the
away from one another, consistent with the seismic evidence margin’s edge in northern Taiwan, the continental crust quickly
for extension across the basin. The eastward motion of the Ilan rebounds and regains much of its normal thickness (Rau and
Basin shows the same sense of motion as the strike-slip fault Wu, 1995). Thus SE-directed motion of the basement south of
data derived from the fault plane solutions of the seismic data. the Ilan Basin reflects collapse of the flank of the Taiwan orogen
It should be noted that the major strike-slip structures (Fig. 11) and motion of the Backbone Range toward the Ryukyu Trench.
are south of the GPS stations on the Ilan Plain. The GPS data Extension in the Okinawa Trough, driven by southward motion
thus indicate a broader zone of crustal extrusion than that based of the Ryukyu forearc toward the trench also provides space for
121˚30’ 121˚36’ 121˚42’ 121˚48’ 121˚54’
25˚00’

D E F

?
?
24˚54’ ?
?

? ?
?
?
?
?
?
? ?
24˚48’

C ? ?
A B
?
? ?

?
24˚42’ ? ? ? ?
? ? ?
?? ?
?
? ?
?
?
?
?
?
24˚36’ ?

? ?
? ?
?

km ?
?

0 5 10
24˚30’

0 50 100 150
Hypocentral Depth (km)
Figure 13. Map showing the location of earthquake hypocenters around the Ilan Basin, with the dominant extensional
character of recent events at depths <20 km, whereas deeper events are thrust motions linked to the Ryukyu subduction
zone and the Taiwan collision. Parallel lines A, B, C and D, E, F denote cross sections shown in Figure 14.
272 Clift et al.

SE NW SE NW SE NW
Depth (km) 4 A B C
0
-4
0
Altitude (km)

25

50
0 25 50 0 25 50 0 25 50
Distance (km) Distance (km) Distance (km)
S N S N S N
4
Depth (km)

D E F
0
-4
0
Altitude (km)

25

50
0 25 50 0 25 50 0 25 50
Distance (km) Distance (km) Distance (km)

Figure 14. Cross sections down to 50 km depth across the basin show concentrated seismicity ~13–8 km depth. Those in
the 13–50 km range are evidently associated with collision of the Philippine Sea and Eurasian plates, and the subduction
events are at depths >50 km in this region. Section locations are shown in Figure 13.

the eastward motion of material in the Ilan Basin, driven by the with the extension in the Ilan Basin being the product of post-
gravitational potential of the thickened crust in central Taiwan. orogenic collapse, triggered by subduction polarity reversal, as
This motion is partially accommodated by strike-slip tectonism argued by Teng (1996), and not the result of extension propa-
as seen in the seismic data, as well as by extension, especially gating from the Okinawa Trough, driven by trench forces. The
that focused on the major fault bounding the Ilan Plain to the extensional fault that controls the Ilan Basin does not represent
north, inferred to be the Lishan Fault, reversed in direction in a westward propagating rift of oceanic origin cutting across the
comparison with its thrust sense in the Central Ranges. older orogenic fabric, but rather a reversal of motion on thrusts
as compressional stresses are released toward the northern end of
DISCUSSION the orogen. The major bounding fault can be clearly traced as an
extension of the Lishan Fault, which is a major thrust structure
The fact that the Lishan Fault can be traced along strike from in the Central Ranges but is overturned and moving as an exten-
a major thrust fault into an extensional detachment is consistent sional structure where it reaches the Ilan Plain. The initiation of
Figure 15. Map showing the motions of crustal
blocks around the Ilan Basin relative to stable
Eurasia, recorded by Chang et al. (2003) using
Ilan Basin GPS methods over the period 1990–1995.

Velocity 50 mm/yr

South of Taiwan
Chinese passive Accretionary North Luzon Luzon
Active oceanic subduction margin Complex Trough Arc
NW SE
sealevel

oceanic crust Arc lower crust


continental crust and mantle
Lishan Fault
thrust
Central Taiwan
Compression and
metamorphism sealevel
Figure 16. Schematic depiction of the origin of
the Ilan Basin as a result of gravitational col-
lapse of the Taiwan Central Ranges during a
SW-migrating collision of the Luzon Arc and
mainland China.

Northeastern Taiwan Lishan Fault


Exhumation and Taipei Basin detachment
formation of Ilan Basin
Ilan Basin
sealevel
274 Clift et al.

the Ryukyu Trench removes the compressive force of the Philip- Here we propose that the Ilan Basin can best be under-
pine Sea plate from the orogen and allows the edifice to extend stood in the context of a migrating arc collision, and especially
as the trench provides a free edge toward which material can be as the culmination of gravitational collapse of the resultant oro-
displaced. In turn this suggests that at least the southwestern end gen, having taken place over relatively short periods of geologic
of the Okinawa Trough has also formed as a result of migrat- time (Fig. 17). Collision of the modern Taiwan section of the
ing arc-continent collision and subsequent orogenic collapse and Luzon Arc with China is thought to have begun ca. 6–9 Ma (e.g.,
is not generated by subduction slab rollback (cf. Suppe, 1984). Dorsey, 1988; Teng, 1990; Sibuet et al., 2002; Malavieille et al.,
This is an important revision of the generally held belief that the 2002; Huang et al., 2006) and is already finished and in a state
Okinawa Trough is formed by trench forces, principally rollback of rapid exhumation around the Ilan Basin. Byrne and Crespi
of the trench, much as suggested for the Mariana Trough or the (1997) reported extension throughout the Backbone and Central
Lau Basin (e.g., Hawkins, 1974). However, in the rollback model Ranges. The stronger extension seen around the Ilan Basin can
the presence of the westernmost propagating end of the rift at be understood as an extension of this, made possible as the new
the northern tip of Taiwan would be coincidental, whereas rift- Ryukyu Trench form, removing the compressive stresses of col-
ing related to collision and subduction polarity reversal would lision and allowing the flank of the Taiwan orogen to move later-
predict this position. In practice, little evidence exists for active ally into that space.
rift propagation. The Okinawa Trough is not V-shaped as seen in Figure 18 shows the relationship between tectonically driven
other western Pacific basins, where a propagating rift culminates rock uplift rates and exhumation rates in Taiwan, assuming that
in seafloor spreading. The Okinawa Trough appears to be open- “hard collision” in northern Taiwan initiated ca. 5 Ma, whereas in
ing to the west, yet the older eastern parts of the basin are no the south hard collision between the Luzon Arc and China is just
wider than that closest to Taiwan. In the Okinawa Trough, rifting starting at the modern coast. Initial collision between the Luzon
is followed by a cessation of extension, whereas in contrast roll- forearc and the Chinese margin begins farther south, with the devel-
back and arc rifting are normally followed by further extension in opment of an accretionary prism, which progressively overthrusts
the form of seafloor spreading. the North Luzon Trough (forearc basin; Lundberg et al., 1997).
Evidence that the westward migration of the Okinawa Regional trends in rock uplift rates can be determined from the cur-
Trough is matched by arc migration was provided by Shinjo rent elevations and the age of the collision, together with estimates
(1999), who noted that middle Miocene volcanic rocks from the for the amount of exhumation derived from the metamorphic-
southern Ryukyu Arc are not subduction related but instead are grade and fission-track data (e.g., Dadson et al., 2003). Although
similar to intraplate volcanism seen in China. Such an observa- in some areas modern rates of uplift have been determined by dat-
tion implies that the Ryukyu Arc is new and that the subduction- ing terraces (e.g., Lin, 1969; Peng et al., 1977; Vita-Finzi and Lin,
related Ryukyu volcanic front has propagated into the area since 1998), these terraces are necessarily limited to the coastal regions,
that time. This hypothesis is also supported by the geochemistry mostly in the Coastal Ranges of eastern Taiwan.
of recent volcanic rocks from the southernmost Okinawa Trough Exhumation rates driven by erosion reach a peak in the
(Chung et al., 2000). In a collapse model the Okinawa Trough south of the island, because rates of rock uplift are highest dur-
becomes increasingly younger to the south, the opposite of the ing the most intense period of collisional compression between
rollback model in which the Okinawa Trough spreading centers Luzon and China; these are partly balanced by erosion driven
are propagating into the basin away from Taiwan. largely by precipitation but also by tectonic extension (Crespi et
It is noteworthy that active magmatism and faulting that cuts al., 1996; Teng et al., 2000). Exhumation and vertical uplift rates
right to the seafloor in the Okinawa Trough is restricted to that decrease abruptly toward the northern end of the Central Ranges,
part of the basin closest to Taiwan (Sibuet et al., 1998). In contrast, especially around the Ilan Basin, although active motion along a
middle to late Miocene (6–9 Ma) extension ages are recorded in detachment reversing the Lishan Fault causes increased exhuma-
the northern Okinawa Trough (Letouzey and Kimura, 1985). The tion in the Hsüehshan Range. The rates of burial and exhumation
northern Okinawa Trough thus is either the remnant of an earlier are some of the highest known on Earth. Exhumation rates at
arc-continent collision, as favored by Clift et al. (2003), or has a Nanga Parbat in the Pakistan Himalaya reach 5 mm/yr (Zeitler et
separate origin from that part of the basin closer to Taiwan. The al., 1993), slightly less than the rates seen in the Central Ranges
basement of the southern Okinawa Trough is inferred to be the (Dadson et al., 2003). Although exhumation rates are lower in
extended remnants of the Taiwan orogen, equivalent to the Back- the Hsüehshan Range, reaching >1.6 mm/yr adjacent to the
bone and Hsüehshan Ranges. The recognition of a continuous Lishan Fault, these levels are still relatively high, far exceeding
Taiwan-Sinzi folded zone under the SE edge of the East China rates of 0.33 mm/yr in the Cascade Range of the northwestern
Sea (Hsiao et al., 1999) would suggest a continuous migration of United States (Reiners et al., 2003), somewhat more than the
the orogen from Sinzi at ca. 12 Ma to present-day Taiwan. Vol- ~1 mm/yr found in the Alaska Range (Fitzgerald et al., 1993),
canism is the manifestation of the new arc volcanic front to the and comparable to the 2.5–0.8 mm/yr measured by Tippett and
Ryukyu subduction zone, which overlies the nonvolcanic forearc Kamp (1993) in the southern Alps of New Zealand. Clearly arc-
ridge only in the central and northern Ryukyu Arc, where active continent collision forms some of the most dynamic geology on
extension has ceased. the planet.
Migrating arc-continent collision in the Ilan Basin, Taiwan 275

East China Sea t


ldbel
Fo
Si nzi
n-
wa
Tai
orogenic h
Tr oug
collapse wa
O kina

Arc Figure 17. Shaded bathymetric map of the


kyu
Ryu Taiwan region, showing the collisional oro-
collisional gen, the opposing subduction polarities,
orogen and the Okinawa Trough opening in the
wake of orogenic collapse. The numbered
lines adjacent to the plate boundary show
the inferred time of peak arc collision be-
+4 m.y.
tween Luzon and China. Map is labeled to
show the different stages of arc-continent
collision along strike. Wide white dashed
collision line shows location of the Taiwan-Sinzi
migration Fold Belt, interpreted as remnants of the
early former collisional orogen. Black dashed
collision 7 cm/yr line shows the location of the modern arc
volcanic front, focused by extension in the
Okinawa Trough close to Taiwan.

-5 m.y. Philippine Sea


oceanic
arc

South China Sea

Luzon

CONCLUSIONS start of collapse onshore, the process reached its culmination off-
shore in the Okinawa Trough, indicating that at least some of this
A variety of geological and geophysical data demonstrates basin has formed as a result of collision and not by trench rollback
active extension along a NNW-SSE axis across the Ilan Basin of forces as previously believed (cf. Suppe, 1984). We propose that
northern Taiwan. The structure of the basin appears to be largely mountain building and rifting of the Ilan Plain and the Okinawa
controlled by a SE-dipping detachment fault, likely dipping at Trough are all results of a single common process, arc-continent
~30°, that causes uplift of the Hsüehshan Range to the north of the collision, rather than being the unique interaction of a collisional
basin and preferential fast subsidence of the northern Ilan Basin. orogen and a propagating backarc rift. Like the Alboran Sea in the
Apatite fission-track ages confirm geomorphic evidence for faster western Mediterranean the Ilan-Okinawa Trough shows how far
exhumation of the northern margin of the basin, reaching rates of postorogenic collapse may drive extension and basin formation
at least 1.6 ± 0.3 km/m.y. The main basin-bounding extensional (e.g., Platt and Vissers, 1989), although in this case without the
fault is mapped as a continuation of the Lishan Fault, a major thrust need to invoke delamination of the mantle lithosphere.
structure from central Taiwan. The southern Backbone Range is Because arc-continent collision is a common process in the
shown by seismic and GPS data to be moving southeast toward plate-tectonic cycle we anticipate that collapse basins similar to
the newly formed western Ryukyu Trench. Formation of a free the Ilan Basin should be a common feature and should be recog-
edge in the trench and release of the E-W compressive stresses nizable in ancient collision zones. Although the Ordovician South
allow the Taiwan orogen to collapse in the Ilan Basin–Okinawa Mayo Trough of the Irish Caledonides has been interpreted to be
Trough region. Because the Taiwan orogen is migrating to the such a basin (e.g., Clift et al., 2003), the scarcity of other exam-
southwest along the passive margin of southern China, we sug- ples likely reflects a lack of understanding of existing sequences
gest that the extension and associated basin formation must also rather than the absence of gravitational collapse in ancient arc
move in this direction. Although the Ilan Basin was formed by the collisional events.
276 Clift et al.

Date of collision (Ma)


5 4 3 2 1 0

Line
of se
ction

3000 3000
Okinawa Tr.

Ilan Basin
Altitude (m)

Altitude (m)
2000 2000

1000 1000 Figure 18. Summary of the trends in


rock uplift and exhumation rate along
the length of the island of Taiwan, de-
0 0
picted also as a time progression of
hard collision during initial arc-conti-
-1000 -1000 nent collision passing into extension,
20 12 exhumation, and collapse, culminating

Exhumation Rate (mm/yr.)


cum. exhumation in formation of the Okinawa Trough.
10 Exhumation rates are from Dadson et
Exhumation (km)

15 al. (2003). Gray line shows long-wave-


8 length topography.
Cumulative

10 6

4
5 exhumation rate
2

0
25
Rate of rock uplift (mm/yr.)

cum. uplift
Cumulative rock uplift (km)

20
10

15
5
10
0
5
rate of uplift
-5
0

-5 -10
400 300 200 100 0
Distance (km)

ACKNOWLEDGMENTS fieldwork from an NSC (Taiwan) grant of 93–2116-M-008–001.


F.T.W. acknowledges support from the U.S. National Science
P.C. thanks the University of Aberdeen, College of Physical Sci- Foundation’s Continental Dynamics Program. We thank Don
ences; the Alexander von Humboldt Foundation; and the Joint Fisher, Neil Lundberg, Dave Topping, and Patrick Barnard for
Oceanographic Institutions for contributions to the cost of this their helpful reviews, as well as editor Dave Scholl for his help
research. A.T.L. is grateful for financial support for conducting in improving this manuscript.
Migrating arc-continent collision in the Ilan Basin, Taiwan 277

Hsu, S.-K., and Sibuet, J.-C., 1995, Is Taiwan the result of arc-continent or arc–
REFERENCES CITED arc collision?: Earth and Planetary Science Letters, v. 136, p. 315–324,
doi: 10.1016/0012-821X(95)00190-N.
Barr, T.D., and Dahlen, F.A., 1988, Thermodynamic efficiency of brittle fric- Huang, C.Y., Yuan, P.B., Lin, C.W., and Wang, T.K., 2000, Geodynamic pro-
tional mountain building: Science, v. 242, p. 749–752, doi: 10.1126/sci- cesses of Taiwan arc-continent collision and comparison with analogs in
ence.242.4879.749. Timor, Papua New Guinea, Urals and Corsica: Tectonophysics, v. 325,
Byrne, T., and Crespi, J., 1997, Kinematics of the Taiwan arc-continent colli- p. 1–21, doi: 10.1016/S0040-1951(00)00128-1.
sion and implications for orogenic process: Tapei, Taiwan, International Huang, C.Y., Yuan, P.B., and Tsao, S.H., 2006, Temporal and spatial records of
Conference and Sino-American Symposium on Tectonics of East Asia, active arc-continent collision in Taiwan: A synthesis: Geological Society
p. 38. of America Bulletin, v. 118, p. 274–288, doi: 10.1130/B25527.1.
Casey, J.F., and Dewey, J.F., 1984, Initiation of subduction zones along trans- Hurford, A.J., 1990, Standardization of fission-track dating calibration: Recom-
form and accreting plate boundaries, triple-junction evolution, and forearc mendation by the Fission-track Working Group of the IUGS Subcommis-
spreading centres—Implications for ophiolitic geology and obduction, in sion on Geochronology: Chemical Geology, v. 80, p. 177–178.
Gass, I.G., et al., Ophiolites and Oceanic Lithosphere: Geological Society Konstantinovskaya, E.A., 1999, Geodynamics of island arc–continent collision
[London] Special Publication 13, p. 269–290. in the western Pacific margin: Geotectonics, v. 33, p. 15–34.
Chang, C.-P., Chang, T.-Y., Angelier, A., Kao, H., Lee, J.-C., and Yu, S.-B., Kusznir, N.J., Marsden, G., and Egan, S.S., 1991, A flexural cantilever simple
2003, Strain and stress field in Taiwan oblique convergent system: Con- shear/pure shear model of continental extension, in Roberts, A.M., et al.,
straints from GPS observation and tectonic data: Earth and Planetary Sci- eds., The Geometry of Normal Faults: Geological Society [London] Spe-
ence Letters, v. 214, p. 115–127, doi: 10.1016/S0012-821X(03)00360-1. cial Publication 56, p. 41–61.
Chemenda, A.I., Yang, R.K., Hsieh, C.H., and Groholsky, A.L., 1997, Evolu- Kusznir, N.J., Roberts, A.M., and Morley, C.K., 1995, Forward and reverse
tionary model for the Taiwan collision based on physical modeling: Tecto- modeling of rift basin formation, in Lambiase, J.J., ed., Hydrocarbon
nophysics, v. 274, p. 253–274, doi: 10.1016/S0040-1951(97)00025-5. Habitat in Rift Basins: Geological Society [London] Special Publication
Chen, C.H., 1984, Determination of lower greenschist facies boundary by K- 80, p. 33–56.
mica-chlorite crystallinity in the Central Range, Taiwan: Proceedings of Lai, T.H., and Hsieh, M.L., 2003, Late-Quaternary vertical rock-movement
the Geological Society of China, v. 27, p. 41–53. rates of the coastal plains of Taiwan: Abstracts, Annual Meeting of Geo-
Chen, J., and Jahn, B.M., 1998, Crustal evolution of southeastern China: Nd and logical Society of China, p. 119.
Sr isotopic evidence: Tectonophysics, v. 284, p. 101–133, doi: 10.1016/ Lallemand, S., Front, Y., Bijwaard, H., and Kao, H., 2001, New insights on
S0040-1951(97)00186-8. 3-D plates interaction near Taiwan from tomography and tectonic
Chung, S.L., Wang, S.L., Shinjo, R., Lee, C.S., and Chen, C.H., 2000, Initia- implications: Tectonophysics, v. 335, p. 229–253, doi: 10.1016/S0040-
tion of arc magmatism in an embryonic continental rifting zone of the 1951(01)00071-3.
southernmost part of Okinawa Trough: Terra Nova, v. 12, p. 225–230, Lee, J.C., Chu, H.-T., Angelier, J., Chan, Y.-C., Hu, J.-C., Lu, C.-Y., and Rau,
doi: 10.1046/j.1365-3121.2000.00298.x. R.-J., 2002, Geometry and structure of northern surface ruptures of the
Clark, M.B., Fisher, D., and Lu, C.-Y., 1993, The Hsueshan Range of Taiwan: A 1999 Mw = 7.6 Chi-Chi, Taiwan earthquake: Influence from inherited
crustal scale pop-up structure: Tectonics, v. 12, p. 205–217. fold belt structures: Journal of Structural Geology, v. 24, p. 173–192, doi:
Clift, P.D., and Vannucchi, P., 2004, Controls on tectonic accretion versus 10.1016/S0191-8141(01)00056-6.
erosion in subduction zones: Implications for the origin and recycling Leeder, M.R., and Gawthorpe, R.L., 1987, Sedimentary models for extensional
of the continental crust: Reviews of Geophysics, v. 42, p. RG2001, doi: tilt-block/half-graben basins, in Coward, M.P., et al., eds., Continental
10.1029/2003RG000127. Extensional Tectonics: Geological Society [London] Special Publication
Clift, P.D., Schouten, H., and Draut, A.E., 2003, A general model of arc-conti- 28, p. 139–152.
nent collision and subduction polarity reversal from Taiwan and the Irish Letouzey, J., and Kimura, M., 1985, Okinawa Trough genesis; structure and
Caledonides, in Larter, R.D., and Leat, P.T., eds., Intra-Oceanic Subduc- evolution of a backarc basin developed in a continent: Marine and Petro-
tion Systems; Tectonic and Magmatic Processes: Geological Society leum Geology, v. 2, p. 111–130, doi: 10.1016/0264-8172(85)90002-9.
[London] Special Publication 219, p. 81–98. Lin, A.T., and Watts, A.B., 2002, Origin of the West Taiwan basin by orogenic
Crespi, J., Chan, Y.C., and Swaim, M., 1996, Synorogenic extension and loading and flexure of a rifted continental margin: Journal of Geophysical
exhumation of the Taiwan hinterland: Geology, v. 24, p. 247–250, doi: Research, v. 107, p. 2185, doi: 10.1029/2001JB000669.
10.1130/0091-7613(1996)024<0247:SEAEOT>2.3.CO;2. Lin, A.T., Watts, A.B., and Hesselbo, S.P., 2003, Cenozoic stratigraphy and sub-
Dadson, S., Hovius, N., Chen, H., Dade, W.B., Hsieh, M.L., Willett, S., Hu, sidence history of the South China Seamargin in the Taiwan region: Basin
J.C., Horng, M.J., Chen, M.C., Stark, C.P., Lague, D., and Lin, J.C., 2003, Research, v. 15, p. 453–478, doi: 10.1046/j.1365-2117.2003.00215.x.
Links between erosion, runoff variability and seismicity in the Taiwan Lin, C.C., 1969, Holocene geology of Taiwan: Acta Geologica Taiwanica, v.
orogen: Nature, v. 426, p. 648–651, doi: 10.1038/nature02150. 13, p. 83–126.
Davis, D.W., Sewell, R.J., and Campbell, S.D.G., 1997, U-Pb dating of Meso- Liu, T.K., Chen, Y.G., Chen, W.S., and Jiang, S.H., 2000, Rates of cooling and
zoic igneous rocks from Hong Kong: Geological Society [London] Jour- denudation of the Early Penglai Orogeny, Taiwan, as assessed by fission-
nal, v. 154, p. 1067–1076. track constraints: Tectonophysics, v. 320, p. 69–82, doi: 10.1016/S0040-
Dorsey, B.J., 1988, Provenance evolution and un-roofing history of modern 1951(00)00028-7.
arc-continent collision: Evidence from petrography of Plio-Pleistocene Liu, T.K., Hsieh, S., Chen, Y.G., and Chen, W.S., 2001, Thermo-kinematic evo-
sandstones, eastern Taiwan: Journal of Sedimentary Petrology, v. 58, lution of the Taiwan oblique-collision mountain belt as revealed by zircon
p. 208–218. fission track dating: Earth and Planetary Science Letters, v. 186, p. 45–56,
Fitzgerald, P.G., Stump, E., and Redfield, T.F., 1993, Late Cenozoic uplift of doi: 10.1016/S0012-821X(01)00232-1.
Denali and its relation to relative plate motion and fault morphology: Sci- Lo, C.H., and Yu, T.F., 1996, 40Ar/39Ar dating of high-pressure rocks in the
ence, v. 259, p. 497–499, doi: 10.1126/science.259.5094.497. Tananao basement complex, Taiwan: Journal of the Geological Society of
Galbraith, R.F., 1990, The radial plot: Graphical assessment of spread in ages: China, v. 39, p. 13–30.
Nuclear Tracks, v. 17, p. 207–214. Lu, C.-Y., and Hsü, K.-J., 1992, Tectonic evolution of the Taiwan mountain
Green, P.F., Duddy, I.R., Laslett, G.M., Hegarty, K.A., Gleadow, A.J., and Lov- belt: Petroleum Geology of Taiwan, v. 27, p. 21–46.
ering, J., 1989, Thermal annealing of fission-tracks in apatite; 4, Quantita- Lundberg, N., Reed, D.L., Liu, C.-S., and Lieske, J., 1997, Forearc-basin clo-
tive modeling techniques and extension to geological timescales: Chemi- sure and arc accretion in the submarine suture zone south of Taiwan: Tec-
cal Geology, v. 79, p. 155–182. tonophysics, v. 274, p. 5–23, doi: 10.1016/S0040-1951(96)00295-8.
Hawkins, J.W., 1974, Geology of the Lau Basin, a marginal sea behind the Malavieille, J., Lallemand, S.E., Dominguez, S., Deschamps, A., Lu, C.-Y., Liu,
Tonga arc, in Burk, C.A., and Drake, C.L., eds., The Geology of Conti- C.-S., and Schnurle, P., and the ACT scientific crew, 2002, Arc-continent
nental Margins: New York, Springer-Verlag, p. 505–520. collision in Taiwan: New marine observations and tectonic evolution, in
Hsiao, L.-Y., Lin, K.-A., Huang, S.T., and Teng, L.S., 1999, Structural charac- Byrne, T.B., and Liu, C.-S., eds., Geology and Geophysics of an Arc-Con-
teristics of the southern Taiwan-Sinzi folded zone: Petroleum Geology of tinent Collision, Taiwan: Geological Society of America Special Paper
Taiwan, v. 32, p. 133–153. 358, p. 187–211.
278 Clift et al.

Peng, T.-H., Li, Y.-H., and Wu, F.T., 1977, Tectonic uplift rates of the Taiwan Teng, L.S., 1996, Extensional collapse of the northern Taiwan mountain belt:
Island since the early Holocene, in Ho, C.S., A Special Volume Dedicated Geology, v. 24, p. 949–952, doi: 10.1130/0091-7613(1996)024<0949:
to Biq Chingchang on the Occasion of His Retirement: Taipei, Taiwan, ECOTNT>2.3.CO;2.
Geological Society of China Memoir 2, p. 57–69. Teng, L.S., and Lee, C.-T., 1996, Geodynamic appraisal of seismogenic faults
Platt, J.P., and Vissers, R.L.M., 1989, Extensional collapse of thickened conti- in northeast Taiwan: Proceedings of the Geological Society of China,
nental lithosphere: A working hypothesis for the Alboran sea and Gibraltar v. 39, p. 125–142.
arc: Geology, v. 17, p. 540–543, doi: 10.1130/0091-7613(1989)017<0540: Teng, L.S., Wang, Y., Tang, C.-H., Huang, C.-Y., Huang, T.-C., Yu, M.-S., and
ECOTCL>2.3.CO;2. Ke, A., 1991, Tectonic aspects of the Paleogene depositional basin of
Rau, R.-J., and Wu, F.T., 1995, Tomographic imaging of lithospheric structures northern Taiwan: Proceedings of the Geological Society of China, v. 34,
under Taiwan: Earth and Planetary Science Letters, v. 133, p. 517–532, p. 313–336.
doi: 10.1016/0012-821X(95)00076-O. Teng, L.S., Lee, C.-T., Tsai, Y.-B., and Hsiao, L.-Y., 2000, Slab breakoff as a
Reiners, P.W., Ehlers, T.A., Mitchell, S.G., and Montgomery, D.R., 2003, Cou- mechanism for flipping of subduction polarity in Taiwan: Geology, v. 28,
pled spatial variations in precipitation and long-term erosion rates across p. 155–158, doi: 10.1130/0091-7613(2000)28<155:SBAAMF>2.0.CO;2.
the Washington Cascades: Nature, v. 426, p. 645–647, doi: 10.1038/ Tippett, J.M., and Kamp, P.J.J., 1993, Fission-track analysis of the Late Ceno-
nature02111. zoic vertical kinematics of continental Pacific crust, South Island, New
Roberts, A.M., Yielding, G., Kusznir, N.J., Walker, I., and Dorn-Lopez, D., Zealand: Journal of Geophysical Research, v. 98, p. 16,119–16,148.
1993, Mesozoic extension in the North Sea: Constraints from flexural Vita-Finzi, C., and Lin, J.C., 1998, Serial reverse and strike slip on imbricate
backstripping, forward modeling and fault populations, in Parker, J.R., faults: The Coastal Range of east Taiwan: Geology, v. 26, p. 279–282,
ed., Petroleum Geology of NW Europe: Geological Society [London], doi: 10.1130/0091-7613(1998)026<0279:SRASSO>2.3.CO;2.
Proceedings of the 4th Conference, p. 1123–1136. von Huene, R., and Scholl, D.W., 1991, Observations at convergent margins
Rudnick, R.L., and Fountain, D.M., 1995, Nature and composition of the con- concerning sediment subduction and the growth of continental crust:
tinental crust; a lower crustal perspective: Reviews of Geophysics, v. 33, Reviews of Geophysics, v. 29, p. 279–316.
p. 267–309, doi: 10.1029/95RG01302. Waldhauser, F., 2001, HypoDD: A computer program to compute double-dif-
Shinjo, R., 1999, Geochemistry of high Mg andesites and the tectonic evolution ference hypocenter locations: U.S. Geological Survey Open-File Report
of the Okinawa Trough–Ryukyu Arc system: Chemical Geology, v. 157, 01-113 (Menlo Park, California).
p. 69–88, doi: 10.1016/S0009-2541(98)00199-5. Waldhauser, F., and Ellsworth, W.L., 2000, A double-difference earthquake
Shyu, J.B.H., Sieh, K., Chen, Y.G., and Liu, C.S., 2005, Neotectonic architec- location algorithm: Method and application to the northern Hayward
ture of Taiwan and its implications for future large earthquakes: Journal of fault: Bulletin of the Seismological Society of America, v. 90, p. 1353–
Geophysical Research, v. 110, p. B08402, doi: 10.1029/2004JB003251. 1368, doi: 10.1785/0120000006.
Sibuet, J.-C., and Hsu, S.K., 1997, Geodynamics of the Taiwan arc–arc Wang, K.L., Chung, S.L., Chen, C.-H., Shinjo, R., Yang, T.F., and Chen, C.H.,
collision: Tectonophysics, v. 274, p. 221–251, doi: 10.1016/S0040- 1999, Post-collisional magmatism around northern Taiwan and its relation
1951(96)00305-8. with opening of the Okinawa Trough: Tectonophysics, v. 308, p. 363–376,
Sibuet, J.-C., Deffontaines, B., Hsu, S.K., Thareau, N., Le Formal, J.P., and Liu, doi: 10.1016/S0040-1951(99)00111-0.
C.S., 1998, Okinawa Trough backarc basin; early tectonic and magmatic Wang, P.L., Lin, L.H., and Lo, C.H., 1998, 40Ar/39Ar dating of mylonitiza-
evolution: Journal of Geophysical Research, v. 103, p. 30,245–30,267, tion in the Tananao schist, eastern Taiwan: Geological Society of China
doi: 10.1029/98JB01823. Journal, v. 41, p. 159–183.
Sibuet, J.-C., Hsu, S.K., Le Pichon, X., le Formal, J.P., Reed, D., Moore, G., and Willett, S.D., Fisher, D., Fuller, C., Yeh, E.-C., and Lu, C.-Y., 2003, Erosion
Liu, C., 2002, East Asia plate tectonics since 15 Ma: Constraints from the rates and orogenic-wedge kinematics in Taiwan inferred from fission-
Taiwan region: Tectonophysics, v. 344, p. 103–134, doi: 10.1016/S0040- track thermochronometry: Geology, v. 31, p. 945–948, doi: 10.1130/
1951(01)00202-5. G19702.1.
Suppe, J., 1981, Mechanics of mountain building and metamorphism in Tai- Wu, F.T., 1978, Recent tectonics in Taiwan: Journal of Physics of the Earth,
wan, in Ho, C.S., and Ernst, W.G., Papers Presented at the ROC-USA v. 26, p. S265–S299.
Seminar on Plate Tectonics and Metamorphic Geology: Taipei, Taiwan, Wu, F.T., Rau, R.J., and Salzberg, D.H., 1997, Taiwan tectonics: Thin-skinned or
Geological Society of China Memoir 4, p. 67–89. lithospheric collision: Tectonophysics, v. 274, p. 191–220, doi: 10.1016/
Suppe, J., 1984, Kinematics of arc-continent collision, flipping of subduction, S0040-1951(96)00304-6.
and backarc spreading near Taiwan, in Tsan, S.F., et al., eds., A Special Yu, H.-S., and Chou, Y.-W., 2001, Characteristics and development of the
Volume Dedicated to Chun-Sun Ho on the Occasion of His Retirement: flexural forebulge and basal unconformity of western Taiwan fore-
Geological Society of China Memoir 6, p. 21–33. land basin: Tectonophysics, v. 333, p. 277–291, doi: 10.1016/S0040-
Suppe, J., Wang, Y., Liou, J.G., and Ernst, W.G., 1976, Observation of some 1951(00)00279-1.
contacts between basement and Cenozoic cover in the Central Range, Tai- Zeitler, P.K., Chamberlain, C.P., and Smith, H.A., 1993, Synchronous anatexis,
wan: Proceedings of the Geological Society of China, v. 19, p. 59–70. metamorphism and rapid denudation at Nanga Parbat (Pakistan Himalaya):
Tagami, T., Galbraith, R.F., Yamada, G.M., and Laslett, G.M., 1998, Revised Geology, v. 21, p. 347–350, doi: 10.1130/0091-7613(1993)021<0347:
annealing kinetics of fission-tracks in zircon and geological implications, SAMARD>2.3.CO;2.
in Van den Haute, P., and De Corte, F., eds., Advances in Fission-track
Geochronology: Amsterdam, Kluwer Academic Press, p. 99–112.
Teng, L.S., 1990, Geotectonic evolution of late Cenozoic arc-continent col-
lision in Taiwan: Tectonophysics, v. 183, p. 57–76, doi: 10.1016/0040-
1951(90)90188-E. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

The Guerrero Composite Terrane of western Mexico: Collision and subsequent rifting in
a supra-subduction zone

E. Centeno-García*
Instituto de Geología, Universidad Nacional Autónoma de México, Ciudad Universitaria, México D.F. 04510, México

M. Guerrero-Suastegui
O. Talavera-Mendoza
Unidad Académica de Ciencias de la Tierra, Universidad Autónoma de Guerrero, AP 197, Taxco el Viejo, Guerrero, México

ABSTRACT

The Guerrero Composite Terrane of western Mexico is the second largest ter-
rane in North America. Mostly characterized by submarine volcanism and formed
by five terranes, the Guerrero records vast and complex subduction-related processes
influenced by major translation and rifting. It is composed of the Teloloapan, Gua-
najuato, Arcelia, Tahue, and Zihuatanejo Terranes. The Teloloapan Terrane is made
up of Lower Cretaceous island-arc (IA) andesitic to basaltic submarine lava flows,
interbedded with limestone and shallow-marine volcaniclastic rocks. The Guanajuato
and Arcelia Terranes are characterized by Lower Cretaceous supra-subduction
ophiolite successions formed by deep-marine volcanic and sedimentary rocks with
mid-oceanic-ridge basalt (MORB), oceanic-island basalt (OIB), and island-arc basalt
(IAB) signatures. These two terranes are placed between the continent and the more
evolved arc assemblages of the Zihuatanejo Terrane. The Tahue Terrane is composed
of Paleozoic accreted arc and eugeoclinal sedimentary rocks, Triassic rift-related
metaigneous rocks, and overlain unconformably by pillow basalts, limestone, and
volcaniclastic rocks. The Zihuatanejo Terrane was formed by Triassic ocean-flank to
ocean-floor assemblages accreted in Early Jurassic time (subduction complexes). The
subduction complexes are overlain by Middle Jurassic–evolved volcanic arc rocks,
which are in turn unconformably overlain by Early and Late Cretaceous subaerial
and marine arc-related volcano-sedimentary assemblages.
Mesozoic stratigraphy at the paleocontinental margin of Mexico (Oaxaquia and
Mixteca Terranes) is formed by Triassic submarine fan turbidites accreted during
Early Jurassic time; Middle Jurassic–evolved volcanic arc rocks are unconformably
covered by a Late Jurassic to Cretaceous calcareous platform.
Six stages in the tectonic evolution are proposed on the basis of the stratigraphic
and deformational events recorded in western Mexico: (1) A passive or rifting margin
developed along the western margin of continental Mexico throughout the Triassic. A

*centeno@servidor.unam.mx

Centeno-García, E., Guerrero-Suastegui, M., and Talavera-Mendoza, O., 2008, The Guerrero Composite Terrane of western Mexico: Collision and subsequent rift-
ing in a supra-subduction zone, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones:
Geological Society of America Special Paper 436, p. 279–308, doi: 10.1130/2008.2436(13). For permission to copy, contact editing@geosociety.org. ©2008 The
Geological Society of America. All rights reserved.

279
280 Centeno-García et al.

thick siliciclastic turbiditic succession of the Potosi Submarine Fan was accumulated
on the paleo-continental shelf-slope and extended to the west in a marginal oceanic
basin. (2) Subduction began in the Early Jurassic, and the turbidites of the Potosi
Fan with slivers of the oceanic crust were accreted, forming a wide subduction prism.
(3) Exhumation of the accretionary prism and development of a Middle Jurassic con-
tinental arc onto the paleo-continental margin (Oaxaquia and Mixteca Terrane) took
place, and also in the Zihuatanejo Terrane. (4) Intra-arc strike-slip faulting and rift-
ing of the Middle Jurassic continental arc took place along with migration of the
subduction toward the west and development of a calcareous platform in Oaxaquia
and the Mixteca Terrane (continental Mexico). (5) Drifting of the previously accreted
Tahue and Zihuatanejo Terranes formed a series of marginal arc-backarc systems,
or one continuously drifting arc with intra-arc and backarc basins during Early to
middle Cretaceous time. (6) Deformation of the arc assemblages, and development of
Santonian to Maastrichtian foreland and other basins, date the final amalgamation of
the Guerrero Composite Terrane with the continental margin.

Keywords: Guerrero Terrane, Mexico, tectonics, Triassic subduction complex, Creta-


ceous arc volcanism.

INTRODUCTION suggested that the Guerrero Composite Terrane might represent


one or more complex systems of two or three peripheral arcs that
The present configuration of continental Mexico was built developed relatively close to the continent (Campa and Ramírez,
after accretion of basement remnants and oceanic terranes. Dur- 1979; Ramírez-Espinosa et al., 1991; Mendoza and Suastegui,
ing most of their Mesozoic history, Proterozoic to Paleozoic 2000; Centeno-García et al., 2003; Centeno-García, 2005). Some
accreted terranes formed a relatively narrow neck of land adja- models even proposed that the arc was autochthonous and was
cent to the North American craton. This was bordered on its built upon Proterozoic continental crust of nuclear Mexico (de
eastern side by rifting and on its western side by active subduc- Cserna, 1978; Elías-Herrera and Sánchez-Zavala, 1990). In other
tion. Thus Mexico is probably one of the most suitable regions words, there is a model for each likely possibility, but each lacks
in North America for studying the interaction between these two strong supporting evidence.
differing tectonic scenarios. We suggest in this paper, based on New findings on the stratigraphy, discussed in this paper,
evidence recorded in the stratigraphy of the Guerrero Composite suggest a more complex evolution, implying a series of accre-
Terrane and surrounding terranes, that the almost continuously tions to the continent followed by rifting, and later by collision.
subducting Pacific margin of Mexico was directly influenced by In this paper we attempt to present our insights into the evolu-
extensional tectonics associated with the breakup of Pangea and tion of western Mexico gained from examining the stratigraphy
the formation of the Gulf of Mexico. and structure, and the geochemical and geochronological data, of
The Guerrero Composite Terrane (Campa and Coney, 1983) such a vast area. However, we discuss in this paper only strati-
constitutes approximately one-third of Mexico. As originally graphic units and localities that are keys for reconstructing the
described, it is the largest of all the Mexican terranes and probably tectonic evolution. This paper synthesizes the work done by
the second largest of the North America Cordillera after Wrangel- many authors. Although there is the need for more geochrono-
lia (Campa and Coney, 1983; Centeno-García et al., 1993a). The logical and detailed field work, we consider that the preliminary
Guerrero Composite Terrane is characterized mostly by subma- tectonic model presented in this paper is consistent with the evi-
rine and locally subaerial volcanic and sedimentary successions dence collected to date.
that range in age from Jurassic (Tithonian) to middle–Late Creta-
ceous (Cenomanian), and scarce exposures of older rocks. A wide OVERVIEW OF THE GUERRERO COMPOSITE AND
variety of models has been proposed for the origin of the Guerrero NEIGHBORING TERRANES
Composite Terrane. Like other terranes of the North America Cor-
dillera, it was first interpreted as an exotic terrane formed by a far- The stratigraphy of western Mexico is synthesized in this
traveled Cretaceous oceanic arc. Some authors have suggested that paper under the framework of tectono-stratigraphic terranes,
it was an oceanic arc terrane that was accreted to nuclear Mexico which are regions that share the same geological history and
in Late Cretaceous time via a westward-dipping subduction zone are bounded by major faults. As mentioned before, by the early
that closed a major ocean basin (Lapierre et. al., 1992; Tardy et Mesozoic, the Paleozoic and Proterozoic terranes were already
al., 1994; Dickinson and Lawton, 2001, etc.). Other authors have accreted to the southern part of the North American craton.
Guerrero Composite Terrane of western Mexico 281

Those that already formed part of the continental margin dur- al., 1995; Ramírez-Ramírez, 1992; Lawlor et al., 1999; Solari
ing the Mesozoic were Oaxaquia and the Mixteca, Parral, and et al., 2003; Keppie et al., 2003). It is covered by Paleozoic
Cortes Terranes (Fig. 1). Terranes accreted or displaced during sedimentary rocks (Fig. 2) that are capped by Permian volca-
the Mesozoic were those of the Guerrero Composite, the Central, nic and volcaniclastic rocks (McKee et al., 1999; Stewart et al.,
as well as terranes of the western Baja California Peninsula. The 1999; Rosales-Lagarde et al., 2005). Triassic (Carnian–Norian)
latter will not be reviewed in this paper. A brief summary of the sedimentary rocks (La Ballena Formation) are exposed at the
stratigraphy is described as follows; more detailed descriptions of western margin of Oaxaquia (Labarthe et al., 1982; Silva-
key areas and events are discussed later. Romo, 1993; Tristán-Gonzalez and Torres-Hernández, 1994;
Centeno-García and Silva-Romo, 1997; Barboza-Gudiño et al.,
OAXAQUIA 1998, 1999, 2004; Bartolini et al., 2002). These rocks are made
up of a thick succession of turbidites (Fig. 2) deposited in a
At the end of the Paleozoic, Proterozoic basement terranes submarine fan environment named the Potosi Fan (Centeno-
of Gondwanan affinity were already accreted to the southern part García, 2005).
of the North American craton. The largest of these is the Oaxa- Triassic rocks of the Potosi Fan were deformed prior to deposi-
quia block (Fig. 1), a crustal fragment, subcontinent in size, of tion of Jurassic volcanic-volcaniclastic rocks (Centeno-García and
Grenville affinity (Ortega-Gutiérrez et al., 1995). This crustal Silva-Romo, 1997). They are interpreted as a Jurassic continental
block forms the backbone of eastern Mexico and is referred to arc and rest unconformably on the Triassic Potosi Fan. Jurassic
herein as continental Mexico for the Mesozoic. Oaxaquia has arc strata are made up of subaerial andesitic-rhyolitic lava flows,
a Precambrian (1157–900 Ma) crystalline basement (gneisses interbedded with volcaniclastic rocks (Silva-Romo, 1993). The
and anorthosites; Patchett and Ruíz, 1987; Ortega-Gutiérrez et arc sequence changes transitionally upsection to shallow-marine

115° 110° 105° 100° 95° 90°

USA Cenozoic volcanism


C
a
Te b o M ÉXICO Oaxaquia
Mo
r r rc j av North Cortes Terrane
an a e-
30° e So America 30°
no Parral Terrane
ra La
Ba
bia Central Terrane
Me

Cortes Fa
O ult Mixteca Terrane
ga

Terrane Sa a x
sh

n M aq
uia Guerrero Composite Terrane
ea

arc
r

os
Fa
ult Tahue Terrane
Arcelia Terrane N
Mojave-Sonora
Megashear Zihuatanejo 25°
25°
G

Early Mesozoic Co

Teloloapan
ue

Central
Fol

Fig. 6
Terrane Guanajuato
rr

d
and
er
o

O
Thr
C

ax
ust

.4
om

g aq
Fi
Bel
po

ui
nt

GULF OF
t Fr
in
PA

si

a
en

on
te

ta

MÉXICO
l
C

t
M
Te

ar
IF

gi
rr

20°
IC

20°
an
e
O
C

Mixteca
E

Fig. 8
A
N

A
IC

AL
ÉX

EM
M

T
UA
G

115° 110° 105° 100° 95° 90°

Figure 1. Map showing main tectono-stratigraphic terranes, major faults mentioned in the text, and locations of Figures 4, 6, and 8.
282 Centeno-García et al.

GUERRERO COMPOSITE TERRANE

Guanajuato
Teloloapan

Oaxaquia
Caborca
Mixteca

Central
Arcelia

Cortes
Parral
Tahue
Zihuatanejo
Maastrichtian
Coast Huetamo Northern basic-andesitic
Campanian submarine volcanism
Santonian andesitic
Coniacian submarine volcanism
Turonian
❖ rhyolitic-andesitic
submarine volcanism
Cenomanian
? rhyolitic-andesitic
Albian ❖ continental volcanism
Aptian
volcaniclastic rocks
terrestrial and shallow
Barremian
marine sedimentary
? ? ●

rocks
Hauterivian siliciclastic and
Valanginian ●
volcaniclastic rocks
Berriasian limestone, shale
● and evaporites
Tithonian ●

siliciclastic turbidites
Kimmeridgian
● ●
?
Oxfordian ● accretionary prism
Callovian ●
Bathonian metamorphic arc
Bajocian volcanic and
Aalenian ● sedimentary rocks
Toarcian gneiss and schist
Pliensbachian

Sinemurian

Hettangian
Rhaetian

Norian

Carnian
Ladinian
Early Triassic

Paleozoic
Proterozoic

Figure 2. Simplified stratigraphic columns for Oaxaquia and terranes mentioned in the text. They show the age range (in Ma) of sedimentation
and magmatism for western and central Mexico. Geochronological data are represented as follows: Black circles are U/Pb ages, and diamonds
are Ar/Ar and K/Ar ages.

volcaniclastic rocks, limestone, and some evaporites (Fig. 2; American seaway. A major change upsection from calcareous to
Silva-Romo, 1993; Tristán-González and Torres-Hernández, clastic sedimentation occurred at the uppermost part of the Cre-
1994; Barboza-Gudiño et al., 2004). Calcareous sedimentation taceous, forming a thick succession of sandstone, shale, and con-
in Oaxaquia ranges in age from late Oxfordian–Kimmeridgian to glomerate (Caracol Formation; Silva-Romo, 1993). Oaxaquia is
Turonian and is interpreted as the southern extension of the North overthrust by the Guerrero Composite Terrane (Fig. 1).
Guerrero Composite Terrane of western Mexico 283

MIXTECA TERRANE Paleozoic “miogeosyncline” of Nevada. It was transferred toward


the south by Middle to Late Jurassic time via the Mohave-Sonora
The basement of the eastern Mixteca Terrane is made up of megashear (Anderson and Silver, 1979, 2005; Stewart et al.,
pre-Mississippian polydeformed metamorphic rocks of the Acat- 1990). The Cortes Terrane is interpreted as an autochthonous ter-
lán Complex (Ortega-Gutiérrez, 1981; Ruíz et al., 1988; Yañez et rane to North America, which probably evolved at the margin of
al., 1991). This complex is considered to be the result of complex the Caborca Terrane (Stewart et al., 1990). It is made up of a thick
interactions between Gondwana and Laurentia previous and dur- succession of Paleozoic deep-marine turbidites that were thrust
ing the assembling of Pangea (Ortega-Gutiérrez et al., 1999). It over platform limestone of the Caborca Terrane (Figs. 1 and 2).
is unconformably overlain by Permian sedimentary rocks, which The Cortes Terrane is interpreted as continental-slope depos-
are in turn overlain unconformably by Middle Jurassic volcanic its, and it is considered the southern extension of the Paleozoic
and sedimentary rocks (Fig. 2; García-Díaz et al., 2004). At the Cordilleran “eugeoclinal” deposits from Nevada and California
western part of the terrane, near the limit with the Guerrero Com- (Poole and Madrid, 1988; Coney and Campa, 1987; Stewart et al.,
posite Terrane, partly metamorphosed volcanic and volcaniclastic 1990). The previously deformed Paleozoic deep-marine rocks of
rocks are exposed (Taxco Schist and Chapolapa Formation; de the Cortes Terrane are overlain by Triassic (Carnian–Norian) ter-
Cserna and Fries, 1981; Talavera-Mendoza, 1993; Campa and Iri- restrial and marine sedimentary rocks (Stewart et al., 1990; Stew-
ondo, 2004). The Taxco Schist is made up of andesitic to rhyolitic art and Roldán-Quintana, 1991). The Triassic rocks are overlain
lavas and volcaniclastic rocks of Early Cretaceous age (Talavera- by Cretaceous red beds and volcanic rocks (Stewart and Roldán-
Mendoza, 1993; Campa and Iriondo, 2004). The Taxco Schist Quintana, 1991). Contact relationships between the Cortes and
is unconformably overlain by a thick limestone succession of Guerrero Composite Terranes have not been well constrained,
Albian to Cenomanian age and by Turonian–Maastrichtian clastic but the contact is inferred to be a Late Cretaceous thrust fault.
rocks (Mexcala Formation; Campa and Ramírez, 1979; Talavera-
Mendoza et al., 1995). Contacts between the Mixteca Terrane and CENTRAL TERRANE
Oaxaquia, as well as between the Mixteca and Guerrero Com-
posite Terranes, are partially exposed. The Mixteca Terrane is on The nature of the basement of the Central Terrane is unknown,
strike-slip fault contact with Oaxaquia, and rocks of the Guerrero but it is assumed to be different from the Proterozoic basement of
Composite Terrane are thrust over the Mixteca Terrane. Oaxaquia because its oldest exposed rocks near its contact are
a subduction-related accretionary complex (Taray Formation;
PARRAL TERRANE Anderson et al., 1990; Diaz-Salgado et al., 2003; Anderson et al.,
2005; Centeno-García, 2005). The subduction zone on which the
The Parral Terrane (Figs. 1 and 2) was first defined by Taray Formation was deformed was probably constructed along
Pacheco et al. (1984) and Coney and Campa (1987) and was rede- the Oaxaquia continental margin between Late Permian and Early
fined by Centeno-García (2005). The basement of the Parral Ter- Jurassic time (Diaz-Salgado et al., 2003; Anderson et al., 2005).
rane is formed by Devonian to Carboniferous metamorphic rocks The complex is unconformably overlain by Oxfordian subaerial
(Pescadito Schist; Eguiluz and Campa, 1982; Araujo and Arenas, rhyolitic to andesitic volcanic rocks and red beds (Jones et al.,
1986; Zaldivar and Garduño, 1984). These Paleozoic metamor- 1995). These rocks change transitionally to shallow-marine lime-
phic rocks are unconformably overlain by red beds and volca- stone that ranges in age from Late Jurassic to Late Cretaceous
nic successions (Nazas Formation; Pantoja-Alor, 1963), which (Córdoba-Méndez, 1964). The location of the northern and east-
change transitionally to Tithonian limestone (Araujo and Arenas, ern contact between the Central Terrane and Oaxaquia is inferred
1986; Contreras-Montero et al., 1988). Cretaceous calcareous on the basis of the location of the last exposures of Paleozoic–
and clastic sedimentation of the Parral Terrane is laterally con- early Mesozoic rocks, and a contrast in deformation styles of Cre-
tinuous with the calcareous-clastic deposits that cover Oaxaquia taceous rocks in both (Fig. 1). The contact between the Central
and the Central Terrane. Relationships among the Parral, Cen- and Guerrero Composite Terranes has not been studied in detail
tral, and Cortes Terranes and the Parral Terrane and Oaxaquia are but is inferred on the basis of the distribution of the northernmost
unknown, because the contacts are covered by Cretaceous lime- exposures of Cretaceous marine volcanic rocks that belong to the
stone or by Cenozoic volcanic successions. Therefore, the exact Guerrero Composite Terrane. Structural trends on both sides of
locations of their boundaries are unknown but are inferred by the the contact suggest that the Central Terrane is overthrust by the
difference in styles of deformation of the Cretaceous rocks. Guerrero Composite Terrane to the south (Fig. 1). The thrusting is
inferred to have occurred about Late Cretaceous time.
CABORCA AND CORTES TERRANES
GUERRERO COMPOSITE TERRANE
The Caborca Terrane has a Proterozoic basement older than
1.7 Ga (Anderson and Silver, 1981), covered by a thick Paleozoic Areas with large volumes of Lower Cretaceous volcanic
sedimentary succession. It has been interpreted to be a displaced and volcaniclastic rocks, located toward the west of Oaxa-
block of continental North America, originally located along the quia and the Mixteca Terrane, were originally grouped as the
284 Centeno-García et al.

Guerrero Terrane by Campa and Coney (1983) and were, 10 up of large volumes of Triassic (Norian) quartz-rich turbidites
years later, divided into the Tahue, Nahuatl, and Tepehuano (sandstone and shale) that are tectonically imbricated (Campa
Terranes by Sedlock et al. (1993). Subsequent regional map- et al., 1982; Centeno-García et al., 1993a, 1993b). The turbi-
ping has shown that the divisions proposed by Campa and dites form a matrix within which are blocks and slabs of pillow
Coney (1983) are closer to the field locations of faults delim- basalts, diabase, banded gabbros, chert, and limestone (Fig. 2).
iting the terranes than those of Sedlock et al. (1993). There- These rocks have received different names at different outcrops:
fore, more recent reviews of the terrane distribution of Mexico Zacatecas Formation, Arteaga Complex, and Las Ollas Complex
(e.g., Centeno-García, 2005) have been based on Campa and (Burckhardt and Scalia, 1906; Ranson et al., 1982; Cuevas-Pérez,
Coney (1983). The Guerrero is a composite terrane, formed by 1983; Monod and Calvet, 1991; Centeno-García and Silva-
at least five terranes: Tahue, Zihuatanejo, Guanajuato, Arcelia, Romo, 1997; Talavera-Mendoza, 2000; Centeno-García et al.,
and Teloloapan (Figs. 1 and 2; Talavera-Mendoza et al., 1995; 2003). The deformation of these rocks varies from gently folded
Mendoza and Suastegui, 2000; Centeno-García et al., 2003; strata to highly sheared block-in-matrix textures, and their meta-
Centeno-García, 2005). Their stratigraphy is briefly described morphism ranges from none to high-greenschist–amphibolite
from NNW to ESE (Fig. 1): facies (Centeno-García et al., 2003). Blueschist facies have been
reported only in one locality (Las Ollas Complex; Talavera-
Tahue Terrane Mendoza, 2000). These lithologies are interpreted to constitute
an Upper Triassic(?)–Lower Jurassic subduction-related accre-
The Tahue Terrane contains the oldest rocks found so far tionary complex.
within the Guerrero Composite Terrane (Fig. 2; Centeno-García, Scattered exposures of rocks of Middle to Late Jurassic–
2005). These rocks comprise Ordovician marine rhyolitic- evolved arc volcanism lie along the Pacific Coast of the Zihua-
andesitic lavas and clastic and calcareous rocks, all deformed tanejo Terrane. These rocks are made up of submarine rhyo-
and metamorphosed to low-greenschist facies (El Fuerte Com- litic lavas and volcaniclastic rocks, and granitoids that were
plex; Mullan, 1978; Roldán-Quintana et al., 1993; Poole and emplaced in rocks of the accretionary complex (Bissig et al.,
Perry, 1998). These rocks may have originated as an oceanic 2003; Centeno-García et al., 2003). The Middle to Upper Juras-
arc that apparently was accreted previous to the deposition of sic arc rocks were in turn deformed and exhumed previous to the
Pennsylvanian–Permian deep-marine sedimentary rocks (San deposition of uppermost Jurassic–Cretaceous arc-related strata
José de Gracia Formation; Carrillo-Martínez, 1971; Gastil et (Centeno-García et al., 2003).
al., 1991; Arredondo-Guerrero and Centeno-García, 2003; The Cretaceous arc succession ranges from Berriasian to
Centeno-García, 2005). These deep-marine turbidites are strongly Cenomanian in age, and it includes andesitic, basaltic, and some
deformed but do not show the metamorphism of the El Fuerte rhyolitic volcanic and volcaniclastic rocks, interbedded with
Complex; thus an unconformable contact relationship between limestone, evaporites, and some red beds (Grajales and López,
these two units is inferred. Paleozoic rocks of the Tahue Terrane 1984). The arc succession contains abundant fossils such as rud-
are unconformably overlain by Cretaceous marine arc volcanic ists, gastropods, microfossils, fossil logs, and vertebrates.
rocks and are interpreted as part of the Guerrero Arc (Ortega- This arc succession was deformed prior to the intrusion of
Gutiérrez et al., 1979; Henry and Fredrikson, 1987; Roldán- large granitoids of latest Cretaceous to Paleogene age (Schaaf
Quintana et al., 1993; Freydier et al., 1995). These rocks are also et al., 2000). Also, uppermost Cretaceous (Santonian to Maas-
cut by mafic and ultramafic intrusions that are part of the same trichtian) red beds and volcanic rocks rest unconformably on all
Cretaceous arc magmatism (Henry and Fredrikson, 1987; Gas- previous units (Altamira Areyán, 2002; Benammi et al., 2005).
til et al., 1999; Arredondo-Guerrero and Centeno-García, 2003). The contact between the Zihuatanejo Terrane and Oaxaquia is
Therefore, the Paleozoic units form the basement upon which exposed at its northern limit, where Cretaceous arc rocks of the
the arc was built. The Tahue Terrane also contains metamorphic Zihuatanejo Terrane are thrust over shallow-marine limestone of
rocks of Triassic age (Keppie et al., 2006). The contact relation- Oaxaquia. Its contact with the Arcelia and Guanajuato Terranes
ship between the Cortes and Tahue Terranes has not been studied is inferred to be an east-verging thrust, but it is covered by upper-
in detail, but it is inferred to be a thrust (Fig. 1; Roldán-Quintana most Cretaceous and Cenozoic red beds and volcanic rocks.
et al., 1993). The contact between the Tahue and Zihuatanejo Ter-
ranes is not exposed. Guanajuato Terrane

Zihuatanejo Terrane The Guanajuato Terrane has been interpreted as a com-


plete crustal section through a primitive island arc that appears
The Zihuatanejo Terrane is the largest of all terranes that to lack an older basement (Ortiz-Hernandez et al., 1991; Ortiz-
form the Guerrero Composite Terrane (Fig. 1). It extends north Hernandez, 1992). It has also been interpreted as the remains of
of the Mexican Volcanic Belt and along the Pacific Coast of Mex- an oceanic basin that lay between the Guerrero arc and the conti-
ico (Centeno-García et al., 1993a, 1993b; Talavera-Mendoza et nental margin (Freydier et al., 2000). This terrane was formed by
al., 1995; Mendoza and Suastegui, 2000). Its basement is made a series of tectonic slivers that placed lower crust rocks (gabbro,
Guerrero Composite Terrane of western Mexico 285

tonalite, serpentinite, wehrlite, and dike swarms) on pillow carbonates or Upper Cretaceous clastic sediments that belong
basalts, rhyolitic tuffs, volcanic turbidites, chert, and black detri- to the Mixteca Terrane (Fig. 1; Campa and Ramirez, 1979). The
tal limestone (Quintero-Legorreta, 1992; Ortiz-Hernandez et al., nature of its basement remains unknown. Metamorphic rocks
1992; Lapierre et al., 1992; Monod et al., 1990; Martínez-Reyes, that are exposed near the northwestern boundary of the Teloloa-
1992; Ortiz-Hernandez et al., 2003). These rocks were poorly pan Terrane with the Arcelia Terrane have been interpreted as a
dated as Tithonian–Hauterivian in age (Ortiz-Hernandez et al., possible basement for the former (Elías-Herrera and Sánchez-
2003; Hall and Mortensen, 2003). Previously deformed volca- Zavala, 1990; Sanchez-Zavala, 1993). The rocks in this area are
nic turbidites are unconformably overlain by Aptian–Albian of uncertain age and origin.
limestone (Ortiz-Hernandez et al., 2003). This suggests that
sedimentation and at least one phase of deformation occurred TECTONIC MODEL
previous to the Aptian–Albian (Ortiz-Hernandez et al., 2003).
At present the Guanajuato Terrane is thrust over the calcareous The most abundant rocks of the Guerrero Composite Terrane
platform of Oaxaquia (Ortiz-Hernández et al., 2002). Contact are marine, and rarely subaerial, arc volcanic and sedimentary
relationships between the Guanajuato and Zihuatanejo Terranes successions that range in age from latest Jurassic (Tithonian) to
have not been constrained. middle Late Cretaceous (Cenomanian). The composition of the
few scattered exposures of older units suggests a complex earlier
Arcelia Terrane tectonic evolution. These older rocks were not taken into con-
sideration for the tectonic models proposed by previous authors
The Arcelia Terrane is made up of basaltic pillow lavas and (de Cserna, 1978; Campa and Ramirez, 1979; Elías-Herrera and
ultramafic bodies, black shale and chert, and volcanic turbidites, Sánchez-Zavala, 1990; Tardy et al., 1994; Lapierre, et. al., 1992;
all intensively deformed and partly metamorphosed (Ramírez- Dickinson and Lawton, 2001, etc.). Based on the available infor-
Espinosa et al., 1991; Talavera-Mendoza et al., 1995). It is char- mation, we identified six main tectonic stages in the evolution of
acterized by Early Cretaceous deep-marine primitive arc or arc- the Guerrero Composite Terrane. These stages are represented in
related oceanic facies and shows the least evolved magmatism Figure 3 and are briefly described in this section. Detailed discus-
of all the arc successions of the Guerrero Composite Terrane sion of the data that support the reconstruction of each stage is
(Talavera-Mendoza et al., 1995; Mendoza and Suastegui, 2000). presented in the following section.
The Arcelia Terrane appears to lack an older basement (Talavera-
Mendoza et al., 1995; Mendoza and Suastegui, 2000). Rocks of Stage I: Collision of a Paleozoic Oceanic Arc?—Basement
the Arcelia Terrane apparently were thrust over the assemblages of the Tahue Terrane
of the Teloloapan Terrane, and were in turn overthrust by rocks
of the Zihuatanejo Terrane. However, these contacts are inferred The basement of the Tahue Terrane (Fig. 3) is composed of
because they are covered by younger red beds. the early Paleozoic accreted volcanic-sedimentary rocks of the
El Fuerte Metamorphic Complex. There are not enough data
Teloloapan Terrane available to constrain the origin of this complex. Preliminary
interpretations considered these rocks as remnants of Gondwa-
The Teloloapan Terrane consists of two distinct regions: the nan crust accreted during the formation of Pangea (Poole et al.,
eastern region is characterized by shallow-marine volcanic and 2005). In this model, metamorphic rocks of El Fuerte could be
sedimentary deposits (Fig. 2), and the western region by deeper the western continuation of basement rocks of the Parral Terrane
volcanic and sedimentary facies (Guerrero-Suastegui et al., 1991; (Figs. 1 and 2). An alternative interpretation is that the El Fuerte
Ramírez-Espinoza et al., 1991; Talavera-Mendoza et al., 1995; Complex may be a displaced fragment of the early Paleozoic arc
Mendoza and Suastegui, 2000; Guerrero-Suastegui, 2004). Both (Antler Arc) that collided with the western continental margin of
are marine arc assemblages, which vary in composition from North America during late Paleozoic time (Burchfiel et al., 1992;
basalt-andesite to scarce dacite-rhyolite (Talavera-Mendoza Sánchez-Zavala et al., 1999; Dickinson, 2004; Centeno-García,
et al., 1995). This unit contains microfossils (radiolarians and 2005). Carboniferous deep-marine turbidites (San José de Gracia
coccoliths), gastropods, and bivalves that range in age from Formation) that apparently cover the lower Paleozoic arc rocks
Hauterivian to Aptian; these rocks change transitionally upsec- unconformably may be correlative with deep-marine sedimen-
tion to Aptian–Albian island-arc carbonates (Guerrero-Suastegui tary rocks exposed in the eastern peninsular ranges of Baja Cali-
et al., 1991; Ramírez-Espinoza et al., 1991; Talavera-Mendoza et fornia and the southwestern Cordillera of North America (Gastil
al., 1995). The Teloloapan Terrane (Fig. 1) is exposed in the east- et al., 1991; Centeno-García, 2005).
ernmost parts of the Guerrero Composite Terrane. It is charac- In either of the two scenarios, deformed Paleozoic rocks of
terized structurally by a complex thrust-fault system that verges the Tahue Terrane are the basement upon which Cretaceous vol-
eastward. Its Lower Cretaceous rocks are severely deformed and canism was built, indicating an earlier history of accretion of the
metamorphosed in low-grade greenschist facies. The Teloloapan Guerrero Composite Terrane than was previously interpreted by
Terrane overrides either Lower to Middle Cretaceous platform other authors.
W Tahue
(Guerrero Composite Terrane) Cortes
E
Passive margin sediments Stage I
Arc collision and
Caborca development of a passive
margin
Paleozoic
Vizcaíno? Zihuatanejo (Guerrero Composite) Oaxaquia
and Central terranes Continent margin
Siliciclastic turbidites share the same provenance Stage II
n
Potosí Submarine Fan
Vizcaino?
Marginal oceanic basin
Oaxaquia and Mixteca
with active rift volcanism
Arteaga Basin

Triassic
Carnian-Norian
Vizcaíno? Guerrero Central Oaxaquia
Potosi Fan Stage III
Taray Complex (deformed) Accretion via subduction
Arteaga Complex wide accretionary prism
?
? Oaxaquia and Mixteca Early Jurassic
pre 180-163 Ma

Guerrero Central Oaxaquia


transtension? Stage IV
Tumbiscatío Continental arc and
158 Ma
Granitoid contemporaneous strike-slip
163-158 Ma and extension (roll-back of
+ the subducting plate?)
? +Oaxaquia and Mixteca
Middle to
Late Jurassic
Sub
duc
tion

Guerrero Composite Terrane Central Oaxaquia (north)


and Mixteca
Zihuatanejo/Tahue Arcelia
Guanajuato Teloloapan Stage V
transtension? Formation of a multiple arcs
Calcareous K platform system or a single arc with
intra-arc/back arc rifting
Cretaceous
? ? ? Oaxaquia and Mixteca Berriasian-Cenomanian
Sub
duc
tion

Guerrero Composite Terrane Central / Oaxaquia Stage VI a


Zihuatanejo/Tahue Arcelia Teloloapan and Mixteca (south) Deformation and locally
Guanajuato
metamorphism, cut by
granitoids 105 Ma compression 105 Ma granitoids in the
Zihuatanejo terrane
Oaxaquia and Mixteca Cretaceous
Cenomanian?-pre Santonian

Red Beds Stage VI b


Marine turbidites Active continental arc in
the west and syntectonic
marine turbidites (foreland)
in the east and deformation
Cretaceous
Santonian?-Maastrichtian

Figure 3. Tectonic models for the evolution of western Mexico, showing the alternating stages of subduction-collision and rifting.
Guerrero Composite Terrane of western Mexico 287

Stage II: Late Triassic Passive Margin—Deposition of the in the Zacatecas Formation, and in the Arteaga and Las Ollas
Potosi Fan Complexes (Zihuatanejo Terrane). These last three units formed
in the distal ocean-floor zone of the Potosi Fan. The presence
The paleo-continental edge of Mexico lay approximately of mélanges (Arteaga Complex and Taray Formation) as well as
at the western boundary of the Oaxaquia and Mixteca Terranes blueschist in the Las Ollas Complex (Zihuatanejo Terrane) indi-
in the early Mesozoic (Fig. 1; Centeno-García, 2005). Thus the cates that deformation occurred in a subduction zone. During this
Central and Guerrero Composite Terranes (Fig. 1) were accreted deformational event the turbidites of the Potosi Submarine Fan,
or displaced to their present position during the Mesozoic. Sedi- with slivers of the oceanic crust and its sedimentary cover, were
mentation along the western continental margin of Oaxaquia was accreted to the continent. This accretionary prism apparently
dominated by large volumes of siliciclastic turbidites (quartz-rich was very wide, as suggested by the large areas that are floored
sandstone and shale) that were deposited in the distal continental by it. Whether the subducting slab was dipping toward the west
shelf or at the continental slope at least during Carnian–Norian (under an oceanic arc) or the east (under continental Mexico)
time (Fig. 3). Accretionary complexes that form the basement of has not been constrained. There are two isolated reports of dated
the Central Terrane and parts of the Guerrero Composite Terrane Early Jurassic volcanic rocks in Oaxaquia (Barboza-Gudiño et
(Zihuatanejo Terrane) are formed largely (up to 60% of the total al., 2004; Fastovsky et al., 2005), but whether they are part of a
area of exposures) by similar quartz-rich sandstone and shale continental arc or not is not known. Evidence of contemporane-
turbidites that made up the matrix within which blocks of vari- ous oceanic-arc magmatism is exposed in the Vizcaíno Penin-
able composition are embedded. These turbidites in the accreted sula of Baja California (Kimbrough and Moore, 2003), where
terranes contain fossils of the same age as those from turbidites Triassic–Jurassic volcanic rocks have geochemical signatures of
deposited at the continental slope of Oaxaquia. primitive arc affinity. It is possible that the rocks in the Vizcaíno
Detrital zircon ages obtained from turbidites from all the Peninsula represent a displaced fragment of an oceanic arc that
localities of the Carnian–Norian turbidites, from Oaxaquia to accreted to the Arteaga and Las Ollas Complexes of the western
basal accretionary complexes of the Central and Zihuatanejo Ter- Guerrero Composite Terrane, which in turn accreted to the Taray,
ranes, show the same populations, which suggest that the fan tur- Zacatecas, and La Ballena Formations, but this model needs to be
bidites spread into a marginal oceanic basin that was later accreted supported by more evidence.
to the continental margin. These siliciclastic rocks are grouped as
the Potosi Fan (Centeno-García, 2005) and are important because Stage IV: Late Jurassic Continental Arc—Overlapping
they can be traced from Oaxaquia to the present Pacific Coast Assemblage for Guerrero Composite Terrane, Central
of Mexico, and they tie together the Central Terrane, the west- Terrane, Oaxaquia, and Mixteca Terrane
ernmost part of the Guerrero Composite Terrane (Zihuatanejo
Terrane), and the continental margin of southern North America Subaerial volcanic and sedimentary rocks, as well as shallow
(Oaxaquia) during Late Triassic time. Thus, the Potosi Subma- porphyritic intrusives, dikes, and sills, overlie or cut previously
rine Fan may have been a large sedimentary feature, probably deformed Triassic sedimentary rocks in Oaxaquia and rocks of
close to the dimensions of the present Bengal Fan. the accretionary prism in the Central Terrane. These rocks range
There is no evidence of Triassic magmatism in continental in age from 174 to 158 Ma (Jones et al., 1995; Barboza-Gudiño
Mexico, and detrital zircon geochronology of the fan turbidites et al., 2004). A common attribute of all the outcrops of these
show that the youngest age populations are much older than dep- rocks is that they are mostly rhyolitic in composition, with minor
ositional ages in all the studied localities of the Potosi Fan (Fig. 3; dacitic-andesitic lava flows and tuffs, and show evolved-arc geo-
Centeno-García et al., 2005; Centeno-García, 2005). Therefore, chemical signatures (Centeno-García and Silva-Romo, 1997;
the Potosi Fan probably was deposited across a passive margin, Centeno-García, 2002; Centeno-García and Díaz-Salgado, 2002).
or at least a margin that had no active subduction along the length Coeval volcanic rocks have been reported in the Mixteca Terrane
of the fan at the time of deposition. as well, suggesting that arc volcanism was widespread in conti-
nental Mexico at that time (García-Díaz et al., 2004). Rocks of
Stage III: Accretion of the Potosi Fan to the Continental similar age range and similar evolved-arc geochemical signatures
Margin via Subduction—Basement of the Central and are exposed in the western Zihuatanejo Terrane of the Guerrero
Zihuatanejo Terranes Composite Terrane (Bissig et al., 2003; Centeno-García et al.,
2003). This suggests that the Guerrero Composite Terrane may
All the Triassic units of central and western Mexico are have been incorporated into the continental margin by that time.
strongly deformed and partially metamorphosed, indicating that a Summarizing the data described above: (1) Triassic base-
major compressional event occurred during latest Triassic–Early ment rocks of the Zihuatanejo Terrane (Guerrero Composite Ter-
Jurassic time. This event is characterized by tight folding, shear- rane) share a provenance linkage with rocks of the same age in
ing, and axial cleavage in the continent-slope deposits of the Oaxaquia and the Central Terrane; (2) all Triassic rocks, from
Potosi Fan in Oaxaquia (La Ballena Formation), and block- those deposited on the paleo-continent’s margin of Mexico to
in-matrix texture in the Taray Formation (Central Terrane), those within the accreted terranes, were deformed previous to the
288 Centeno-García et al.

development of a Late Jurassic continental arc; and (3) evolved compositions of most of the Upper Jurassic–Cretaceous igneous
Upper Jurassic continental-arc volcanism was widespread among rocks of the different arc assemblages of the Guerrero Composite
continental Mexico and accreted terranes (Central and Zihua- Terrane suggest primitive sources, with little or no influence on
tanejo Terranes). On the basis of these facts, we propose in this an evolved continental crust (e.g., Ortiz-Hernandez et al., 1991;
paper that the first accretion of the Guerrero Terrane occurred Lapierre et al., 1992; Centeno-García et al., 1993a; Freydier et al.,
during latest Triassic–Early Jurassic time instead of near the 1995; Gastil et al., 1999; Talavera-Mendoza et al., 1995; Men-
end of the Cretaceous, as previously proposed by other authors. doza and Suastegui, 2000, among others). Basalts with ocean-
Therefore, the Late Jurassic magmatic event represents an over- island (OI) and mid-oceanic-ridge basalt (MORB) signatures of
lapping assemblage that stitches all the terranes of central and the Arcelia and Guanajuato terranes (Lapierre et al., 1992; Ortiz-
western Mexico for that period. Hernandez et al., 2003; Mendoza and Suastegui, 2000) suggest
the influence of a mantle source for the magmatism.
Stage V: Late Jurassic–Early Cretaceous Intra-Arc Strike Regional differences in the strata suggest abrupt lateral
Slip(?)–Rifting of the Continental Arc—Drifting of the changes in the depositional environments from shallow marine
Guerrero Composite Terrane to deep marine. Also, lateral differences in thickness of the suc-
cessions suggest that they may have been deposited in alternate
It has been proposed that major lateral displacements subsiding basins and basement highs where the deposits draped
occurred during the activity of the Jurassic continental arc of thinly or were absent. These major geological differences suggest
stage IV (Anderson and Silver, 2005). Therefore, the arc was that intra-arc rifting was considerable and was probably associ-
originally in a more northerly position, and it was displaced, via ated with a complex paleogeography of marginal arc and backarc
the Mojave-Sonora Megashear, to its present position in central systems in western Mexico. Whether or not the different terranes
Mexico prior to, or at, the early stage of development of the cal- of the Guerrero Composite Terrane were formed in a single arc
careous platform (Anderson and Silver, 2005). has not been constrained. Some authors proposed that the Guer-
Whether this major strike-slip system existed or not has been rero Terrane formed from a complex system of two or three arcs
widely discussed (see GSA Special Paper 393). We consider that (Ramírez-Espinosa et al., 1991; Mendoza and Suastegui, 2000).
extensive geological evidence of major tectonism during and However, no Cretaceous subduction-related accretionary prisms
after arc volcanism exists (see following discussion). The cessa- have been identified within any of the terranes of the Guerrero
tion of magmatism in the Central Terrane and Oaxaquia suggests Composite Terrane.
a change in the location of the subduction zone. Then, a major
regional calcareous platform developed over the arc and other Stage VI: Final Accretion of the Guerrero Composite
older rocks. This major transgression initiated the deposition Terrane, and Development of a New Continental Arc
of limestone on Oaxaquia, and on the Mixteca and Central Ter-
ranes. Calcareous sedimentation in central and eastern Mexico A major Late Cretaceous–early Paleogene orogenic phase is
was characterized by high subsidence rates (Goldhammer, 1999). recorded throughout Mexico, coeval to the Sevier and Laramide
Arc magmatism continued only in a small area in the western orogenies in western North America. This event is associated with
Mixteca Terrane and became widespread in the Guerrero Com- the Mexican Fold and Thrust Belt of the Sierra Madre Oriental.
posite Terrane. Although there is some overlap in age ranges of Apparently, final amalgamation of the Guerrero Composite Ter-
arc volcanism among the terranes that form the Guerrero Com- rane occurred during this orogenic event, and volcanic and sedi-
posite Terrane, there is a general trend from older ages in eastern mentary rocks of the Teloloapan, Guanajuato, Zihuatanejo, and
Oaxaquia and the Central Terrane to younger ages in the western Tahue Terranes were thrust over the calcareous platform rocks
Guerrero Composite Terrane (Fig. 3). This suggests a possible of Oaxaquia and the Central, Cortes, and Mixteca Terranes. The
W-SW migration of the subduction zone. We propose that dur- amount of tectonic transport apparently is significant, as xeno-
ing and after the continental arc activity (Late Jurassic–Early liths of Precambrian continental crust were found in Cenozoic
Cretaceous time), large amounts of extension and lateral trans- volcanic rocks that erupted onto accreted rocks of the Guanajuato
lations may have occurred (see inferred faults in Fig. 1). This Terrane (Urrutia-Fucugauchi and Uribe-Cifuentes, 1999). Sig-
extensional-transtensional(?) event split the continental arc, ini- nificant tectonic transport is also suggested by the amount of
tiating the drifting of parts of previously accreted oceanic rocks shortening that produced tight folding and major thrusting within
(basements of the Tahue and Zihuatanejo Terranes) and the gen- the northern Zihuatanejo Terrane and the Arcelia and Teloloapan
eration of new oceanic crust (Guanajuato and Arcelia Terranes). Terranes (Salinas-Prieto et al., 2000). In contrast, deformation of
With the data available, it seems that volcanic activity at the Cretaceous rocks in the southern parts of the Zihuatanejo Terrane
northern Zihuatanejo Terrane and at the Guanajuato and Teloloa- formed wide regional anticlines, and some overturned folds and
pan Terranes was restricted to latest Jurassic–Early Cretaceous minor thrust faults locally. The structures generally trend NW-
time (Fig. 3). In contrast, in the Arcelia, Tahue, and southern SE, although locally some structures trend N-S and E-W.
Zihuatanejo Terranes, arc volcanism apparently continued up Santonian terrestrial sedimentation covers unconformably
to Albian–Cenomanian time (Fig. 3). Geochemical and isotopic the previously deformed arc assemblages of the Zihuatanejo
Guerrero Composite Terrane of western Mexico 289

Terrane (Benammi et al., 2005). Synorogenic sedimentary basins Ordovician to Devonian–Early Mississippian and were deformed
(Caracol Formation in Oaxaquia, and Mexcala Formation in the during the Mississippian (Poole and Madrid, 1988; Stewart et al.,
Mixteca Terrane) containing clasts derived from the Guerrero 1990). These rocks are in turn overlain by Upper Carboniferous
Composite Terrane suggest that these terranes were deformed and Permian turbidites (Fig. 5; Poole and Madrid, 1988; Stew-
and exhumed by that time. In addition, synorogenic sedimenta- art et al., 1990; Poole et al., 2005). They all were deposited in
tion overlaps the Arcelia and Teloloapan terranes (Miahuatepec a deep-marine environment and are interpreted to be part of the
Formation), which suggests that these two terranes were also Paleozoic continental slope-rise deposits of western North Amer-
amalgamated during the same orogenic event (Mendoza and Sua- ica (Poole and Madrid, 1988; Stewart et al., 1990). All these units
stegui, 2000; Guerrero-Suastegui, 2004). All these synorogenic of the Cortes Terrane were deformed and thrust over the Caborca
basins range in age from Turonian to Maastrichtian. In addition, Terrane by Late Permian to Early Triassic time, and they are
Paleocene granitoids along the coast cut the previously folded unconformably covered by Upper Triassic terrestrial and marine
units of the Zihuatanejo Terrane and suggest a Late Cretaceous– sedimentary rocks (Stewart et al., 1990). Therefore, the Caborca
early Paleogene deformation. and Cortes Terranes were assembled by early Mesozoic time.
Therefore, final amalgamation of the Guerrero Composite The nature of the contact between the Cortes and Tahue Ter-
Terrane occurred between Santonian and Turonian–Maastrich- ranes (the latter belonging to the Guerrero Composite Terrane)
tian time. has not been mapped in detail. It is inferred to be a thrust that
verges toward the north, and it is probably north of El Fuerte
DISCUSSION town in Sinaloa State (Fig. 5), based on the northernmost expo-
sures of Cretaceous marine volcanic rocks of the Guerrero Ter-
This section summarizes the stratigraphic, structural, and rane (Servais et al., 1982; Henry and Fredrikson, 1987; Roldán-
geochemical data that support the proposed stages for the tec- Quintana et al., 1993; Freydier et al., 1995).
tonic evolution of western Mexico. The oldest Paleozoic rocks of the Tahue Terrane (Guer-
rero Composite Terrane) are exposed in the area of El Fuerte
Stage I: Origin of the Basement of the Tahue Terrane (El Fuerte Complex; Figs. 4 and 5). The El Fuerte Complex is
formed by marine rhyolitic to andesitic lava flows and volca-
Exposures of pre-Cretaceous rocks in northwest Mexico niclastic rocks, interbedded with quartz-rich sandstone, shale,
are scattered; thus contact relationships among them can only and thin-bedded limestone (Mullan, 1978; Roldán-Quintana et
be indirectly inferred (Figs. 1 and 4). Approximate distribution al., 1993; Poole and Perry, 1998). All these various components
of the contacts among the terranes of western Mexico (Caborca, are deformed and metamorphosed to greenschist facies (Mul-
Cortes, and Tahue; Figs. 1 and 4) was outlined on the basis of lan, 1978; Roldán-Quintana et al., 1993). Sedimentary rocks of
the geographic distribution of pre-Cretaceous outcrops and lat- the El Fuerte Complex contain Ordovician conodonts (Poole
eral changes in the isotopic signatures of Cretaceous–Paleogene and Perry, 1998). Preliminary geochemical analyses indicate a
granitoids (Valencia-Moreno et al., 2001). Thus the nature of the calc-alkaline island-arc affinity for the volcanic rocks of the El
contacts and the amount of displacement among different base- Fuerte Complex, and are similar to those from coeval Paleo-
ments are unknown. In this section the main stratigraphic units zoic arc rocks in the Klamath Mountains of Northern California
that define the terranes are described following a NW to SE tran- (Lapierre et al., 1987). However, more detailed geochemical
sect throughout the Paleozoic rocks of the Caborca, Cortes, and and geochronological work needs to be done to constrain their
Tahue Terranes (Guerrero Composite Terrane). origin and relationships.
At the southern margin of the Caborca Terrane a thick shel- Upper Paleozoic deep-marine sedimentary rocks are ex-
fal limestone succession is exposed that contains Carboniferous– posed south of El Fuerte, in San José de Gracia town, Mazat-
Permian fusulinids and other shallow-marine fossil fauna (Stew- lán City, and other scattered localities in Sinaloa State (Figs. 4
art et al., 1990). These rocks are overridden by a north-verging and 5). These rocks belong to the San José de Gracia Formation
major thrust fault that places deeper marine sedimentary rocks (Carrillo-Martínez, 1971; Gastil et al., 1991; Arredondo-Guerrero
of the Cortes Terrane on the shelfal rocks of the Caborca Ter- and Centeno-García, 2003) and are made up of quartz-rich sand-
rane (Fig. 1; Coney and Campa, 1987; Poole and Madrid, 1988; stone and shale turbidites, thin-bedded calcareous debris flows,
Stewart et al., 1990). black shale, and chert. The turbidites contain olistoliths of lime-
Basal metamorphic rocks are not exposed in the Cortes Ter- stone with chert nodules, which in turn contain Middle Pennsyl-
rane, but its basement has been interpreted as thinned Proterozoic vanian to Early Permian fossils at the San José de Gracia locality
rocks, perhaps the same as in the Caborca Terrane, or else Pro- (Carrillo-Martínez, 1971; Gastil et al., 1991). The San José de
terozoic metamorphic rocks different from those of the Caborca Gracia Formation has been interpreted as deposits in a deep-
Terrane (McDowell et al., 1999; Valencia-Moreno et al., 1999; marine environment (Gastil et al., 1991). The contact between
Valencia-Moreno et al., 2001). The deep-marine sedimentary the El Fuerte Complex and the San José de Gracia Formation
rocks of the Cortes Terrane are sandstone and shale turbidites, is not exposed. However, major differences in deformation and
graptolitic shale, chert, and layered barite that range in age from metamorphism (turbidites of the San José de Gracia Formation
Hermosillo

e
a Terran
Caborc a n e
e rr
Cortes T

Caborca Terrane
Limestone (Lower Cretaceous)
Ciudad Shallow marine limestone
Obregón S O N O R A and shale (Paleozoic)
Cortes Terrane
Navojoa ne
erra e Terrestrial and marine
27°
s T rran sedimentary rocks (Triassic)
rte Te
Co ue
T a h Deep Marine turbidites, chert
and barite (Paleozoic)

Sonobari
El Fuerte

C
H
San Jose de Gracia
N

I
26°

H
U
A
Sinaloa de Leyva-Porohui

H
U
A
D
U
R
25° A

Overlapping units
N

CULIACÁN
Recent sediments
G
O

Recent basalts
Miocene-Pliocene
Volcanic and sedimentary
(conglomerate-sandstone)
Granitoids (Paleocene-Oligocene)
Upper Cretaceous-Oligocene
Tarahumara Formation and Lower and Upper 24°
Volcanic Supergroup

Tahue Terrane (Guerrero Composite Terrane)


Granitoids (Upper Jurassic-Lower Cretaceous)
Cretaceous Guerrero Arc
MAZATLÁN
Marine volcanic and sedimentary rocks
San Francisco Gneiss (Triassic) 23°
San Jose de Gracia Fm. (Carboniferous)
El Fuerte Complex (Ordovician)
km
0 25 50 100

109° 108° 107° 106°

Figure 4. Geologic map of Sinaloa and the southern Sonora states, showing the geology of Caborca, Cortes, and Tahue Terranes (after
Carrillo-Martínez, 1971; Mullan, 1978; Gastil et al., 1978; Henry and Fredrikson, 1987; Stewart and Roldán-Quintana, 1991; Ortega et
al., 1992; and our own work).
WEST EAST
western Oaxaquia
Maastrichtian Caracol Fm.
Campanian shale and sandstone
foreland-basin fill
Santonian GUERRERO COMPOSITE TERRANE
Coniacian
Turonian Tahue Terrane Zihuatanejo Terrane Guanajuato Terrane Central Terrane
Cenomanian
La Borda, Limestone
Albian El Saucito and Titonian to Aptian
ChilitosFms. Tuna Mansa
Aptian Lower Cretaceous diorite
Aptian? Cerro Pelon Transitional Limestone
volcanic turbidites tonalite
Barremian volcanic turbidites Santa Ana from continental to Tithonian to Aptian
limestone, tuff and calcareous debris flows, dike swarm marine shale,
chert, pillowed lavas tuff and chert Esperanza Fm. sandstone, evaporites,
Hauterivian and dikes deformed Transitional
pillowed lavas and dikes limestone, chert, limestone
Valanginian and in part deformed and in part volcaniclastic turbidites Kimmeridgian-Tithonian from continental to
Berriasian metamorphosed metamorphosed felsic lavas 146 Ma marine shale,
Tithonian felsic and La Luz sandstone, evaporites,
basic lavas basaltic flows La Perlita limestone
Kimmeridgian and tuff limestone
150-147 Ma Albian-Aptian felsic lavas and Oxfordian-Tithonian
Oxfordian
Callovian
massive limestone, volcaniclastics felsic lavas and
Bathonian
black limestone continental volcaniclastic rocks
Bajocian unconformity shales and sanstone Oxfordian continental
Aalenian
unconformity unknown age
Arperos Fm.
Toarcian Hauterivian-Tithonian? Felsic lavas and
Pliensbachian deep marine volcaniclastic volcaniclastic rocks
unconformity turbidites,chert, limestone 189-172 Ma
Sinemurian

Hettangian
unconformity unconformity
Rhaetian Francisco gneiss
Upper Triassic Taray Fm. La Ballena Fm.
Norian Zacatecas Fm. Paleozoic-Triassic? Carnian-Norian
migmatized gneisses Norian-Carnian
and amphibolites matrix is quartz-rich turbidites deformed quartz-rich
matrix is quartz-rich turbidites blocks of pillow lavas, chert, turbidites
Carnian blocks of pillow lavas serpentinite and limestone continent-slope facies
accretionary complex accretionary complex
Ladinian El Fuerte Complex ?
Early Triassic Ordovician
metamorphosed
volcanic and quartz-rich linked by provenance (Potosí Fan)
Paleozoic turbidites, limestone,
chert, basalts and rhyolites
Figure 5. Simplified stratigraphic columns for Oaxaquia and terranes north of the Transmexican Volcanic Belt. The columns are in an east-west order. They include the Tahue Terrane,
which is part of the Guerrero Composite Terrane. Vertical scale shows the age range (in Ma).
292 Centeno-García et al.

are strongly deformed but not metamorphosed) indicate that the Sonobari (Figs. 4 and 5; Mullan, 1978; Keppie et al., 2006).
contact is probably an angular unconformity. The Francisco Gneiss is made up of migmatized gneisses and
Both units of the Tahue Terrane (El Fuerte Complex and amphibolites that have within-plate and continental tholeiite geo-
San José de Gracia Formation) are important because they can chemical signatures (Keppie et al., 2006). This suggests that the
constrain the paleogeography of the northern Guerrero Terrane. Tahue and Cortes terranes may have been geographically sepa-
Preliminary single-grain, detrital zircon geochronology from rated by that time.
the quartz-rich sandstone from the turbidites of the San José de
Gracia Formation shows populations that have a North American La Ballena Formation
affinity (Centeno-García et al., unpublished data) and are similar Triassic rocks of Oaxaquia crop out on its western margin,
to those from Paleozoic rocks in Baja California, in the Cortes near its boundary with the Guerrero Composite Terrane (Figs. 1
Terrane, and in Nevada (Gehrels et al., 2002). and 6). They are grouped as the La Ballena Formation (Silva-
The stratigraphy, geochemistry, and provenance of the Paleo- Romo, 1993; Silva-Romo et al., 2000), and their largest expo-
zoic rocks suggest that the Tahue Terrane (Guerrero Composite sures are in the Peñón Blanco, Charcas, and Real de Catorce
Terrane) was linked to the tectonic evolution of the western conti- areas (Fig. 6; Silva-Romo, 1993; Tristán-González and Torres-
nental margin of North America, probably up to Permian–Triassic Hernández, 1994; Barboza-Gudiño et al., 2004). The La Ballena
time. After that, there were major differences in the composition Formation is made up of quartz-rich sandstone and shale, and
of the Mesozoic sedimentary cover of the Caborca-Cortes Ter- scarce conglomerates deposited as small channel-fill lenses. The
ranes with respect to that of the Tahue Terrane. Therefore, it is sedimentary structures of these Triassic rocks indicate deposi-
likely that a fragment of previously accreted island-arc and con- tion mostly by turbidity currents, although some debris flows
tinent-margin assemblages drifted from the continental margin and large slumps are present. This sequence contains abundant
sometime in the early Mesozoic. trace fossils and ammonites and bivalves of Late Triassic (Car-
Contact relationships between the Paleozoic sedimentary nian) age at the Peñón Blanco and Charcas areas (Cantu-Chapa,
rocks of the San José de Gracia Formation (Tahue Terrane) and 1969; Silva-Romo et al., 2000; Bartolini et al., 2002). Sedimen-
the Triassic subduction-related complex of the Zihuatanejo Ter- tary structures and fossil fauna suggest that the deposition of this
rane are unknown because the contact is covered by younger unit occurred in a submarine fan that developed on an external
rocks. However, the Tahue and Zihuatanejo Terranes share simi- platform or continental slope setting. These rocks form part of
lar Cretaceous volcanic and sedimentary cover. the Potosi Submarine Fan (Centeno-García, 2005). The origi-
nal thickness is unknown, but up to 4640 m was penetrated by
Stages II and III: Triassic Potosi Fan and Its Accretion to exploration drilling without reaching the base of the succession
the Continental Margin (López-Infanzón, 1986).

There are few exposures of Triassic rocks in Mexico, and Taray Formation
they are limited to the Caborca and Cortes Terranes, western Similar marine siliciclastic rocks crop out at the Pico de
Oaxaquia, the Central and Zihuatanejo Terranes, and a small out- Teyra region in the Central Terrane (Figs. 5 and 6). They belong
crop in the Vizcaíno Peninsula in Baja California. Triassic rocks to the Taray Formation, made up of highly deformed quartz-rich
have not been found in the Mixteca Terrane or in other terranes turbidites (sandstone and shale) interbedded with some black
of Mexico. In this section we briefly describe Triassic rocks of chert and scarce detrital limestone that contains fragments of cri-
the Cortes and Tahue Terranes (Barranca Group and Francisco noids, gastropods, corals, bivalves, and bryozoans (Diaz-Salgado
Gneiss) and focus on the marine Triassic rocks of Oaxaquia (La et al., 2003). The Taray siliciclastic turbidites form a matrix
Ballena Formation), the Central Terrane (Taray Formation), and within which blocks of black and green chert, pillow basalts, ser-
the Guerrero Composite Terrane (Zacatecas Formation, and the pentinite, and crystallized limestone can be found (Figs. 5 and 6;
Arteaga and Las Ollas Complexes). Diaz-Salgado et al., 2003). The age of this unit remains undeter-
mined; however, there are reports of fusulinids from one of the
Barranca Group and Francisco Gneiss limestone blocks (Anderson et al., 1990). The youngest detrital
Triassic (Carnian–Norian) sedimentary rocks of the Cortes zircons collected from the sedimentary matrix are Late Permian
Terrane are made up of fluvial sandstone and shale that contain in age (Diaz-Salgado et al., 2003). There is also a report of molds
abundant coal layers (Barranca Group; Stewart and Roldán- of bivalves of possible Carnian age (Barboza-Gudiño et al., 1999;
Quintana, 1991). These sediments were deposited unconform- Bartolini et al., 2002). Thus deposition of the sedimentary matrix
ably on previously deformed Paleozoic deep-marine rocks. The should have occurred between the Late Permian to the Late Trias-
Triassic fluvial deposits change transitionally up the column to sic. The Taray Formation has a block-in-matrix structural style,
shallow-marine siliciclastic deposits. These rocks have no evi- formed by centimeter-size blocks to blocks of hundreds of meters
dence of contemporaneous volcanism. In contrast, Triassic rocks in size, all in a highly sheared sedimentary matrix. This charac-
of the Tahue Terrane (Guerrero Composite Terrane) are made up teristic is typical of a subduction accretionary complex (Ander-
of metamorphosed igneous rocks of the Francisco Gneiss near son et al., 1990, 2005; Diaz-Salgado et al., 2003).
Guerrero Composite Terrane of western Mexico 293

102° 101° 100°

Cenozoic cover
Caopas Oaxaquia and Central Terrane
U/Pb 158 Ma Concepción
Ju-K Calcareous Platform
del Oro
Pico Ju felsic and intermediate lava
de flows, epiclastic rocks and redbeds
Teyra
Tr Quartz-rich shale and sandstone
turbidites (Potosí Fan)
24° Huiznopala Tr(?) Accretionary complex
Granjeno blocks of pillow basalt, ultramafic,
chert, and marble in a sedimentary
quartz-rich matrix
Real de Matehuala
Central Catorce
Huizachal
Proterozoic and Paleozoic
U/Pb190 Ma metamorphic and sedimentary
terrane rocks

Guanajuato Terrane
Ju-K pillow basalts, deep marine
23° Charcas volcanic turbidites, chert, and mafic
and ultramafic plutons

Oaxaquia Zihuatanejo Terrane


Zacatecas Ju-K pillow basalts, volcanic
Peñón turbidites, detrital limestone
U/Pb 147-150 Ma Blanco
Tr Accretionary complex
blocks of pillow basalt, in a
San Luis Potosí sedimentary quartz-rich matrix
22°
Gu ter
an ran

Zihuatanejo 100 km
aju e

terrane
at
o

U/Pb146 Ma
Guanajuato

Figure 6. Geologic map of central Mexico, showing main stratigraphic units of Oaxaquia and the Central, Zihuatanejo, and Guanajuato Terranes
(modified from Ortega et al., 1992). Tr—Triassic; Ju—Jurassic; K—Cretaceous.

Zacatecas Formation structures associated with two distinct deformational events, one
The oldest rocks of the Zihuatanejo Terrane in its northern- of them prior to the deformation that is recorded in the Creta-
most exposure are Triassic in age as well (Fig. 6). They make ceous rocks as well. The small size of the outcrop prohibits con-
up the Zacatecas Formation, which crops out in a small tectonic straints on the tectonic origin of the Zacatecas Formation, but its
window at the western margin of Zacatecas City (Fig. 6; Burck- lava flows and siliciclastic turbidites are similar to those from the
hardt and Scalia, 1906; Ranson et al., 1982; Cuevas-Pérez, 1983; Arteaga Accretionary Complex, which is exposed in the southern
Monod and Calvet, 1991). This formation is made up of quartz- part of the Zihuatanejo Terrane.
rich turbidites (sandstone and shale) that contain blocks of pil-
low basalts that have MORB geochemical signatures (Fig. 7; Arteaga Complex
Centeno-García and Silva-Romo, 1997). The Zacatecas Forma- More exposures of Triassic(?) rocks are found in the south-
tion contains fossil ammonites and bivalves of Late Triassic (Car- ern part of the Zihuatanejo Terrane (Fig. 7). Their largest outcrops
nian) age (Burckhardt and Scalia, 1906; Bartolini et al., 2002). are located in the Arteaga, Placeres del Oro, and Tiquicheo areas
Its contact with the La Borda Formation of Late Jurassic(?)–Cre- (Arteaga Complex) and near Zihuatanejo City (Las Ollas Complex)
taceous age is inferred to have been originally an unconformity, (Fig. 8; Centeno-García et al., 1993a, 1993b; Talavera-Mendoza
but it was sheared and detached during Late Cretaceous thrust- et al., 1995; Mendoza and Suastegui, 2000). The Arteaga Com-
ing and folding (Fig. 3). Rocks of the Zacatecas Formation show plex is made up of quartz-rich turbidites (sandstone and shale),
WEST EAST
Mixteca Terrane
GUERRERO COMPOSITE TERRANE (western Continent-margin)
Maastrichtian
Mexcala Fm.
Campanian Zihuatanejo Terrane Arcelia Terrane TeloloapanTerrane shale and sandstone
Santonian Coastal region Huetamo region foreland-basin fill
Coniacian subaerial to marine
Turonian
shale, sandstone,
Cenomanian evaporites, limestone, Comburindio and
Albian
basaltic-andesitic Mal Paso Fms.
lava flows shallow marine
Aptian Albian-Cenomanian Cuautla and
limestone,volcaniclastic Limestone Morelos Fms.
subaerial few marine rocks, few andesitic-basaltic Aptian-Albian Limestone
Barremian volcaniclastic rocks, lava flows Villa de Ayala Fm. Aptian-Albian
felsic lavas Aptian Albian MORB basaltic pillow unconformity
Hauterivian Aptian lavas, IA lavas and Taxco Schist,
Angao and San felsic and
Valanginian volcaniclastic rocks Zicapa and
Lucas Fms. dikes, chert, deep andesitic
Berriasian and limestone volcanic turbidites, chert marine volcaniclastic Chapolapa
Tithonian age not well felsic and basic lavas turbidites lavas and felsic lavas and
Kimmeridgian constrained volcaniclastics volcaniclastics
unconformity shallowing upward Early Cretaceous shallow marine in the
Oxfordian Neocomian continental affinity
Callovian
Cuale east, deep marine in 127-133 Ma
Bathonian felsic lavas and the west
Bajocian Tumbscatío granitoids Las Lluvias
Aalenian
145-137 Ma
163-155 Ma Ignimbrite
Toarcian 168-179 Ma
unconformity
Pliensbachian

Sinemurian

Hettangian
Rhaetian
Arteaga Complex
Norian Norian-Carnian
matrix is quartz-rich turbidites Acatlan Complex
Carnian blocks of pillow lavas, chert, mid-Paleozoic
serpentinite and limestone poly deformed sedimentary
Ladinian accretionary complex and volcanic rocks
Early Triassic continent-continent collision
Paleozoic

Figure 7. Simplified stratigraphic columns for the terranes described in the text that are south of the Transmexican Volcanic Belt, except for the Guanajuato Terrane, and include the
Mixteca Terrane and the Teloloapan, Arcelia, and Zihuatanejo Terranes (Guerrero Composite Terrane). Vertical scale shows the age range (in Ma). MORB—mid-oceanic-ridge basalt;
IA—island arc.
Zihuatanejo Terrane Teloloapan Terrane Mixteca Terrane Overlapping assemblages
Huetamo area, Cretaceous Lower Cretaceous arc assemblages Aptian-Albian Calcareous Miahuatepec Formation
arc assemblage, mostly marine shallow marine andesitic to basaltic Platform Upper Cretaceous clastic rocks
volcaniclastic rocks, limestone and lava flows and volcaniclastic rocks, foreland basin-fill (overlaps
some pryroclastic and lava flows massive and reefal limestone (east) Lower Cretaceous continental Arcelia and Teloloapan terranes)
Coastal Cretaceous arc assemblage deep marine lava flows, calcareous arc assemblages, marine and Mexcala Formation
shallow marine and terrestrial rhyolite, debris flows and volcanic turbidites (west) terrestrial rhyolite to andesite lava Turonian-Maastrichtian clastic rocks
andesite and some basalt, limestone, flows and epiclastic rocks, foreland basin-fill (overlaps
volcaniclastic and basement-derived Arcelia Terrane quartz-rich clastic rocks Mixteco and Teloloapan terranes)
clastic rocks Cretaceous IAB pillow lavas, Paleozoic Acatlán Complex CutzamalaFormation
Middle to Upper Jurassic fine-grained volcanic turbidites Santonian-Maastrichtian red beds
plutons and chert (deep marine) (overlaps Zihuatanejo and Arcelia
Las Ollas Subduction Complex Xolapa Terrane terranes)
Cretaceous MORB pillow lavas,
(age unknown) fine-grained volcanic turbidites Jurassic to Cretaceous Upper Cretaceous to Paleogene
and chert (deep marine) migmatites, gneisses, and plutons granitoids
Arteaga Subduction Complex
(Triassic) Cenozoic volcanic and
sedimentary rocks
104° 103° 102° 101°

Zapotitlán Nevado de Colima


Observed terrane boundary
(thrust fault)
Tecalitlán
Minatitlán
Zitácuaro inferred terrane boundary
Jilotlán (thrust fault)
Colima Pijuamo inferred terrane boundary
Tepalcatepec (strike-slip fault)
Manzanillo Nueva
Italia Rivers Roads
Tejupilco 19°
Tecoman
Z I H U ATA N E J O T E R R A N E State limit
Coalcoman
Aguililla 100°
TERRANE

U/Pb 163+3 Ma
Aquila Infiernillo Dam Huetamo
Taxco
Tumbiscatío
e

Balsas
Ar/Ar 152-158 Ma River Teloloapan
N Arteaga Iguala
Arcelia
RRANE
marine

deep marin

iver
PA as R
CIF Bals

IC
shallow

OC MIXTECA TERRANE
18°
ARCELIA

EA
N Playa Azul Zicapa
Lázaro Cárdenas
TELOLOAPAN TE

10 0 20 40
Zihuatanejo
km Chilpancingo
Chapolapa
XOLAPA TERRANE

Figure 8. Geologic map of southwestern Mexico, showing the simplified geology of the Mixteca, Teloloapan, Arcelia, and Zihuatanejo Terranes (after Campa and Ramirez, 1979;
Ortega et al., 1992; Talavera-Mendoza et al., 1995; Corona-Chávez and Israde-Alcántara, 1999; Mendoza and Suastegui, 2000; Centeno-García et al., 2003). IAB—island-arc basalt;
MORB—mid-oceanic-ridge basalt.
296 Centeno-García et al.

black and green chert, and mafic tuff that form a matrix that con- and Cortes Terranes (González-León et al., 2005) but are simi-
tains blocks and slabs of pillow basalts, diabase, banded gabbros, lar to those from Triassic fluvial sedimentary rocks of Arizona
chert, and limestone, all deformed in a block-in-matrix structural (Anderson, 2006). This suggests that at the end of the Triassic
style (Centeno-García et al., 2003). Chert layers contain radio- the terranes of central and western Mexico may have been to the
larians of Triassic (Ladinian–Carnian) age (Campa et al., 1982). north of their present locations.
Pillow basalts and gabbros have oceanic geochemical signatures Based on this evidence, we propose that the margin of the
(MORB; Centeno-García et al., 1993a; Centeno-García et al., western paleo-continent of Mexico was passive or rifting at the
2003). Sedimentary structures preserved in some exposures of end of the Triassic. This passive margin received abundant clastic
unmetamorphosed turbidites, along with the affinity of the few sedimentation, forming the large Potosi Fan. Sediments of this
fossils found in the sedimentary rocks of the matrix, suggest that fan were deposited on oceanic crust (Arteaga Basin in Fig. 3).
the sequence was deposited in a deep-ocean environment. Appar- When subduction started, slivers of the ocean floor were tectoni-
ently the quartz-rich turbidites were contemporaneous with oce- cally mixed with the already existing passive-margin quartz-rich
anic magmatic activity, as they are interbedded with volcaniclas- turbidites that were forming the Taray and Zacatecas Formations
tic rocks (Centeno-García et al., 2003). The block-in-matrix style as well as the Arteaga and Las Ollas Complexes. Whether the
of deformation of the Arteaga Complex, as well as its lithological ocean basin that was covered by sediments of the Potosi Fan was
associations, indicate that it was formed in a subduction accre- an active marginal oceanic basin, a marginal backarc basin, or an
tionary prism. Metamorphism ranges from none to amphibolite open ocean–continent flank is still uncertain. The only potential
facies; blueschist facies has not been found in the area. evidence of association of the Potosi Fan sediments with tholei-
itic oceanic volcanism is the volcaniclastic rocks interbedded
Las Ollas Complex with the siliciclastic turbidites in the Arteaga Complex, as the
The Las Ollas Complex forms part of the Zihuatanejo volcaniclastic rocks have geochemical signatures between primi-
Terrane and is exposed near Zihuatanejo City (Figs. 7 and 8; tive island arc and MORB (Centeno-García et al., 2003).
Talavera-Mendoza, 2000). This complex is a tectonic mélange At least two phases of deformation are found in all the Trias-
formed by highly sheared blocks of metabasalt, banded and mas- sic rocks of Oaxaquia and the Central and Zihuatanejo Terranes.
sive gabbro, metadolerite, ultramafic rocks, and shale and quartz- The first event comprised strong shearing and tight folding, and
rich sandstone (Talavera-Mendoza, 2000). These blocks are the block-in-matrix structures. A second event was recorded
enveloped in a highly sheared clastic (quartz-rich sandstone) or only in the Arteaga Complex. This event deformed the Jurassic
serpentinitic matrix (Talavera-Mendoza, 2000). Blueschist facies granitoids as well, and it is characterized by a mylonitic fabric.
were reported by Talavera-Mendoza (1993, 2000). Geochemical The third event was common to all the Triassic units and is also
compositions of the basalts are typical of MORB and primitive recorded in the Jurassic and Cretaceous cover sediments, and it
oceanic-arc magmas (Talavera-Mendoza, 2000). 40Ar/39Ar and is characterized by axial cleavage, open to tight folding, reverse
K/Ar ages obtained from amphibole from several metagabbro faulting, and thrusting.
blocks range from 223 Ma to 96 Ma (Permian to early Ceno- The time of accretion of the Central Terrane with Oaxaquia
manian) (Delgado, 1982; A. Iriondo, 2003, personal commun.). is assumed to have been prior to the Middle Jurassic, because the
This has been interpreted to be the subduction complex of the La Ballena Formation of Oaxaquia and the Taray Formation of
Cretaceous arc (Vidal-Serratos, 1991; Talavera-Mendoza, 1993); the Central Terrane were deformed and locally metamorphosed
however, its contact relationships with Cretaceous arc-related prior to deposition of Upper Jurassic terrestrial volcanic and clas-
rocks, and similarities with the Arteaga Complex, suggest an ear- tic formations (Tristán-González and Torres-Hernández, 1994;
lier origin. Jones et al., 1995; Silva-Romo et al., 2000). The Zihuatanejo
Quartz-rich turbidites from the La Ballena Formation of Terrane (Guerrero Composite Terrane) was also accreted at that
Oaxaquia (continental Mexico), the matrix of the Taray Forma- time, because the Arteaga Complex is cut by granitoids of Middle
tion of the Central Terrane, and the Arteaga and Las Ollas Com- Jurassic age as well (Centeno-García et al., 2003).
plexes and the Zacatecas Formation of the Zihuatanejo Terrane The subduction zone that formed the Taray and Zacatecas
(Guerrero Composite Terrane) have similar and distinctive com- Formations, and the Arteaga and Las Ollas Complexes, was
positions and detrital-zircon age populations (Centeno-García probably constructed along the continental margin of Oaxaquia
et al., 2005; Talavera-Mendoza et al., 2007). Therefore, Triassic in Early Jurassic time. Whether the subducting slab was dipping
sedimentation of the central and western terranes of Mexico is to the east or to the west has not been determined.
linked by provenance. The youngest zircon age populations from
all the samples (latest Permian) are much older than the depo- Stage IV: Jurassic Continental Arc of Western Mexico
sitional ages of the turbidites (Carnian–Norian), which means
that there was no active volcanism at that time. In other words, Erosion and exhumation of the accreted continental slope
there is no evidence of Late Triassic continental arc volcanism in sediments and the accretionary complexes occurred prior to the
Mexico. Zircon age populations of the Potosi Fan are different initiation of Middle to Late Jurassic magmatism. This is indicated
from those of Triassic quartz-rich sandstone from the Caborca by the major angular unconformity that separates the Jurassic arc
Guerrero Composite Terrane of western Mexico 297

succession from the deformed Triassic rocks of Oaxaquia and Díaz-Salgado, 2004). The Caopas Formation was formed by
the Central and Zihuatanejo Terranes. Jurassic arc magmatism shallow porphyritic intrusives. Felsic volcanic rocks of the Rodeo
has also been identified in the Mixteca Terrane. The Jurassic arc Formation yielded a K-Ar age of 183 Ma (López-Infanzón, 1986),
rocks have different names at different locations; they are hereby and the Caopas Formation a U/Pb age of 158 Ma (Jones et al.,
described by their occurrence in different terranes: 1995). Terrestrial volcaniclastic rocks of the Rodeo Formation
are interpreted to have been deformed previous to the deposition
Nazas, Huizachal, and La Joya Formations in Oaxaquia of late Oxfordian limestone (Anderson et al., 1991; Bartolini et
The La Ballena Formation (Oaxaquia) is unconformably al., 2002). However, in another locality nearby the Nazas Forma-
overlain by the volcanic rocks and red beds of the Nazas Forma- tion changes transitionally upward to shallow-marine calcareous
tion in the Peñón Blanco, Charcas, and Real de Catorce areas rocks that range in age from Late Jurassic to Late Cretaceous
(Figs. 5 and 6; Silva-Romo, 1993; Tristán-González and Torres- (Córdoba-Méndez, 1964; Díaz-Salgado, 2004).
Hernández, 1994; Barboza-Gudiño et al., 2004). The Nazas For- All these volcanic-sedimentary units are interpreted in this
mation is made up of dacitic and minor rhyolitic and andesitic work as the first overlapping succession that stitches the Central
lava flows and pyroclastic flows, dikes, and porphyritic shallow Terrane with Oaxaquia. The Caopas and Rodeo Formations, as
intrusives. The volcanic rocks are interbedded with conglomer- well as the Nazas Formation, are interpreted as continental intra-
ate, sandstone, and scarce paleosols. Conglomerate is formed arc assemblages (Jones et al., 1995).
mostly by volcanic clasts and a few clasts of sandstone and shale
derived from the underlying La Ballena Formation. The volca- Las Lluvias Ignimbrite of the Mixteca Terrane
niclastic conglomerate and sandstone form lens-shaped bedding Jurassic arc volcanism was also recorded in the Mixteca
with low-angle cross-bedding, interbedded with some debris Terrane in which ignimbrites, interbedded with fluvial and
flows, suggesting that they were deposited in a terrestrial (fluvial shallow-marine siliciclastic deposits, yielded U/Pb ages of 168.2
and alluvial fan) environment. ± 1.2 Ma, 177.3 ± 1.5 Ma, and 179.1 ± 1.5 Ma (Campa and Iri-
Although their age has not been well constrained at all ondo, 2003).
the exposures, there is a report of U/Pb ages as old as 189 Ma
at a subaerial volcanic-sedimentary succession in Huizachal Cuale Assemblage and Tumbiscatio Granitoids of the
(Huizachal Formation; Fastovsky et al., 2005), which might not Zihuatanejo Terrane
belong to the same volcanic arc event (Figs. 5 and 6). Rocks of Evidence of coeval Jurassic magmatism has been found
the Nazas Formation at Real de Catorce yielded U/Pb ages of in two localities in the Zihuatanejo Terrane (Guerrero Com-
172 ± 5 Ma (Barboza-Gudiño et al., 2004). The Nazas Forma- posite Terrane). One of the exposures is NE of Puerto Vallarta
tion changes transitionally upward to shallow-marine volcanicla- City, in the Cuale mining district, and the other locality is in the
stic rocks, some evaporites, and thin-bedded limestone, which in Tumbiscatio region, both along the Pacific Coast (Figs. 7 and
turn become a thick limestone succession in the Peñón Blanco 8). Rocks at Cuale contain volcanogenic massive sulfide (VMS)
and Charcas areas (Fig. 6). The basal part of this limestone suc- deposits and are composed of submarine rhyolitic lavas and tuffs,
cession contains late Oxfordian–Kimmeridgian fossil faunas volcanic sandstone with evolved-arc geochemical affinity (Bis-
(Centeno-García and Silva-Romo, 1997). In contrast, there is sig et al., 2003), and shale that yielded U/Pb ages of 162.4 and
an internal angular unconformity in the Real de Catorce local- 155.9 Ma (Bissig et al., 2003). These rocks are strongly deformed
ity (Fig. 6), which separates in two units, the terrestrial volcanic and partially metamorphosed, and their contact with Cretaceous
and sedimentary successions (Nazas and La Joya Formations; unmetamorphosed marine volcanic and sedimentary successions
Barboza-Gudiño et al., 2004). The upper La Joya Formation has not been determined.
changes transitionally upward to shallow-marine volcanic sand- Two Jurassic granitoids crop out in the Tumbiscatio region.
stone and shale interbedded with thin limestone strata. The oldest They were emplaced in previously deformed sedimentary rocks
fossils reported from the base of the limestone succession in Real of the Arteaga Complex, and vary in composition from grano-
de Catorce are Oxfordian in age (Barboza-Gudiño et al., 2004). diorite to granite to quartz monzonite. Their geochemical com-
positions are typical of calc-alkaline subduction-related gran-
Caopas, Rodeo, and Nazas Formations of the Central Terrane ites, which are more evolved than granitoids of Cretaceous and
The volcanic cover of the Taray Formation (Central Terrane; Cenozoic ages from the same area. Both granitoids show intense
Figs. 5 and 6) belongs to the Caopas, Rodeo, and Nazas Forma- shearing and internal deformation. Grajales and López (1984)
tions (Córdoba-Méndez, 1964; López-Infanzón, 1986; Jones et obtained one K/Ar date of Late Jurassic age (158 Ma). U/Pb
al., 1995). The Rodeo and Nazas Formations are lateral equiva- isotopic analysis yielded a 163 Ma age, and Ar/Ar ages are 158
lents of the same rocks but named differently in separate out- and 152.4 Ma (Centeno-García et al., 2003). The igneous rocks
crops (Díaz-Salgado, 2004). Both units are made up of rhyolitic of the Cuale and Tumbiscatio regions have strong similarities
to andesitic lava flows and dikes, and pyroclastic deposits that are in geochemical composition and age with volcanic rocks of the
interbedded with fluvial sedimentary rocks, mostly sandstone and Central Terrane (Caopas, Rodeo, and Nazas); thus we suggest
conglomerate (Anderson et al., 1990, 1991; Jones et al., 1995; that they probably originated in the same volcanic arc. Therefore,
298 Centeno-García et al.

the Arteaga Complex was probably accreted to the continental Although much detailed work needs to be done in order to
margin, either near or along the strike from central Mexico. reconstruct the paleogeography of western Mexico during the Cre-
The Jurassic volcanic event did not produce a thick strati- taceous, the available evidence indicates three important features:
graphic column and apparently did not have large volumes of 1. Magmatism prograded generally east to west through
volcanic products. The column changes transitionally upward to time, from the oldest ages in the Oaxaquia and Mixteca
shallow-marine calcareous rocks. Therefore, the lithologic asso- Terranes to the youngest ages in the coastal Zihuatanejo
ciations and vertical facies changes of this volcanic-sedimentary Terrane. There is some overlap of age ranges for the vol-
event are similar to those of a continental rift. However, the scarce canism among the different terranes, e.g., volcanism of
geochemical analyses from its volcanic rocks suggest an arc set- the Mixteca Terrane overlaps in age with part of the vol-
ting (Jones et al., 1995). These rocks have been interpreted as the canism of the Teloloapan Terrane (Guerrero Composite
southern continuation of the Jurassic continental arc that devel- Terrane). However, on a large scale, Albian–Cenomanian
oped along the southwestern margin of North America (Jones et volcanism is absent in the Mixteca and the Teloloapan
al., 1995). Terranes, and it is widespread in the coastal region of the
It has been proposed that major strike-slip faults were prob- Zihuatanejo and Arcelia Terranes.
ably active during the arc activity (Mojave-Sonora Megashear; 2. Magma chemistry changed through time toward a more
Jones et al., 1995). This could explain the fact that the Potosi Fan primitive melt. The Middle Jurassic volcanic and intru-
is south of its possible continental fluvial correlative in Arizona, sive rocks in all the terranes show mostly felsic-evolved
as well as the southward displacement of the Tahue Terrane. continental-arc geochemical signatures, including the
Whether or not the Jurassic volcanic event was coeval with Mixteca Terrane and Oaxaquia. In contrast the Cretaceous
a major transform fault has not been well documented. The volcanic rocks of the Guerrero Composite Terrane range
evidence in favor of an important synsedimentary deformation from tholeiitic basalts to andesites, with few rhyolites.
involving major extension is as follows: (1) Minor synsedimen- They show more primitive island-arc (IA) geochemical
tary normal faults and local angular unconformities are present signatures overall, and some even have MORB to oceanic-
within the Jurassic arc volcanic and sedimentary successions, island basalt (OIB) signatures. The Mixteca Terrane is the
and pre-Cretaceous mylonitic shearing is recorded in the Juras- exception to this trend; its magmatism remained evolved,
sic granitoids of the Tumbiscatio region (Zihuatanejo Ter- with continental arc signatures, into the Cretaceous.
rane). (2) Arc magmatism suddenly ceased in the Central Ter- 3. Within different assemblages of the Guerrero Composite
rane and Oaxaquia, followed by a rapid transgression recorded Terrane there are major differences in the stratigraphy,
in a few meters of transitional sedimentation. (3) Subsidence sediment composition, and depositional environments.
rates apparently were significant during the early stages of the And the Guerrero Composite Terrane overall is different
Oxfordian–Kimmeridgian marine sedimentation, because the from the volcanic-sedimentary rocks of the Mixteca Ter-
calcareous rocks show evidence of deeper sedimentation at rane to the east. In their present distribution, areas with
higher stratigraphic levels as well as overall rapid sedimenta- shallow-marine and terrestrial volcanic-sedimentary suc-
tion. (4) Although fault planes have been obliterated by younger cessions alternate with areas with deep-marine volcanic-
deformational events, they have been inferred by the rapid lat- sedimentary successions, and suggest a complex paleo-
eral changes in thickness and facies of the calcareous succession geography for that time.
through the interval from the end of the Jurassic to the Early These three features are hereby interpreted as evidence of
Cretaceous. (5) In addition, major regional lineaments have intra-arc rifting-translation. We propose, as a hypothesis to be
been identified in central and eastern Oaxaquia, including the tested, that the subduction zone might have migrated to the west.
San Marcos and La Babia Faults (Fig. 1) (Goldhammer, 1999; This would have produced thinning of the crust, which in turn
Chávez-Cabello et al., 2005). would have originated more primitive IA geochemical signatures
of the magmas and promoted the development of deep basins.
Stage V: Rifting of the Guerrero Terranes and Formation Whether the amount of extension was large enough to develop
of a Complex Arc System oceanic basins and several parallel subduction zones has not
been determined.
In this section we list the main stratigraphic features of the The stratigraphy, depositional environments, age, and geo-
volcanic-sedimentary successions of the Guerrero Composite chemical affinities of the main units are summarized by terrane.
Terrane and the Mixteca Terrane. Arc volcanism was absent in First, those of southern Mexico are described, following a section
Oaxaquia and the Central Terrane through the end of the Juras- from east to west. Next, Cretaceous rocks of the northern terranes
sic and the Cretaceous. During this period, oceanic crust was are described from east to west as well.
emplaced toward the east of Oaxaquia in the Gulf of Mexico,
and continuous subsidence prevailed throughout the Early Creta- Mixteca Terrane
ceous, resulting in a thick calcareous platform that covered all the Three localities with Early Cretaceous volcanism have been
Central Terrane and Oaxaquia. identified in the western Mixteca Terrane near the contact with
Guerrero Composite Terrane of western Mexico 299

the Guerrero Composite Terrane: the Taxco Schist and the Chapo- Tejupilco volcanic-sedimentary sequence might represent an arc
lapa and Zicapa Formations (Fries, 1960; de Cserna and Fries, assemblage older than the rest of the Guerrero Terrane magma-
1981; Talavera-Mendoza, 1993; Campa and Iriondo, 2003; Fitz tism. They based this conclusion on U-Pb dates from associ-
et al., 2002). The Taxco Schist is made up of submarine andesitic ated sulfide deposits. The ages they obtained vary broadly from
to rhyolitic lava flows and tuffs interbedded with epiclastic rocks Carnian (227 Ma) to Oxfordian (156 Ma). However, the same
and quartz-rich sandstone and shale (de Cserna and Fries, 1981; volcanic-sedimentary rocks were considered a part of the Cre-
Talavera-Mendoza, 1993). Its volcanic rocks have a continental- taceous arc assemblage by other authors (Campa and Ramirez,
arc geochemical affinity, more evolved than contemporaneous 1979; Talavera-Mendoza et al., 1995).
magmatism from the Guerrero Composite Terrane (Talavera- The arc assemblage of the Teloloapan Terrane consists of
Mendoza; 1993; Centeno-García et al., 1993a). The Zicapa For- two distinct regions with different volcanic and sedimentary
mation is made up of dacitic to rhyolitic lava flows interbedded rocks. The eastern region is characterized by shallow-marine
with fluvial deposits (Fitz et al., 2002). The Chapolapa Formation deposits, and the western region is composed of deeper facies
is composed mostly of marine lava flows and epiclastic rocks. (Guerrero-Suastegui et al., 1991; Ramírez-Espinoza et al., 1991;
The abundance of quartzites within the volcanic-sedimentary Talavera-Mendoza et al., 1995; Mendoza and Suastegui, 2000;
successions of the Taxco Schist and Zicapa Formation suggests Guerrero-Suastegui, 2004).The stratigraphy of the eastern region,
that a crystalline basement was exposed during the arc activity. from base to top, is made up of basaltic to andesitic pillow and
U/Pb dating of lavas from the Taxco Schist by sensitive massive lava flows, volcanic breccias, and pyroclastic flow depos-
high-resolution ion microprobe (SHRIMP) methods yielded its (Villa de Ayala Formation; Talavera-Mendoza et al., 1995).
130–131 Ma ages (Campa and Iriondo, 2004), and from the These deposits are interbedded with epiclastic sandstone and
volcanic-volcaniclastic rocks of the Zicapa Formation, 127 Ma conglomerate. Primary structures in the volcaniclastic rocks sug-
(Fitz et al., 2002). Lava flows from the Chapolapa Formation gest a marine depositional environment (Guerrero-Suastegui et
have 129–133 Ma SHRIMP U/Pb ages. The Taxco Schist shows al., 1991; Guerrero-Suastegui, 2004). Storm deposits, coral frag-
one phase of deformation and metamorphism prior to the deposi- ments, and other fossils suggest shallow and warm waters. This
tion of Aptian–Albian carbonates. Thus, Early Cretaceous vol- unit contains fossil gastropods and bivalves that range in age from
canic rocks of the Mixteca Terrane are unconformably covered Hauterivian to Aptian (Guerrero-Suastegui et al., 1991; Ramírez-
by a carbonate platform that ranges in age from Early to middle Espinoza et al., 1991; Talavera-Mendoza et al., 1995).
Cretaceous (Fries, 1960). Geochemical analyses of volcanic rocks of the Villa de Ayala
The limestone succession in the western Mixteca Terrane Formation of the Teloloapan Terrane indicate that the magma-
changes upward to a thick clastic succession (Mexcala Forma- tism is calc-alkaline and similar to that of active intraoceanic arcs
tion) of Turonian to Maastrichtian age (Guerrero-Suastegui, (Talavera-Mendoza, 1993; Talavera-Mendoza et al., 1995; Lapi-
2004). The Mexcala Formation is made up of alternating sand- erre et al., 1992; Mendoza and Suastegui, 2000; Centeno-García
stone, shale, and conglomerate, deposited in deltaic and subma- et al., 1993a). The base of the Villa de Ayala Formation is not
rine-fan environments (Figs. 7 and 8). It is a synorogenic deposit exposed. The maximum thickness is considered to be ~3000 m
(foreland basin-fill) associated with regional thrusting and fold- (Guerrero-Suastegui, 2004). The volcanic succession of this for-
ing of both the Guerrero Composite and Mixteca Terranes at the mation changes transitionally upward to thick, massive reefal
end of the Cretaceous. Therefore, the Mexcala Formation is the limestone of the Teloloapan Formation. At the base the Teloloa-
first overlapping assemblage that stitches the Guerrero Composite pan Formation is composed of intertidal limestone interbedded
Terrane and the Mixteca Terrane, and marks the final amalgama- with volcaniclastic rocks containing rudists and nerineas of late
tion of the Guerrero Composite Terrane to continental Mexico. Aptian–early Albian age (Guerrero-Suastegui et al., 1991, 1993;
Guerrero-Suastegui, 2004). Thus magmatism ceased prior to the
Teloloapan Terrane late Aptian (Guerrero-Suastegui et al., 1991; Mendoza and Sua-
The Teloloapan Terrane (Figs. 1 and 8) is exposed in the east- stegui, 2000; Guerrero-Suastegui, 2004). The Teloloapan Forma-
ernmost parts of the Guerrero Composite Terrane. This terrane is tion grades upward into the Pachivia Formation of Turonian age,
characterized structurally by a complex east-vergent thrust-fault which is made up of shale and fine-grained sandstone and shale.
system. Its rocks are severely deformed and metamorphosed to The Pachivia Formation is the western equivalent of the Mexcala
low-grade greenschist facies. The Teloloapan Terrane overrides Formation of the Mixteca Terrane and indicates that the Telo-
either Cretaceous platform carbonates or Upper Cretaceous silic- loapan and Mixteca Terranes were already in close proximity
iclastic rocks that belong to the Mixteca Terrane (Figs. 7 and 8; (Guerrero-Suastegui et al., 1991; Talavera-Mendoza et al., 1995;
Talavera-Mendoza et al., 1995). Guerrero-Suastegui, 2004).
The nature of the basement of the Teloloapan Terrane The stratigraphy of the western part of the Teloloapan Ter-
remains unknown. Metamorphic rocks of the Tejupilco area rane comprises submarine basaltic, andesitic, and felsic lava
(Fig. 8) were interpreted as a possible basement for the Telo- flows and volcaniclastic rocks (Villa de Ayala Formation) depos-
loapan Terrane by Elías-Herrera and Sánchez-Zavala (1990), ited in deeper water conditions than the sediments of the eastern
and Sanchez-Zavala (1993). These authors suggested that the Teloloapan Terrane. It is in transitional contact upsection with
300 Centeno-García et al.

the Acapetlahuaya Formation, composed of thin-bedded volca- The arc volcanism of the Mixteca and Teloloapan Terranes
nic shale and sandstone at the base, at some localities interbed- has been interpreted as part of a single arc-backarc system in
ded with dark, thinly laminated limestone. It changes transition- which volcanism of the Mixteca Terrane would be the backarc
ally upward to shale, with little or no volcanic material at the basin (Cabral-Cano et al., 2000; Monod et al., 1994). An alter-
top (Campa and Ramirez, 1979; Guerrero-Suastegui, 2004). This native interpretation is that these two terranes belong to differ-
unit has been highly tectonized, making it difficult to calculate ent arcs, separated by a double-dipping subduction of an oceanic
its original thickness and contact relationships. Apparently, the basin (Guerrero-Suastegui, 2004).
Acapetlahuaya Formation changes laterally toward the west and
overlies transitionally the volcaniclastic deposits of the Villa de Arcelia Terrane
Ayala Formation. Its upper contact with the Amatepec Forma- Thrust over the Teloloapan Terrane is the Arcelia Terrane
tion is highly tectonized. The Acapetlahuaya Formation contains (Guerrero Composite Terrane), which shows deeper marine
ammonoids, and radiolarians that are late Aptian in age (Campa facies and less evolved magmatism than the rest of the arc suc-
et al., 1974; Guerrero-Suastegui et al., 1993; Talavera-Mendoza cessions of the Guerrero Composite Terrane (Talavera-Mendoza
et al., 1995; Guerrero-Suastegui, 2004). et al., 1995; Mendoza and Suastegui, 2000). This terrane is made
The Amatepec Formation is made up of thin-bedded black up of basaltic pillow lavas and ultramafic bodies, black shale and
detrital limestone and is devoid of volcanic material. It is inter- chert, and fine-grained volcanic turbidites (Fig. 7), all intensively
preted as deep-basin–slope deposits. This formation is tightly deformed and partly metamorphosed (Ramírez-Espinoza et al.,
folded and overlies either the Villa de Ayala or the Acapetlahuaya 1991; Talavera-Mendoza et al., 1995). The chert layers contain
Formation. It is late Albian to early Cenomanian in age, based on radiolarians reported as Albian–Cenomanian in age (Dávila and
calcispherulids, planktonic foraminifers, and radiolarians (Campa Guerrero, 1990; Ramírez-Espinoza et al., 1991). Ar/Ar and K/Ar
and Ramirez, 1979; Guerrero-Suastegui et al., 1991, 1993; ages (93.4–105 Ma; Delgado et al., 1990; Ortiz and Lapierre,
Talavera-Mendoza, 1993; Talavera-Mendoza et al., 1995). The 1991; Elías-Herrera, 1993) are compatible with biochronology,
deep-marine limestone is overlain by turbiditic sandstone-shale but detrital zircon ages from volcanic turbidites are older (mean
successions of the Miahuatepec Formation (Talavera-Mendoza et age, 130 Ma; Talavera-Mendoza et al., 2007). Geochemical sig-
al., 1995). Fossils have not been found, but it is at least post–early natures of the Arcelia magmas are similar to those in recent prim-
Cenomanian because of its stratigraphic position. The Miahuate- itive IAs and oceanic basins (MORB) (Talavera-Mendoza, 1993;
pec Formation was deposited, during the amalgamation of the Talavera-Mendoza et al., 1995; Mendoza and Suastegui, 2000).
Zihuatanejo, Arcelia, and Teloloapan Terranes, in a thrust-related There are no exposures of older rocks in the Arcelia Terrane, and
basin (Guerrero-Suastegui et al., 1991; Ramírez-Espinoza et al., no clasts of older metamorphic or sedimentary rocks have been
1991; Talavera-Mendoza et al., 1995; Mendoza and Suastegui, found in its sedimentary strata. Mendoza and Suastegui (2000)
2000; Guerrero-Suastegui, 2004). suggest that this terrane is entirely oceanic, that it may have origi-
The ages of magmatism of the Teloloapan Terrane have nated as an independent oceanic arc and backarc basin, and that
been poorly constrained by the limited fossils found in the vol- it represents partly developed oceanic crust. An alternative inter-
caniclastic levels. A few U/Pb isotopic ages from felsic lavas at pretation is that the Arcelia Terrane could also be a backarc basin
the base of the succession range in age from 137.4 to 145.9 Ma of the Zihuatanejo Terrane (Centeno-García et al., 2003a).
(Tithonian–Hauterivian; Mortensen et al., 2003). Thus, magma-
tism of the Teloloapan Terrane is in part contemporaneous with Southern Part of the Zihuatanejo Terrane
that of the Mixteca, Guanajuato, and Zihuatanejo Terranes. Uppermost Jurassic–Cretaceous volcanic-sedimentary assem-
There are three distinctive differences in the Cretaceous stra- blages of the Zihuatanejo Terrane can be grouped in three main
tigraphy between the Mixteca and Teloloapan Terranes: (1) Vol- regions: northern Zihuatanejo Terrane (Zacatecas area), Huetamo
canism of the Mixteca Terrane is more evolved, and its isotopic area, and coastal Zihuatanejo-Colima region (Figs. 7 and 8). The
signatures show influence of old continental crust in the magma uppermost Jurassic to Cretaceous strata of the southern Zihuatanejo
generation. In contrast, volcanism of the Teloloapan Terrane is Terrane are not as strongly deformed as those in other terranes, and
more primitive and has no traces of contamination by old con- original contact relationships and complete stratigraphic columns
tinental crust (Centeno-García et al., 1993a; Talavera-Mendoza are well preserved. The strata are characterized by numerous lat-
et al., 1995; Mendoza and Suastegui, 2000). (2) Metamorphic eral facies changes and internal erosional and angular unconformi-
and quartz clasts are abundant (up to 70%) in the sandstones that ties. The geographic distribution of the facies is highly irregular,
are interbedded with volcanic rocks in the Mixteca Terrane but and it has not yet been determined in detail. Therefore, compiling,
are absent throughout the stratigraphic column of the Teloloapan correlating, and synthesizing the stratigraphy of the area is difficult
Terrane. (3) Magmatism ceased in the Mixteca Terrane before because it varies considerably from one locality to another. The
the Aptian, and part of the volcanic-sedimentary succession was stratigraphy of the northern Zihuatanejo Terrane (Zacatecas area)
deformed and metamorphosed (Taxco Schist). In contrast, vol- is described later.
canism continued in the Teloloapan Terrane until Aptian–Albian The stratigraphic column of the southern Zihuatanejo Ter-
time, and no internal deformation has been identified. rane in the Huetamo area is made up of Triassic basement rocks
Guerrero Composite Terrane of western Mexico 301

of the Arteaga Complex, overlain by uppermost Jurassic to Cre- and ignimbrites) and minor andesitic and dacitic lavas (Tecalitlán
taceous volcanic and sedimentary cover. These rocks are thrust Formation, Titzupa-La Unión assemblage, Playitas Formation,
over the Arcelia Terrane (Figs. 7 and 8). Arc-related rocks of the etc.; Ferrusquía et al., 1978; Grajales and López, 1984; Pantoja
Huetamo region (Figs. 7 and 8) overall have been formed by a and Estrada, 1986; Centeno-García et al., 2003). These units are
thick succession of alternating shale, sandstone, and conglom- interbedded with epiclastic deposits such as tuff, volcanic shale,
erate, with scattered basaltic pillow lavas, submarine ignimbrite and sandstone, and some conglomerate. The assemblage also
flows, and other intermediate pyroclastic and epiclastic flows in contains thin beds of limestone containing orbitolinids, gastro-
the lower parts of the succession (Angao and San Lucas For- pods, and some pelecypods of late Albian–Cenomanian age (Fer-
mations; Pantoja, 1959; Guerrero-Suastegui, 1997). These arc- rusquía et al., 1978; Grajales and López, 1984). Raindrop marks,
related rocks lie unconformably on the Arteaga Complex (Figs. 7 desiccation polygons, and dinosaur footprints can be found in this
and 8). Fossils of Late Jurassic age have been reported from the succession (Ferrusquía et al., 1978). The lower parts of the Cre-
Angao Formation (Pantoja, 1959), although the major exposures taceous succession are missing in the Arteaga region, where non-
of volcaniclastic rhythmic sedimentary rocks are Berriasian to marine and shallow-marine volcanic and volcaniclastic rocks of
upper Aptian (Guerrero-Suastegui, 1997). Their depositional Aptian–Albian age rest unconformably on the Arteaga Complex.
environment changes upsection from deep to shallow marine. Overall, Cretaceous volcanic rocks of the southern Zihua-
The volcaniclastic rocks of the San Lucas Formation change tanejo Terrane show geochemical and isotopic signatures that
toward the top to thick limestone zones with fossil ammonites, suggest a transitional composition between oceanic island arcs
orbitolinids, and rudists of late Aptian–early Albian age (El and active continental margins (Centeno-García, 1994; Freydier
Cajon and Mal Paso Formations; Guerrero-Suastegui, 1997; et al., 1997; Mendoza and Suastegui, 2000). The high potassium
Pantoja-Alor and Caballero, 2003). This sequence alternates with content, abundance of felsic lavas, and trace element abundances
or changes laterally into marine and terrestrial volcanic sandstone of these volcanic rocks are similar to those observed in IAs where
and conglomerate (Comburindio Formation; Guerrero-Suastegui, the crust is thick (>~20 km), allowing magmatic differentiation
1997; Pantoja-Alor and Caballero, 2003). The conglomerate is (Centeno-García, 1994).
covered by massive, thick packets of limestone (Huetamo For- Rocks of the southern Zihuatanejo Terrane are distinctive
mation) that contain fossils of late Albian–Cenomanian age. This from the rest of the terranes because they were deposited in
unit is found only in the central parts of the Huetamo region (Pan- shallow-marine and fluvial environments, contain fossil verte-
toja, 1990). brates, and show calc-alkaline volcanism more evolved than that
The arc succession of the Zihuatanejo Terrane in the Huetamo of the Teloloapan and Arcelia Terranes. Sedimentary rocks inter-
area was deformed prior to the deposition of a thick, subaerial bedded with the volcanic flows contain clasts of their basement
red-bed succession that is interbedded with volcanic rocks (Cut- rocks, made up of sandstone, quartz, and mylonitic granite. Thus
zamala Formation of Campa and Ramirez, 1979) and is related its stratigraphy is similar to that of arcs constructed on intermedi-
to a continental arc of Santonian–Maastrichtian age (Altamira- ate crust with a previous history of accretions. The presence of
Areyán, 2002; Benammi et al., 2005). fossil vertebrates suggests proximity to the continent.
The oldest Cretaceous rocks of the Zihuatanejo Terrane in
the Zihuatanejo-Colima region of coastal Mexico that have been Northern Guerrero Terrane
penetrated by drilling are Berriasian–Hauterivian in age (Alberca
Formation; Cuevas, 1981). The lower member of the Alberca Following a section from east to west in the northern part of
Formation is made up of interbedded black shale, sandstone, and the Guerrero Terrane, the main stratigraphic characteristic is an
limestone, and some tuff. The upper member is composed mostly absence of rocks similar to those of the Mixteca or Teloloapan Ter-
of andesitic-basaltic lava flows interbedded with limestone and rane. Instead, deep-marine volcanic-sedimentary successions of
shale. The Alberca Formation changes transitionally upward to the Guanajuato Terrane were thrust directly over limestone of the
andesitic and basaltic lava flows, with some rhyolitic flows, inter- calcareous platform of Oaxaquia. Contact relationships between
bedded with pyroclastic (intermediate tuffs and ignimbrites) and the Guanajuato Terrane and the northern Zihuatanejo Terrane are
epiclastic deposits. It contains limestone packets interbedded with unconstrained because the contact is covered by younger units. It
subaerial conglomerate and sandstone, red siltstone, and some is inferred that the Guanajuato Terrane is overthrust by the Zihua-
evaporites, and continues into limestone with scarce basaltic pil- tanejo Terrane on the basis of regional vergence of the structures.
low lavas at the top (Tecalitlán, Tepalcatepec, and Madrid Forma- Contact relationships between the Tahue and Zihuatanejo Ter-
tions). The age range of these units, based on their fossil content, ranes are unknown because the contact is covered by overlapping
is Barremian to Cenomanian (Grajales and López, 1984). Cenozoic assemblages.
Along the west coast between the cities of Colima and
Zihuatanejo are exposures of an important succession of red Guanajuato Terrane
beds, alternating with lesser amounts of limestone in compari- The succession at the Guanajuato Terrane has been described
son with other areas of the Guerrero Composite Terrane. The as a complete stratigraphic column of an accreted volcanic arc,
assemblage is made up of rhyolitic lavas (lava flows, breccias, as its assemblages vary from the roots of the arc (gabbros and
302 Centeno-García et al.

diabases, and dike swarms) to pillow basalts, interbedded with unconformably on the Arperos Formation and suggests that sedi-
thin-bedded siltstone, shale, chert, and fine-grained volcanic mentation and at least one phase of deformation occurred prior
sandstone (Figs. 6 and 7; Ortiz-Hernandez et al., 1991; Ortiz- to the Aptian–Albian (Ortiz-Hernandez et al., 2003). Whether or
Hernandez, 1992). However, all the different stratigraphic levels not this deformation is related to the accretion of the Guanajuato
are in the form of tectonic slivers (Fig. 7), with the deepest mafic Terrane to the continental margin has not been determined. At
levels (gabbro, tonalite, serpentinite, wehrlite, dike swarms) present the Guanajuato Terrane is thrust over the calcareous
thrust over the upper stratigraphic levels (pillow basalt and vol- platform of Oaxaquia in the San Miguel de Allende area (Ortiz-
canic turbidites). Hernández et al., 2002).
The uppermost thrust sheet is made up of ultramafic-mafic Rocks of the Guanajuato Terrane have been correlated with
rocks of the Cerro Pelón tonalite and the Tuna Mansa diorite. the Arcelia Terrane, and both were interpreted as having formed
These ultramafic rocks are thrust over a succession incorporat- part of an oceanic arc independent of the Zihuatanejo and other
ing a diabasic feeder dike swarm, basaltic pillow lavas (La Luz arc terranes (Ortiz-Hernandez et al., 1992). Also, these rocks
basalts), rhyolitic tuffs (Cubilete Tuff), and a deep-marine vol- are considered relicts of an oceanic basin consumed by subduc-
caniclastic succession made up of sandstone and shale turbi- tion related to the arc of the Zihuatanejo Terrane (Lapierre et al.,
dites, chert, and black detrital limestone (Esperanza Formation; 1992; Tardy et al., 1994). An alternative preliminary interpreta-
Quintero-Legorreta, 1992; Ortiz-Hernandez et al., 1992; Ortiz- tion, based on provenance and stratigraphy, is that the Guanajuato
Hernandez et al., 2003). Basalts of this assemblage show geo- Terrane may have been the backarc basin of the Zihuatanejo Ter-
chemical signatures similar to present primitive volcanic island rane (Centeno-García et al., 2003).
arcs (Ortiz-Hernandez, 1992).
The third and lowermost structural level (Fig. 7) is com- Zihuatanejo Terrane
posed of a thick turbidite succession of volcanic graywackes, The Upper Jurassic–Cretaceous stratigraphy of the Zacate-
quartzites, micritic limestone, radiolarian chert, black shale, cas area in the northern Zihuatanejo Terrane is very different than
and rare conglomerate resting on basaltic pillow lavas (Arperos the stratigraphy of the neighboring Central Terrane and Oaxa-
Formation; Ortiz-Hernandez et al., 1992; Lapierre et al., 1992; quia (Figs. 5–7). Whereas the strata in the northern Zihuatanejo
Quintero-Legorreta, 1992; Monod et al., 1990; Martínez-Reyes, Terrane are mostly composed of volcanic and volcaniclastic
1992; Ortiz-Hernandez et al., 2003). Pillow basalts at the base rocks, northern Oaxaquia and the Central Terrane were covered
of the Arperos Formation are more alkaline than the La Luz by a thick, shallow-marine calcareous platform during the Late
basalts and show OI geochemical signatures (Ortiz-Hernandez Jurassic–Cretaceous (Centeno-García and Silva-Romo, 1997).
et al., 2003). The Arperos Formation is unconformably overlain This suggests that the Zihuatanejo Terrane was probably under-
by the Aptian–Albian La Perlita Limestone (Ortiz-Hernandez et going dislocation from the continental margin during that time.
al., 2003). The arc stratigraphy of the Zacatecas area is formed by the
It is difficult to reconstruct the role of the Guanajuato Ter- La Borda, Chilitos, and El Saucito Formations (de Cserna, 1976;
rane in the tectonic evolution of western Mexico because of the Yta et al., 1990; Olvera-Carranza et al., 2001; Olvera-Carranza,
lack of enough geochronological data. The only U/Pb zircon 2002). These three formations are made up of pillow basalts and
age reported from the area comes from the El Gordo volcano- volcanic breccias, interbedded with thin-bedded siltstone, shale,
genic massive sulfide ore deposit (Hall and Mortensen, 2003), chert, and volcanic sandstone and conglomerate, with scarce fel-
which is considered part of the lowermost succession by Hall sic tuff beds and detrital limestone (Centeno-García and Silva-
and Mortensen, (2003), but it is at the stratigraphic level of the Romo, 1997; Olvera-Carranza, 2002). The chert layers contain
second thrust sheet (Cubilete tuff?) in the stratigraphy proposed radiolarian fossils of Neocomian(?) to Aptian–Albian(?) age (Yta
by Ortiz-Hernandez et al. (1992). The age of a rhyolite from El et al., 1990; Olvera-Carranza, 2002). However, older U/Pb ages
Gordo volcanogenic massive sulfide ore deposit reported by Hall have been reported (150–148 Ma) from the base of the succes-
and Mortensen (2003) yielded a 146.1 Ma U/Pb age. There are sion (Danielson, 2000; Mortensen et al., 2003). Lapierre et al.
also reports of badly preserved radiolarians from the Arperos For- (1992) and Freydier et al. (1995) characterized this magmatism
mation that are not in good enough condition to be age indicators as primitive IA and OI basalts. Sedimentary structures and fos-
(possibly Valanginian–Turonian in age), but a report of nanno- sil content suggest that the La Borda, El Saucito, and Chilitos
fossils suggests a Tithonian–Hauterivian age (Ortiz-Hernandez Formations were deposited as distal turbidites and grain flows in
et al., 2003). Other ages reported from the Guanajuato area are a volcaniclastic submarine apron (Centeno-García et al., 2003).
from K/Ar analyses and seem to have been reset by later thermal These Jurassic–Cretaceous arc successions contain important
events (Ortiz-Hernandez et al., 1992, 2003). The sedimentary volcanogenic, massive sulfide ore deposits (Yta et al., 1990; Dan-
rocks of the La Luz and Arperos Formations seem to be distal ielson, 2000; Mortensen et al., 2003).
volcanic turbidite deposits, but the abundance of limestone asso-
ciated with the pillow lavas suggests that deposition occurred Tahue Terane
above the carbonate compensation depth (Ortiz-Hernandez et al., Cretaceous successions of the Tahue Terrane are exposed
2003). Aptian–Albian limestone of the La Perlita Formation rests mostly in the Sinaloa de Leyva–Porohui region (Fig. 4). They
Guerrero Composite Terrane of western Mexico 303

were formed by submarine pillow lavas, volcaniclastic rocks, • The siliciclastic rocks of the Potosi Fan extended to the
shale, and limestone. They contain Albian ammonites (Ortega- west in a marginal oceanic basin (Arteaga Basin) that at
Gutiérrez et al., 1979; Freydier et al., 1995; Gastil et al., 1999), present forms the basement of the Zihuatanejo Terrane of
but Ar/Ar ages from the lavas are younger (86 Ma; Gastil et al., the Guerrero Composite Terrane.
1999), suggesting resetting. The basaltic lavas show MORB and • The first compressional event that deformed the Triassic
OIB geochemical affinities, but the volcaniclastic rocks are more rocks originated tight folding, shearing, and axial cleavage
evolved and show IA geochemical signatures (Freydier et al., in the La Ballena Formation, and block-in-matrix texture
1995; Gastil et al., 1999). Although this volcanic succession has in the Taray and Zacatecas Formations and the Arteaga
been interpreted as the northern continuation of the Arperos For- Complex. This deformation was related to subduction
mation of the Guanajuato Terrane, and part of a major oceanic along the early Mesozoic continental margin. It may have
basin that originally lay between the Guerrero arc and the con- started sometime between the Late Triassic and Early
tinent (Tardy et al., 1994; Lapierre, et. al., 1992; Dickinson and Jurassic, accreting the turbidites of the Potosi Submarine
Lawton, 2001), the stratigraphy does not support such a tectonic Fan, with slivers of the oceanic crust, to the continent.
scenario because (1) the Cretaceous volcanic rocks rest uncon- • Whether the subducting slab was dipping toward the west
formably on a Paleozoic basement, (2) the stratigraphy and facies or the east is not well constrained, but the accretionary
associations are not indicative of deep-pelagic sedimentation and prism apparently was very wide. Evidence of contempo-
oceanic-ridge volcanism, and (3) the Guanajuato successions raneous oceanic arc magmatism is found in the Vizcaíno
apparently are older than the arc assemblages of the Tahue and Peninsula, where a volcanic sequence of primitive arc affin-
other parts of the Guerrero Composite Terrane. ity is exposed. It is possible that the rocks in the Vizcaino
Peninsula represent a displaced fragment of an oceanic arc
SUMMARY that accreted to the Arteaga Complex of the Guerrero Com-
posite Terrane, but this model needs more evidence.
• The stratigraphy of the Guerrero Composite Terrane of • Arc-related volcanic and sedimentary rocks unconform-
western Mexico is characterized by a series of terranes ably overlie the deformed Triassic rocks of Oaxaquia and
whose basements were formed by Paleozoic to Triassic the Central and Guerrero Composite Terranes. They are
fragments of oceanic arcs, continental slope sediments, characterized by continental rhyolitic to andesitic lava
and ocean floor assemblages that were accreted to the con- flows, interbedded with fluvial and alluvial deposits. The
tinent and consecutively rifted and translated. succession shows minor angular unconformities, probably
• Metamorphosed Ordovician volcanic and marine sedimen- related to tilting. These rocks have been interpreted as the
tary rocks and a thick succession of deep-marine turbidites southern continuation of the Jurassic continental arc that
of the NW Guerrero Composite Terrane (Tahue Terrane) developed along the southwestern margin of the United
make up the record of a middle Paleozoic collision and States. Magmatism was active from ca. 163 to 155 Ma
development of a Carboniferous to Permian passive mar- (Callovian–Oxfordian), although older volcanic rocks
gin. These rocks might be equivalent to the early Paleozoic have been reported for eastern Mexico (189 Ma). The
Antler Arc and eugeoclinal sedimentation in the SW Cor- Jurassic arc shows more evolved geochemical signatures
dillera of North America. than the subsequent volcanic events.
• The continental margin during the early Mesozoic was • During and after the continental arc activity (Late
located in the middle of Mexico, approximately along the Jurassic–Early Cretaceous), large amounts of extension
boundary between Oaxaquia and the Central–Guerrero and lateral translations probably occurred, as suggested by
Composite Terranes. This continental margin was active the changes in the stratigraphy. It has been proposed that
during the Permian–Carboniferous, when a continental arc major strike-slip faults were probably active during the
developed in Oaxaquia. arc activity (Mojave-Sonora Megashear). Arc magmatism
• Permian–Carboniferous arc-related magmatism ceased, ceased in central Mexico, and considerable subsidence
and a passive or rifted margin developed along the west- and extension is evidenced by the fast deepening of the
ern continental margin of Mexico, extending throughout calcareous platform that developed over the arc rocks.
the Triassic. This development is suggested by the thick • Major stratigraphic, geochemical, and isotopic differences
submarine siliciclastic turbidite succession that accumu- are evident in the different Cretaceous stratigraphic assem-
lated on the western paleo-continental shelf–slope region blages among the Guerrero terranes. They are, from east to
(Potosi Submarine Fan). The siliciclastic fan turbidites are west: Andesitic-basaltic submarine lava flows and tuff (IA
mostly continent-derived, quartz-rich sandstone, siltstone, geochemical signatures), interbedded with limestone and
and shale, containing fossils of Carnian–Norian age. shallow-marine volcaniclastics (Teloloapan Terrane) that
• The Potosi Fan is interpreted as passive-margin deposits, were thrust over contemporaneous but more evolved arc
as there is no evidence of contemporaneous magmatism successions and the calcareous platform of southern con-
either in the stratigraphy or in the provenance. tinental Mexico (Mixteca Terrane). Ophiolite successions,
304 Centeno-García et al.

with deep-marine volcanic and sedimentary rocks with Anderson, T.H., McKee, J.W., and Jones, N.W., 1991, A northwest trending,
MORB, OIB, and IA signatures (Guanajuato and Arcelia Jurassic fold nappe, northernmost Zacatecas, Mexico: Tectonics, v. 10,
p. 383–401.
Terranes), are placed between the continent and the more Anderson, T.H., Jones, N.W., and McKee, J.W., 2005, The Taray Formation:
evolved arc in the north (Zihuatanejo Terrane) and between Jurassic(?) mélange in northern Mexico—Tectonic implications, in
the two shallow-marine arcs (Teloloapan and Zihuatanejo Anderson, T.H., et al., eds., The Mojave-Sonora Megashear Hypothesis:
Development, Assessment, and Alternatives: Geological Society of Amer-
Terranes) in the south. ica Special Paper 393, p. 427–455.
• These major geological differences suggest that intra-arc Araujo, M.J., and Arenas, P.R., 1986, Estudio tectonico-sedimentario, en el mar
rifting was considerable and originated a series of marginal mexicano Estados de Chihuahua y Durango: Boletín Sociedad Geológica
Mexicana, v. 47–2, p. 43–71.
arc-backarc systems in western Mexico, with complex Arredondo-Guerrero, P., and Centeno-García, E., 2003, Geology of the Mazat-
paleogeography. Two possible scenarios can be proposed for lán region, southern Sinaloa state, Mexico: Geological Society of America
the Cretaceous paleogeography of western Mexico: (1) that Abstracts with Programs, v. 35, no. 4, p. 71.
Barboza-Gudiño, J.R., Tristan-Gonzalez, M., and Torres-Hernandez, J.R.,
there was one single rifting arc, with westward migration 1998, The Late Triassic–Early Jurassic active continental margin of west-
of the magmatism and development of deep-marine intra- ern North America in northeastern Mexico: Geofisica Internacional, v. 37,
arc and backarc basins (Guanajuato and Arcelia Terranes); p. 283–292.
Barboza-Gudiño, J.R., Tristan-Gonzalez, M., and Torres-Hernandez, J.R.,
and (2) that rifting during the end of the Jurassic was large 1999, Tectonic setting of pre-Oxfordian units from central and northeast-
enough to allow the formation of multiple marginal island ern Mexico: A review, in Bartolini, C., et al., eds., Mesozoic Sedimentary
arcs, separated by oceanic backarc basins. and Tectonic History of North-Central Mexico: Geological Society of
America Special Paper 340, p. 197–210.
• The proposed timing of the final amalgamation of the Barboza-Gudiño, J.R., Hoppe, M., Gómez-Anguiano, M., and Martínez-
Guerrero terranes to the margin of older terranes that form Macías, P.R., 2004, Aportaciones para la interpretación estratigráfica y
the eastern part of Mexico is Turonian to Maastrichtian, estructural de la porción noroccidental de la Sierra de Catorce, San Luis
Potosí, México: Universidad Nacional Autónoma de México, Instituto de
as suggested by the age span of foreland basins associated Geología: Revista Mexicana de Ciencias Geológicas, v. 21, p. 299–319.
with the deformation of the arc. Overlapping the previ- Bartolini, C., Cantú-Chapa, A., Lang, H., and Barboza-Gudiño, R., 2002, The
ously deformed Arcelia and Zihuatanejo Terranes, a new Triassic Zacatecas Formation in Central Mexico, paleotectonic, paleo-
geographic and paleobiogeographic implications, in Bartolini C., et al.,
arc developed along the coast by Santonian time. eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins
and Petroleum Systems: American Association of Petroleum Geologists
ACKNOWLEDGMENTS Memoir 75, p. 295–315.
Benammi, M., Centeno-García, E., Martinez-Hernandez, E., Morales-Gamez,
M., Urrutia-Fucugauchi, J., and Tolson, G., 2005, Presencia de dinosau-
This paper is a contribution to PAPIIT projects IN109605–3 rios en la Barranca Los Bonetes en el sur de México (Región de Tiquicheo,
and IN116599, funded by the Universidad Nacional Autónoma Estado de Michoacán) y sus implicaciones cronoestratigráficas: Universi-
dad Nacional Autónoma de México, Instituto de Geología: Revista Mexi-
de México (UNAM), and to projects UC-MEXUS Exotic ver- cana de Ciencias Geológicas, v. 22, p. 429–435.
sus Fringing Arc Models: Implications for the Growth of Con- Bissig, T., Mortensen, J.K., and Hall, B., 2003, The volcano-sedimentary set-
tinents, and SEP/2003 C02 42642. Special thanks are due J. ting of the Kuroko type Vhms District of Cuale, Jalisco, Mexico: Geologi-
cal Society of America Abstracts with Programs, v. 35, no. 4, p. 61.
Stock, C. Busby, C. Vita-Finzi, and A.E. Draut for their reviews Burchfiel, B.C., Cowan, D.S., and Davis, G.A., 1992, Tectonic overview of the
and comments, which greatly improved the paper. Cordilleran orogen in the western United States, in Burchfiel, B.C. et al.,
eds., The Cordilleran Orogen, Conterminous United States: Boulder, Col-
orado, Geological Society of America, The Geology of North America, v.
REFERENCES CITED G-3, p. 407–479.
Altamira Areyán, A., 2002, Las litofacies y sus implicaciones de la cuenca sedi- Burckhardt, C., and Scalia, S., 1906, Géologie des environs de Zacatecas:
mentaria Cutzamala-Tiquicheo, Estado de Guerrero y Michoacán, México Mexico, International Geological Congress, 10th, Excursion Guidebook
[M.S. thesis]: México, Universidad Nacional Autónoma de México, Insti- 16, 26 p.
tuto de Geología, 79 p. Cabral-Cano, E., Lang, H.R., and Harrison, C.G.A., 2000, Stratigraphic assess-
Anderson, C.E., 2006, U-Pb ages of detrital zircons in the Holbrook Member ment of the Arcelia–Teloloapan area, southern Mexico: Implications for
of the Moenkopi Formation near Winslow, Arizona: Geological Society of southern Mexico’s post-Neocomian tectonic evolution: Journal of South
America Abstracts with Programs, v. 38, no. 5, p. 10. American Earth Sciences, v. 13, p. 443–457.
Anderson, T.H., and Silver, L.T., 1979, The role of the Mojave-Sonora Campa, M.F., and Coney, P.J., 1983, Tectono-stratigraphic terranes and mineral
megashear in the tectonic evolution of northern Sonora, in Anderson, T.H., resource distributions in Mexico: Canadian Journal of Earth Sciences,
and Roldán-Quintana, J., eds., Geology of Northern Sonora: San Diego, v. 20, p. 1040–1051.
California, Geological Society of America Annual Meeting, Guidebook, Campa, M.F., and Iriondo, A., 2003, Early Cretaceous protolith ages for metavol-
Field Trip 27, p. 59–68. canic rocks from Taxco and Taxco Viejo in southern Mexico: Geological
Anderson, T.H., and Silver, L.T., 1981, An overview of Precambrian rocks in Society of America Abstracts with Programs, v. 35, no. 4, p. 71.
Sonora, México: Universidad Nacional Autónoma de México, Instituto de Campa, M.F., and Iriondo, A., 2004, Significado de dataciones Cretácicas de
Geología: Revista Mexicana de Ciencias Geológicas, v. 5, p. 131–139. los arcos volcánicos de Taxco, Taxco Viejo y Chapolapa, en la evolución
Anderson, T.H., and Silver, L.T., 2005, The Mojave-Sonora megashear—Field de la plataforma Guerrero-Morelos: Unión Geofísica Mexicana: Reunión
and analytical studies leading to the conception and evolution of the Nacional de Ciencias de la Tierra, GEOS, v. 24, p. 173.
hypothesis, in Anderson, T.H., et al., eds., The Mojave-Sonora Megashear Campa, M.F., and Ramirez, J., 1979, La Evolución Geológica y la Metalogéne-
Hypothesis: Development, Assessment, and Alternatives: Geological sis de Guerrero: Universidad Autónoma de Guerrero, Ser. Técnico-Cientí-
Society of America Special Paper 393, p. 1–50. fica, v. 1, 84 p.
Anderson, T.H., McKee, J.W., and Jones, N.W., 1990, Jurassic(?) melange in Campa, M.F., Campos, M., Flores, R., and Oviedo, R., 1974, La sequencia
north-central Mexico: Geological Society of America Abstracts with Pro- Mesozoíca volcanico-sedimentaria metamorfizada de Ixtapan de la Sal,
grams, v. 22, no. 3, p. 3. Mex.-Teloloapan, Gro.: Soc. Geol. Mex. Bull., v. 35, p. 7–28.
Guerrero Composite Terrane of western Mexico 305

Campa, M.F., Ramirez, J., and Bloome, C., 1982, La secuencia volcanico-sedi- Corona-Chávez, P., and Israde-Alcántara, I., 1999, Carta Geológica del Estado
mentaria metamorfizada del Triasico (Ladiniano-Carnico) de la region de de Michoacán: Map, scale 1:250,000.
Tumbiscatio, Michoacan: Sociedad Geológica Mexicana, VI Convención Cuevas, S.F., 1981, Prospecto Tpalcatepec: México, D.F., Petróleos Mexicanos,
Geológica Nacional, 6 Resúmenes, 48 p. IGPR-164, Internal report (unpublished).
Cantu-Chapa, A., 1969, Una nueva localidad Triásico Superior en México: Cuevas-Pérez, E., 1983, The geological evolution of the Mesozoic in the State
Revista Instituto Mexicano del Petróleo, v. 1, p. 71–72. of Zacatecas Mexico: Zentralblatt für Geologie und Palaeontologie, Teil
Carrillo-Martínez, 1971, Geología de la Hoja San José de Gracia, Sinaloa I: Allgemeine, Angewandte: Regionale und Historische Geologie, v. 1,
[senior thesis]: Mexico City, Universidad Nacional Autónoma de México, p. 190–201.
Facultad de Ingeniería, 154 p. Danielson, T.J., 2000, Age, paleotectonic setting, and common Pb isotope sig-
Centeno-García, E., 1994, Tectonic evolution of the Guerrero Terrane, western nature of the volcanogenic massive sulfide deposit, southeastern Zacate-
Mexico [Ph.D. thesis]: Tucson, University of Arizona, 220 p. cas State, central Mexico [M.S. thesis]: Vancouver, University of British
Centeno-García, E., 2002, Overview of Mesozoic subduction-related magmatic Columbia, 120 p.
events of Mexico and their relationship with the rest of the cordillera: Geo- Dávila, V.M., and Guerrero, M., 1990, Una edad basada en radiolarios para la
logical Society of America Abstracts with Programs v. 34, no. 5, p. A-97. secuencia volcánica-sedimentaria de Arcelia, Estado de Guerrero: Socie-
Centeno-García, E., 2005, Review of Upper Paleozoic and Lower Mesozoic dad Geológica Mexicana, 10th Convención Geológica Nacional, Libro de
stratigraphy and depositional environments of central and west Mexico: Resúmenes, p. 83.
Constraints on terrane analysis and paleogeography, in Anderson, T.H., De Cserna, Z., 1976, Geology of the Fresnillo area, Zacatecas, México: Geo-
et al., eds., The Mojave-Sonora Megashear Hypothesis: Development, logical Society of America Bulletin, v. 87, p. 1191–1199.
Assessment, and Alternatives: Geological Society of America Special De Cserna, Z., 1978, Notas sobre la geología de la región comprendida entre
Paper 393, p. 233–258. Iguala, Ciudad Altamirano y Temascaltepec, Estados de Guerrero y
Centeno-García, E., and Díaz-Salgado, C., 2002, Estratigrafía y Geoquímica México: Sociedad Geológica Mexicana, Libro Guía de la Excursión
de las rocas volcánicas de la Formación Huizachal en la Región de Aram- geológica a Tierra Caliente, p. 1–25.
berri, Estado de Nuevo León, III Reunión Nacional de Ciencias de la De Cserna, Z., and Fries, C., Jr., 1981, Hoja Taxco 14Q-h(7), and Resumen de
Tierra, Puerto Vallarta: Actas INAGEQ, v. 8, p. 244. la Geología de la Hoja Taxco, Estados de Guerrero, México y Morelos:
Centeno-Garcia, E., and Silva-Romo, G., 1997, Petrogenesis and tectonic Universidad Nacional Autónoma de México, Insituto de Geología, Carta
evolution of central México during Triassic-Jurrasic time: Universidad Geologica de Mexico: Map Ser., scale 1:100,000, and 47 p. text.
Nacional Autónoma de México, Instituto de Geología: Revista Mexicana Delgado, A.L., López, M.M., York, D., and Hall, C.M., 1990, Geology and geo-
de Ciencias Geológicas, v. 14, p. 244–260. chronology of ultramafic localities in the Cuicateco and Tierra Caliente
Centeno-García, E., Ruíz, J., Coney, P., Patchett, J.P., and Ortega, G.F., 1993a, Complexes, southern Mexico: Geological Society of America Abstracts
Guerrero Terrane of Mexico: Its role in the Southern Cordillera from new with Programs v. 22, no. 7, p. 326.
geochemical data: Geology, v. 21, p. 419–422. Delgado, L.A., 1982, Descripción preliminar de la geología y mecánica
Centeno-García, E., García-Díaz, J.L., Guerrero-Suastegui, M., Ramírez-Espi- de emplazamiento del complejo ultrabásico de Loma Baya, Guer-
noza, J., Salinas-Prieto, J.C., and Talavera-Mendoza, O., 1993b, Geology rero, México: Universidad Nacional Autónoma de México, Instituto de
of the southern part of the Guerrero Terrane, Ciudad Altamirano-Teloloa- Geofísica: Geofísica Internacional, v. 25, p. 537–558.
pan area, in Ortega, G.F., et al., eds., Terrane Geology of Southern México: Díaz Salgado, C., 2004, Caracterización tectónica y procedencia de la for-
Universidad Nacional Autónoma de México, Instituto de Geología, First mación Taray, región de Pico de Teyra, estado de Zacatecas [M.S. thesis]:
Circum-Pacific and Circum-Atlantic Terrane Conference, Gto. Mexico: Universidad Nacional Autónoma de México, Posgrado en Ciencias de la
Guidebook of Field Trip B, p. 22–33. Tierra, Instituto de Geología, 95 p.
Centeno-García, E., Olvera-Carranza, K., Corona-Esquivel, R., Camprubí, A., Diaz-Salgado, C., Centeno-García, E., and Gehrels, G., 2003, Stratigraphy,
Tritlla, J., and Sanchez-Martinez, S., 2003a, Depositional environment depositional environments, and tectonic significance of the Taray Forma-
and paleogeographic distribution of the Jurassic-cretaceous arc in the tion, northern Zacatecas state, Mexico: Geological Society of America
western and northern Guerrero Terrane, Mexico: GSA 99th Cordilleran Abstracts with Programs, v. 35, no. 4, p. 71.
Section Annual Meeting Abstracts with Programs, v. 35, n. 4, p. 76. Dickinson, W., 2004, Evolution of the North American Cordillera: Annual
Centeno-García, E., Corona-Chavez, P., Talavera-Mendoza, O., and Iriondo, Review of Earth and Planetary Sciences, v. 32, p. 13–45, doi: 10.1146/
A., 2003b, Geology and tectonic evolution of the Western Guerrero ter- annurev.earth.32.101802.120257.
rane—A transect from Puerto Vallarta to Zihuatanejo, México, in Geo- Dickinson, W.R., and Lawton, T.F., 2001, Carboniferous to Cretaceous assem-
logic Transects across Cordilleran México: Guidebook for Field Trips of bly and fragmentation of Mexico: Geological Society of America Bul-
the 99th GSA Cordilleran Section Meeting, UNAM Instituto de Geologia letin, v. 113, p. 1142–1160.
Publicación Especial no. 1, p. 201–228. Eguiluz, A.S., and Campa, M.F., 1982, Problemas tectónicos del sector de San
Centeno-García, E., Gehrels, G., Diaz-Salgado, C., and Talavera-Mendoza, O., Pedro El Gallo, en los Estados de Chihuahua y Durango: Boletín Asoci-
2005, Zircon provenance of Triassic (Paleozoic?) turbidites from central ación Mexicana de Geólogos Petroleros, v. 34, p. 5–42.
and western Mexico: Implications for the early evolution of the Guerrero Elías-Herrera, M., 1993, Geology of the Valle de Bravo and Zacazonapan areas,
Arc: Geological Society of America Abstracts with Programs, v. 37, no. south-central Mexico, in Ortega-Gutiérrez, F., et al., eds., Proceedings of
4, p. 12. the First Circum-Pacific and Circum-Atlantic Terrane Conference: Gua-
Chávez-Cabello, G., Aranda-Gómez, J.J., Molina-Garza, R.S., Cossío-Tor- najuato, Mexico, p. 12–21.
res, T., Irving, R., Arvizu-Gutiérrez I.R., and González-Naranjo, G.A., Elías-Herrera, M., and Sánchez-Zavala, J.L., 1992, Tectonic implications of
2005, La falla San Marcos: Una estructura jurásica de basamento mul- a mylonitic granite in the lower structural levels of the Tierra Caliente
tirreactivada del noreste de México, in Alaniz-Álvarez, S.A., and Nieto- Complex (Guerrero Terrane), Southern México: Universidad Nacional
Samaniego, A.F., eds., Boletín de la Sociedad Geológica Mexicana, Volu- Autónoma de México, Instituto de Geología: Revista Mexicana de Cien-
men Conmemorativo Del Centenario, Grandes Fronteras Tectónicas De cias Geológicas, v. 9, p. 113–125.
México, v. 57, p. 27–52. Fastovsky, D.E., Hermes, O.D., Strater, N.H., Bowring, S.A., Clark, J.M., Mon-
Coney, P.J., and Campa, M.F., 1987, Lithotectonic terrane map of México (west tellano, M., and Hernandez R., 2005, Pre–Late Jurassic, fossil-bearing
of the 91st meridian): U.S. Geological Survey Miscellaneous Field Stud- volcanic and sedimentary red beds of Huizachal Canyon, Tamaulipas,
ies Map and Report MF-1874-D, scale 1:10,000,000, 1 sheet. Mexico, in Anderson, T.H., et al., eds., The Mojave-Sonora Megashear
Contreras-Montero, B., Martínez-Cortes, A., and Gómez-Luna, M.A., 1988, Hypothesis: Development, Assessment, and Alternatives: Geological
Bioestratigrafía y sedimentología del Jurásico Superior en San Pedro del Society of America Special Paper 393, p. 259–282.
Gallo, Durango, México: Revista del Instituto Mexicano del Petróleo, v. Ferrusquía, I., Applegate, S.P., and Espinosa, L., 1978, Rocas volcanosedi-
20, p. 5–29. mentarias mesozoicas y huellas de dinosaurios en la región surocciden-
Córdoba-Méndez, D.A., 1964, Geology of Apizolaya Quadrangle (east half), tal pacífica de México: Universidad Nacional Autónoma de México,
Northern Zacatecas, México [M.A. thesis]: University of Texas at Austin, Instituto de Geología: Revista Mexicana de Ciencias Geológicas, v. 2,
111 p. p. 150–162.
306 Centeno-García et al.

Fitz, D.E., Campa, M.F., and López, M.M., 2002, Fechamiento de lavas ande- Convención sobre la evolución Geológica Mexicana, 1er Congreso Mexi-
síticas de la Formación Zicapa, en el límite oriental de la Plataforma cano de Mineralogía, Pachuca, Memoir, p. 67–70.
Guerrero Morelos: Actas INAGEQ, v. 8, p. 178. Guerrero-Suastegui, M., Ramírez Espinosa, J., Gómez Luna, M.E., González
Freydier, C., Lapierre, H., Tardy, M., Coulon, C., Martinez-Reyes, J., and Casildo, V., and Martínez Cortes, A., 1993, Depósitos de tormenta y fauna
Orsini, J.B., 1995, Les formations magmatiques de Porohui (Sinaloa); fósil asociada del Albiano Superior (Formación Teloloapan) Noroeste del
temoins de l’evolution géodynamique mesozoique et tertiaire des Cordil- Estado de Guerrero: Sociedad Mexicana de Paleontología, IV Congreso
leres mexicaines: Comptes Rendus de l’Académie des Sciences, Ser. 2: Nacional de Paleontología, Memorias, p. 93–97.
Sciences de la Terre et des Planètes, v. 321, p. 529–536. Hall, B.V., and Mortensen, J.K., 2003, Bimodal-siliciclastic massive sulfide
Freydier, C., Lapierre, H., Briqueu, L., Tardy, M., Coulon, Ch., and Martínez, deposits of the Leon-Guanajuato District, Central Mexico: Geological
J., 1997, Volcaniclastic sequences with continental affinities within the Society of America Abstracts with Programs, v. 35, no. 4, p. 61.
Late Jurassic–Early Cretaceous intraoceanic Arc terrane (western Mex- Henry, C.D., and Fredrikson, G., 1987, Geology of part of southern Sinaloa,
ico): Journal of Geology, v. 105, p. 483–502. Mexico, adjacent to the Gulf of California: Geological Society of America
Freydier, C., Lapierre, H., Ruiz, J., Tardy, M., Martinez-R.J., and Coulon, C., Map and Chart Ser., v. MCH063, p. 1–14.
2000, The Early Cretaceous Arperos basin: An oceanic domain dividing Jones, N.W., McKee, J.W., Anderson, T.H., and Silver, L.T., 1995, Jurassic vol-
the Guerrero arc from nuclear Mexico evidenced by the geochemistry of canic rocks in northeastern Mexico: A possible remnant of a Cordilleran
the lavas and sediments: Journal of South American Earth Sciences, v. 13, magmatic arc, in Jaques-Ayala, C., et al., eds., Studies on the Mesozoic of
p. 325–336. Sonora and Adjacent Areas: Geological Society of America Special Paper
Fries, C., Jr., 1960, Geología del Estado de Morelos y de partes adyacentes de 301, p. 179–190.
México y Guerrero, región central meridional de México: Universidad Keppie, J.D., Dostal, J., Cameron, K.L., Solari, L.A., Ortega-Gutiérrez, F., and
Nacional Autónoma de México, Instituto de Geología, Boletín, no. 60, Lopez, R., 2003, Geochronology and geochemistry of Grenvillian igne-
236 p. ous suites in the northern Oaxacan Complex, southern Mexico: Tectonic
García-Díaz, J.L., Tardy, M., Campa Uranga, M.F., and Lapierre H., 2004, implications: Precambrian Research, v. 120, p. 365–389.
Geología de la Sierra Madre del Sur en la región de Chilpancingo y Oli- Keppie, J.D., Dostal, J., Miller, B.V., Ortega-Rivera, A., Roldán-Quintana, J.,
nalá, Guerrero, una contribución al conocimiento de la evolución geo- and Lee, J.W.K., 2006, Geochronology and geochemistry of the Fran-
dinámica del margen Pacífico mexicano a partir del Jurásico: Unión cisco Gneiss: Triassic continental rift tholeiites on the Mexican margin of
Geofísica Mexicana, Reunión Nacional de Ciencias de la Tierra, GEOS, Pangea metamorphosed and exhumed in a Tertiary core complex: Interna-
v. 24, p. 173. tional Geology Review, v. 48, p. 1–16.
Gastil, G., Miller, R., Anderson, P., Crocker, J., Campbell, M., Buch, P., Kimbrough, D.L., and Moore, T.E., 2003, Ophiolite and volcanic arc assem-
Lothringer, C., Leier-Engelhardt, P., DeLattre, M., Hoobs, J., and Roldán- blages on the Vizcaino Peninsula and Cedros Island, Baja California Sur,
Quintana, J., 1991, The relation between the Paleozoic strata on opposite Mexico: Mesozoic forearc lithosphere of the Cordilleran magmatic arc,
sides of the Gulf of California, in Pérez-Segura, E., and Jacques-Ayala, in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
C., eds., Studies of Sonoran Geology: Geological Society of America and the Southwestern USA, A Volume in Honor of R. Gordon Gastil:
Special Paper 254, p. 1–7. Geological Society of America Special Paper 374, p. 43–71.
Gastil, G., Rector, R., Hazelton, G., Al-Riyami, R., Hanes, J., Farrar, E., Boeh- Labarthe, G., Tristán, M., and Aguillón, R.A., 1982, Estudio geológico-minero
nel, H., Ortega-Rivera, A., and Guzman, J.G., 1999, Late Cretaceous pil- del área de Peñón Blanco, estados de San Luis Potosí y Zacatecas: Insti-
low basalt, siliceous tuff and calc-turbidite near Porohui, northern Sinaloa, tuto de Geología y Metalurgia, Universidad Autónoma de San Luis Potosí,
Mexico, in Bartolini, C., et al., eds., Mesozoic sedimentary and tectonic Folleto Técnico no. 76, 80 p.
history of north-central Mexico: Geological Society of America Special Lapierre, H., Brouxel, M., Albarède, F., Coulon, C., Lecuyer, C., Martin, P.,
Paper 340, p. 145–150. Mascle, G., and Rouer, O., 1987, Paleozoic and Lower Mesozoic mag-
Gastil, R.G., Morgan, G.J., Krummenacher, D., 1978, Mesozoic history of mas from the eastern Klamath Mountains (North California) and the
peninsular California and related areas east of the Gulf of California, in geodynamic evolution of northwestern America: Tectonophysics, v. 140,
Howell, D.G., McDougall, K.A., eds., Mesozoic Paleogeography of the p. 155–177.
western United States: Los Angeles, California, Pacific Section, Society Lapierre, H., Ortiz, L.E., Abouchami, W., Monod, O., Coulon, C., and Zim-
of the Economic Paleontologists and Mineralogists, p. 107–115. mermann, J.L., 1992, A crustal section of an intra-oceanic island arc: The
Gehrels, G.E., Stewart, J.H., and Ketner, K.B., 2002, Cordilleran-margin Late Jurassic–Early Cretaceous Guanajuato magmatic sequence, central
quartzites in Baja California—Implications for tectonic transport: Earth Mexico: Earth and Planetary Science Letters, v. 108, p. 61–77.
and Planetary Science Letters, v. 199, p. 202–210. Lawlor, P.J., Ortega-Gutiérrez, F., Cameron, K.L., Ochoa-Camarillo, H., Lopez,
Goldhammer, R.K., 1999, Mesozoic sequence stratigraphy and paleogeographic R., and Sampson, D.E., 1999, U–Pb geochronology, geochemistry, and
evolution of northeast Mexico, in Bartolini, C., et al., eds., Mesozoic Sedi- provenance of the Grenvillian Huiznopala Gneiss of Eastern Mexico: Pre-
mentary and Tectonic History of North-Central Mexico: Geological Soci- cambrian Research, v. 94, p. 73–99.
ety of America Special Paper 340, p. 197–210. López-Infanzón, M., 1986, Petrologia y radiometria de rocas igneas y metamor-
González-León, C., Stanley, G.D., Gehrels, G., and Centeno-Garcia, E., 2005, ficas de Mexico: Boletin de la Asociacion Mexicana de Geologos Petrole-
New data on the lithostratigraphy, zircon and isotope provenance, and ros, v. 38, p. 59–98.
paleogeographic setting of the El Antimonio Group, Sonora, Mexico, in Martínez-Reyes, J., 1992, Mapa geológico de la sierra de Guanajuato, escala
Anderson, T.H., et al., eds., The Mojave-Sonora Megashear Hypothesis: 1:100,000, con Resumen de la geología de la sierra de Guanajuato [map]:
Development, Assessment, and Alternatives: Geological Society of Amer- Universidad Nacional Autónoma de México, Instituto de Geología, Carta
ica Special Paper 393, p. 259–282. Geológica de México Ser., scale 1:100,000.
Grajales, M., and López, M., 1984, Estudio petrogenético de las rocas ígneas y McDowell, F.W., Housh, T.B., and Wark, D.A., 1999, Nature of the crust beneath
metamorficas en el Prospecto Tomatlan-Guerrero-Jalisco: Instituto Mexi- west-central Chihuahua, Mexico, based upon Sr, Nd, and Pb isotopic com-
cano del Petróleo, Subdirección de Tecnología y Exploración, Proyecto positions at the Tomochic volcanic center: Geological Society of America
C-1160 (unpublished). Bulletin, v. 111, p. 823–830, doi: 10.1130/0016–7606(1999)111<0823:
Guerrero-Suastegui, M., 1997, Depositional history and sedimentary petrology NOTCBW>2.3.CO;2.
of the Huetamo sequence, Southwestern Mexico [M.S. thesis]: University McKee, J.W., Jones, N.W., and Anderson, T.H., 1999, Late Paleozoic and
of Texas at El Paso, 95 p. early Mesozoic history of the Las Delicias terrane, Coahuila, Mexico, in
Guerrero-Suastegui, M., 2004, Depositional and tectonic history of the Guer- Bartolini, C., et al., eds., Mesozoic Sedimentary and Tectonic History of
rero Terrane, Sierra Madre de Sur; with emphasis on sedimentary suc- North-Central Mexico: Geological Society of America Special Paper 340,
cessions of the Teloloapan area, southwestern Mexico [Ph.D. thesis]: St. p. 161–189.
John’s, Newfoundland, Memorial University, 600 p. Mendoza, O.T., and Suastegui, M.G., 2000, Geochemistry and isotopic compo-
Guerrero-Suastegui, M., Ramírez-Espinosa, J., Talavera-Mendoza, O., and sition of the Guerrero Terrane (western México): Implications for the tec-
Campa-Uranga, M.F., 1991, El desarrollo carbonatado del Cretácico Infe- tonomagmatic evolution of southwestern North America during the Late
rior asociado al arco de Teloloapan, Noroccidente del Estado de Guerrero: Mesozoic: Journal of South American Earth Sciences, v. 13, p. 297–324.
Guerrero Composite Terrane of western Mexico 307

Monod, O., and Calvet, P., 1991, Structural and stratigraphic reinterpretation versidad Nacional Autónoma de México, Instituto de Geología: Revista
of the Triassic units near Zacatecas, Zac., Central Mexico: Evidence of a Mexicana de Ciencias Geológicas, v. 19, p. 81–90.
Laramide nappe pile: Zentralblatt für Geologie und Palaeontologie, Teil Ortiz-Hernandez, E.L., Acevedo-Sandoval, O.A., and Flores-Castro, K., 2003,
I: Allgemeine, Angewandte: Regionale und Historische Geologie, v. 1, Early Cretaceous intraplate seamounts from Guanajuato, central México,
p. 1533–1544. geochemical and mineralogical data: Universidad Nacional Autónoma de
Monod, O., Lapierre, H., Chiodi, M., Martínez-Reyes, J., Calvet, Ph., Ortiz- México, Instituto de Geología: Revista Mexicana de Ciencias Geológicas,
Hernández, E., and Zimmermann, J.L., 1990, Réconstitution d’un arc v. 20, p. 27–40.
insulaire intra-océanique au Mexique central: La séquence volcanopluto- Pacheco, G.C., Castro, M.R., and Gómez, M.A., 1984, Confluencia de terrenos
nique de Guanajuato (Crétacé inférieur): Comptes Rendus de l’Académie estratotectónicos en Santa María del Oro, Durango, México: Revista del
des Sciences de Paris, Sér. 2, v. 310, p. 45–51. Instituto Mexicano del Petróleo, v. 16, p. 7–20.
Monod, O., Faure, M., and Salinas, J.C., 1994, Intra-arc opening and closure Pantoja, A.J., 1959, Estudio Geológico de Reconocimiento de la región de
of a marginal sea: The case of the Guerrero Terrane (SW Mexico): Island Huetamo, Estado de Michoacán: Consejo de Recursos Naturales no
Arc, v. 3, p. 25–34, doi: 10.1111/j.1440-1738.1994.tb00002.x. renovables: Boletin (Instituto de Estudios de Poblacion y Desarrollo
Mortensen, J.K., Hall, B.V., Bissig, T., Friedman, R.M., Danielson, T., Oliver, [Dominican Republic]), v. 50, p. 1–33.
J., Rhys, D.A., and Ross, K.V., 2003, U-Pb zircon age and Pb isotopic Pantoja, A.J., 1990, Redefinición de las unidades estratigráficas de la secuencia
constraints on the age and origin of volcanogenic massive sulfide deposits Mesozoica de la región de Huetamo-Altamirano, Estados de Michoacán y
in the Guerrero Terrane of central Mexico: Geological Society of America Guerrero: Convención Geológica Nacional, Resúmenes, p. 66.
Abstracts with Programs, v. 35, no. 4, p. 61. Pantoja, A.J., and Estrada, B.S., 1986, Estratigrafía de los alrededores de la
Mullan, H.S., 1978, Evolution of part of the Nevadan Orogen in northwestern mina de fierro de El Encino, Jalisco: Sociedad Geológica Mexicana: Bole-
Mexico: Geological Society of America Bulletin, v. 89, p. 1175–1188, tin (Instituto de Estudios de Poblacion y Desarrollo [Dominican Repub-
doi: 10.1130/0016-7606(1978)89<1175:EOPOTN>2.0.CO;2. lic]), v. 47, p. 1–15.
Olvera-Carranza, K., 2002, Estudio estratigráfico-estructural del sector central Pantoja-Alor, J., 1963, Hoja Biseca 13R-k(3), con resumen de la geología de
de la Sierra de Zacatecas, México [B.E. thesis]: Universidad Autónoma de la Hoja San Pedro del Gallo, Estado de Durango: Universidad Nacional
Nuevo León, Facultad de Ciencias de la Tierra, 70 p. Autónoma de México, Instituto de Geología: Carta Geológica de México
Olvera-Carranza, K., Centeno, E., and Camprubí, A., 2001, Deformation Ser., scale 1:100,000.
and distribution of massive sulphide deposits in Zacatecas, Mexico, in Pantoja-Alor, J. and Gómez-Caballero, J.A., 2003, Main geologic and biostrati-
Pietrzynski, A., et al., eds., Mineral Deposits at the Beginning of the 21st graphic features of the Cretaceous of southwestern Mexico (Guerrero
Century: Lisse, Swets and Zeitlinger, p. 313–316. terrane), guidebook for field trips of the 99th GSA Cordilleran Section
Ortega, G.F., Mitre, S.L.M., Roldán, Q.J., Aranda, G.J., Morán, Z.D., Alaníz, Meeting: UNAM Instituto de Geologia Publicación Especial, no. 1, p.
A.S., and Nieto, S.A., 1992, Carta Geológica de la República Mexicana: 229–260.
Universidad Nacional Autónoma de México, Instituto de Geología: Map, Patchett, P.J., and Ruíz, J., 1987, Nd isotopic ages of crust formation and meta-
scale 1:2,000,000, 5th edition. morphism in the Precambrian of eastern and southern México: Contri-
Ortega-Gutiérrez, F., 1981, Metamorphic belts of southern Mexico and their butions to Mineralogy and Petrology, v. 96, p. 523–528, doi: 10.1007/
tectonic significance: Geofísica Internacional, v. 20, p. 177–202. BF01166697.
Ortega-Gutiérrez, F., Elías-Herrera, M., Reyes-Salas, M., Ortega-Gutiérrez, F., Poole, F.G., and Madrid, R.J., 1988, Allochthonous Paleozoic eugeoclinal rocks
Prieto-Vélez, R., Zúñiga, Y., and Flores, S., 1979, Una secuencia volcano- of the Barita de Sonora mine area, central Sonora, México, in Rodríguez-
plutónica-sedimentaria cretácica en el norte de Sinaloa; ¿un complejo Torres, ed., El Paleozoico de la región central del Estado de Sonora:
ofiolítico?: Universidad Nacional Autónoma de México, Instituto de Libreto Guía de la Excursión Geológica para el Segundo Simposio sobre
Geología: Revista Mexicana de Ciencias Geológicas, v. 3, p. 1–8. la Geología y Minería del Estado de Sonora, Excursiones de Campo, Uni-
Ortega-Gutiérrez, F., Ruíz, J., and Centeno-García, E., 1995, Oaxaquia—A versidad Nacional Autónoma de México, Hermosillo, Sonora, p. 32–41.
Proterozoic microcontinent accreted to North America during the Poole, F.G., and Perry, W.J., Jr., 1998, Laurentia-Gondwana continental margins
Late Paleozoic: Geology, v. 23, p. 1127–1130, doi: 10.1130/0091- in Northern Mexico, and their late Paleozoic collision, IGCP Project 376
7613(1995)023<1127:OAPMAT>2.3.CO;2. annual meeting: Instituto de Geología, Universidad Nacional Autónoma
Ortega-Gutiérrez, F., Elias-Herrera, M., Reyes-Salas, M., Macias-Romo, C., de México, México City, México, Laurentia-Gondwanan connections
and Lopez, R., 1999, Late Ordovician–Early Silurian continental colli- before Pangea, Program and Abstracts, p. 27.
sional orogeny in southern Mexico and its bearing on Gondwana–Lauren- Poole, F.G., Perry, W.J., Madrid, R.J., and Amaya-Martínez, R., 2005, Tectonic
tia connections: Geology, v. 27, p. 719–722. synthesis of the Ouachita-Marathon-Sonora orogenic margin of southern
Ortiz, E., and Lapierre, H., 1991, Las secuencias toleíticas de Guanajuato y Laurentia: Stratigraphic and structural implications for timing of defor-
Arcelia, México centro-meridional: Remanentes de un arco insular intra- mational events and plate-tectonic model, in Anderson, T.H., et al., eds.,
oceánico del Jurásico superior–Cretácico inferior: Zentralblatt für Geolo- The Mojave-Sonora Megashear Hypothesis: Development, Assessment,
gie und Palaeontologie, Teil I: Allgemeine, Angewandte: Regionale und and Alternatives: Geological Society of America Special Paper 393, p.
Historische Geologie, v. 6, p. 1503–1517. 543–596.
Ortiz-Hernandez, E.L., 1992, L’arc intra-océanique allocthone Jurassique supéri- Quintero-Legorreta, O., 1992, Geología de la región de Comanja, Estados
eur–Crétacé inférieur du domaine Cordillerain Mexicain (Guerrero Ter- de Guanajuato y Jalisco: Universidad Nacional Autónoma de México,
rane): Pétrographie, géochimie et mineralisations associées des segments de Instituto de Geología: Revista Mexicana de Ciencias Geológicas, v. 10,
Guanajuato et de Palmar Chico–Arcelia consequences paleogéographiques p. 6–25.
[thesis]: Université Joseph Fourier-Grenoble, France, 109 p. Ramírez-Espinosa, J., Campa, M.F., Talavera, O., and Guerrero, M., 1991, Car-
Ortiz-Hernandez, E.L., Yta, M., Talavera, O., Lapierre, H., Monod, O., and acterización de los arcos insulares de la Sierra Madre del Sur y sus impli-
Tardy, M., 1991, Origine intra-océanique des formations volcano-pluto- caciones tectónicas: Convención sobre la evolución Geológica Mexicana,
niques d’arc du Jurassique supérieur–Crétacé inférieur du Mexique cen- 1er Congreso Mexicano de Mineralogía, Pachuca, Memoir, p. 163–166.
tro-méridional: Comptes Rendus de l’Académie des Sciences (Paris) (Ser. Ramírez-Ramírez, C., 1992, Pre-Mesozoic geology of Huizachal-Peregrina
2), v. 305, p. 1093–1098. Anticlinorium, Ciudad Victoria, Tamaulipas, and adjacent parts of eastern
Ortiz-Hernandez, E.L., Chiodi, M., Lapierre, H., Monod, O., and Calvet, P., Mexico [Ph.D. thesis]: University of Texas at Austin, 450 p.
1992, El arco intraoceánico alóctono (Cretácico Inferior) de Guanajuato- Ranson, W.A., Fernandez, L.A., Simmons, W.B., Jr., and de la Vega, E.S., 1982,
características petrográficas, geoquímicas, estructurales e isotópicas del Petrology of the metamorphic rocks of Zacatecas, México: Sociedad
complejo filonianao y de las lavas basálticas asociadas, implicaciones Geológica Mexicana, v. 43, p. 37–59.
geodinámicas: Universidad Nacional Autónoma de México, Instituto de Roldán-Quintana, J., Gonzalez-Leon, C.M., and Amaya-Martinez, R., 1993,
Geología: Revista Mexicana de Ciencias Geológicas, v. 9, p. 126–145. Geologic constraints on the northern limit of the Guerrero Terrane in
Ortiz-Hernández, E.L., Flores-Castro, K., and Acevedo-Sandoval, O.A., 2002, northwestern Mexico, in Ortega Gutiérrez, F., et al., eds., First Circum-
Petrographic and geochemical characteristics of upper Aptian calc-alka- Pacific and Circum-Atlantic Terrane Conference: Guanajuato, México,
line volcanism in San Miguel de Allende, Guanajuato state, Mexico: Uni- Proceedings, p. 124–127.
308 Centeno-García et al.

Rosales-Lagarde, L., Centeno-García, E., Dostal, J., Sour-Tovar, F., Ochoa- l’évolution géodynamique des cordillères mexicaines [Ph.D. thesis]: Uni-
Camarillo, H., and Quiroz-Barroso, S., 2005, The Tuzancoa Formation of versité Joseph Fourier-Grenoble I, France, 462 p.
Hidalgo State: Evidence of a Carboniferous–Permian submarine arc built Talavera-Mendoza, M.O., 2000, Mélanges in southern México: Geochemistry
on continental crust in eastern Mexico: International Geology Review, and metamorphism of Las Ollas complex (Guerrero Terrane): Canadian
v. 47, p. 901–919. Journal of Earth Sciences, v. 37, p. 1309–1320.
Ruíz, J., Patchett, P.J., and Ortega-Gutiérrez, F., 1988, Proterozoic and Pha- Talavera-Mendoza, O., Ramirez-Espinosa, J., and Guerrero-Suástegui, M.,
nerozoic basement terranes of México from Nd isotopic studies: Geologi- 1995, Petrology and geochemistry of the Teloloapan subterrane, a Lower
cal Society of America Bulletin, v. 100, p. 274–281, doi: 10.1130/0016– Cretaceous evolved intra-oceanic island-arc: Geofísica Internacional,
7606(1988)100<0274:PAPBTO>2.3.CO;2. v. 34, p. 3–22.
Salinas-Prieto, J.C., Monod, O., and Faure, M., 2000, Ductile deformations of Talavera-Mendoza, O., Ruiz, J., Gehrels, G., Valencia, V., and Centeno-García,
opposite vergence in the eastern part of the Guerrero Terrane (SW Mex- E., 2007, Detrital zircon U/Pb geochronology of southern Guerrero and
ico): Journal of South American Earth Sciences, v. 13, p. 389–402. western Mixteca arc successions (southern Mexico): new insights for the
Sanchez-Zavala, J.L., 1993, Secuencia volcanosedimentaria Jurásico Superior- tectonic evolution of southwestern North America during Late Mesozoic:
Cretácico Arcelia Otzoloapan (Terreno Guerrero), area Valle de Bravo- Geolgical Society of America Bulletin, v. 119, no. 9/10, p. 1052 1065,
Zacazonapan, Estado de México: Petrografía, Geoquímica, Metamor- doi: 10.1130/B26016.1.
fismo e Interpretación Tectónica [M.S. thesis]: Universidad Nacional Tardy, M., Lapierre, H., Freydier, C., Coulon, C., Gill, J.B., Mercier de Lepinay,
Autónoma de México, Facultad de Ciencias, Mexico, 91 p. B., Beck, C., Martinez, J., Talavera, M., Ortiz, E., Stein, G., Bourdier,
Sánchez-Zavala, J.L., Centeno-García, E., and Ortega-Gutiérrez, F., 1999, J.L., and Yta, M., 1994, The Guerrero suspect terrane (western Mexico)
Review of Paleozoic stratigraphy of Mexico and its role in the Gondwana- and coeval arc terranes (the Greater Antilles and the Western Cordillera of
Laurentia connections, in Ramos, V.A., and Keppie, J.D., eds., Laurentia- Colombia): A late Mesozoic intra-oceanic arc accreted to cratonal Amer-
Gondwana Connections before Pangea: Geological Society of America ica during the Cretaceous: Tectonophysics, v. 234, p. 49–73.
Special Paper 336, p. 211–226. Tristán-González, M., and Torres-Hernández, J.R., 1994, Geología del área
Schaaf, P., Böhnel, H., and Pérez-Venzor, P.A., 2000, Pre-Miocene palaeogeog- de Charcas, Estado de San Luis Potosí, 1994: Universidad Nacional
raphy of the Los Cabos Block, Baja California Sur: Geochronological and Autónoma de México, Instituto de Geología: Revista Mexicana de Cien-
palaeomagnetic constraints: Tectonophysics, v. 318, p. 53–69. cias Geológicas, v. 11, p. 117–138.
Sedlock, R.L., Ortega-Gutiérrez, F., and Speed, R.C., 1993, Tectonostrati- Urrutia-Fucugauchi, J., and Uribe-Cifuentes, R.M., 1999, Lower crustal xeno-
graphic Terranes and Tectonic Evolution of Mexico: Geological Society liths from the Valle de Santiago Maar Field, Michoacan-Guanajuato
of America Special Paper 278, 153 p. Volcanic Field, Central Mexico: International Geology Review, v. 41,
Servais, M., Rojo, Y.R., and Colorado, L.D., 1982, Estudio de las rocas basicas p. 1067–1081.
y ultrabasicas de Sinaloa y Guanajuato; postulacion de un paleogolfo de Valencia-Moreno, M., Ruíz, J., and Roldán-Quintana, J., 1999, Geochemistry
Baja California y de una digitacion tethysiana en Mexico central: Geomi- of Laramide granitic rocks across the southern margin of the Paleozoic
met, v. 3, p. 53–71. North American continent, Central Sonora, Mexico: International Geol-
Silva-Romo, G., 1993, Estudio de la Estratigrafía y Estructuras Tectónicas de ogy Review, v. 41, p. 845–857.
la Sierra de Salinas, Estados de San Luis Potosí y Zacatecas [M.S. the- Valencia-Moreno, M., Ruíz, J., Barton, M.D., Patchett, P.J., Zürcher, L., Hod-
sis]: Universidad Nacional Autónoma de México, Facultad de Ciencias, kinson, D., and Roldán-Quintana, J., 2001, A chemical and isotopic study
México, 111 p. of the Laramide granitic belt of northwestern Mexico: Identification of the
Silva-Romo, G., Arellano Gil, J., Mendoza Rosales, C., and Nieto Obregon, southern edge of the North American Precambrian basement: Geological
J., 2000, A submarine fan in the Mesa Central, Mexico: Journal of South Society of America Bulletin, v. 113, p. 1409–1422, doi: 10.1130/0016–
American Earth Sciences, v. 13, p. 429–442. 7606(2001)113<1409:ACAISO>2.0.CO;2.
Solari, L.A., Keppie, J.D., Ortega-Gutiérrez, F., Cameron, K.L., Lopez, R., and Vidal-Serratos, R., 1991, Estratigrafía y tectónica de la región de Zihuatanejo,
Hames, W.E., 2003, ~990 and ~1,100 Grenvillian tectonothermal events Estado de Guerrero, Sierra Madre del Sur: Convención sobre la evolución
in the northern Oaxacan Complex, southern Mexico: Roots of an orogen: Geológica Mexicana, 1er Congreso Mexicano de Mineralogía, Pachuca,
Tectonophysics, v. 365, p. 257–282. Memoir, p. 231–233.
Stewart, J.H., and Roldán-Quintana, J., 1991, Upper Triassic Barranca Group: Yañez, P., Ruiz, J., Patchett, P.J., Ortega-Gutiérrez, F., and Gehrels, G., 1991,
Nonmarine and shallow-marine rift-basin deposits of northwestern Mex- Isotopic studies of the Acatlan Complex, southern Mexico: Implications
ico, in Pérez-Segura, E., and Jacques-Ayala, C., eds., Studies of Sonoran for Paleozoic North American tectonics: Geological Society of America
Geology: Geological Society of America Special Paper 254, p. 19–36. Bulletin, v. 103, p. 817–828.
Stewart, J.H., Poole, F.G., Ketner, K.B., Madrid, R.J., Roldán-Quintana, J., Yta, M., Lapierre, H., Monod, O., and Wever, P., 1990, Magmatic and struc-
and Amaya-Martínez, R., 1990, Tectonics and stratigraphy of the Paleo- tural characteristics of the Lower Cretaceous arc-volcano-sedimentary
zoic and Triassic southern margin of North America, Sonora, México, in sequence of Saucito-Zacatecas-Fresnillo (central Mexico), geodynamic
Gehrels, G.E., and Spencer, J.E., eds., Geologic Excursions through the implications: Munich, Geowisenschaftliches Lateinamerika, Kolloquium,
Sonoran Desert Region, Arizona and Sonora: Arizona Geological Survey Memoir, p. 21.11–23.11.
Special Paper 7, p. 183–202. Zaldivar, R.J., and Garduño, M.V.H., 1984, Estudio estratigráfico y estructural
Stewart, J.H., Blodgett, R.B., Boucot, A.J., Carter, J.L., and Lopez, R., 1999, de las rocas del Paleozoico Superior de Santa Maria del Oro, Durango,
Exotic Paleozoic strata of Gondwanan provenance near Ciudad Victoria, y sus implicaciones tectónicas [abstract]: Reunión Anual, Sociedad
Tamaulipas, Mexico, in Ramos, V.A., and Keppie, J.D., eds., Laurentia- Geológica de México, p. 37–38.
Gondwana Connections before Pangea: Geological Society of America
Special Paper 336, p. 227–252.
Talavera-Mendoza, O., 1993, Les formations orogéniques mésozoïques du
Guerrero (Mexique méridional): Contribution à la connaissance de MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Tectonic architecture of an arc-arc collision zone, Newfoundland Appalachians

Alexandre Zagorevski*
Cees R. van Staal*
Vicki McNicoll
Neil Rogers
Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario, K1A 0E8, Canada

Pablo Valverde-Vaquero
Instituto Geológico y Minero de España (IGME), La Calera 1, Tres Cantos (Madrid), 28760, Spain

ABSTRACT

The Appalachian-Caledonian orogen records a complex history of the closure of


the Cambrian−Ordovician Iapetus Ocean. The Dunnage Zone of Newfoundland pre-
serves evidence of an Ordovician arc-arc collision between the Red Indian Lake Arc,
which forms part of the peri-Laurentian Annieopsquotch accretionary tract (ca. 480–
460 Ma), and the peri-Gondwanan Victoria Arc (ca. 473–453 Ma). Despite the similar-
ity in age, the coeval arc systems can be differentiated on the basis of the contrasts
that are apparent across the suture zone, the Red Indian Line. These contrasts include
structural and tectonic history, stratigraphy, basement characteristics, radiogenic lead
in mineral deposits, and fauna. The arc-arc collision is considered in terms of modern
analogues (Molucca and Solomon Seas) in the southwest Pacific, and the timing is
constrained by stratigraphic relations in the two arc systems. The Victoria Arc occu-
pied a lower-plate setting during the collision and underwent subsidence during the
collision, similar to the Australian active margin and Halmahera arcs in the south-
west Pacific. The timing of the subsidence is constrained by three new ages of volcanic
rocks in the Victoria Arc (457 ± 2; 456.8 ± 3.1; 457 ± 3.6 Ma) that immediately predate
or are coeval with deposition of the Caradoc black shale. In contrast the Red Indian
Lake Arc contains a sub-Silurian unconformity and a distinct lack of Caradoc black
shale, suggesting uplift during the collision. The emergent peri-Laurentian terranes
provided detritus into the newly created basin above the Victoria Arc. The evidence of
this basin is preserved in the Badger Group, which stratigraphically overlies the peri-
Gondwanan Victoria Arc but incorporated peri-Laurentian detritus. Thus the Badger
Group forms a successor basin(s) over the Red Indian Line. Following the collision,
subduction stepped back into an outboard basin, the Exploits-Tetagouche backarc,
closing the Iapetus Ocean along the Dog Bay Line in the Silurian. Correlative tracts in

*Email, Zagorevski: azagorev@nrcan.gc.ca; Present address, van Staal: Geological Survey of Canada, 625 Robson Street, Vancouver, British Columbia V6B
5J3, Canada.

Zagorevski, A., van Staal, C.R., McNicoll, V., Rogers, N., and Valverde-Vaquero, P., 2008, Tectonic architecture of an arc-arc collision zone, Newfoundland Appa-
lachians, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of
America Special Paper 436, p. 309–333, doi: 10.1130/2008.2436(14). For permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of
America. All rights reserved.

309
310 Zagorevski et al.

the Northern Appalachians and British Caledonides support the Ordovician arc-arc
collision; however, the evidence is less obvious than in Newfoundland.

Keywords: arc-arc collision, Dunnage Zone, U/Pb geochronology, Sandbian and Katian
black shale, Red Indian Line.

INTRODUCTION the seismic-reflection data of van der Velden et al. (2004). This
tectonic model is considered in the context of the tectonics and
The Cambrian–Ordovician closure of the main tract of the sedimentation in recent active arc-arc collisions in the south-
Iapetus Ocean generated a diverse set of arc-backarc terranes and west Pacific, specifically the Huon-Finisterre (Solomon Sea)
microcontinents that were accreted sequentially to the Lauren- and Halmahera-Sangihe (Molucca Sea) arc-continent and arc-
tian margin, leading to significant growth of the eastern margin arc collisions (e.g., Abers and McCaffrey, 1994; Pubellier et al.,
(present coordinates) of North America and the formation of 1999; Whitmore et al., 1999).
the Appalachian-Caledonian orogen. The reconstruction of the
paleogeography of the Iapetus Ocean requires identification of Modern Arc-Arc Collisions
the various terranes and their tectonic setting and has important
implications for a wide variety of topics ranging from the distri- Prior to reviewing the evidence for the arc-arc collision in
bution of mineral deposits to the tectonic development of conver- central Newfoundland we briefly introduce two modern ana-
gent margins and continental growth. logues and review the key evidence that would enable the rec-
Newfoundland lies in a critical position in the Appalachians ognition of arc-arc collisions in ancient orogens. We have cho-
because it forms the geological link between the Northern Appa- sen the Huon-Finisterre arc-continent collision in the Solomon
lachian and the Caledonide segments of this once-continuous Sea (Abbott et al., 1994; Abers and McCaffrey, 1994) and the
orogen. As such the development of ideas in Newfoundland has Halmahera-Sangihe arc-arc collision in the Molucca Sea (Lalle-
greatly influenced the evolution of thought on the tectonic devel- mand et al., 1998; Pubellier et al., 1999) because they are active
opment of the Appalachian-Caledonian orogen as a whole. In and preserve different stages of collision along the strike of the
addition to occupying a key position in the orogen, Newfound- orogen. In addition, we believe that they most closely resemble
land is also perhaps the best constrained portion of the Appala- the relationships in Newfoundland.
chians, featuring excellent coastal exposures, detailed mapping,
high-density geochemical, isotopic, geochronological, and geo- Solomon Sea
physical data that include two crustal-scale seismic-reflection The Solomon Sea plate is currently subducting under the Bis-
transects in central Newfoundland (van der Velden et al., 2004). marck Arc to the north (New Britain Trench) and under the active
The central mobile belt of Newfoundland comprises a Australian continental margin to the south (Trobriand Trough:
complex tectonic collage of arc-backarc terranes that formed Fig. 1; e.g., Abers and McCaffrey, 1994). In the western Solo-
within the Iapetan realm outboard of Laurentia and Gondwana. mon Sea the Australian-Bismarck-Solomon-plate triple junction
The juxtaposition of these terranes was piecemeal, involving forms the site of a modern arc-arc collision where the Bismarck
several arc-continent and arc-arc collisions, with paleogeo- Arc is colliding with the active Australian margin. The collision is
graphic complexity similar to the modern arcs in the southwest diachronous from the northwest to the southeast and has occurred
Pacific (e.g., van Staal et al., 1998). Similar to the southwest northeast of the triple junction, forming the Huon and Finisterre
Pacific (Hall, 2002), the accretionary processes in the Iapetus Ranges on the Huon Peninsula. The deformation is accommo-
operated on very short time scales, with the accretion of many dated by southwest-directed thrusting of the Bismarck Arc and
arc-backarc terranes that occurred within 5–10 m.y. following its accretionary prism over the Australian active margin along the
their formation (Lissenberg et al., 2005a; Zagorevski et al., Ramu-Markham Fault, a continuation of the New Britain Trench
2006, 2007a). Consequently, the resolution of these processes on land. An important out-of-sequence thrust, the Gain Thrust,
requires highly detailed geochronology, which is available in cuts across the Ramu-Markham Fault and emplaces the volcanic
central Newfoundland. rocks of the Bismarck Arc over its accretionary prism (Fig. 2).
This paper reviews the evidence for an Ordovician arc-arc Continuation of movement on the Gain thrust may in time struc-
collision recorded in the central mobile belt of the Newfound- turally mask the accretionary prism, leading to the juxtaposition
land Appalachians. Three new U-Pb zircon geochronologi- of the Bismarck Arc and active Australian margin.
cal ages constrain the timing of volcanism and sedimentation As a result of the collision and loading by the Bismarck Arc,
in the Exploits Subzone and the age of the arc-arc collision. parts of the Australian continental shelf underwent rapid subsid-
Select structural data for central Newfoundland is presented ence (2 km over the last 348 m.y.: Galewsky et al., 1996) indi-
and integrated into a tectonic model, which also incorporates cated by the presence of the drowned carbonate platform above
144º 146ºº 148ºº 150ºº 152º 154º 156º 120 º 125 º 130 º


Bismarck Sea

Philippine
Sea
South Bismark Plate

Ph


ilip

Bismarck Arc
SA

p
ine
Hu
on
Celebes Sea SP

Tre
-
Ra Fin New Britain

n
i
Ma nge ster

North Sulawes
nch

ch

i
r s re Tre Trench
Fa kham Mark ri tain
ult ham
C wB HA
d anyoncp Ne Solomon Sea



Trob
Pa

r iand Molucca
pu

MS Trou Sea
an

gh


Pe
ni

Sul
ns
ul

North
a

awe
Woodlark Basin Banda
Sea

si
Australian
10º

Banda

10º
Plate



Sea
144º 146º 148º 150º 152º 154º 156º 120 º 125 º 130 º
Subduction zone/
Volcano Fault
reverse fault

Figure 1. Simplified tectonic map of Solomon Sea and Molucca Sea. Digital base modified from Bird (2003); boundaries modified from Abers
and McCaffrey (1994), Pubellier et al. (1999), Milsom (2001), and Bird (2003). Abbreviations: dcp—drowned carbonate platform; HA—Halma-
hera Arc; SA—Sangihe Arc; SP—Snellius Plateau; MS—Morobe Shelf.

A Molucca Sea collision zone B Solomon Sea collision zone


Ramu-Markham Gain Bismarck
Celebes Sangihe Pujada Snellius Fault Thrust Arc
Sea Arc Ridge Plateau

7ºN Australian

?
Active Margin
Figure 2. Synthetic sections of the Mo-
148º10′ E lucca Sea and Solomon Sea collision
zones, demonstrating key tectonic rela-
Molucca Slab
tionships. (A) Based on Lallemand et
Sea Plate
al. (1998) and Pubellier et al. (1999).
(B) Modified best gravity fit model of
Markham
Cotabato Sangihe Sangihe Philippine Trobriand New Britain Abbott et al. (1994). Projection of the
Canyon
Trench Arc forearc Trench Trough Trench slab at depth at latitude 148°N is based
turbidites
on Abers and McCaffrey (1994). Large
s
black arrows indicate uplift of the Huon-
Snelliu u Finisterre terrane and subsidence of the
a
5ºN P la t e ? Solomon ? Morobe Shelf west of the cross section.
Sea Plate
Molucca
Sea Plate
148º55′ E

sedimentary overriding subducting accretionary oceanic upper recent


basin arc arc wedge crust mantle volcano
312 Zagorevski et al.

the Morobe Shelf (Fig. 1; Galewsky and Silver, 1997). The degree should have undergone several tectono-thermal events relating
of subsidence of the Australian margin decreases to the east of to protracted arc volcanism, based on the reconstructions of Hall
the Australia-Bismarck-Solomon-plate triple junction, where the (2002). (2) The geochemistry of the volcanic rocks of the Halma-
collision has not yet occurred (Galewsky and Silver, 1997). The hera and Sangihe Arcs shows distinct characteristics (Macpher-
emergent orogen to the west of the triple junction supplies abun- son et al., 2003), which may be utilized in their separation.
dant sediment into the Solomon Sea. The detritus is dominantly (3) The pre- and syncollisional stratigraphy may prove extremely
transported as turbidity currents parallel to the trench along the useful in differentiation of the arc sequences. Both the Australian
Markham Canyon and is deposited on the subducting Molucca margin and the Halmahera Arc–Snellius Plateau were drowned
Sea plate and Australian margin (Galewsky and Silver, 1997). prior to and during the collision, while the overriding arcs and
forearc basins were undergoing uplift and erosion. Emergence
Molucca Sea of the overriding arc and collisional orogen supplies syntectonic
The Molucca Sea is the site of the collision between the fac- sediment into the basin created above the subducting arc. The
ing Sangihe and Halmahera Arcs, which formed as a result of the appearance of sediment with provenance consistent with deriva-
double subduction of the Molucca Sea plate to the west and east, tion from the overriding arc thus would provide information on
respectively (Fig. 1; e.g., Hall, 2002). The collision initiated in the timing of the emergence of the orogen. Finally, the delamina-
the northern Molucca Sea during the Pliocene and migrated to tion of the subducting oceanic plate may lead to extensive syn- to
the south, where the collision is still in its infancy (Pubellier et postcollisional magmatism, marking the end of the collision, such
al., 1999; Hall, 2000). The Molucca Sea plate has been entirely as in Mindanao (Sajona et al., 2000).
overridden by the collisional complex derived primarily from the The recognition of ancient arc-arc collision zones similar to
Sangihe forearc. In the southern Molucca Sea, where the collision those in the southwest Pacific thus would require an approach
is the least advanced, the Halmahera forearc is in part overridden that integrates stratigraphy, geochemistry, geochronology, struc-
by the Sangihe forearc. To the north, the Halmahera forearc is tural geology, and geophysics, combined with a good understand-
completely overridden by the Sangihe forearc, and the Snellius ing of basement-cover relationships and affinity of the various
Plateau, the continuation of the Halmahera Arc to the north, is terranes. We believe that the quality, quantity, and distribution
subducting at present (Lallemand et al., 1998; Pubellier et al., of data make the Newfoundland Appalachians unique among
1999; Hall, 2000). As a result of subduction, the Snellius Plateau ancient orogens in that they permit detailed reconstruction of the
underwent rapid subsidence, as evidenced by the presence of a tectonic architecture of an Ordovician arc-arc collision.
drowned carbonate shelf, whereas parts of the overriding colli-
sional complex have been uplifted above sea level (Talaud Island GEOLOGICAL SETTING
and Pujada Ridge: Figs. 1, 2).
The Molucca Sea plate is entirely overridden by the colli- The complexity of the closure of the Iapetus Ocean is
sional complex but is still actively subducting in the south Molucca reflected in the zonal division of the northern Appalachians,
Sea. To the north, evidence from deep earthquakes suggests that where five zones have been defined: Humber, Dunnage, Gander,
the Molucca Sea plate has delaminated from the overriding plate Avalon, and Meguma (Fig. 3; Williams, 1995b). The Humber
beneath Mindanao (Lallemand et al., 1998). Mindanao is char- Zone represents the Cambrian–Ordovician Laurentian margin.
acterized by widespread Pliocene and younger syn- to postcol- The Dunnage Zone (the central mobile belt) lies outboard of the
lisional basaltic to dacitic magmatism. The compositions of the Humber Zone and contains allochthonous ensialic and ensimatic
magmas range from calc-alkaline to shoshonitic. The diversity of arc-backarc complexes that formed within the Iapetus Ocean. Far-
the magmas and the geochemistry has been related to the delami- ther outboard, the Gander, Avalon, and Meguma Zones represent
nation of the Molucca Sea plate (Sajona et al., 2000). microcontinents derived from Gondwana (see van Staal, 2005,
for a review). The margin of Laurentia experienced three oro-
Recognition of Ancient Arc-Arc Collisions genic episodes from Ordovician to Early Devonian time, namely
The juxtaposition of coeval arc terranes, such as those the Taconic, Salinic, and Acadian. These episodes occurred as a
described for the Molucca and Solomon Seas, presents a dif- result of the successive arrival of the Dashwoods, Gander, and
ficulty with respect to their identification and interpretation in Avalon microcontinents (van Staal, 2005).
ancient orogens. In the case of both collision zones described The Dunnage Zone is subdivided into the peri-Laurentian
above, the facing arc terranes may be differentiated on the basis Notre Dame and the peri-Gondwanan Exploits Subzones (Fig. 3;
of the (1) tectonic history and basement characteristics, (2) geo- Williams et al., 1988). Other subsequent subdivisions (Williams,
chemistry, and (3) stratigraphic record. (1) Prior to the collision 1995a) reflect minor regional differences internal to these two
the Halmahera Arc-backarc underwent a period of shortening that subzones that are beyond the scope of this paper. The Notre Dame
resulted in thrusting of the backarc rocks onto the arc and a gap Subzone comprises predominantly Early to Middle Ordovician
in volcanism, whereas the Sangihe Arc was not shortened during Notre Dame Arc plutons that intruded a ribbon microcontinent,
this time. The Australian margin in the Solomon Sea area was not referred to as the Dashwoods microcontinent (Waldron and van
active until recently, whereas the basement to the Bismarck Arc Staal, 2001), as well as accreted peri-Laurentian terranes. The
Tectonic architecture of an arc-arc collision zone 313

59° 57° 55° 53°

Laurentian
51° N 51°
Margin
Humber Zone

Dunnage Zone Dunnage Zone


50° 50°
Notre Dame Subzone
Gulf of
e
on

St. Lawrence Notre Dame


rZ

Bay Gondwanan
be

Margin
m

e e
Hu

on m

L
bz D a

RI Dunnage Zone:
Su tre

49° 49°
o

Exploits Subzone
N

e
Exploits Zon
Subzone Gander Zone
er
nd

Z Figure 3. Tectono-stratigraphic subdivi-


LS
Ga

V sions of Newfoundland (after Williams,


ne

Avalon Zone
Zo

1995a) and northern Appalachians


n
o

(from Hibbard et al., 2004). 1—Popelo-


al

Meguma Zone
Av

gan inlier; 2—Bathurst Supergroup;


3—Chain Lakes Massif and Jim Pond
Formation; 4—Shelburne Falls Arc;
50 0 50 100
47° 5—Bronson Hill Arc; RIL—Red Indian
Kilometres
Line; VLSZ—Victoria Lake shear zone;
BVBL—Baie Verte – Brompton Line.
59° 57° 55° 53°
45 50
°W
°W

°W

°N °N
70

60
65

BVBL
RIL
1
3
2

50
CF BVBL °N
4 5 RIL

BBF
CB
40 F
°N DHF
°W
°W

40 45
°W

°N °N
60
65
70

Exploits Subzone contains peri-Gondwanan ensialic and ensi- basins and seaways underlain by oceanic crust (e.g., Exploits-
matic arc-backarc complexes. The two subzones are differenti- Tetagouche backarc basin: van Staal et al., 1998). In Notre
ated on the basis of faunal, paleomagnetic, isotopic, and struc- Dame Bay the Red Indian Line is marked by late brittle faults
tural contrasts (Williams et al., 1988). and locally by Ordovician mélange (McConnell et al., 2002;
Williams et al., 1988), although the exact nature and location of
Notre Dame–Exploits Relationships in Newfoundland this boundary is debated (e.g., Arnott et al., 1985; Wasowski et
al., 1986; van der Voo et al., 1991). In central Newfoundland the
The identification of the marked geological contrasts surface trace of the Red Indian Line is marked by a mélange-
between the Notre Dame and Exploits Subzones has led to rec- phyllonite belt (Fig. 3; Rogers and van Staal, 2002; Zagorevski
ognition of the Red Indian Line (Fig. 3; Williams et al., 1988), et al., 2006, 2007b), which coincides with the surface projec-
which is now accepted as the main suture zone in the north- tion of a major crustal-scale reflector, interpreted to be a fault,
ern Appalachians, along which several thousand kilometers of in seismic-reflection profiles (van der Velden et al., 2004). The
Iapetus oceanic lithosphere has been consumed (e.g., van Staal, surface trace of the Red Indian Line is cut out in central New-
2005). Following the formation of the Red Indian Line the con- foundland by the Devonian Victoria Lake shear zone (Fig. 3;
vergence continued by means of subduction in Iapetan marginal Valverde-Vaquero and van Staal, 2002), which is a continuation
314 Zagorevski et al.

of the Noel-Paul’s Line (Williams et al., 1988). At depth the Red Subduction in the Annieopsquotch accretionary tract initi-
Indian Line is truncated by a seismic reflector interpreted to be a ated at ca. 480 Ma, leading to the formation of the Annieops-
Devonian wedging-related fault (van der Velden et al., 2004). quotch ophiolite belt (Dunning and Krogh, 1985; Dunning,
1987; Lissenberg et al., 2005b). This may have occurred in a
Notre Dame Subzone reentrant in the Dashwoods margin (Waldron and van Staal,
2001; Zagorevski et al., 2006). Subsequently, the subduction
The Notre Dame Subzone in Newfoundland comprises developed under the Dashwoods margin, forming a continen-
several composite terranes formed outboard of the Lauren- tal arc and backarc by ca. 473 Ma (Buchans Group and Lloyds
tian margin. The western margin and central part of the Notre River ophiolite complex: Dunning et al., 1987; Swinden et al.,
Dame Subzone is dominated by the Dashwoods microconti- 1997; Zagorevski et al., 2006). This arc rifted, forming an oce-
nent (Waldron and van Staal, 2001) and its Notre Dame Arc anic basin (Pickett, 1987) and a younger arc, represented by the
suprastructure. The eastern margin of the Notre Dame Sub- Red Indian Lake Group in central Newfoundland (Zagorevski
zone contains a tectonic collage of continental and intraoce- et al., 2006).
anic supra-subduction zone complexes collectively referred to
as the Annieopsquotch accretionary tract (Figs. 4–6; van Staal Red Indian Lake Group
et al., 1998; Lissenberg et al., 2005b; Zagorevski et al., 2006). The Red Indian Lake Group comprises mafic, andesitic, and
The Annieopsquotch accretionary tract formed above a west- felsic volcanic rocks, epiclastic sedimentary rocks, red shale, and
dipping subduction zone and was facing the main Iapetan basin. ferruginous chert. Tholeiitic mafic volcanic rocks are predominant
As such, it contains the most outboard arc-backarc complexes at the stratigraphic base, while calc-alkaline mafic, andesitic, and
of the Notre Dame Subzone and provides constraints on the felsic volcanic rocks become more common at the stratigraphic
development of the Notre Dame Subzone and its juxtaposition top, suggesting a maturing arc sequence (Zagorevski et al., 2006).
with the Exploits Subzone. Negative εNd values of felsic volcanic rocks (−6.4 to −7.7) and

N 57° 55° L
RI
12
11 13
10
49°
B
U 49°
R
G

5
6
8
4 9
7
3
1

48°
48°
2

57° 55°

Annieopsquotch Badger Victoria-Exploits forearc, Penobscot


accretionary tract Group arc and backarc volcanic Arc
and sedimentary rocks

Figure 4. Distribution of Annieopsquotch accretionary tract, Victoria Arc, Penobscot basement, and Badger Group in
Newfoundland (after Colman-Sadd et al., 1990). 1—Dashwoods Subzone; 2—Windsor Point Group; 3—Annieops-
quotch ophiolite belt; 4—Lloyds River ophiolite complex; 5—Buchans Group; 6—Red Indian Lake Group; 7—Pats
Pond Group and Wigwam Brook Group; 8—Sutherlands Pond group; 9—Tally Pond Group and Noel Paul’s Brook
group; 10—Exploits Group; 11—Wild Bight Group; 12—Summerford Group; 13—Dunnage mélange; GRUB—Gan-
der River Ultrabasic Belt; RIL—Red Indian Line.
Red Indian Exploits Subzone
Notre Dame Subzone Line
Llandovery Wenlock
Silurian

Springdale Group and equivalents Botwood Group


430
Silurian
440

Badger Group
*
Llanvirn Caradoc Ash.
450

Ordovician
Upper

Widsor Point
Group (2)

WBG
460

NPBg

VICTORIA ARC
Figure 5. Relationships between selected
Ordovician Ordovician

WwBG
tectono-stratigraphic units in New-
Middle
Age (Ma)

Sutherlands
Pond group (8)
foundland. Subdivision of the Ordovi-
Summerford Group
(accreted seamount; 12)
470

Red Indian
Lake Group (6)
Arenig

cian Period into epochs follows McKer-


row and van Staal (2000). Numbers in
Dashwoods
Subzone(1)

Hgs parentheses refer to the locations shown


Buchans
Group (5)
Lloyds River
Ophiolite Complex (4)
Lower
480

in Figure 4. Hgs—Harpoon gabbro


Tremadoc

suite; NPBg—Noel Paul’s Brook group;


Dunnage
melange (13)

Penobscot orogeny
Annieopsquotch
ophiolite belt (3)

WwBG—Wigwam Brook Group. See


text for discussion.
490

Cambrian

*
Upper

Pats Pond
Group (7)

PENOBSCOT ARC
Wild Bight
Group (11)
Exploits
Group
500

(10)
Annieopsquotch accretionary
Cambrian

tract
Middle

Tally
Pond
510

Group
(9)
Victoria Lake Supergroup
Ophiolite Peri-Gondwanan Penobscot
Black shale
volcanic rocks arc
Peri-Laurentian Sedimentary
Limestone Plutonic rocks
volcanic rocks rocks
U/Pb zircon age Hiatus

Map Buchans
Unclassified Red Indian Location
contact Line
Lithoprobe Red Indian Lake
Thrust fault seismic 48o45’
reflection 56o30’
profiles
Notre Dame Subzone
Annieopsquotch
accretionary tract o
57 15' Figure 6. Simplified geology of cen-
tral Newfoundland (after Zagorevski et
Exploits Subzone S1
al., 2006). Stereonets are lower hemi-
Victoria Lake Supergroup n=162
sphere, equal area, contoured using
Victoria arc 48 30’
o
the Kamb method, with the number of
o
57 00' points indicated. Stereonets of S1 are
Penobscot arc for the indicated portions of the mapped
ak
e area, south and southwest of Red Indian
sL S1 Lake. The S2 stereonet is for the entire
10 km d
L loy n=289 mapped area south and southwest of
57o45' Red Indian Lake.
to ria Lake
Vic
o
S2
48 15'
n=190
57o30’ S1
n=284
316 Zagorevski et al.

abundant Mesoproterozoic zircon inheritance suggest that this arc as the Victoria Arc in central Newfoundland. The Victoria Arc is
formed on continental crust. The range of U/Pb zircon ages of the a correlative of the Popelogan arc of New Brunswick, Canada
volcanic rocks of the Red Indian Lake Group (462 +2/–9, 463 (Fig. 3: Popelogan inlier; van Staal et al., 1998).
± 3, 464.8 ± 3.5, 465 ± 2 Ma: Zagorevski et al., 2006) indicates
that arc volcanism is predominantly Llanvirn (i.e., 459–465 Ma: Victoria Arc
McKerrow and van Staal, 2000). The Red Indian Lake Group is The tectonic setting of the Victoria Arc, and its correla-
the most outboard and the youngest sequence identified within tive Popelogan arc in New Brunswick, are well constrained in
the Annieopsquotch accretionary tract. Correlatives of the Red the northern Appalachians (e.g., van Staal et al., 1998). The
Indian Lake Group likely occur throughout the Annieopsquotch earliest Victoria Arc volcanism in Newfoundland occurred at
accretionary tract and may include the undated Crescent Lake ca. 473 Ma (Fig. 4; Wild Bight Group: MacLachlan and Dun-
Formation (Fig. 3; Bostock, 1988) and parts of the Cottrell’s Cove ning, 1998b). Coeval eruption of arc- and backarc-related vol-
Group (Arenig: in O’Brien, 2003; i.e., 465–480 Ma: McKerrow canic rocks (Exploits Group: O’Brien et al., 1997; Wild Bight
and van Staal, 2000) in the Notre Dame Bay region. The Red Group: MacLachlan and Dunning, 1998b) indicates an exten-
Indian Lake Group and its correlatives are henceforth referred to sional phase of arc magmatism, which led to the rifting of the
as the Red Indian Lake Arc, which comprises the youngest arc Victoria and Popelogan arcs from Ganderia and the opening of
sequence in the Annieopsquotch accretionary tract. the Exploits-Tetagouche backarc basin in Newfoundland and
New Brunswick (Fig. 3; O’Brien et al., 1997; MacLachlan and
Notre Dame Subzone Cover Dunning, 1998b; van Staal et al., 1998; Valverde-Vaquero et al.,
The Notre Dame Subzone is unconformably overlain by the 2006; Bathurst Supergroup: van Staal et al., 2003). Abundant fos-
volcaniclastic and sedimentary rocks of the spatially restricted sil and some age-dating evidence indicate that the Victoria Arc
Windsor Point Group in southwestern Newfoundland (Fig. 4; was active until at least middle Llanvirn time (e.g., O’Brien et
453 +5/–4 Ma: Dube et al., 1996); however, the most wide- al., 1997). For example, Victoria Arc–related basalts of the Sops
spread unconformity is sub-Silurian. The Notre Dame Subzone Head Complex in Notre Dame Bay form peperitic contacts with
is regionally overlain unconformably by the rocks of the Silurian Llanvirn limestone (McConnell et al., 2002).
Springdale Group or its equivalents (Chandler et al., 1987; Dun- In central Newfoundland the Victoria Arc is represented by
ning et al., 1990). In central Newfoundland these rocks comprise the Sutherlands Pond group (ca. 457–462 Ma: Dunning et al.,
felsic to andesitic volcanic rocks, breccia, polymictic conglom- 1987; Rogers et al., 2005; see following), Noel Paul’s Brook group
erate, and red sandstone (e.g., van Staal et al., 2005a, b). The (ca. 457–465 Ma: van Staal et al., 2005c), and Wigwam Brook
provenance of the polymictic conglomerate suggests derivation Group (453 ± 4 Ma: Zagorevski et al., 2007a). The youngest dated
predominantly from the Notre Dame Subzone. Some clasts bear Victoria Arc–related volcanic rocks in the lower Wigwam Brook
a predepositional foliation, indicating deformation and exhuma- Group (van Staal et al., 2005c; Zagorevski et al., 2007a) comprise
tion of the Ordovician tectonites by Silurian time. a coarsening-up sequence of felsic tuff and breccia that are inter-
calated with turbiditic sandstone, conglomerate, black shale, and
Exploits Subzone basalt with arclike chemistry. The upper Wigwam Brook Group
comprises turbiditic sandstone and black shale, suggesting cessa-
To the east of the Red Indian Line the Exploits Subzone in tion of arc magmatism (Zagorevski et al., 2007a).
Newfoundland can be subdivided into two distinct arc-related
sequences (Fig. 4; Jenner and Swinden, 1993; O’Brien et al., Exploits Backarc
1997; MacLachlan and Dunning, 1998a, b; van Staal et al., Similarly to the Victoria Arc, the Exploits-Tetagouche back-
1998; Rogers et al., 2006; Zagorevski et al., 2007a). The older arc was magmatically active until the late Llanvirn (e.g., Law-
sequence, the Penobscot arc and backarc, contains ca. 513– rence Head Formation: O’Brien et al., 1997). The coeval sedi-
485 Ma volcaniclastic and sedimentary rocks that formed above mentation in the Exploits backarc was continuous from the early
an east-dipping subduction zone along the margin of Ganderia Arenig until the Caradoc (Caradoc: 449–459 Ma: McKerrow and
(Rogers et al., 2006; Zagorevski et al., 2007a), a postulated peri- van Staal, 2000) and mostly consisted of deep-marine sediments
Gondwanan microcontinent (van Staal et al., 1996). A gap in interrupted by occasional pulses in magmatism, such as the erup-
magmatism in the Exploits Subzone (485–478 Ma) is correlated tion of backarc-basin basalts of the Laurence Head Formation
with the collision of the Penobscot arc with Ganderia at ca. 480– (O’Brien et al., 1997). The end of the Llanvirn was marked by
485 Ma and obduction of the Penobscot backarc ophiolites (Jen- cessation of magmatism, deposition of chert and regionally dis-
ner and Swinden, 1993) onto the Newfoundland Gander Zone continuous limestone, followed by the black shale of the Law-
(Fig. 3; e.g., Colman-Sadd et al., 1992; Tucker et al., 1994; van rence Harbour Formation (Williams, 1995a).
Staal et al., 1998). The Penobscot arc is stratigraphically over- In central Newfoundland, backarc volcanism is preserved in
lain by ca. 473–453 Ma volcano-sedimentary rocks (Fig. 4; e.g., the Red Cross Group, which comprises felsic volcanic rocks (466
MacLachlan and Dunning, 1998b; O’Brien et al., 1997; Evans ± 3 Ma: Valverde-Vaquero et al., 2006), basalt, turbiditic sandstone,
and Kean, 2002; Zagorevski et al., 2007a), which are referred to black shale, and calcareous shale locally intruded by gabbro (457
Tectonic architecture of an arc-arc collision zone 317

± 3 Ma: Valverde-Vaquero et al., 2006). Similar to the rest of the accretionary tract, for example, the thickness of the individual
Exploits Subzone, the upper parts of the Red Cross Group com- tectonic panels varies between 0 and 5 km, indicating partial to
monly contain abundant black shale and calcareous shale. total excision of tectono-stratigraphic units along strike. Detailed
investigations of the structural geology indicate that the macro-
Exploits Subzone Cover scopic simplicity is deceptive and that the rocks have undergone
One of the key characteristics of the Exploits Subzone is multiple phases of deformation. Seven deformation episodes
the presence of the regionally transgressive Caradoc black shale (D1 through D7) have been recognized in central Newfoundland,
cover, accompanied by the general cessation of volcanism (Fig. 5; based on overprinting relationships as well as available age and
van der Pluijm et al., 1987; Williams, 1995a). In the Notre Dame stratigraphic constraints (Zagorevski et al., 2007b). The regional
Bay area (Fig. 3), this time was marked by deposition of the black geometry of the rocks is primarily the result of D1, D2, and D4 epi-
shale of the Lawrence Harbour and Shoal Arm Formations (e.g., sodes, although only D1 structures formed during the arc-arc col-
Williams et al., 1995; O’Brien et al., 1997). In central Newfound- lision. Hence, only these are discussed below. D3, 5–7 so far have
land the Wigwam Brook Group (Zagorevski et al., 2007a) and been shown to be present only on a mesoscopic scale (Zagorevski
Stanley Waters formation (see following discussion; van Staal et et al., 2007b) and therefore are not imperative for this chapter.
al., 2005c) mirror the relationships in Notre Dame Bay and indi- The oldest phase of deformation identified in central New-
cate cessation of volcanism accompanied-followed by deposition foundland (D1) was heterogeneous and led to strain localization
of the Caradoc black shale. into northwest-dipping brittle to ductile shear zones ten to several
The cessation of arc magmatism and deposition of black shale hundred meters in thickness. These deformation zones are char-
are commonly assumed to have occurred at the Llanvirn–Caradoc acterized by mylonite, phyllonite, and/or black shale mélange that
boundary (e.g., O’Brien, 2003). However, in the Wigwam Brook mark the boundaries between the various tectono-stratigraphic
Group this transition occurred in the middle Caradoc (453 ± 4 Ma: units (Figs. 6, 7). In proximity to the Red Indian Line in cen-
Zagorevski et al., 2007a). Volcanism of similar age may also have tral Newfoundland, D1 shear zones accommodated SSE-directed
been the source of felsic volcanic boudins, which are common in motion (Lissenberg, 2005; Zagorevski et al., 2007b).
the mélange belts of central Newfoundland (Rogers and van Staal, D1 shear zones are overprinted by upright moderate- to
2002; Valverde-Vaquero et al., 2006). The cessation of volcanism shallow-plunging asymmetric F2 folds and axial planar S2 folia-
and the subsidence of the Victoria Arc thus appear to have been tion. S2 is commonly a composite foliation formed by transposi-
somewhat diachronous along strike, and may have started earlier tion of S1 by tight to isoclinal F2 folds (Figs. 6, 7). The composite
in Notre Dame Bay. Alternatively this may be an artifact of sparse S1–2 is the regionally dominant foliation. The enveloping surface
dating in Notre Dame Bay or of local basin architecture such that of the F2 folds dips northwest, suggesting that the D1 shear zones
parts of the basin were effectively isolated from the volcaniclastic were originally shallow northwest-dipping structures prior to the
input and received only background sedimentation. D2 structures. In addition, the emplacement of “old over young”
Regionally the Ashgill to Wenlock (424–449 Ma: McKer- rocks suggests that the D1 shear zones represent thrust faults
row and van Staal, 2000) Badger Group (Williams et al., 1993) (Fig. 8; Zagorevski et al., 2007b). This interpretation is supported
stratigraphically overlies the Caradoc black shale and comprises by areas of weak D2 strain, such as Buchans (Fig. 6), where a
an upward-coarsening sedimentary sequence of deep-marine tur- southeast-directed antiformal thrust stack is developed (Calon
bidites to shallow-marine conglomerate. Structural and sedimen- and Green, 1987).
tological investigations of the Badger Group indicate deposition D2 strain is commonly concentrated along D1 shear zones
in a syntectonic setting (e.g., Kusky et al., 1987; Williams et al., where the D1 tectonites provided a strong planar anisotropy. This
1995). The detrital provenance of the sedimentary rocks suggests resulted in transposition of S1 and reactivation of D2-steepened
derivation from the Notre Dame Subzone (see Williams et al., D1 shear zones as southeast-directed reverse faults. Similar to D1
1995), confirmed by the U/Pb detrital zircon data (McNicoll et shear zones, D2 reverse faults also accommodated SSE-directed
al., 2001). The syntectonic setting and provenance have led to the translation. The differentiation of D1 and D2 structures in areas
interpretation of the Badger Group as a successor basin or basins of high D2 strain thus can be tenuous in the absence of overprint-
over the Red Indian Line in the forearc region (arc-trench gap) ing relationships.
of the Late Ordovician to Early Silurian subduction complex D4 structures include S4 foliation, F4 folds, and D4 shear
formed during the closure of the Tetagouche-Exploits backarc zones that overprint D1 and D2 structures. Although the D4 struc-
basin (Kusky et al., 1987; Pickering, 1987; van Staal et al., 1998; tures timewise are not related to the arc-arc collision discussed
Valverde-Vaquero et al., 2006). in this paper, they are locally paramount to the kinematic inter-
pretation of the geometry of the deformed rocks. D4 structures
STRUCTURAL GEOLOGY were associated with formation of a northwest-directed thrust
and fold belt in central Newfoundland. An excellent example of
Central Newfoundland displays a relatively simple macro- a D4 structure is the Devonian Victoria Lake shear zone (Fig. 6;
scopic structure with internally folded structural panels bounded Valverde-Vaquero et al., 2006). This shear zone emplaces the
by curviplanar shear zones (Fig. 6). In the Annieopsquotch metamorphic tectonites of the Gander Zone over the Exploits
318 Zagorevski et al.

A B C D

S1 S1

S5
F1
S2
S4 (composite) S2

Figure 7. Representative relationships between D1 and D2 structures. (A) Horizontal outcrop of polydeformed black shale mélange with a basaltic
knocker marking the boundary between the Sutherlands Pond group and the Penobscot arc. (B) Horizontal outcrop of highly deformed Ordovi-
cian limestone along the same boundary, with sinistral boudinage of a vein. (C) Vertical photomicrograph of F2-folded D1 metabasic tectonite
along the boundary between Lloyds River ophiolite complex and Annieopsquotch ophiolite belt (see Fig. 5). (D) Vertical outcrop demonstrating
the relationships between F1 folded quartz veins and upright F2 folds in felsic tuff of the Red Indian Lake Group.

A B Notre AAT RIL Victoria


RIL
N Dame Arc Lake Sg.

10
Ganderia
Laurentia 20
B
RIL 30

Moho 40

C Notre
50 km

C Dame Arc RIL

10
Ganderia
20
Laurentia Laurentia 30
40
50 km

Figure 8. Interpretation of D1-D2 relationships in central Newfoundland. (A) Interpretation of the three-dimensional geometry of the Red Indian
Line, based on the asymmetry of F2 folds, which indicate a northwest-dipping enveloping surface. Inset shows a close-up of the central area with
folded shear zones (light gray) and folded unconformity between the Pats Pond and Wigwam Brook groups (dark gray). Location of seismic-
reflection profiles is indicated by panels B and C. Interpretation of the D1-D2 relationship in panel A is consistent with the seismic-reflection
profiles shown in panels B and C (simplified from van der Velden et al., 2004). AAT—Annieopsquotch accretionary tract; RIL—Red Indian
Line; Sg—supergroup.

Subzone and in part buries the D1 and D2 shear zones, including 468 ± 2 Ma U/Pb zircon; ca. 464–458 Ma U/Pb zircon: Lissen-
the Red Indian Line. berg et al., 2005a). The oldest ages of D1 deformation predate
the youngest peri-Laurentian arc volcanism in the Annieops-
Ages of D1 and D2 Deformation Zones quotch accretionary tract (Fig. 5; i.e., Red Indian Lake Group:
Zagorevski et al., 2006). Therefore, Arenig to early Llanvirn
The ages of the D1 zone have been obtained from the syn- deformation is strictly peri-Laurentian and is not related to the
tectonic intrusions in the westernmost units of the Annieops- accretion of any peri-Gondwanan terranes. This deformation is
quotch accretionary tract. These syntectonic intrusions span probably a result of back thrusting that initiated following the
from Arenig to early Caradoc (ca. 470 Ma Ar/Ar hornblende; Taconic collision of the Dashwoods microcontinent with the
Tectonic architecture of an arc-arc collision zone 319

Laurentian margin at ca. 470 Ma (Waldron and van Staal, 2001; (ca. 429 Ma: Dunning et al., 1990), their intimate association
van Staal et al., 2006). with D2 reverse faults, and the presence of clasts containing pre-
The earliest constraints on the start of deformation that depositional S1 foliation indicate that they were deposited during
affected both the Notre Dame and Exploits Subzones are pro- the D2 deformational episode, following the exhumation of the
vided by the stratigraphic relationships. The dismemberment of D1 tectonites.
the Caradoc black shale unit into mélange and broken formation
along the Red Indian Line, without incorporation of Ashgill-age U/Pb GEOCHRONOLOGY
rocks, suggests that the D1 deformation that affected both sub-
zones was under way in the late Caradoc (also observed in the U-Pb thermal ionization mass spectrometry (TIMS) and
Sops Head–Boones Point Complex: McConnell et al., 2002). sensitive high-resolution ion microprobe (SHRIMP II) analyses
Hence, although the D1 deformation was protracted, lasting from were conducted at the Geological Survey of Canada (GSC) in
at least 470 Ma until at least 450 Ma, only the youngest part of order to constrain the timing of the arc volcanism and black shale
the D1 deformation is related to the accretion of peri-Gondwa- deposition in the Exploits Subzone. U-Pb TIMS analytical meth-
nan terranes. ods are outlined in Parrish et al. (1987), with treatment of analyti-
The upper limit of D1 deformation and the span of D2 defor- cal errors after Roddick et al. (1987). SHRIMP II analyses were
mation are constrained by the Silurian plutonic rocks (435–427 conducted using analytical procedures described by Stern (1997),
Ma: Dunning et al., 1990; Whalen et al., 2006; Zagorevski et with standards and U-Pb calibration methods following Stern and
al., 2007b). Some of these plutons stitch major D1 shear zones Amelin (2003). U-Pb TIMS and SHRIMP analyses are presented
and contain rare enclaves of D1 tectonites. Despite displaying in Tables 1 and 2, respectively, and are plotted in concordia dia-
cross-cutting relationships, the Silurian plutonic rocks are also grams with errors at the 2σ level (Fig. 9). Further details on the
deformed along the shear zones at amphibolite facies within U-Pb analytical techniques are presented in the Appendix. Geo-
error of the age of intrusion and as such are interpreted as syn- chronology sample locations are indicated in Figure 10.
D2 plutons (Zagorevski et al., 2007b). Hence, D2 is constrained
by these plutons to have started by at least 435 Ma and contin- Penny Brook Formation
ued to at least 426 Ma (Zagorevski et al., 2007b). The age of the
D2 deformation is also constrained by the unconformable Silu- A sample (RAX05–900, Z8777; UTM zone 21, 594708 E,
rian cover of the Notre Dame Subzone. The age of the red beds 5474803 N; NAD 83: Fig. 10) from a thin seam of felsic tuff,

Penny Brook 0.082 505


520 Stanley Waters formation
Formation Fine grained felsic tuff
0.082 Felsic Tuff 495 835
500 Ma
0.078 480 485 Z3
0.078
Pb/238U

460 475
0.074
440 465
206

0.070 Z2 Z1
Concordia age = 0.074 Z1
420 455
457.5 ± 2.7 Ma LI = 457 ± 3.6 Ma
0.066 SHRIMP II data (MSWD=0.1)
400
TIMS data 100 µ m
0.062
0.4 0.5 0.6 0.7 0.8 0.54 0.57 0.60 0.63 Figure 9. U/Pb SHRIMP II and TIMS
Sutherlands 0.082 505 concordia diagrams (2σ, decay con-
520 Sutherlands Pond
Pond group group stants included). MSWD—mean square
0.082 Rhyolite 500 Felsic volcanic of weighted deviates.
Z4A
485
Pb/238U

0.078 480 0.078


475
460
0.074
206

465
440
0.070 0.074
Concordia age = 455 Z5A
420
456.8 ± 3.1 Ma 457 ± 2 Ma
0.066 SHRIMP II data TIMS data
400
0.062 0.070
0.4 0.5 0.6 0.7 0.8 0.54 0.57 0.60 0.63 0.66 0.69
207 235 207 235
Pb/ U Pb/ U
320 Zagorevski et al.

Badger Bay
Sutherlands Pond gr.
Victoria Lake Mafic volcanic Red Indian
5477000 Supergroup 5358000 Felsic volcanic Lake
Stanley
Badger Waters Red Indian
Group Wigwam
Shoal Lake Group Brook Group
l’s
au p 457±4 Ma ck te 462+4/-2 Ma1
Arm
el P rou ba i
Fm. No ok g p le Su 5399000
ip ive
Bro Cr rus 457±1 Ma
t 457±3 Ma
In
5475000
Wild Bight 457±3 Ma Victoria Lake
Group 5355000 Supergroup
5396000
Victoria Lake
593000 596000 480000 483000 518000 524000

Figure 10. Simplified geology of the geochronology sample locations (after McConnell et al., 2002; Rogers et al., 2005; van Staal et al., 2005c).
For explanation of the fills, see Figure 5. From Dunning et al. (1987).

contained within a section of sandstone, was collected from the anchored by fraction Z1 with a concordant age of 457 ± 3.6 Ma.
Penny Brook Formation at the top of the Wild Bight Group ~30 m This concordant age provides the best estimate for the extrusion
below the stratigraphic contact with the Shoal Arm Formation age of the tuff.
(B. O’Brien, 2005, personal commun.). The sample contained
a small number of euhedral prismatic zircons, many of which Sutherlands Pond Group
contain abundant inclusions. Backscatter SEM images reveal
faint oscillatory zoning in many of the grains. As the sample did The first Sutherlands Pond group sample (RAX01–904,
not yield much zircon, and the presence of xenocrystic zircon in z7157; UTM zone 21, 517905 E, 5398510 N: Fig. 10) was col-
this thin tuff unit was a possibility, the sample was analyzed by lected from a quartz-feldspar phyric rhyolite unit that contains
SHRIMP. A concordia age, utilizing all of the SHRIMP analy- small black quartz phenocrysts. This rhyolite occurs as small
ses, is calculated to be 457.5 ± 2.7 Ma (mean square of weighted beds within a rhyolitic breccia sequence that typifies the Tims
deviates [MSWD] of concordance and equivalence = 0.82; n = Creek formation (Rogers et al., 2005). The sample contains abun-
18) (Fig. 9; Table 1). This date is interpreted to be the crystalliza- dant euhedral, prismatic zircons. Three-multigrain zircon TIMS
tion age of the felsic tuff. analyses from this sample are highly discordant (15%–57%)
owing to the presence of inherited zircon (Table 2, not plotted).
Noel Paul’s Brook Group Backscatter and cathodoluminescence scanning electron micro-
scope (SEM) images of the zircons reveal the presence of inher-
Crystal tuff from the Stanley Waters formation of the ited cores rimmed by thick, oscillatory-zoned magmatic rims. A
Noel Paul’s Brook group (van Staal et al., 2005c) was sampled concordia age, calculated from the SHRIMP analyses of mag-
along the eastern shore of the Stanley Waters arm of Victoria matic grains and rims (n = 21), is 456.8 ± 3.1 Ma (MSWD of
Lake (VL01–9104; Z7334; UTM zone 21, 418768 E, 5356382 concordance and equivalence = 1.6) (Table 1; Fig. 9). Inherited
N; Fig. 10). The Stanley Waters formation forms part of a nar- cores and entirely xenocrystic grains analyzed from this rhyo-
row but regionally extensive belt of coarse volcanogenic gray- lite range in age from ca. 970 to 1300 Ma (analyses in Table 2
wacke, fine-grained crystal tuff, and minor intercalations of marked by an asterisk, not plotted). The date of 456.8 ± 3.1 Ma is
green volcanogenic siltstone and shale (van Staal et al., 2005c). interpreted as the crystallization age of the rhyolite.
This sequence is directly adjacent to the ca. 565 Ma Valentine A second sample from the Sutherlands Pond group
Lake monzonite (Evans and Kean, 2002). Although the contact (MRB01–06, z7098; UTM zone 21, 521960 E, 5398445 N:
is not exposed, it is presumed to be a reverse fault. The volca- Fig. 10) was collected from an aphanitic, altered, brecciated
nogenic siltstone grades into Caradoc black shale, which caps felsic volcanic rock. A very small number of fair quality, small
the sequence. Therefore, the tuff is interpreted to represent the euhedral zircons with abundant inclusions and fractures were
upper part of a volcanogenic sequence at the base of the Caradoc retrieved from this sample. There was only enough material to
black shale, which forms a regional cover for the Victoria Lake analyze two multigrain fractions. One TIMS analysis is 34% dis-
Supergroup. This sample yielded abundant stubby to elongated cordant and contains a large inherited component (Z4A; Table 2;
prismatic zircon grains with melt inclusions through their cores Fig. 9). The other analysis (Z5A) is concordant, with an age of
(Fig. 9). The three zircon fractions analyzed define a mixing line 456.9 ± 1.3 Ma. Although there is only one concordant analysis,
(MSWD = 0.114) with an upper intercept of 835 ± 170 Ma and the age of 457 ± 2 Ma is consistent with the crystallization age
a lower intercept of 456 ± 12 Ma (Table 1; Fig. 9). This line is of the RAX01–904 sample from the Sutherlands Pond group.
TABLE 1. U/Pb SHRIMP ANALYTICAL DATA
Ages (Ma)
204 204 208 207 206 207 206 207
U Th Th Pb* Pb Pb ± 204Pb Pb ± 208Pb Pb ±207Pb Pb ± 206Pb Corr Pb ± 207Pb Pb ± 206Pb Pb ± 207Pb
206 206 206 206 235 235 238 238 206 206 238 238 206 206
Spot name (ppm) (ppm) U (ppm) (ppb) Pb Pb f(206)204 Pb Pb U U U U Coeff Pb Pb U U Pb Pb
RAX05-900 (Z8777): Felsic tuff, Penny Brook Formation
8777-7.1 116 33 0.297 8 4 0.000539 0.000380 0.0093 0.0830 0.0166 0.5217 0.0643 0.0730 0.0011 0.244 0.0518 0.0062 454 7 278 278
8777-10.1 316 86 0.282 23 1 0.000058 0.000128 0.0010 0.0910 0.0053 0.5737 0.0283 0.0738 0.0009 0.362 0.0564 0.0026 459 5 467 106
8777-11.1 287 101 0.363 21 0 0.000010 0.000010 0.0002 0.1197 0.0028 0.5618 0.0111 0.0730 0.0009 0.692 0.0558 0.0008 454 5 444 32
8777-13.1 462 223 0.498 35 7 0.000229 0.000063 0.0040 0.1553 0.0048 0.5493 0.0141 0.0732 0.0009 0.561 0.0544 0.0012 456 5 388 49
8777-14.1 290 144 0.512 22 3 0.000162 0.000142 0.0028 0.1613 0.0063 0.5744 0.0264 0.0732 0.0010 0.422 0.0569 0.0024 455 6 489 95
8777-15.1 214 94 0.454 16 5 0.000353 0.000137 0.0061 0.1280 0.0064 0.5520 0.0268 0.0723 0.0010 0.408 0.0554 0.0025 450 6 426 103
8777-16.1 277 131 0.488 21 3 0.000179 0.000163 0.0031 0.1546 0.0097 0.5521 0.0304 0.0728 0.0011 0.385 0.0550 0.0028 453 6 412 119
8777-17.1 191 72 0.390 14 3 0.000213 0.000181 0.0037 0.1235 0.0094 0.5641 0.0322 0.0734 0.0010 0.346 0.0557 0.0030 457 6 442 125
8777-18.1 359 115 0.331 27 3 0.000148 0.000092 0.0026 0.1035 0.0046 0.5682 0.0202 0.0742 0.0009 0.465 0.0556 0.0018 461 6 435 72
8777-19.1 369 214 0.599 30 3 0.000134 0.000073 0.0023 0.1807 0.0080 0.5711 0.0227 0.0755 0.0012 0.511 0.0548 0.0019 469 7 406 79
8777-20.1 441 237 0.555 34 6 0.000223 0.000076 0.0039 0.1702 0.0058 0.5481 0.0162 0.0735 0.0009 0.523 0.0541 0.0014 457 5 375 58
8777-27.1 282 114 0.418 21 4 0.000226 0.000193 0.0039 0.1370 0.0098 0.5627 0.0327 0.0721 0.0009 0.336 0.0566 0.0031 449 5 476 127
8777-24.1 266 102 0.398 19 3 0.000164 0.000138 0.0028 0.1173 0.0072 0.5629 0.0260 0.0725 0.0010 0.409 0.0563 0.0024 451 6 464 97
8777-29.1 372 169 0.470 29 1 0.000041 0.000069 0.0007 0.1518 0.0040 0.5749 0.0166 0.0748 0.0009 0.535 0.0557 0.0014 465 6 441 56
8777-28.1 293 170 0.598 23 1 0.000055 0.000099 0.0010 0.1866 0.0052 0.5672 0.0200 0.0738 0.0009 0.460 0.0557 0.0018 459 5 441 72
8777-45.1 166 70 0.435 13 4 0.000368 0.000217 0.0064 0.1333 0.0094 0.5549 0.0380 0.0741 0.0010 0.323 0.0543 0.0036 461 6 385 154
8777-44.1 200 73 0.378 15 3 0.000272 0.000138 0.0047 0.1129 0.0063 0.5684 0.0263 0.0745 0.0010 0.408 0.0554 0.0024 463 6 427 98
8777-41.1 108 30 0.289 7 7 0.001031 0.000365 0.0179 0.0703 0.0143 0.4621 0.0601 0.0715 0.0013 0.258 0.0469 0.0059 445 8 43 278
8777-39.1 771 453 0.606 63 2 0.000037 0.000054 0.0007 0.1904 0.0031 0.5951 0.0153 0.0759 0.0010 0.597 0.0569 0.0012 472 6 486 47
RAX01-904 (Z7157): Rhyolite, Sutherlands Pond group
7157-46.1 47 35 0.773 4 4 0.001226 0.000525 0.0213 0.2544 0.0211 0.5945 0.0848 0.0711 0.0015 0.267 0.0607 0.0084 443 9 627 330
7157-87.1 432 277 0.663 34 3 0.000114 0.000036 0.0020 0.2109 0.0033 0.5508 0.0109 0.0712 0.0009 0.725 0.0561 0.0008 443 5 456 31
7157-43.1 559 369 0.682 44 2 0.000059 0.000049 0.0010 0.2135 0.0030 0.5584 0.0140 0.0714 0.0009 0.619 0.0568 0.0011 444 6 482 45
7157-39.1 495 228 0.475 37 1 0.000049 0.000033 0.0009 0.1524 0.0020 0.5660 0.0099 0.0718 0.0008 0.754 0.0572 0.0007 447 5 498 26
7157-84.1 152 104 0.710 12 4 0.000420 0.000106 0.0073 0.2295 0.0053 0.5736 0.0204 0.0719 0.0009 0.478 0.0579 0.0018 448 6 524 71
7157-55.1 385 251 0.673 30 2 0.000094 0.000046 0.0016 0.2140 0.0028 0.5592 0.0113 0.0722 0.0008 0.663 0.0561 0.0009 450 5 458 34
7157-78.1 123 66 0.559 9 5 0.000630 0.000139 0.0109 0.1678 0.0062 0.5335 0.0258 0.0723 0.0011 0.435 0.0535 0.0024 450 7 351 102
7157-86.1 126 35 0.285 9 2 0.000296 0.000151 0.0051 0.0992 0.0062 0.5832 0.0283 0.0723 0.0012 0.458 0.0585 0.0025 450 7 548 98
7157-99.1 364 232 0.658 29 5 0.000227 0.000067 0.0039 0.2040 0.0035 0.5530 0.0152 0.0727 0.0010 0.609 0.0552 0.0012 452 6 420 50
7157-83.1 274 177 0.665 22 2 0.000117 0.000054 0.0020 0.2081 0.0033 0.5723 0.0129 0.0728 0.0009 0.627 0.0571 0.0010 453 5 494 39
7157-92.1 269 171 0.656 21 3 0.000163 0.000059 0.0028 0.2079 0.0049 0.5714 0.0165 0.0729 0.0010 0.586 0.0569 0.0013 454 6 486 53
(continued)
TABLE 1. U/Pb SHRIMP ANALYTICAL DATA (CONTINUED)
Ages (Ma)
204 204 208 207 206 207 206 207
U Th Th Pb* Pb Pb ± 204Pb Pb ± 208Pb Pb ±207Pb Pb ± 206Pb Corr Pb ± 207Pb Pb ± 206Pb Pb ± 207Pb
206 206 204 206 206 235 235 238 238 206 206 238 238 206 206
Spot name (ppm) (ppm) U (ppm) (ppb) Pb Pb f(206) Pb Pb U U U U Coeff Pb Pb U U Pb Pb
RAX01-904 (Z7157): Rhyolite, Sutherlands Pond group (continued)
7157-77.1 281 152 0.557 22 5 0.000274 0.000081 0.0048 0.1739 0.0037 0.5568 0.0215 0.0729 0.0008 0.408 0.0554 0.0020 454 5 427 81
7157-69.1 208 68 0.339 15 5 0.000386 0.000102 0.0067 0.1034 0.0043 0.5683 0.0195 0.0737 0.0009 0.461 0.0559 0.0017 458 5 450 70
7157-20.1 312 166 0.550 25 1 0.000044 0.000093 0.0008 0.1796 0.0055 0.5716 0.0176 0.0739 0.0008 0.466 0.0561 0.0015 460 5 457 62
7157-45.3 425 135 0.329 31 5 0.000199 0.000048 0.0035 0.1039 0.0024 0.5726 0.0131 0.0740 0.0009 0.631 0.0561 0.0010 460 5 457 40
7157-91.1 69 3 0.040 5 3 0.000650 0.000306 0.0113 0.0112 0.0115 0.5582 0.0520 0.0740 0.0012 0.291 0.0547 0.0049 460 7 400 215
7157-37.1 287 118 0.426 22 4 0.000228 0.000086 0.0040 0.1332 0.0037 0.5638 0.0171 0.0741 0.0009 0.509 0.0552 0.0015 461 5 420 60
7157-101.1 229 95 0.431 17 4 0.000248 0.000083 0.0043 0.1351 0.0037 0.5666 0.0177 0.0743 0.0009 0.480 0.0553 0.0015 462 5 423 63
7157-14.1 416 227 0.563 33 2 0.000081 0.000047 0.0014 0.1776 0.0025 0.5741 0.0111 0.0745 0.0008 0.648 0.0559 0.0008 463 5 449 33
7157-4.1 498 346 0.717 41 0 0.000010 0.000010 0.0002 0.2272 0.0019 0.5791 0.0079 0.0752 0.0008 0.844 0.0558 0.0004 468 5 446 17
7157-6.1 112 63 0.586 9 2 0.000230 0.000138 0.0040 0.1869 0.0063 0.5954 0.0274 0.0761 0.0010 0.391 0.0567 0.0024 473 6 481 97
7157-7.1 142 100 0.723 12 1 0.000110 0.000100 0.0019 0.2293 0.0050 0.6135 0.0208 0.0767 0.0009 0.464 0.0580 0.0018 476 5 531 68
7157-5.1* 73 51 0.718 13 1 0.000090 0.000095 0.0016 0.2278 0.0047 1.5942 0.0493 0.1614 0.0021 0.527 0.0717 0.0019 964 12 976 55
7157-100.1* 69 92 1.375 15 7 0.000717 0.000162 0.0124 0.4230 0.0079 1.5824 0.0687 0.1624 0.0027 0.498 0.0707 0.0027 970 15 947 80
7157-79.1* 67 49 0.757 13 1 0.000123 0.000095 0.0021 0.2431 0.0052 1.7060 0.0463 0.1680 0.0021 0.555 0.0737 0.0017 1001 11 1032 47
7157-88.1* 50 27 0.544 9 5 0.000678 0.000211 0.0118 0.1539 0.0090 1.6470 0.0895 0.1686 0.0030 0.442 0.0708 0.0035 1004 17 953 104
7157-18.1* 124 44 0.366 22 3 0.000162 0.000049 0.0028 0.1089 0.0032 1.7502 0.0324 0.1729 0.0020 0.701 0.0734 0.0010 1028 11 1026 27
7157-36.1* 95 28 0.302 17 5 0.000380 0.000086 0.0066 0.0891 0.0039 1.7502 0.0451 0.1748 0.0023 0.601 0.0726 0.0015 1039 12 1003 43
7157-101.2* 117 38 0.339 21 2 0.000128 0.000098 0.0022 0.1050 0.0040 1.8462 0.0515 0.1791 0.0026 0.618 0.0748 0.0017 1062 14 1062 45
7157-37.2* 75 13 0.182 13 6 0.000474 0.000118 0.0082 0.0551 0.0047 1.8560 0.0606 0.1850 0.0027 0.554 0.0728 0.0020 1094 15 1007 57
7157-87.2* 340 165 0.501 70 1 0.000021 0.000030 0.0004 0.1527 0.0016 2.0648 0.0307 0.1932 0.0023 0.849 0.0775 0.0006 1139 12 1134 16
7157-80.1* 129 66 0.526 28 3 0.000121 0.000066 0.0021 0.1605 0.0032 2.2415 0.0476 0.2008 0.0027 0.722 0.0810 0.0012 1180 15 1220 29
7157-45.1* 81 25 0.320 19 2 0.000148 0.000069 0.0026 0.0975 0.0044 2.8578 0.0708 0.2226 0.0038 0.762 0.0931 0.0015 1296 20 1490 31
Note: See Stern (1997). Uncertainties reported at 1σ (absolute) and are calculated by numerical propagation of all known sources of error; f(206)204 refers to mole fraction of total 206Pb that is due to common Pb, calculated
using the 204Pb method; common Pb composition used is the surface blank.
TABLE 2. U-Pb TIMS ANALYTICAL DATA
Isotopic ratios†† Ages (Ma)##
Fract.* Description† 206 208 207 206 207 206 207 207
Wt. U Pb§ Pb# Pb** Pb Pb ± 1SE Pb ± 1SE Corr.§§ Pb ± 1SE Pb Pb Pb Disc
204 206 235 238 206 238 235 206
(ug) (ppm) (ppm) Pb (pg) Pb U Abs U Abs Coeff. Pb Abs U ± 2SE U ± 2SE Pb ± 2SE (%)
MRB01-06 (Z7098): Brecciated felsic volcanic rock, Sutherlands Pond group
Z4A (42) Co,Clr,Eu,St,aIn,fFr,M1° 15 251 22 1983 10 0.19 0.70183 0.00134 0.07973 0.00008 0.769 0.06385 0.00008 494.5 1.0 539.9 1.6 736.5 5.5 34.1
Z5A (48) Co,Clr,Eu,St,aIn,fFr,M3° 18 239 19 2503 8 0.21 0.56821 0.00102 0.07344 0.00011 0.696 0.05611 0.00007 456.9 1.3 456.9 1.3 456.8 5.8 0.0

RAX01-904 (Z7157): Rhyolite, Sutherlands Pond group


Z1 (10) Co,Clr,Eu,Pr,rIn,Dia 26 128 15 7290 3 0.18 1.10375 0.00233 0.11263 0.00025 0.859 0.07107 0.00008 688.0 2.9 755.2 2.3 959.5 4.7 29.8
Z2 (7) Co,Clr,Eu,Pr,rIn,Dia 30 129 12 2313 9 0.14 0.74132 0.00191 0.08806 0.00019 0.793 0.06106 0.00010 544.1 2.2 563.2 2.2 641.2 6.7 15.8
Z3A (4) Co,Clr,Eu,St,rIn,Dia 29 206 33 9106 6 0.22 2.14224 0.00323 0.13993 0.00019 0.953 0.11103 0.00005 844.3 2.1 1162.5 2.1 1816.4 1.7 57.0

VL01-9104 (Z7334): Felsic Tuff, Noel Paul’s Brook group


Z1 (18) Co,Clr,Pr,Eu,aIn,NM5° 6 219 17 1262 5 0.12 0.58395 0.00231 0.07490 0.00011 0.500 0.05654 0.00020 465.6 1.3 467.0 3.0 473.8 15.3 1.8
Z2 (19) Co,Clr,Pr,Eu,aIn,NM5° 10 579 45 2367 11 0.15 0.58595 0.01076 0.07446 0.00124 0.572 0.05708 0.00093 463.0 14.9 468.3 13.8 494.4 70.1 6.6
Z3 (20) Co,Clr,Pr,Eu,aIn,NM5° 20 327 26 3228 10 0.13 0.63209 0.00456 0.07926 0.00056 0.967 0.05784 0.00011 491.7 6.6 497.4 5.7 523.7 8.1 6.3
*All fractions are zircon and have been abraded following the method of Krogh (1982). Number in parentheses refers to the number of grains in the analysis.

Zircon descriptions: Co—colorless; Clr—clear; fFr—few fractures; aIn—abundant inclusions; rIn—rare inclusions; Eu—euhedral; Pr—prismatic; St—stubby prism; Dia—diamagnetic; M1°—magnetic @ 1.8A, 1°SS; M3°—
magnetic @ 1.8A, 3°SS; NM5°—nonmagnetic @ 1.8A, 5°SS.
§
Radiogenic Pb.
#
Measured ratio, corrected for spike and fractionation.
**Total common Pb in analysis corrected for fractionation and spike.
††
Corrected for blank Pb and U and common Pb; errors quoted are 1σ absolute; procedural blank values for this study ranged from 0.1 to 0.3 pg for U and 2 to 10 pg for Pb; Pb blank isotopic composition is based on the
analysis of procedural blanks; corrections for common Pb were made using Stacey and Kramers (1975) compositions.
§§
Correlation coefficient.
##
Corrected for blank and common Pb; errors quoted are 2σ in Ma.
324 Zagorevski et al.

Hence we interpret the 457 ± 2 Ma as the crystallization age of isotopic characteristics of felsic volcanic rocks. The lowest εNd val-
the Sutherlands Pond group felsic volcanic rock. ues recorded in the volcanic rocks of the Exploits Subzone occur
in the youngest rocks (Wigwam Brook Group: εNd −4), whereas
DISCUSSION most of the rocks have generally maintained positive to slightly
negative εNd values (εNd +8 to −1) with TDM ages generally in the
Identification of four new ages of Caradoc peri-Gondwanan 1.30–0.6 Ga range (Fig. 11; Zagorevski et al., 2006). In contrast to
volcanism in the Victoria Arc, in addition to a previously obtained the Exploits Subzone, volcanic rocks of the Notre Dame Subzone
age (453 ± 3 Ma: Zagorevski et al., 2007a), demonstrates that Cara- display consistently lower εNd values, with εNd in felsic volcanic
doc volcanism was much more widespread than previously postu- rocks commonly in the −3 to −10 range (Whalen et al., 1997; Lis-
lated (e.g., O’Brien et al., 1997; MacLachlan et al., 2001), allowing senberg et al., 2005a; Zagorevski et al., 2006), indicating a greater
more precise timing constraints to be placed on the deposition of the contribution of continental crust and/or more mature continental
cover and the cessation of magmatism in the Victoria Arc. The age crust (Fig. 11).
obtained for the Penny Brook Formation extends the age range of Zircon inheritance supports the Sm/Nd isotope data and
volcanism in Notre Dame Bay until 457.5 ± 2.7 Ma and constrains indicates that the volcanic rocks of the Exploits Subzone have
the maximum age of sedimentation in the overlying manganiferrous been deposited on continental basement that underwent tectono-
chert, shale, and black shale of the Caradoc Shoal Arm Formation, magmatic events at ca. 560 Ma and ca. 900–1200 Ma (Zagor-
which forms part of the Victoria Arc cover sequence. The volcanic evski et al., 2007a). This is supported by the zircon inheritance in
rocks from the Sutherlands Pond and Noel Paul’s Brook groups are the Sutherlands Pond group rhyolite (ca. 970–1200 Ma: Table 2).
the same age (457 ± 2; 456.8 ± 3.1; 457 ± 3.6 Ma). Because they This basement is similar to the Proterozoic Crippleback Igneous
predate deposition of most of the black shale, these ages constrain Suite and Sandy Lake Group (ca. 560 Ma, TDM ~1300: Kerr et al.,
the maximum age of deposition of the cover of the Victoria Arc 1995; Rogers et al., 2006), which stratigraphically underlie the
independently from sparse fossil evidence. oldest parts of the Penobscot arc in Newfoundland (Rogers et al.,
2006). Similar relationships occur in the broadly correlative New
Contrasting Ordovician Arc Systems: Red Indian Lake River Belt in New Brunswick (Johnson and McLeod, 1996). This
and Victoria Arcs basement likely represents a fragment of Ganderia (van Staal et
al., 2004), a Gondwana-derived microcontinent, which is charac-
The ages obtained in this study and by Zagorevski (2006) terized by zircon provenance in the 0.54–0.55, 0.6–0.8, 1.0–1.55,
indicate that the Victoria Arc was magmatically active from and 2.5–2.7 Ga age ranges (van Staal et al., 1996), εNd values
ca. 473 Ma until 453 Ma. The Victoria Arc magmatism was greater than −4, and TDM ages of late Precambrian igneous rocks
coeval with the Annieopsquotch accretionary tract, where mag- spanning 0.9 to 1.35 Ga (Kerr et al., 1995; Rogers et al., 2006).
matism initiated at ca. 480 Ma (Annieopsquotch ophiolite belt:
Dunning and Krogh, 1985; Lissenberg et al., 2005b) and con-
tinued until ca. 460 Ma (Red Indian Lake Arc: Zagorevski et al., Notre Dame/ Exploits Subzone Gander
2006). The subsequent juxtaposition of these coeval arc systems Dashwoods sz. Zone
along the RIL can make their separation very difficult in the 8
absence of a multidisciplinary data set.
Williams et al. (1988) noted the broad contrasts in stratig-
raphy, structure, fauna, plutonic rocks, radiogenic lead in min-
3
eral deposits, magnetic anomalies, and gravity anomalies as the
basis for separation of the Dunnage Zone into the Notre Dame Age (Ma)
εNd

and Exploits Subzones. Detailed mapping in central Newfound- 480 460 460 500 540
land has revealed additional criteria that can be utilized for the -2
separation of the arc systems, namely the zircon provenance and
the Sm/Nd isotopic characteristics of volcanic rocks. In the fol-
lowing sections we discuss the differences in the basement char-
-7 Legend
acteristics, tectonic history, and cover sequences between the
Mafic
Red Indian Lake and Victoria Arcs. These will then be utilized
Felsic
to reconstruct the Ordovician arc-arc collision in Newfoundland Gneiss
based, in part, on modern analogues in the southwest Pacific. -12

Basement Characteristics Figure 11. Sm/Nd isotopic evolution of the Notre Dame and Exploits
Subzones. Compiled from Jenner and Swinden (1993), Kerr et al.
(1995), Swinden et al. (1997), Whalen et al. (1997), MacLachlan and
The Red Indian Lake and Victoria Arcs can be effectively dif- Dunning (1998a, b), Rogers (2004), Lissenberg et al. (2005b), Rogers
ferentiated on the basis of the zircon inheritance and the Sm/Nd et al. (2006), and Zagorevski et al. (2006, 2007a).
Tectonic architecture of an arc-arc collision zone 325

The ensialic plutonic and volcanic rocks of the Annieops- SSE-directed emplacement of the Red Indian Lake Arc over the
quotch accretionary tract and the Red Indian Lake Arc have abun- Victoria Arc. This observation is confirmed by seismic-reflection
dant inheritance in the 935–1845 Ma age range (Dec et al., 1997; studies of Thurlow et al. (1992) and van der Velden et al. (2004),
Zagorevski et al., 2006), consistent with the presence of Laurentian which clearly show the Red Indian Line to emplace the Annie-
basement at depth (e.g., Cawood et al., 2001; Cawood and Nem- opsquotch accretionary tract and the Red Indian Lake Arc over
chin, 2001; Cawood et al., 1995). The presence of inheritance in the Exploits Subzone and the Victoria Arc (Fig. 8). Interpreta-
the 1.0–1.55 Ga age ranges is common to both subzones and thus tion of the migrated seismic reflection data by van der Velden et
is not a reliable differentiation tool. However, zircon inheritance al. (2004) suggests that the Victoria Arc and its basement were
in the 1.7–1.9 Ga age range is generally associated with the Lau- subducted to a depth of at least 18 km and underthrust the Annie-
rentian margin (Cawood and Nemchin, 2001). In addition, zircon opsquotch accretionary tract by at least 60 km (Fig. 8).
inheritance in the 520–565 Ma age range appears to be restricted
to the peri-Gondwanan terranes in the Dunnage Zone, owing Cover Sequences of the Notre Dame and Exploits Subzones
to the extensive magmatism of this age in the Ganderia-derived
basement (Dunning and O’Brien, 1989; Evans et al., 1990; Rog- The contrast between the cover sequences of the Notre Dame
ers et al., 2006). Although ca. 550 Ma magmatism (e.g., Cawood and Exploits Subzones is highly distinctive in the Newfoundland
et al., 2001, and references therein) related to the opening of Iape- Appalachians and can serve as a reliable tool to differentiate the
tus is present in Laurentia (Cawood et al., 2001), it is generally two arc systems. Adhering to the original observations of Williams
restricted to the Humber Zone. et al. (1988) the Notre Dame Subzone, and therefore the Annie-
Sm/Nd isotopic signatures and the presence of locally abun- opsquotch accretionary tract, contain a sub-Silurian unconformity
dant zircon inheritance in felsic volcanic rocks indicate the con- that is overlain by Silurian continental red sandstones and volcanic
tribution of continental basement to both the Red Indian Lake and rocks. In contrast, the rocks of the Exploits Subzone, and thus the
Victoria Arcs. The contrasts in the Sm/Nd isotopic characteristics Victoria Arc, are overlain by Caradoc black shale and display a gen-
and the presence of specific age ranges of zircon indicate that the erally continuous sedimentation through the Ordovician-Silurian
nature of the arc basement is fundamentally different and reflects boundary (i.e., Badger Group: Williams et al., 1995). Although
the peri-Laurentian and peri-Gondwanan derivation of the Red several models for the depositional setting of the Badger Group
Indian Lake and Victoria Arcs, respectively (Fig. 11). have been proposed (Wasowski et al., 1986), the potential tectonic
causes for the contrasts between the subzones have not yet been
Tectonic History addressed. We now briefly discuss several key observations that
allow a comparison with modern arc-arc collisions.
The Annieopsquotch accretionary tract and the Victoria Arc Parts of the Victoria Arc were covered by late Llanvirn lime-
underwent distinctly different tectonic histories in the Early to stone (Figs. 5, 7; e.g., Williams, 1995b; O’Brien et al., 1997;
Middle Ordovician. The Annieopsquotch accretionary tract McConnell et al., 2002), suggesting relatively shallow water
underwent deformation throughout much of its history, starting prior to the collision. The deposition of limestone was locally
at ca. 470 Ma and lasting until the collision with the Victoria Arc accompanied by formation of mélange, suggesting initiation of
(Fig. 5; Lissenberg et al., 2005a). Some of this deformation may deformation in the Caradoc (e.g., Sops Head Complex: McCon-
be related to the simultaneous collision between the Dashwoods nell et al., 2002). This was followed by widespread deposition
microcontinent and the Laurentian margin (Waldron and van of Caradoc black shale (Fig. 5; e.g., Williams, 1995b). In central
Staal, 2001). In contrast, the initiation of magmatism in the Vic- Newfoundland, Caradoc black shale of the Laurence Harbour
toria Arc at ca. 473 Ma followed a short orogenic episode (Fig. 5; Formation unconformably overlies the volcanic rocks of the Tally
Penobscot orogeny: Colman-Sadd et al., 1992) in the Exploits Pond Group (ca. 513 Ma: Rogers et al., 2005, 2006), suggesting
Subzone and Gander Zone. From the onset of magmatism, the that the Penobscot basement of the Victoria Arc was locally emer-
continental Victoria Arc was extensional (e.g., MacLachlan and gent prior to the Caradoc. The deposition of black shale above the
Dunning, 1998b; O’Brien et al., 1997; Rogers et al., 2003), open- Victoria Arc and its basement, followed by deposition of deep-
ing a wide Tetagouche-Exploits oceanic backarc basin in New- marine turbidites of the Badger Group (Williams et al., 1993),
foundland and New Brunswick (van Staal et al., 1998). support rapid submergence of the Victoria Arc (e.g., Williams,
The earliest deformation that is common to both the Red 1995b). The zircon provenance of the turbidites in the lower Bad-
Indian Lake and Victoria Arcs started during Caradoc time. This ger Group suggests derivation from the Laurentian margin or the
is constrained by the involvement of the Caradoc black shale and Notre Dame Subzone (McNicoll et al., 2001), indicating that the
volcanic rocks in the mélange zones that mark the D1 thrusts, collision of the Victoria Arc and the Annieopsquotch accretion-
including the Red Indian Line (e.g., McConnell et al., 2002; ary tract had already started by Early Ashgill time.
Rogers and van Staal, 2002). The D1 mélange zones most likely In modern arc-arc collisions, the downgoing plate (i.e., Snel-
represent the syncollisional terrane boundaries formed during lius Plateau and Morobe Shelf) undergoes rapid subsidence,
the accretion of the Victoria Arc to the Laurentian margin. Shear as indicated by drowned carbonate platforms (e.g., Abers and
sense indicators along the D1 shear zones suggest sinistral oblique McCaffrey, 1994; Galewsky and Silver, 1997; Pubellier et al.,
326 Zagorevski et al.

1999). At the same time the overriding plate is rapidly uplifted and 13). Consistent with this interpretation, structural and seismic-
forms an emergent orogen (e.g., Talaud Island, Huon-Finisterre reflection studies indicate that the Victoria Arc occupies a lower
Ranges: Fig. 1). The detritus derived from the emergent orogen is plate setting with respect to the Red Indian Lake Arc (Fig. 8; e.g.,
transported parallel to the trench and deposited as thick turbidite van der Velden et al., 2004). Hence it was partially subducted
sequences such as the Markam Canyon turbidites (Galewsky and under the Red Indian Lake Arc, resulting in loading of the Vic-
Silver, 1997; Whitmore et al., 1999). This detritus would have its toria Arc crust, rapid subsidence, and widespread deposition of
source in the overriding plate. black shale above the arc and backarc. This is consistent with the
The concurrence of the subsidence of the Victoria Arc and absence of the Upper Wild Bight Group equivalents in central
the influx of the peri-Laurentian detritus suggest that this was Newfoundland, which were probably overridden by the Annie-
caused by the collision with the Red Indian Lake Arc (Figs. 12, opsquotch accretionary tract during the arc-arc collision (Fig. 8)

A 465–460 Ma Red Indian Iapetus Dunnage Victoria


Exploits-
Lake group Tetagouche
(AAT) Ocean melange arc
backarc

Accreted
terranes
Iapetus
Ocean Plate

B 460–455 Ma Uplift of peri-Laurentian Loading of Exploits-


terranes RIL Victoria arc Tetagouche
backarc
SHC

forearc
Molucca Figure 12. Tectonic model for the
Iapetus
subduction? Sea Plate
Ocean Plate Caradoc arc-arc collision in central
Newfoundland (based on van Staal et
al., 1998). AAT—Annieopsquotch ac-
cretionary tract; DBL—Dog Bay Line;
RIL—Red Indian Line; SHC—Sops
C 455–450 Ma Emergent peri-Laurentian Caradoc black Head Complex; WPG—Windsor Point
terranes shale and Group.
RIL volcanic rocks Future
WPG DBL

initiation of
subduction
Slab
detachment

Iapetus
Ocean Plate

D 450–430 Ma Sub-Silurian unconformity RIL Badger Future


basin DBL
Tectonic architecture of an arc-arc collision zone 327

c. 450 Ma

Figure 13. Caradoc paleogeography of


Laurentia arc-arc collision in the Appalachian-
Caledonian orogen, slightly modified
Baltica from van Staal et al. (1998). Original
RILA figure courtesy of Conall MacNiocaill.
ETBAB—Exploits-Tetagouche backarc
VPA basin; RILA—Red Indian Lake Arc;
AB
ETB a VPA—Victoria-Popelogan Arc.
deri
Gan
lonia
Ava
Gondwana

similar to the Snellius Plateau (Fig. 2). The loading and subsid- peri-Laurentian provenance and their deposition on the subsided
ence of the Victoria Arc would have occurred contemporaneously Victoria Arc (McNicoll et al., 2001) signaled the emergence of the
with the Caradoc arc-arc collision, as indicated by the formation Notre Dame Subzone and Laurentia during the arc-arc collision.
of the Caradoc black shale mélange (Fig. 12; e.g., Sops Head
Complex: McConnell et al., 2002). Syncollisional Magmatism
Any subduction of the buoyant Victoria Arc crust and its
continental basement would have most likely resulted in the The presence of abundant Caradoc volcanism in the Exploits
emergence of the Annieopsquotch accretionary tract and the Subzone indicates that volcanism and deformation were coeval.
Notre Dame Subzone above sea level (e.g., Huon-Finisterre Caradoc felsic volcanic rocks in the Victoria Arc achieve distinctly
Range: see previous discussion). Although no unconformities of low εNd values (εNd −4 in the Wigwam Brook Group: Zagorevski
the right age have been yet described within the Annieopsquotch et al., 2006; εNd −4 in the Sutherlands Pond group: Rogers, 2004)
accretionary tract in central Newfoundland, an upper Caradoc that are atypical for the Exploits Subzone but common in the
unconformity has been defined in southwestern Newfoundland Notre Dame Subzone (Fig. 11). The presence of Caradoc black
below the Windsor Point Group (453 +5/–4 Ma: Dube et al., shale and moderately low εNd values distinguishes the Caradoc
1996), which unconformably overlies the Notre Dame Subzone volcanic rocks of the Exploits Subzone from the peri-Laurentian
plutons (ca. 469–488 Ma Cape Ray Igneous Complex: Dube et felsic volcanic rocks of the Red Indian Lake Group (εNd −6.4 to
al., 1996). We interpret this unconformity and the lack of late −7.7; Fig. 11; Zagorevski et al., 2006); however, the distinctly low
Caradoc rocks in the Annieopsquotch accretionary tract to imply εNd values in the Caradoc volcanic rocks are rather atypical for the
that the Annieopsquotch accretionary tract had indeed breached Exploits Subzone, which generally contained positive to slightly
sea level, leading to nondeposition and/or erosion through- negative εNd values throughout its history (Fig. 11). This may have
out the Caradoc (Fig. 12). Hence, the influx of turbidites with resulted from tapping of a larger proportion or a different part of
328 Zagorevski et al.

the Ganderian basement during magma generation, leading to a be poorly preserved owing to metamorphism and deformation as
rapid change in εNd values in the Caradoc. well as the presence of extensive cover sequences.
An alternate solution is more appealing in terms of an arc- In addition to these obstacles, the age of deformation, tec-
arc collision model, although it would be difficult to prove. The tonic relationships, nature of the basement, and preservation of
low εNd in the Sutherlands Pond group rhyolite may be a result of tectonic elements can vary significantly along the strike of the
subduction of peri-Laurentian sediment below the Victoria Arc. collision zone. First, both the Molucca and Solomon Sea arc-arc
Comparison with the geochemical evolution of the Halmahera- collisions are diachronous along strike of the orogen; hence the
Sangihe arc collision zone (Macpherson et al., 2003) indicates duration of arc volcanism, the age of the cover sequences, and the
that this is a feasible, albeit not a unique, solution. In both the timing of deformation will also be diachronous (Fig. 1). Second,
Halmahera and Sangihe Arcs, the geochemistry of volcanic rocks the polarity of thrusting can change along strike and may result
that erupted prior to and during the collision indicates increased in stratigraphic differences in the cover sequences. In the cen-
sediment flux through the subduction zone. The increased sedi- tral Molucca collision zone, for example, the Halmahera Arc is
ment flux could be a result of the collision of the facing accretion- subducted under the Sangihe Arc. At the same time, back thrusts
ary prisms, as is probably the case in the Molucca Sea, or deposi- emplace the Halmahera Arc over the Sangihe Arc in the north-
tion of emergent orogen-derived sediment along active trenches ern Molucca Sea (Fig. 2). Hence the polarity of thrusting changes
(e.g., Markham Canyon turbidites in the Trobriand Trough). A from north to south. Third, long-range correlation of the Appa-
similar explanation can be utilized for the Wigwam Brook Group lachian terranes may require adjustments to the local basement
felsic tuff; however, since the tuff may be epiclastic in nature, characteristics, as the nature of the basement is unlikely to remain
the low εNd values could also have been achieved by mixing with exactly the same along the whole strike of the orogen. Finally, the
sediment of low εNd values. Hence, the most likely source would absence of tectonic elements can result from either strike-slip exci-
still be the emergent peri-Laurentian terranes. sion (e.g., Elders, 1987) or the local peculiarity of the colliding arc
Although the discussion of the chemical characteristics of systems. Analogue experiments suggest that the behavior of the
the volcanic rocks is beyond the scope of this paper the (pre-) overriding arc during the collision will depend on the strength of
Caradocian volcanic rocks have a range of compositions from the arc lithosphere and the retroarc region; i.e., the arc and forearc
arclike to nonarc, which suggests eruption in an extensional may be deformed but preserved, the forearc may be subducted,
suprasubduction zone setting (e.g., Penny Brook Formation: the forearc and part of the arc may be subducted, or the whole arc
McConnell and O’Brien, 2000; Sutherlands Pond group: Evans may be subducted (Boutelier et al., 2003). The lack of preserva-
and Kean, 2002; Rogers, 2004; Rogers et al., 2005; Diversion tion of the forearc basin to the Red Indian Lake Arc, for example,
Lake Group: Swinden et al., 1989; Evans and Kean, 2002; Rog- indicates that the forearc from the central Newfoundland section
ers et al., 2005; Sops Head Complex: McConnell et al., 2002). An was removed by either strike-slip translation or subduction.
alternate setting for the syncollisional extensionlike magmatism
may be above a delaminated Iapetus slab (Fig. 12), similar to the Northern Appalachians
Pliocene and younger volcanic rocks in Mindanao (Sajona et al., Despite the potential problems with preservation and long-
2000). This is an appealing model because it would also explain range correlations, equivalents of the Annieopsquotch accretion-
the change of isotopic characteristics of the younger volcanic ary tract and Victoria Arc are recognized elsewhere in the north-
rocks in the Exploits Subzone, as both peri-Gondwanan and peri- ern Appalachians. The remnants of the Dashwoods microconti-
Laurentian sources can be involved. nent and its Notre Dame Arc suprastructure in New England are
preserved in the Shelburne Falls Arc (ca. 470–485 Ma: Karabi-
Along-Strike Variations nos et al., 1998) and the Chain Lakes Massif (>473 Ma: Gerbi et
al., 2006) both of which have clear peri-Laurentian affinities. The
Peri-Laurentian magmatism in the Annieopsquotch accre- Annieopsquotch accretionary tract is largely obscured in Atlantic
tionary tract (ca. 460–480 Ma: Dunning and Krogh, 1985; Canada by cover sequences; however, correlatives of the Annie-
Zagorevski et al., 2006) was in part coeval with the Victoria Arc opsquotch accretionary tract include the Boil Mountain Complex
(ca. 453–473 Ma: e.g., MacLachlan and Dunning, 1998b; Zago- and the Jim Pond Formation (ca. 477–485 Ma: Coish and Rog-
revski et al., 2007a), requiring subduction on both sides of the ers, 1987; Gerbi et al., 2006), which are adjacent to the Chain
Iapetus Ocean, which culminated in a Molucca Sea–type arc-arc Lakes Massif (Fig. 3) as well as to the Caucomgomoc inlier (van
collision (Fig. 13; van Staal et al., 1998). The key evidence that Staal et al., 1998).
can be used to infer the presence of distinct and separated arc The equivalents of the Victoria Arc (Fig. 3: Popelogan Inlier:
sequences in Newfoundland include structural and tectonic his- Wilson, 2003; Munsungun inlier: Ayuso and Schulz, 2003; Ayuso
tory, stratigraphy, basement characteristics, radiogenic lead in et al., 2003) and backarc (Fig. 3: Bathurst Supergroup: van Staal
mineral deposits, fauna, zircon provenance, and Sm/Nd isotopic et al., 2003) are well preserved in New Brunswick and Maine;
characteristics (see previous discussion; Williams et al., 1988). however, the equivalents of the Victoria Arc in western Maine,
Extrapolation of these interpretations is difficult outside of New- New Hampshire, and Vermont are less clear. The Middle to Late
foundland, where the equivalents of the Dunnage Zone tend to Ordovician Bronson Hill Arc (Moench and Aleinikoff, 2003),
Tectonic architecture of an arc-arc collision zone 329

which lies to the east of the peri-Laurentian Shelburne Falls Arc, by a change of polarity and a step-back of the subduction into the
is a probable correlative of the Munsungun inlier, which is an Tetagouche-Exploits backarc (Fig. 12C, D) and formation of the
equivalent of the Popelogan Arc in Maine (Ayuso and Schulz, Quimby sequence of the Bronson Hill Arc.
2003; Ayuso et al., 2003). The Bronson Hill Arc contains two
distinct sequences: the Ammonoosuc (457–470 Ma) and Qui- British Caledonides
mby (435–456 Ma) sequences, which are thought to have formed A direct correlation between Newfoundland and the British
above subduction zones of opposite polarity (Moench and Aleini- Caledonides is difficult, because the well-preserved sedimen-
koff, 2003). Although the Ammonoosuc sequence has been tec- tary sequences of the late Llanvirn to the Late Silurian Southern
tonically correlated with the peri-Laurentian Shelburne Falls Arc Uplands accretionary complex (e.g., Legget, 1987) are largely
(e.g., Moench and Aleinikoff, 2003), the contact relationships are absent in the northern Appalachians, a feature that may reflect
obscured by the Connecticut Valley Trough and younger deposits, strike-slip excision or subduction. Van Staal et al. (1998) proposed
and the age span of the Ammonoosuc sequence overlaps with the a correlation of the Annieopsquotch accretionary tract with the
age of the Victoria Arc, although relationships with the Penobscot South Connemara Group in Ireland (e.g., Ryan and Dewey, 2004)
Arc basement have not been identified. In addition, the Ammo- and the correlative Northern Belt of the Southern Uplands terrane
noosuc sequence shares several characteristics with the peri- in Scotland. This correlation was in part supported by Armstrong
Gondwanan Victoria Arc (van Staal et al., 1998; Hibbard et al., and Owen (2001), who proposed a new terrane, Novantia, to be
2004), including an important period of extension as exemplified a correlative of the Annieopsquotch accretionary tract. Novantia
by the Chickwolnepy sheeted intrusions (ca. 467 Ma; Fitz, 2002; has been subducted underneath the Southern Uplands; however,
Moench and Aleinikoff, 2003). The cover sequences are also its presence was deduced on the basis of geophysical modeling.
remarkably similar to the Victoria Arc. The type Ammonoosuc Armstrong and Owen (2001) also recognized correlatives of the
volcanic rocks are conformably overlain by the Caradoc sulfidic Victoria Arc in the Grangegeeth Terrane of eastern Ireland and
black shale of the Partridge Formation, followed by an uncon- proposed that these have also been in part subducted under the
formity and deposition of marine graywacke of the Ashgill Qui- Southern Uplands. The juxtaposition of the Novantia and Grang-
mby Formation (Moench and Aleinikoff, 2003). The presence of egeeth terranes in the Ordovician was subsequently supported by
a Caradoc-Ashgill unconformity between the Ammonoosuc and the presence of distinctly peri-Gondwanan (ca. 560 Ma) zircons
Quimby sequences (Moench and Aleinikoff, 2003) differentiates in the Caradoc Portpatrick Formation of the Southern Uplands
the Bronson Hill Arc from the Victoria Arc; however, this can be (Phillips et al., 2003).
easily reconciled by considering the along-strike variations in an Comparison of the Newfoundland arc-arc collision with
arc-arc collision zone. the Southern Uplands illustrates the importance of along-strike
The Quimby sequence, overlying the Ammonoosuc sequence, variation in a collisional setting. The Southern Uplands accre-
comprises marine shales, graywackes, and conglomerates tionary complex contains both the peri-Laurentian and the peri-
intruded by Upper Ordovician to Lower Silurian plutons (Moench Gondwanan provenance, indicating emergence of both Lau-
and Aleinikoff, 2003). The Quimby sequence corresponds to the rentia and the Grangegeeth Terrane during the collision. The
Bronson Hill Arc in southern New Hampshire and central Mas- Badger Group in Newfoundland, on the other hand, contains a
sachusetts (454–442 Ma: Tucker and Robinson, 1990; Karabinos distinctly peri-Laurentian provenance (McNicoll et al., 2001),
et al., 1998). This phase of the Bronson Hill Arc development has supporting the emergence of the peri-Laurentian terranes and
been interpreted to have occurred following an arc collision of the subsidence of the Victoria Arc during the collision.
Shelburne Falls Arc and the Ammonoosuc sequence of the Bron-
son Hill Arc with the Laurentian margin (Karabinos et al., 1998; CONCLUSIONS
Moench and Aleinikoff, 2003). The collision was followed by
establishment of a west-dipping subduction zone along the com- Comparison with the modern analogues indicates that the rec-
posite Laurentian margin following a polarity flip (Karabinos et ognition of ancient arc-arc collision zones requires an integrated
al., 1998; Moench and Aleinikoff, 2003). and multidisciplinary approach. In Newfoundland, two coeval but
The tectonic evolution of New England broadly mirrors the distinct and separated arc systems are preserved in the Dunnage
Ordovician arc-arc collision in Newfoundland. The accretion of Zone: the peri-Laurentian Red Indian Lake Arc (ca. 473–460 Ma)
the Ammonoosuc sequence of the Bronson Hill Arc to the com- and the peri-Gondwanan Victoria Arc (ca. 473–453 Ma). The two
posite Laurentian margin, comprising the peri-Laurentian Shel- arc systems collided above a doubly dipping subduction zone
burne Falls Arc and Jim Pond Formation, occurred at ca. 457 Ma, during early to middle Caradoc time along the Red Indian Line
followed by a magmatic gap lasting several million years, deep- (Figs. 12, 13), the main Iapetus suture zone, marking the end of
marine sedimentation, and probable subduction reversal (Moench the Taconic orogeny and the start of the Salinic orogeny. Follow-
and Aleinikoff, 2003). This closely agrees with the accretion of ing the collision the subduction stepped back into the Exploits-
the Victoria Arc to the Red Indian Lake Arc, where the Victoria Tetagouche backarc basin, leading to the closure of that tract
arc was loaded, leading to subsidence and deep-marine sedimen- along the Silurian Dog Bay Line (Williams et al., 1993; van Staal
tation (Fig. 12B, C). Subsequent convergence was accommodated et al., 1998).
330 Zagorevski et al.

The separation of the Annieopsquotch accretionary tract and heavy liquid techniques. Mineral separates were sorted by mag-
the Victoria Arc is supported by contrasts in structural and tec- netic susceptibility using a FrantzTM isodynamic separator. Mul-
tonic history, stratigraphy, basement characteristics, radiogenic tigrain zircon fractions analyzed were very strongly air abraded
lead in mineral deposits, and fauna that are apparent across the following the method of Krogh (1982). Treatment of analyti-
Red Indian Line (Williams et al., 1988). A brief review of the cal errors follows Roddick et al. (1987) with errors for the ages
northern Appalachians and the British Caledonides reveals great reported at the 2σ level (Table 2). U-Pb concordia diagrams are
similarity in the correlative tracts that also preserve evidence of presented in Figure 9. The data are plotted in concordia diagrams
an Ordovician arc-arc collision, although the interpretation of the with errors at the 2σ level (Fig. 9), using Isoplot v. 2.49 (Ludwig,
polarity of subduction zones (e.g., Armstrong and Owen, 2001; 2001) to generate the plots.
Moench and Aleinikoff, 2003) differs from the models based A concordia age (Ludwig, 1998) is calculated for some of
on Newfoundland (e.g., van Staal et al., 1998). Future research the samples presented in this paper. A concordia age incorporates
should aim to reconcile these differences. The timing of the arc- errors in the decay constants and includes both an evaluation of
arc collision appears to have been restricted to the Caradoc in all concordance and an evaluation of equivalence of the data (how
of the correlative tracts. This similarity may reflect either coeval well the data fit the assumption that they are repeated measure-
collision along the entire length of the tract or the postcollisional ments of the same point). The calculated concordia ages and errors
dispersal of the accreted terranes along the Laurentian margin quoted in the text are at 2σ, with decay constant errors included.
(e.g., Elders, 1987).
ACKNOWLEDGMENTS
APPENDIX: U-Pb GEOCHRONOLOGY ANALYTICAL
TECHNIQUES This paper is a contribution to the Targeted Geoscience Initia-
tive 3–Appalachians (2005–2010) and Red Indian Line Tar-
SHRIMP II analyses were conducted at the Geological geted Geoscience Initiative projects (2000–2003). B. O’Brien
Survey of Canada (GSC) using analytical procedures described (Geological Survey of Newfoundland and Labrador) is thanked
by Stern (1997), with standards and U-Pb calibration methods for the insightful field trip through Notre Dame Bay and cen-
following Stern and Amelin (2003). Zircons from the samples tral Newfoundland that helped to inspire this manuscript. The
were cast in 2.5 cm diameter epoxy mounts (GSC mount 373 for manuscript has benefited from reviews from S. Castonguay,
sample RAX05–900 and GSC mount 254 for sample RAX01– D. Chew, A. Draut, and P. Karabinos. This is Geological Survey
904 along with fragments of the GSC laboratory standard zircon of Canada Contribution 20060169.
(z6266, with 206Pb/238U age = 559 Ma). The midsections of the
zircons were exposed using 9, 6, and 1 µm diamond compound, REFERENCES CITED
and the internal features of the zircons were characterized with
Abbott, L.D., Silver, E.A., and Galewsky, J., 1994, Structural evolution of a
backscatter electrons (BSE) and cathodoluminescence (CL) uti- modern arc-continent collision in Papua New Guinea: Tectonics, v. 13,
lizing a Cambridge Instruments scanning electron microscope p. 1007–1034, doi: 10.1029/94TC01623.
(SEM). Mount surfaces were evaporatively coated with 10 nm Abers, G.A., and McCaffrey, R., 1994, Active arc-continent collision; earth-
quakes, gravity anomalies, and fault kinematics in the Huon-Finisterre
of high-purity Au. Analyses were conducted using a 16O– primary collision zone, Papua New Guinea: Tectonics, v. 13, p. 227–245, doi:
beam, projected onto the zircons at 10 kV. The sputtered area 10.1029/93TC02940.
used for analysis was ca. 25 µm in diameter with a beam current Armstrong, H.A., and Owen, A.W., 2001, Terrane evolution of the paratectonic
Caledonides of northern Britain: Geological Society [London] Journal,
of ~5 nA and 13 nA for RAX05–900 and RAX01–904, respec- v. 158, p. 475–486.
tively. The count rates of 10 isotopes of Zr+, U+, Th+, and Pb+ Arnott, R.J., McKerrow, W.S., and Cocks, L.R.M., 1985, The tectonics and dep-
in zircon were sequentially measured over 6 scans with a single ositional history of the Ordovician and Silurian rocks of Notre Dame Bay,
Newfoundland: Canadian Journal of Earth Sciences, v. 22, p. 607–618.
electron multiplier and a pulse counting system with deadtime of Ayuso, R.A., and Schulz, K.J., 2003, Nd-Pb-Sr isotope geochemistry and
22 ns. Off-line data processing was accomplished using custom- origin of the Ordovician Bald Mountain and Mount Chase massive sul-
ized in-house software. The 1σ external errors of 206Pb/238U ratios fide deposits, northern Maine: Economic Geology Monographs, v. 11,
p. 611–630.
reported in Table 1 incorporate a ±1.0% error in calibrating the Ayuso, R.A., Wooden, J.L., Foley, N.K., Slack, J.F., Sinha, A.K., and Persing,
standard zircon (see Stern and Amelin, 2003). No fractionation H., 2003, Pb isotope geochemistry and U-Pb zircon (SHRIMP-RG) ages
correction was applied to the Pb isotope data; common Pb cor- of the Bald Mountain and Mount Chase massive sulfide deposits, north-
ern Maine; mantle and crustal contributions in the Ordovician: Economic
rection utilized the measured 204Pb/206Pb and compositions mod- Geology Monographs, v. 11, p. 589–609.
eled after Cumming and Richards (1975). The 206Pb/238U ages for Bird, P., 2003, An updated digital model of plate boundaries: Geochemistry,
the analyses have been corrected for common Pb using both the Geophysics, Geosystems—G3, v. 4, p. 52.
Bostock, H.H., 1988, Geology and petrochemistry of the Ordovician volcano-
204 and 207 methods (Stern, 1997), but there is generally no sig- plutonic Robert’s Arm Group, Notre Dame Bay, Newfoundland: Geologi-
nificant difference in the results. cal Survey of Canada Bulletin 369, 84 p.
U-Pb TIMS analytical methods utilized in this study are out- Boutelier, D., Chemenda, A., and Burg, J.P., 2003, Subduction versus accretion
of intra-oceanic volcanic arcs; insight from thermo-mechanical analogue
lined in Parrish et al. (1987). Heavy mineral concentrates were experiments: Earth and Planetary Science Letters, v. 212, p. 31–45, doi:
prepared by standard crushing, grinding, WilfleyTM table, and 10.1016/S0012-821X(03)00239-5.
Tectonic architecture of an arc-arc collision zone 331

Calon, T.J., and Green, F.K., 1987, Preliminary results of a detailed structural Galewsky, J., and Silver, E.A., 1997, Tectonic controls on facies transitions in an
analysis of the Buchans Mine area: Geological Survey of Canada Paper, oblique collision; the western Solomon Sea, Papua New Guinea: Geologi-
v. 86–24, p. 273–288. cal Society of America Bulletin, v. 109, p. 1266–1278, doi: 10.1130/0016-
Cawood, P.A., van Gool, J.A.M., and Dunning, G.R., 1995, Collisional tecton- 7606(1997)109<1266:TCOFTI>2.3.CO;2.
ics along the Laurentian margin of the Newfoundland Appalachians: Geo- Galewsky, J., Silver, E.A., Gallup, C.D., Edwards, R.L., and Potts, D.C., 1996,
logical Association of Canada Special Paper, v. 41, p. 283–301. Foredeep tectonics and carbonate platform dynamics in the Huon Gulf,
Cawood, P.A., and Nemchin, A.A., 2001, Paleogeographic development of the Papua New Guinea: Geology, v. 24, p. 819–822, doi: 10.1130/0091-
East Laurentian margin; constraints from U-Pb dating of detrital zircons 7613(1996)024<0819:FTACPD>2.3.CO;2.
in the Newfoundland Appalachians: Geological Society of America Bul- Gerbi, C.C., Johnson, S.E., and Aleinikoff, J.N., 2006, Origin and orogenic role
letin, v. 113, p. 1234–1246, doi: 10.1130/0016-7606(2001)113<1234: of the Chain Lakes massif, Maine and Quebec: Canadian Journal of Earth
PDOTEL>2.0.CO;2. Sciences, v. 43, p. 339–366, doi: 10.1139/E05-112.
Cawood, P.A., McCausland, P.J.A., and Dunning, G.R., 2001, Opening Iape- Hall, R., 2000, Neogene history of collision in the Halmahera region, Indone-
tus; constraints from the Laurentian margin in Newfoundland: Geologi- sia: Proceedings of the Indonesian Petroleum Association, 27th Annual
cal Society of America Bulletin, v. 113, p. 443–453, doi: 10.1130/0016- Convention, p. 487–493.
7606(2001)113<0443:OICFTL>2.0.CO;2. Hall, R., 2002, Cenozoic geological and plate tectonic evolution of SE Asia and
Chandler, F.W., Sullivan, R.W., and Currie, K.L., 1987, The age of the Spring- the SW Pacific: Computer-based reconstructions, model and animations:
dale Group, western Newfoundland, and correlative rocks; evidence for a Journal of Asian Earth Sciences, v. 20, p. 353–431, doi: 10.1016/S1367-
Llandovery overlap assemblage in the Canadian Appalachians: Transac- 9120(01)00069-4.
tions of the Royal Society of Edinburgh: Earth Sciences, v. 78, p. 41–49. Hibbard, J., van Staal, C., Rankin, D., and Williams, H., 2004, Lithotectonic
Coish, R.A., and Rogers, N.W., 1987, Geochemistry of the Boil Mountain ophi- map of the Appalachian Orogen, Canada–United States of America: Geo-
olitic complex, Northwest Maine, and tectonic implications: Contributions logical Survey of Canada, Map 2096A, scale 1:500,000.
to Mineralogy and Petrology, v. 97, p. 51–65, doi: 10.1007/BF00375214. Jenner, G.A., and Swinden, H.S., 1993, The Pipestone Pond Complex, central
Colman-Sadd, S., Hayes, J., Knight, I., and Pereira, C.P.G., 1990, The geology Newfoundland; complex magmatism in an eastern Dunnage Zone ophiol-
of the Island of Newfoundland; map 90–01: Report of Activities—New- ite: Canadian Journal of Earth Sciences, v. 30, p. 434–448.
foundland Geological Survey Branch, v. 1990, p. 24. Johnson, S.C., and McLeod, M.J., 1996, The New River Belt; a unique segment
Colman-Sadd, S.P., Dunning, G.R., and Dec, T., 1992, Dunnage-Gander rela- along the western margin of the Avalon composite terrane, southern New
tionships and Ordovician orogeny in central Newfoundland; a sediment Brunswick, Canada: Geological Society of America Special Paper 304,
provenance and U/Pb age study: American Journal of Science, v. 292, p. 149–164.
p. 317–355. Karabinos, P., Samson, S.D., Hepburn, J.C., and Stoll, H.M., 1998, Taconian
Cumming, G.L., and Richards, J.R., 1975, Ore lead in a continuously chang- Orogeny in the New England Appalachians; collision between Lau-
ing Earth: Earth and Planetary Science Letters, v. 28, p. 155–171, doi: rentia and the Shelburne Falls Arc: Geology, v. 26, p. 215–218, doi:
10.1016/0012-821X(75)90223-X. 10.1130/0091-7613(1998)026<0215:TOITNE>2.3.CO;2.
Dec, T., Swinden, H.S., and Dunning, R.G., 1997, Lithostratigraphy and geo- Kerr, A., Jenner, G.A., and Fryer, B.J., 1995, Sm-Nd isotopic geochemistry of
chemistry of the Cottrells Cove Group, Buchans-Roberts Arm volcanic Precambrian to Paleozoic granitoid suites and the deep-crustal structure
belt; new constraints for the paleotectonic setting of the Notre Dame Sub- of the southeast margin of the Newfoundland Appalachians: Canadian
zone, Newfoundland Appalachians: Canadian Journal of Earth Sciences, Journal of Earth Sciences, v. 32, p. 224–245.
v. 34, p. 86–103. Krogh, T.E., 1982, Improved accuracy of U-Pb ages by creation of more con-
Dube, B., Dunning, G.R., Lauziere, K., and Roddick, J.C., 1996, New cordant systems using an air abrasion technique: Geochimica et Cosmo-
insights into the Appalachian Orogen from geology and geochronology chimica Acta, v. 46, p. 637–649, doi: 10.1016/0016-7037(82)90165-X.
along the Cape Ray fault zone, Southwest Newfoundland: Geological Kusky, T.M., Kidd, W.S.F., and Bradley, D.C., 1987, Displacement history of
Society of America Bulletin, v. 108, p. 101–116, doi: 10.1130/0016- the Northern Arm Fault, and its bearing on the post-Taconic evolution of
7606(1996)108<0101:NIITAO>2.3.CO;2. north-central Newfoundland: Journal of Geodynamics, v. 7, p. 105–133,
Dunning, G.R., 1987, Geology of the Annieopsquotch Complex, Southwest doi: 10.1016/0264-3707(87)90067-6.
Newfoundland: Canadian Journal of Earth Sciences, v. 24, p. 1162–1174. Lallemand, S.E., Popoff, M., Cadet, J.-P., Bader, A.-G., Pubellier, M., Rangin,
Dunning, G.R., and Krogh, T.E., 1985, Geochronology of ophiolites of the C., and Deffontaines, B., 1998, Genetic relations between the central
Newfoundland Appalachians: Canadian Journal of Earth Sciences, v. 22, and southern Philippine Trench and the Sangihe Trench: Journal of Geo-
p. 1659–1670. physical Research, B, Solid Earth and Planets, v. 103, p. 933–950, doi:
Dunning, G.R., and O’Brien, S.J., 1989, Late Proterozoic–early Paleozoic crust 10.1029/97JB02620.
in the Hermitage Flexure, Newfoundland Appalachians; U/Pb ages and Leggett, J.K., 1987, The Southern Uplands as an accretionary prism; the impor-
tectonic significance: Geology, v. 17, p. 548–551, doi: 10.1130/0091- tance of analogues in reconstructing palaeogeography: Geological Soci-
7613(1989)017<0548:LPEPCI>2.3.CO;2. ety [London] Journal, v. 144, p. 737–752.
Dunning, G.R., Kean, B.F., Thurlow, J.G., and Swinden, H.S., 1987, Geochro- Lissenberg, C.J., 2005, The origin and tectonic evolution of the Annieopsquotch
nology of the Buchans, Roberts Arm, and Victoria Lake groups and Mans- ophiolite belt, Newfoundland Appalachians: Ontario, Canada, University
field Cove Complex, Newfoundland: Canadian Journal of Earth Sciences, of Ottawa, 149 p.
v. 24, p. 1175–1184. Lissenberg, C.J., Zagorevski, A., McNicoll, V.J., van Staal, C.R., and Whalen, J.B.,
Dunning, G.R., O’Brien, S.J., Colman-Sadd, S.P., Blackwood, R.F., Dickson, 2005a, Assembly of the Annieopsquotch accretionary tract, Newfoundland
W.L., O’Neill, P.P., and Krogh, T.E., 1990, Silurian orogeny in the New- Appalachians; age and geodynamic constraints from syn-kinematic intru-
foundland Appalachians: Journal of Geology, v. 98, p. 895–913. sions: Journal of Geology, v. 113, p. 553–570, doi: 10.1086/431909.
Elders, C.F., 1987, The provenance of granite boulders in conglomerates of the Lissenberg, C.J., van Staal, C.R., Bedard, J.H., and Zagorevski, A., 2005b, Geo-
northern and central belts of the Southern Uplands of Scotland: Geologi- chemical constraints on the origin of the Annieopsquotch ophiolite belt,
cal Society [London] Journal, v. 144, p. 853–863. Newfoundland Appalachians: Geological Society of America Bulletin, v.
Evans, D.T.W., and Kean, B.F., 2002, The Victoria Lake Supergroup, central 117, p. 1413–1426, doi: 10.1130/B25731.1.
Newfoundland—Its definition, setting and volcanogenic massive sulphide Ludwig, K.R., 1998, On the treatment of concordant uranium-lead ages: Geo-
mineralization: Newfoundland Department of Mines and Energy, Geo- chimica et Cosmochimica Acta, v. 62, p. 665–676, doi: 10.1016/S0016-
logical Survey, Open File NFLD/2790, 68 p. 7037(98)00059-3.
Evans, D.T.W., Kean, B.F., and Dunning, G.R., 1990, Geological studies, Vic- Ludwig, K.R., 2001, User’s manual for Isoplot/Ex rev. 2.49: A Geochronologi-
toria Lake Group, central Newfoundland: Government of Newfoundland cal Toolkit for Microsoft Excel: Berkeley, California, Berkeley Geochro-
and Labrador, Department of Mines and Energy, Geological Survey, nological Center, Special Publication 1a, 55 p.
Report 90–1, p. 131–144. MacLachlan, K., and Dunning, G., 1998a, U-Pb ages and tectonomagmatic
Fitz, T.J., 2002, Magmatism in an extensional setting within the Bronson Hill relationships of early Ordovician low-Ti tholeiites, boninites and related
belt: The Chickwolnepy Intrusions of northern New Hampshire: Physics plutonic rocks in central Newfoundland, Canada: Contributions to Miner-
and Chemistry of the Earth, v. 27, p. 97–108. alogy and Petrology, v. 133, p. 235–258, doi: 10.1007/s004100050450.
332 Zagorevski et al.

MacLachlan, K., and Dunning, G.R., 1998b, U-Pb ages and tectonomagmatic Rogers, N., van Staal, C.R., and Theriault, R., 2003, Volcanology and tec-
relationships of Middle Ordovician volcanic rocks of the Wild Bight tonic setting of the northern Bathurst mining camp; Part I, Extension
Group, Newfoundland Appalachians: Canadian Journal of Earth Sciences, and rifting of the Popelogan Arc: Economic Geology Monographs, v. 11,
v. 35, p. 998–1017, doi: 10.1139/cjes-35-9-998. p. 157–179.
MacLachlan, K., O’Brien, B.H., and Dunning, G.R., 2001, Redefinition of the Rogers, N., van Staal, C.R., Pollock, J., and Zagorevski, A., 2005, Geology,
Wild Bight Group, Newfoundland; implications for models of island-arc Lake Ambrose and part of Buchans, Newfoundland (NTS 12-A/10 and
evolution in the Exploits Subzone: Canadian Journal of Earth Sciences, part of 12-A/15): Map, scale 1:50,000.
v. 38, p. 889–907, doi: 10.1139/cjes-38-6-889. Rogers, N., van Staal, C.R., McNicoll, V., Pollock, J., and Zagorevski, A.W.,
Macpherson, C.G., Forde, E.J., Hall, R., and Thirlwall, M.F., 2003, Geochemi- 2006, Neoproterozoic and Cambrian arc magmatism along the eastern
cal evolution of magmatism in an arc-arc collision; the Halmahera and margin of the Victoria Lake Supergroup: A remnant of Ganderian base-
Sangihe Arcs, eastern Indonesia: Geological Society [London] Special ment in central Newfoundland? in McCausland, P.J.A., Murphy, J.B., and
Publication 219, p. 207–220. MacNiocaill, C., eds., Endings and Beginnings: Paleogeography of the
McConnell, B., and O’Brien, B., 2000, Stratigraphic and petrogenetic relation- Neoproterozoic-Cambrian Transition: Amsterdam, Elsevier, Precambrian
ships between arc and non-arc basalts in western and southern parts of Research, v. 147, p. 320–341.
the Ordovician Wild Bight Group, Dunnage Zone, Newfoundland: St. Ryan, P.D., and Dewey, J.F., 2004, The South Connemara Group reinterpreted:
John’s, NF, Canada, Government of Newfoundland and Labrador Current A subduction-accretion complex in the Caledonides of Galway Bay, west-
Research Report, v. 2000–1, p. 217–230. ern Ireland: Journal of Geodynamics, v. 37, p. 513–529, doi: 10.1016/
McConnell, B.J., O’Brien, B.H., and Nowlan, G.S., 2002, Late Middle Ordovi- j.jog.2004.02.018.
cian olistostrome formation and magmatism along the Red Indian Line, Sajona, F.G., Maury, R.C., Pubellier, M., Leterrier, J., Bellon, H., and Cotten,
the Laurentian Arc–Gondwanan Arc boundary, at Sops Head, Newfound- J., 2000, Magmatic source enrichment by slab-derived melts in a young
land: Canadian Journal of Earth Sciences, v. 39, p. 1625–1633, doi: post-collision setting, central Mindanao (Philippines): Lithos, v. 54,
10.1139/e02-084. p. 173–206, doi: 10.1016/S0024-4937(00)00019-0.
McKerrow, W.S., and van Staal, C.R., 2000, The Palaeozoic time scale reviewed: Stacey, J.S., and Kramers, J.D., 1975, Approximation of terrestrial lead iso-
Geological Society [London] Special Publication 179, p. 5–8. tope evolution by a two-stage model: Earth and Planetary Science Letters,
McNicoll, V., van Staal, C.R., and Waldron, J.W.F., 2001, Accretionary his- v. 26, p. 207–221, doi: 10.1016/0012-821X(75)90088-6.
tory of the northern Appalachians: SHRIMP study of Ordovician–Silurian Stern, R.A., 1997, The GSC Sensitive High Resolution Ion Microprobe
syntectonic sediments in the Canadian Appalachians: St. John’s, New- (SHRIMP); analytical techniques of zircon U-Th-Pb age determina-
foundland: Geological Association of Canada/Mineralogical Association tions and performance evaluation: Geological Survey of Canada, Current
of Canada Annual Meeting Abstracts, v. 26, p. 100. Research, Report 1997-F, 31 p.
Milsom, J., 2001, Subduction in eastern Indonesia; how many slabs?: Tectono- Stern, R.A., and Amelin, Y., 2003, Assessment of errors in SIMS zircon U-
physics, v. 338, p. 167–178, doi: 10.1016/S0040-1951(01)00137-8. Pb geochronology using a natural zircon standard and NIST SRM 610
Moench, R., and Aleinikoff, J.N., 2003, Stratigraphy, geochronology, and glass: Chemical Geology, v. 197, p. 111–146, doi: 10.1016/S0009-
accretionary terrane settings of two Bronson Hill arc sequences, northern 2541(02)00320-0.
New England: Physics and Chemistry of the Earth, v. 28, p. 113–160. Swinden, H.S., Jenner, G.A., Kean, B.F., and Evans, D.T.W., 1989, Volcanic
O’Brien, B., Swinden, H.S., Dunning, G.R., Williams, S.H., and O’Brien, F.H.C., rock geochemistry as a guide for massive sulphide exploration in central
1997, A peri-Gondwanan arc–back arc complex in Iapetus; Early–Mid Newfoundland: Report of Activities—Mineral Development Division (St
Ordovician evolution of the Exploits Group, Newfoundland: American John’s, 1985), v. 89–1, p. 201–219.
Journal of Science, v. 297, p. 220–272. Swinden, H.S., Jenner, G.A., and Szybinski, Z.A., 1997, Magmatic and tectonic
O’Brien, B.H., 2003, Geology of the central Notre Dame Bay region (Parts of evolution of the Cambrian–Ordovician Laurentian margin of Iapetus; geo-
NTS areas 2E/3,6,11) northeastern Newfoundland: Government of New- chemical and isotopic constraints from the Notre Dame Subzone, New-
foundland and Labrador, Department of Mines and Energy, Geological foundland: Geological Society of America Memoir 191, p. 337–365.
Survey, Report 03–03, 147 p. Thurlow, J.G., Spencer, C.P., Boerner, D.E., Reed, L.E., and Wright, J.A., 1992,
Parrish, R.R., Roddick, J.C., Loveridge, W.D., Sullivan, R.W., and Anonymous, Geological interpretation of a high resolution reflection seismic survey at
1987, Uranium-lead analytical techniques at the Geochronology Labora- the Buchans Mine, Newfoundland: Canadian Journal of Earth Sciences,
tory, Geological Survey of Canada: Geological Survey of Canada Paper, v. 29, p. 2022–2037.
v. 87–2, p. 3–7. Tucker, R.D., and Robinson, P., 1990, Age and setting of the Bronson Hill mag-
Phillips, E.R., Evans, J.A., Stone, P., Horstwood, M.S.A., Floyd, J.D., Smith, matic arc; a re-evaluation based on U-Pb zircon ages in southern New
R.A., Akhurst, M.C., and Barron, H.F., 2003, Detrital Avalonian zircons in England: Geological Society of America Bulletin, v. 102, p. 1404–1419,
the Laurentian Southern Uplands terrane, Scotland: Geology, v. 31, p. 625– doi: 10.1130/0016-7606(1990)102<1404:AASOTB>2.3.CO;2.
628, doi: 10.1130/0091-7613(2003)031<0625:DAZITL>2.0.CO;2. Tucker, R.D., O’Brien, S.J., and O’Brien, B.H., 1994, Age and implications of
Pickering, K.T., 1987, Wet-sediment deformation in the Upper Ordovician Early Ordovician (Arenig) plutonism in the type area of the Bay du Nord
Point Leamington Formation; an active thrust-imbricate system during Group, Dunnage Zone, southern Newfoundland Appalachians: Canadian
sedimentation, Notre Dame Bay, north-central Newfoundland, in Jones, Journal of Earth Sciences, v. 31, p. 351–357.
M.F., and Preston, R.M.F., eds., Deformation of Sediments and Sedi- Valverde-Vaquero, P., and van Staal, C.R., 2002, Geology and magnetic anoma-
mentary Rocks: Geological Society [London] Special Publication 29, lies of the Exploits-Meelpaeg boundary zone in the Victoria Lake area
p. 213–239. (central Newfoundland); regional implications: Newfoundland Geologi-
Pickett, J.W., 1987, Geology and geochemistry of the Skidder Basalt: Geologi- cal Survey Branch, Current Research, Report 02–1, p. 197–209.
cal Survey of Canada Paper, v. 86–24, p. 195–218. Valverde-Vaquero, P., van Staal, C.R., McNicoll, V., and Dunning, G.R., 2006,
Pubellier, M., Bader, A.G., Rangin, C., Deffontaines, B., and Quebral, R., 1999, Mid–Late Ordovician magmatism and metamorphism along the Gander
Upper plate deformation induced by subduction of a volcanic arc; the Snel- margin in central Newfoundland: Geological Society [London] Journal,
lius Plateau (Molucca Sea, Indonesia and Mindanao, Philippines): Tecto- v. 163, p. 347–362, doi: 10.1144/0016-764904-130.
nophysics, v. 304, p. 345–368, doi: 10.1016/S0040-1951(98)00300-X. van der Pluijm, B.A., Karlstrom, K.E., and Williams, P.F., 1987, Fossil evidence
Roddick, J.C., Loveridge, W.D., and Parrish, R.R., 1987, Precise U/Pb dating of for fault-derived stratigraphic repetition in the northeastern Newfoundland
zircon at the sub-nanogram Pb level: Chemical Geology: Isotope Geosci- Appalachians: Canadian Journal of Earth Sciences, v. 24, p. 2337–2350.
ence Section, v. 66, p. 111–121, doi: 10.1016/0168-9622(87)90034-0. van der Velden, A.J., van Staal, C.R., and Cook, F.A., 2004, Crustal structure, fossil
Rogers, N., 2004, Red Indian Line geochemical database: Geological Survey of subduction, and the tectonic evolution of the Newfoundland Appalachians;
Canada Open File 4605. evidence from a reprocessed seismic reflection survey: Geological Society
Rogers, N., and van Staal, C.R., 2002, Toward a Victoria Lake Supergroup; a of America Bulletin, v. 116, p. 1485–1498, doi: 10.1130/B25518.1.
provisional stratigraphic revision of the Red Indian to Victoria lakes area, van der Voo, R., Johnson, R.J.E., van der Pluijm, B.A., and Knutson, L.C.,
central Newfoundland: Newfoundland Geological Survey Branch, Cur- 1991, Paleogeography of some vestiges of Iapetus; paleomagnetism of the
rent Research, Report 02–1, p. 185–195. Ordovician Robert’s Arm, Summerford and Chanceport groups, central
Tectonic architecture of an arc-arc collision zone 333

Newfoundland: Geological Society of America Bulletin, v. 103, p. 1564– the Cambrian–Ordovician Notre Dame Arc for the evolution of the central
1575, doi: 10.1130/0016-7606(1991)103<1564:POSVOI>2.3.CO;2. mobile belt, Newfoundland Appalachians: Geological Society of America
van Staal, C.R., 2005, North America; Northern Appalachians, in Selley, R.C., Memoir 191, p. 367–395.
et al., eds., Encyclopedia of Geology: Oxford, Elsevier, p. 81–92. Whalen, J.B., McNicoll, V.J., van Staal, C.R., Lissenberg, C.J., Longstaffe, F.J.,
van Staal, C.R., Sullivan, R.W., and Whalen, J.B., 1996, Provenance of tec- Jenner, G.A., and van Breemen, O., 2006, Spatial, temporal and geochem-
tonic history of the Gander Zone in the Caledonian/Appalachian Orogen; ical characteristics of Silurian collision-zone magmatism, Newfoundland
implications for the origin and assembly of Avalon: Geological Society of Appalachians: An example of a rapidly evolving magmatic system related
America Special Paper 304, p. 347–367. to slab break-off: Lithos, v. 89, p. 377–404.
van Staal, C.R., Dewey, J.F., MacNiocaill, C., and McKerrow, W.S., 1998, The Whitmore, G.P., Crook, K.A.W., and Johnson, D.P., 1999, Sedimentation in a
Cambrian–Silurian tectonic evolution of the Northern Appalachians and complex convergent margin; the Papua New Guinea collision zone of the
British Caledonides; history of a complex, west and southwest Pacific- western Solomon Sea: Marine Geology, v. 157, p. 19–45, doi: 10.1016/
type segment of Iapetus, in Blundell, D.J., and Scott, A.C., eds., Lyell: S0025-3227(98)00132-7.
The Past Is the Key to the Present: Geological Society [London] Special Williams, H., 1995a, Chapter 2: Temporal and spatial divisions, in Williams,
Publication 143, p. 199–242. H., ed., Geology of the Appalachian-Caledonian Orogen in Canada and
van Staal, C.R., Wilson, R.A., Rogers, N., Fyffe, L.R., Langton, J.P., McCutch- Greenland: Geological Survey of Canada, p. 23–42.
eon, S.R., McNicoll, V., and Ravenhurst, C.E., 2003, Geology and tec- Williams, H., 1995b, Chapter 3: Dunnage Zone—Newfoundland, in Williams,
tonic history of the Bathurst Supergroup, Bathurst mining camp, and its H., ed., Geology of the Appalachian-Caledonian Orogen in Canada and
relationships to coeval rocks in southwestern New Brunswick and adja- Greenland: Geological Survey of Canada, p. 142–166.
cent mine; a synthesis: Economic Geology Monographs, v. 11, p. 37–60. Williams, H., Colman-Sadd, S.P., Swinden, H.S., and Anonymous, 1988, Tec-
van Staal, C.R., McNicoll, V., Valverde-Vaquero, P., Barr, S.M., Fyffe, L.R., tonic-stratigraphic subdivisions of central Newfoundland: Geological
and Reusch, D.N., 2004, Ganderia, Avalonia, and the Salinic and Acadian Survey of Canada Paper, v. 88–1B, p. 91–98.
orogenies: Geological Society of America Abstracts with Programs, v. 36, Williams, H., Currie, K.L., and Piasecki, M.A.J., 1993, The Dog Bay Line; a
no. 2, p. 128–129. major Silurian tectonic boundary in Northeast Newfoundland: Canadian
van Staal, C.R., Lissenberg, C.J., Pehrsson, S., Zagorevski, A., Valverde- Journal of Earth Sciences, v. 30, p. 2481–2494.
Vaquero, P., Herd, R.K., McNicoll, V., and Whalen, J.B., 2005a, Geology, Williams, H., Lafrance, B., Dean, P.L., Williams, P.F., Pickering, K.T., and van
Puddle Pond, Newfoundland (NTS 12-A/05): Map, scale 1:50,000. der Pluijm, B.A., 1995, Chapter 4: Badger belt, in Williams, H., ed., Geol-
van Staal, C.R., Valverde-Vaquero, P., Zagorevski, A., Boutsma, S., Pehrsson, ogy of the Appalachian-Caledonian Orogen in Canada and Greenland:
S., van Noorden, M., and McNicoll, V., 2005b, Geology, King George IV Geological Survey of Canada, p. 403–413.
Lake, Newfoundland (NTS 12-A/04): Map, scale 1:50,000. Wilson, R.A., 2003, Geochemistry and petrogenesis of Ordovician arc-related
van Staal, C.R., Valverde-Vaquero, P., Zagorevski, A., Rogers, N., Lissenberg, mafic volcanic rocks in the Popelogan Inlier, northern New Brunswick:
C.J., and McNicoll, V., 2005c, Geology, Victoria Lake, Newfoundland Canadian Journal of Earth Sciences, v. 40, p. 1171–1189, doi: 10.1139/
(NTS 12-A/06): Map, scale 1:50,000. e03-034.
van Staal, C.R., Whalen, J.B., McNicoll, V.J., Pehrsson, S., Lissenberg, C.J., Zagorevski, A., Rogers, N., McNicoll, V., Lissenberg, C.J., van Staal, C.R., and
Zagorevski, A., van Breemen, O., and Jenner, G.A., 2007, The Notre Valverde-Vaquero, P., 2006, Lower to Middle Ordovician evolution of
Dame arc and the Taconic orogeny in Newfoundland, in Hatcher, R.D., peri-Laurentian arc and back-arc complexes in the Iapetus: Constraints
Jr., Carlson, M.P., McBride, J.H., and Martínez-Catalán, J.R., eds., 4-D from the Annieopsquotch Accretionary Tract, central Newfoundland:
Framework of Continental Crust: Geological Society of America Memoir Geological Society of America Bulletin, v. 118, p. 324–342, doi: 10.1130/
200, p. 511–552, doi: 10.1130/2007.1200(26). B25775.1.
Waldron, J.W.F., and van Staal, C.R., 2001, Taconian Orogeny and the accre- Zagorevski, A., McNicoll, V., van Staal, C., and Rogers, N., 2007a, Upper
tion of the Dashwoods Block; a peri-Laurentian microcontinent in Cambrian to Upper Ordovician peri-Gondwanan island arc activity in the
the Iapetus Ocean: Geology, v. 29, p. 811–814, doi: 10.1130/0091- Victoria Lake Supergroup, central Newfoundland: Tectonic development
7613(2001)029<0811:TOATAO>2.0.CO;2. of the northern Ganderian margin: American Journal of Science, v. 307,
Wasowski, J.J., Jacobi, R.D., Elliott, C.G., Williams, P.F., Arnott, R., McKer- p. 339–370.
row, W.S., and Cocks, L.R.M., 1986, The tectonics and depositional his- Zagorevski, A., van Staal, C.R., and McNicoll, V.J., 2007b, Distinct Taconic,
tory of the Ordovician and Silurian rocks of Notre Dame Bay, Newfound- Salinic and Acadian deformation along the Iapetus suture zone, New-
land; discussions and reply: Canadian Journal of Earth Sciences, v. 23, foundland Appalachians: Canadian Journal of Earth Sciences, v. 44.
p. 583–590. doi:10.1139/E07-037.
Whalen, J.B., Jenner, G.A., Longstaffe, F.J., Gariepy, C., and Fryer, B.J., 1997,
Implications of granitoid geochemical and isotopic (Nd, O, Pb) data from MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

The Catalina Schist: Evidence for middle Cretaceous subduction erosion of southwestern
North America

M. Grove*
Department of Earth & Space Sciences, University of California, Los Angeles, 3806 Geology, Los Angeles, California 90095, USA

G.E. Bebout
Department of Earth & Environmental Sciences, Lehigh University, 31 Williams Drive, Bethlehem, Pennsylvania 18015, USA

C.E. Jacobson
Department of Geological and Atmospheric Sciences, 253 Science I, Iowa State University, Ames, Iowa 50011-3212, USA

A.P. Barth
Department of Earth Sciences, Indiana–Purdue University, 723 West Michigan Street, Indianapolis, Indiana 46202, USA

D.L. Kimbrough
Department of Geological Sciences, San Diego State University, 5500 Campanile Drive, San Diego, California 92182-1020, USA

R.L. King
School of Earth and Environmental Sciences, Washington State University, Pullman, Washington 99164-2812, USA

Haibo Zou
O.M. Lovera
Department of Earth & Space Sciences, University of California, Los Angeles, 3806 Geology, Los Angeles, California 90095, USA

B.J. Mahoney
Department of Geology, University of Wisconsin–Eau Claire, Phillips 157, Eau Claire, Wisconsin 54702-4004, USA

G.E. Gehrels
Department of Geosciences, University of Arizona, Tucson, Gould-Simpson Building 529, Tucson, Arizona 85721, USA

ABSTRACT

The Catalina Schist underlies the inner southern California borderland of south-
western North America. On Santa Catalina Island, amphibolite facies rocks that
recrystallized and partially melted at ca. 115 Ma and at 40 km depth occur atop an
inverted metamorphic stack that juxtaposes progressively lower grade, high-pressure/
temperature (PT) rocks across low-angle faults. This inverted metamorphic sequence

*marty@oro.ess.ucla.edu

Grove, M., Bebout, G.E., Jacobson, C.E., Barth, A.P., Kimbrough, D.L., King, R.L., Zou, H., Lovera, O.M., Mahoney, B.J., and Gehrels, G.E., 2008, The Catalina
Schist: Evidence for middle Cretaceous subduction erosion of southwestern North America, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and
Applications of the Sedimentary Record in Arc Collision Zones: Geological Society of America Special Paper 436, p. 335–361, doi: 10.1130/2008.2436(15). For
permission to copy, contact editing@geosociety.org. ©2008 The Geological Society of America. All rights reserved.

335
336 Grove et al.

has been regarded as having formed within a newly initiated subduction zone. How-
ever, subduction initiation at ca. 115 Ma has been difficult to reconcile with regional
geologic relationships, because the Catalina Schist formed well after emplacement of
the adjacent Peninsular Ranges batholith had begun in earnest. New detrital zircon
U-Pb age results indicate that the Catalina Schist accreted over a ~20 m.y. interval.
The amphibolite unit metasediments formed from latest Neocomian to early Aptian
(122–115 Ma) craton-enriched detritus derived mainly from the pre-Cretaceous wall
rocks and Early Cretaceous volcanic cover of the Peninsular Ranges batholith. In
contrast, lawsonite-blueschist and lower grade rocks derived from Cenomanian sedi-
ments dominated by this batholith’s plutonic and volcanic detritus were accreted
between 97 and 95 Ma. Seismic data and geologic relationships indicate that the
Catalina Schist structurally underlies the western margin of the northern Penin-
sular Ranges batholith. We propose that construction of the Catalina Schist com-
plex involved underthrusting of the Early Cretaceous forearc rocks to a subcrustal
position beneath the western Peninsular Ranges batholith. The heat for amphibolite
facies metamorphism and anatexis observed within the Catalina Schist was supplied
by the western part of the batholith while subduction was continuous along the mar-
gin. Progressive subduction erosion ultimately juxtaposed the high-grade Catalina
Schist with lower grade blueschists accreted above the subduction zone by 95 Ma.
This coincided with an eastern relocation of arc magmatism and emplacement of the
ca. 95 Ma La Posta tonalite-trondjhemite-granodiorite suite of the eastern Peninsular
Ranges batholith. Final assembly of the Catalina Schist marked the initial stage of the
Late Cretaceous–early Tertiary craton-ward shift of arc magmatism and deformation
of southwestern North America that culminated in the Laramide orogeny.

Keywords: Catalina Schist, U-Pb, zircon, subduction erosion.

INTRODUCTION intervening rocks (Page, 1981; Hall, 1991) or by lateral transport


of hundreds (Hill and Dibblee, 1953; Suppe, 1970; Dickinson,
Subduction erosion is a fundamental convergent margin pro- 1983; Dickinson et al., 2005) to thousands of kilometers (Page,
cess that tectonically shortens the forearc region by displacing 1982; McWilliams and Howell, 1982; Vedder et al., 1983; Debi-
rocks from the overriding plate to subcrustal positions and/or into che et al., 1987).
the asthenosphere (Scholl et al., 1980; von Huene and Scholl, Like the Sur-Obispo terrane, seismic data and geologic
1991, 1993). Roughly half of all present-day convergent margins relationships indicate that the high-pressure/temperature (PT)
exhibit geologic and/or geophysical evidence for subduction- Catalina Schist of southern California is also in direct contact
related removal of rocks from the forearc regions of the overrid- with its coeval magmatic arc. This relationship was revealed by
ing plate to positions beneath the arc and/or into the mantle (e.g., middle Miocene extension that exhumed the Catalina Schist from
Clift and Vannucchi, 2004). The globally average rate at which beneath the northwestern margin of the Peninsular Ranges batho-
continental debris is transported toward the mantle is significant lith across an east-dipping Miocene detachment fault (Crouch
(~2.5 km3/year; Scholl and von Huene, 2007). Unfortunately, its and Suppe, 1993) (Fig. 1). The dramatic extension of the border-
consumptive nature means that direct evidence for subduction land occurred during a middle Miocene microplate capture event
erosion in ancient convergent margins is generally ambiguous or (e.g., Nicholson et al., 1994) triggered by Pacific–North Ameri-
lacking. Indirect evidence for subduction erosion such as “miss- can shearing along the margin (e.g., Atwater, 1970). During this
ing” or “telescoped” forearc crust is difficult to distinguish from event, Late Cretaceous forearc strata and underlying Jurassic and
alternative tectonic processes such as margin-parallel strike-slip earliest Cretaceous basement rocks were ripped from the western
faulting (e.g., Karig, 1980). A classic illustration of the difficulty margin of the Peninsular Ranges batholith, rotated clockwise,
in deciding between subduction erosion and margin-parallel and translated northwestward into the outer California border-
strike-slip dispersal of forearc rocks is provided by the Nacimiento land along a detachment system underlain by the Catalina Schist
fault in west-central California (Page, 1970). There, plutonic and (Wright, 1991; Crouch and Suppe, 1993; Bohannon and Geist,
metamorphic rocks of the Salinian block are juxtaposed against 1998; Ingersoll and Rumelhart, 1999).
a high-pressure, low-temperature mélange of the Sur-Obispo Although the Catalina Schist contains abundant blue-
terrane. This anomalous juxtaposition of lithotectonic belts has schist facies rocks typical of subduction zone environments, it
been explained both by underthrusting (subduction erosion) of also has higher temperature lithologies that define an inverted
Figure 1. Geologic setting of the Catalina Schist. Location map in the upper right shows the distribution of high-
pressure/temperature (PT) subduction complexes and Cretaceous–Tertiary batholiths in southwestern North America.
The offshore distribution of subduction complexes is estimated primarily from Crouch and Suppe (1993), Bohannon
and Geist (1998), Sedlock (1988a, b), Bonini and Baldwin (1998), and Fletcher et al. (2007). The box outlines the loca-
tion of the southern California continental borderland. The simplified map of the borderland shown in the lower left is
based upon Bohannon and Geist (1998), whereas cross section Y–Y′ is after Crouch and Suppe (1993). Symbols P and
P* denote formerly contiguous rocks that dextrally sheared and rotated clockwise during middle Miocene rifting. The
box in the lower left shows the location of Catalina Island. The geologic map and cross section X–X′ of Catalina Island
in the upper left are after Platt (1976). The distribution of lawsonite-albite and actinolite-albite rocks is from Altheim et
al. (1997). Note that epidote-blueschist and epidote-amphibolite rocks are distinguished from Platt’s (1976) greenschist
unit. A coherent kilometer-scale mass of epidote-amphibolite facies metagabbro that directly underlies the amphibolite
unit was included by Platt (1976) within his amphibolite unit. K–T—Cretaceous–Tertiary.
338 Grove et al.

metamorphic sequence that includes amphibolite facies rocks relationship to forearc and batholith rocks of the adjacent Penin-
at the highest structural levels (see inset in Fig. 1). These partly sular Ranges batholith. These new results indicate that the anom-
melted rocks underwent peak grade conditions at ca. 115 Ma alously high-T, high-PT amphibolite facies rocks of the Catalina
(Suppe and Armstrong, 1972; Mattinson, 1986; Sorensen and Schist had been deposited, accreted, and metamorphosed at peak
Barton, 1987; Grove and Bebout, 1995; Anczkiewicz et al., 2004). grade conditions ~15–20 m.y. before the lawsonite-blueschist and
A longstanding explanation for the unusually high-temperature, lower-temperature, high-PT rocks of the complex were accreted.
inverted metamorphism of the Catalina Schist is that it reflects We conclude that amphibolite facies, epidote-amphibolite, and
progressive accretion of early subducted rocks beneath unrefri- possibly epidote blueschist units of the Catalina Schist formed
gerated mantle lithosphere during nascent subduction (e.g., Platt, from forearc strata and basement rocks that were underthrust
1975). Although Platt’s (1975) nascent subduction hypothesis and sheared together with lithospheric mantle beneath the west-
was based upon the Catalina Schist, he extended it to explain the ern margin of the Peninsular Ranges batholith in a subduc-
origin of meter- to kilometer-scale high-grade blocks within the tion erosion process. Our model is consistent with recent deep
Franciscan Complex and other accretionary complexes along the seismic-reflection imaging of subparallel megathrusts within
western margin of North America (e.g., Platt, 1975; Cloos, 1985; modern subduction zones (e.g., Calvert, 2004). Continued sub-
Wakabayashi, 1990, 1992, 1999; Anczkiewicz et al., 2004; Wak- duction erosion ultimately juxtaposed the high-grade rocks of the
abayashi and Dumitru, 2007). Catalina Schist with lawsonite-blueschists and lower grade rocks
The nascent subduction hypothesis makes sense for the Fran- that had formed within a subduction zone setting.
ciscan Complex because the blocks are coeval with a major pulse of
Middle Jurassic magmatic activity along the western North Amer- BACKGROUND
ican continental margin (Evernden and Kistler, 1970; Saleeby and
Sharp, 1980; Stern et al., 1981; Chen and Moore, 1982; Wright Catalina Schist High-Pressure/Temperature Complex
and Fahan, 1988; Barton et al., 1988; Staude and Barton, 2001;
Irwin, 2002). The Middle Jurassic orogenesis included formation High-PT rocks distributed along the western margin of
of the Coast Range ophiolite (Hopson et al., 1981; Shervais et southern and Baja California appear to represent the southern
al., 2005), which is considered by some (e.g., Stern and Bloomer, continuation of the better exposed Franciscan Complex of cen-
1992) to be a product of the initiation of Middle Jurassic subduc- tral and northern California (Fig. 1; Woodford, 1924; Suppe and
tion (see also Dickinson et al., 1996). The equivalence in age of Armstrong, 1972; Kilmer, 1979; Blake et al., 1984). Within the
the Coast Range ophiolite and the high-grade blocks of the Fran- southern California region, on-land exposures of a high-PT com-
ciscan Complex corresponds well with a common relationship plex referred to as the Catalina Schist occur on Santa Catalina
exhibited by many of the world’s ophiolites: namely, age equiva- Island (Bailey, 1941; Platt, 1976) and the Palos Verdes Penin-
lence of the ophiolite and its metamorphic sole (Jamieson, 1986; sula (Woodring et al., 1946; Dibblee, 1999). More widespread
Hacker, 1990a, b; Wakabayashi and Dilek, 2000). submarine exposures of the Catalina Schist occur throughout the
Whereas establishment of a new subduction regime dur- inner southern California borderland, and equivalent basement is
ing the Middle Jurassic may generally explain the occurrence of detected in boreholes within the western and southwestern Los
high-grade rocks of this age within the Franciscan Complex of Angeles Basin (Schoellhamer and Woodford, 1951; Yerkes et al.,
central and northern California, an equivalent event at ca. 115 Ma 1965; Yeats, 1968, 1973; Sorensen, 1985, 1988a, 1988b; Wright,
off the southern California margin makes little sense with respect 1991; Crouch and Suppe, 1993; Bohannon and Geist, 1998; ten
to the convergent margin evolution of southwest North America. Brink et al., 2000).
Emplacement of the 750-km-long Peninsular Ranges batho- The Catalina Schist on Santa Catalina Island has been
lith had begun by Middle Jurassic time (Shaw et al., 2003) and described by Woodford (1924), Bailey (1941), Platt (1975, 1976),
was well under way prior to the proposed subduction-initiation Sorensen and Barton (1987), Sorensen (1986, 1988a, b), Sorensen
event at 115 Ma (Silver and Chappell, 1988; Kistler et al., 2003). and Grossman (1989), Bebout and Barton (1989, 1993, 2002),
Furthermore, no middle Cretaceous ophiolite exists within the and Grove and Bebout (1995) and is only briefly summarized
southern California region. Only fragments of older forearc base- here. It consists of metasedimentary, metavolcanic, and ultramafic
ment are preserved. The latter is best represented by Triassic and protoliths that were metamorphosed and sheared together under
Middle Jurassic ophiolitic rocks that border the southern part of amphibolite facies to lawsonite-blueschist facies and lower grade
the Peninsular Ranges batholith (Kimbrough and Moore, 2003). conditions during the middle Cretaceous (Suppe and Armstrong,
Just as in the case of the Franciscan Complex, the Middle Juras- 1972; Platt, 1976; Mattinson, 1986; Sorensen and Barton, 1987;
sic ophiolite of west-central Baja California is associated with Sorensen, 1988a, 1988b; Grove and Bebout, 1995). Individual
high-PT blocks of equivalent age on the Vizcaino Peninsula and tectonic slices (Fig. 1) contain rocks of broadly equivalent meta-
Cedros Island (Fig. 2; Baldwin and Harrison, 1989, 1992). morphic grade. These are juxtaposed across low-angle faults in an
This paper presents new detrital zircon U-Pb results from apparently inverted metamorphic sequence (Platt, 1976; Fig. 1).
metagraywackes of the major metamorphic units of the Cata- The structurally highest unit is an amphibolite facies shear zone
lina Schist that clarify its accretion history. We confirm a genetic composed primarily of intercalated and metasomatically altered
Figure 2. Geologic map of the Peninsular Ranges batholith, modified after Gastil et al. (1975) and Kimbrough et al.
(2001). Generalized stratigraphic relationships of the forearc region in the northern Vizcaino Peninsula modified after
Kimbrough et al. (2001), Kimbrough and Moore (2003), and references cited within these papers. Simplified stratigraphy
of the northern Santa Ana Mountains is based upon references provided within Lovera et al. (1999). Western limit of the
forearc is the restored pre–middle Miocene position based upon Bohannon and Parsons’s (1995) reconstruction and our
interpretation of aeromagnetic data presented by Langenheim and Jachens (2003). PRB—Peninsular Ranges batholith;
TTG—tonalite, trondjhemite, granodiorite.
340 Grove et al.

mafic and former harzburgite/dunite protoliths (Sorensen and Forearc Basement and Strata
Barton, 1987; Sorensen, 1988b; Bebout and Barton, 1993, 2002).
The proportion of sediment, particularly immature graywacke, The relatively complete forearc sequence of the Vizcaino
increases structurally downward. At the structurally lowest levels Peninsula at the southern end of the Peninsular Ranges batholith
the lawsonite-blueschist and lower grade units are sediment dom- provides the basis to interpret the forearc of the highly disrupted
inated. Each of the major metamorphic units is characterized by southern California borderland (Figs. 1, 2). On the Vizcaino Pen-
meter- to kilometer-scale, compositionally heterogeneous shear insula and on Cedros Island, both Late Triassic (221 Ma) and Mid-
zones that appear to have facilitated metasomatic fluid infiltration dle Jurassic (173 Ma) ophiolite sequences (Moore, 1985; Kim-
at near peak-grade conditions (Bebout and Barton, 1989, 1993, brough, 1985; Kimbrough and Moore, 2003) are the basement for
2002; King et al., 2006, 2007). Upper Jurassic, Lower Cretaceous, Upper Cretaceous, and lower
Primary geologic mapping of the Catalina Schist on Santa Cenozoic arc volcanic and forearc sedimentary rocks (Boles,
Catalina Island was performed by Bailey (1941) and Platt (1976). 1978; Kilmer, 1979; Kienast and Rangin, 1982; Boles and Lan-
Although we rely heavily upon Platt’s (1976) mapping of the dis, 1984; Patterson, 1984; Smith and Busby, 1993; Busby et al.,
central part of the island, we recognize different tectonometa- 1998; Kimbrough et al., 2001; Critelli et al., 2002; Busby, 2004).
morphic units than he did. Platt’s (1976) Catalina greenschist High-PT Cretaceous rocks of the Western Baja terrane constitute
unit has been subdivided into epidote-amphibolite and epidote- an accretionary complex that is juxtaposed beneath the Triassic
blueschist units (Grove and Bebout, 1995). Results presented and Jurassic ophiolitic rocks well outboard of the batholithic mar-
in Grove and Bebout (1995) and in this paper indicate that the gin (Suppe and Armstrong, 1972; Moore, 1986; Sedlock, 1988a,
epidote-amphibolite and epidote-blueschist rocks had impor- b; Baldwin and Harrison, 1989; Baldwin and Harrison, 1992).
tant differences in accretion and metamorphic history. In addi- Only fragmentary evidence exists for the early Mesozoic
tion, we have delineated lawsonite-albite and actinolite-albite forearc basement within the southern California area. Within
facies units on the west end of the island. These lower grade the outer California continental borderland, offshore drilling
rocks appear to structurally underlie a lawsonite-blueschist unit and seismic exploration detect mafic basement beneath forearc
(Altheim et al., 1997). strata (Bohannon and Geist, 1998; ten Brink et al., 2000). Sur-
Because the Catalina Schist terrane was significantly face exposures of lower Mesozoic basement rocks that occur on
extended as it was exhumed during the early middle Miocene Santa Cruz Island (the Willows Complex of Weaver and Nolf,
formation of the inner continental borderland (Wright, 1991; 1969; Hill, 1976; Mattinson and Hill, 1976) are correlated with
Crouch and Suppe, 1993; Nicholson et al., 1994; Bohannon and the Coast Range ophiolite (Jones et al., 1976). The associated
Geist, 1998; Ingersoll and Rumelhart, 1999), original contacts Santa Cruz Island Schist and the correlative Santa Monica slate
bounding the major units within the schist are likely to have been of the western Transverse Ranges exhibit distinctly arclike com-
strongly modified or excised altogether by middle Tertiary struc- positions and could be equivalent to accreted Jurassic terranes of
tures. In addition, the extent to which the Santa Catalina Island the Sierran Foothills (Sorensen, 1985, 1988a). “Saussurite” gab-
exposures of Catalina Schist represent the much larger area of bro similar to altered zones of the Willows Complex gabbros is
high-PT rocks that underlie the inner continental borderland is in fault contact with the Catalina Schist on Santa Catalina Island
difficult to assess. Study of widely distributed deposits of the San (Platt, 1976) and within the subsurface of the southwestern Los
Onofre Breccia indicates that the Santa Catalina Island exposures Angeles Basin (Schoellhamer and Woodford, 1951; Yeats, 1973;
are probably broadly representative of the overall terrane (Stu- Sorensen, 1985, 1988b). Saussurite gabbro clasts are a major
art, 1979). The San Onofre Breccia is an early middle Miocene component within the San Onofre Breccia (Stuart, 1979) and
deposit that accumulated as the Catalina Schist was exhumed confirm that low-PT mafic basement was exhumed along with
(Wright, 1991; Crouch and Suppe, 1993). Clast counts performed the Catalina Schist during middle Miocene borderland rifting.
by Stuart (1979) appear to be broadly consistent with an unroof- Cenomanian and younger forearc strata onlap the west-
ing sequence in which amphibolite and epidote-amphibolite ern margin of the northern Peninsular Ranges batholith (Fig. 2;
grade Catalina Schist and low-PT mafic basement from the hang- Woodring and Popenoe, 1942; Yerkes et al., 1965; Flynn, 1970;
ing wall were much more abundant at lower stratigraphic levels Nordstrom, 1970; Peterson and Nordstrom, 1970; Kennedy and
within the San Onofre Breccia. Our own observations confirm Moore, 1971; Sundberg and Cooper, 1978; Schoellhamer et al.,
this relationship for the thick deposits of this breccia that accumu- 1981; Nilsen and Abbott, 1981; Bottjer et al., 1982; Bottjer and
lated along the western margin of the Peninsular Ranges batho- Link, 1984; Fry et al., 1985; Girty, 1987; Bannon et al., 1989).
lith (Fig. 1). In contrast, the stratigraphically younger San Onofre Thick sections of Upper Cretaceous strata also occur throughout
Breccia, including deposits laid down atop extended borderland the outer borderland (Howell and Vedder, 1981; Vedder, 1987;
crust underlain by the Catalina Schist, tends to be dominated by Bohannon and Geist, 1998). The existence of Lower Cretaceous
lawsonite-blueschist and lower-grade detritus. Based upon these forearc strata is far less certain, however. Sedimentary rocks of this
relationships, we are reasonably confident that our sampling of age have not been described anywhere along the western margin
the Catalina Schist on Santa Catalina Island is broadly represen- of the northwestern Peninsular Ranges, the Santa Monica Moun-
tative of the Catalina Schist terrane as a whole. tains, or on the offshore islands. Their existence in the subsurface
The Catalina Schist 341

is mostly inferred in all areas that have been drilled (Howell and of the Catalina Schist collected from Santa Catalina Island. Many
Vedder, 1981; Vedder, 1987; Bohannon and Geist, 1998; ten Brink of the samples examined were previously studied by Grove and
et al., 2000). Our model for the formation of the Catalina Schist Bebout (1995). We focused upon metasedimentary rocks from
explicitly accounts for the scarcity of these Lower Cretaceous each of the major metamorphic units in order to obtain an upper
forearc rocks in the southern California region (see Discussion). bound upon the depositional age of their sedimentary protolith
and to determine sediment provenance. In our U-Pb analysis of
Peninsular Ranges Batholith zircons from the Catalina Schist, we employed secondary ion-
ization mass spectrometry (SIMS) methods using the UCLA
The Peninsular Ranges batholith of southern and Baja Cali- Cameca ims 1270 ion microprobe. The extensive metamorphic
fornia constitutes a classic Cordilleran continental margin batho- recrystallization that affected zircons from the amphibolite and
lith (Fig. 2; Larsen, 1948; Jahns, 1954; Gastil et al., 1975). The epidote-amphibolite units was significantly ameliorated by the
better studied northern segment of this batholith consists of lon- high spatial resolution of the ion microprobe in conjunction with
gitudinal western and eastern zones based on age, petrology, pre- cathodoluminescence (CL) imagery (e.g., only ~1 nanogram of
batholithic wall rock, geophysical parameters, and depth and style sputtered zircon required to yield a U-Pb age). The techniques are
of emplacement (Gastil et al., 1981; Baird and Miesch, 1984; Tay- described in Grove et al. (2003b) with additional details included
lor, 1986; Gromet and Silver, 1987; Silver and Chappell, 1988; in the GSA Data Repository.1
Ague and Brimhall, 1988; Todd et al., 1988; Hill and Silver, 1988; We analyzed conventionally sectioned and polished zircons
Gastil, 1993; Johnson et al., 1999; Lovera et al., 1999; Todd et al., in epoxy mounts. The zircons were hand selected from heavy
2003; Kistler et al., 2003; Langenheim and Jachens, 2003). The mineral concentrates produced from standard crushing, density,
oldest recognized intrusive rocks are gneissic S-type plutons that and magnetic methods. Further details are included in the Data
occupy the medial zone of the Peninsular Ranges batholith and Repository. Because the number of zircons measured from indi-
yield Middle Jurassic emplacement ages (Todd and Shaw, 1985; vidual samples was typically small owing to the low yields real-
Thomson and Girty, 1994; Shaw et al., 2003; Kistler et al., 2003). ized during mineral separation, we pooled data from multiple
Cretaceous plutons as old as 140 Ma also occur within this batho- samples to obtain statistically meaningful results for each of the
lith (Silver and Chappell, 1988; Alsleben et al., 2005; D.L. Kim- major tectonic units of the Catalina Schist.
brough, personal observ.). These are cut by a regionally extensive As noted above, metamorphic zircon growth was a signifi-
dike swarm emplaced at ca. 130–120 Ma (Böhnel et al., 2002; cant issue for some of the grains we examined, particularly for
D.L. Kimbrough, personal observ.). The well-developed west- the metasediments from the amphibolite unit. Anczkiewicz et al.
ern zone of the Peninsular Ranges batholith is composed mainly (2004) also described evidence for metamorphic zircon growth
of 125–100 Ma gabbro to monzogranite plutons with primitive in reconnaissance U-Pb zircon measurements they reported from
island arc geochemical affinities. The eastern zone of the batho- the Catalina Schist amphibolite unit. Although we selected our
lith is defined by a belt of large-volume, 95 ± 3 Ma tonalite, analysis sites on the basis of morphologic and optical criteria
trondjhemite, and low-K granodiorite (TTG) plutons (Gastil et in CL imagery, a significant number of the spot analyses from
al., 1975; Silver and Chappell, 1988; Walawender et al., 1990) zircons of the amphibolite and epidote-amphibolite units over-
that constitute the La Posta TTG suite (Tulloch and Kimbrough, lapped regions affected by metamorphic recrystallization. Meta-
2003; Kimbrough and Grove, 2006). These Na- and Al-rich plu- morphic overgrowths on igneous zircons generally have Th/U
tons have a deep, garnet-present, melt-source signature, as seen <0.1; Kröner et al., 1994; Rubatto, 2002; Williams and Claesson,
in high Sr, Ba, Sr/Y, and La/Yb. Field and thermobarometric data 1987). We have found a cutoff of Th/U = 0.1 to be empirically
indicate emplacement at ~2–6 kbar pressures (Rothstein and well supported by detrital zircon results from the Cenomanian and
Manning, 2003) followed by rapid Cenomanian–Turonian uplift younger, pluton-derived strata that overlie the Peninsular Ranges
and denudation at rates of ~1–2 mm/yr (Krummenacher et al., batholith. In our sampling of these units (Mahoney et al., 2005),
1975; Lovera et al., 1999; Johnson et al., 1999; Kimbrough et al., 1506 out of 1527, or 98.6% of the zircon U-Pb analyses, yielded
2001; Ortega Rivera, 2003; Grove et al., 2003b). A delayed late Th/U values equal to or greater than 0.1 (Fig. 3A). The average
Campanian–Maastrichtian phase of uplift (e.g., Krummenacher Th/U value measured was 0.48 ± 0.41, with 3.3% of the analyses
et al., 1975) appears to be related to Laramide shallow subduction falling within the Th/U = 0.1–0.2 range. Thus a cutoff of Th/U
and removal of the deep crustal and lithospheric roots of the La = 0.1 is useful for identifying detrital igneous zircons adversely
Posta belt (Lovera et al., 1999; Grove et al., 2003a, b). affected by metamorphic recrystallization. Accordingly, we have
excluded all analyses with Th/U <0.1 from the summary plots or
SAMPLING AND METHODS calculations presented below (Fig. 3B).
Complete data tables of U-Pb age measurements, litho-
Detrital Zircon U-Pb Age Measurements logic descriptions, and sample locations are available from the
1
GSA Data Repository item 2008105 is available online at www.geosociety.
A total of 645 U-Pb zircon ages were measured from 33 sam- org/pubs/ft2008.htm, or on request from editing@geosociety.org or Documents
ples of amphibolite through sub-blueschist facies metagraywacke Secretary, GSA, P.O. Box 9140, Boulder, CO 80301, USA.
342 Grove et al.

GSA Data Repository. In this paper, quoted uncertainties are


±1σ errors unless otherwise specified. The U-Pb ages are gener-
ally 206Pb/238U values for <1 Ga zircons and 207Pb/206Pb ages for
older grains. Analyses with age uncertainties >10% have been
discarded. The U-Pb ages of zircons that exhibited resolvable
206
Pb/238U versus 207Pb/235U discordance are 207Pb/206Pb values.

Rb-Sr Measurements

To clarify the significance of previous 40Ar/39Ar phengite


step-heating results (Grove and Bebout, 1995) that had indicated a
possible Middle Jurassic age for garnet-bearing blueschist blocks
within the lawsonite-blueschist mélange, we carried out Rb-Sr
measurements with one of the blocks (330–4B) that had been
analyzed in this previous study. All column chemistry and mass
spectrometer measurements were carried out at UCLA. There,
Rb and Sr were separated in cation exchange columns containing
AG50W-X8 resin, using 2.5N HCl. The Sr isotopic compositions
were measured with a VG54–30 multicollector thermal ionization
mass spectrometer. 87Sr/86Sr ratios were normalized to 86Sr/88Sr
= 0.1194. In this analysis session, the measured 87Sr/86Sr value
for our Sr standard (NBS 987) value was 87Sr/86Sr = 0.710239
± 16 (2σ, n = 13). The Rb and Sr concentrations were measured
by isotope dilution using VG54–30. Model Rb-Sr isochron ages
were calculated using ISOPLOT/Ex Version 3 (Ludwig, 2003).

RESULTS

Amphibolite Facies

Five biotite + muscovite + garnet ± kyanite-bearing meta-


graywackes from the amphibolite unit yielded abundant subhe-
dral to well-rounded zircon. Alternatively, both metachert samples
examined contained only metamorphic zircon. Although exter-
nal overgrowths <1–5 µm or more were common, zircons also
exhibited internal areas of patchy diffuse recrystallization in CL
imagery. Of the 169 analyses, 48 (28%) overlapped metamorphic
zircon based upon Th/U (Fig. 3B). About half of the affected anal-
yses yielded obviously mixed ages. U-Pb ages calculated for the
remaining analyses with Th/U <0.1 fell between 107 and 126 Ma,
with a peak at 116 ± 6 Ma. This result is in good agreement with
independent estimates for the timing of amphibolite unit recrys-
Figure 3. (A) Measured zircon U-Pb age versus Th/U of Cretaceous tallization (ca. 115 Ma; Suppe and Armstrong, 1972; Mattinson,
forearc sedimentary rocks of the Peninsular Ranges batholith. Data, 1986; Grove and Bebout, 1995; Anczkiewicz et al., 2004).
from Mahoney et al. (2005), demonstrate that virtually all detrital igne- The data for zircons with Th/U >0.1 clearly indicate that the
ous zircons derived from the Peninsular Ranges have Th/U >0.1 (see
also Williams and Claesson, 1987). (B) Equivalent plot for zircon re- sedimentary protolith of the amphibolite facies metagraywackes
sults from the Catalina Schist amphibolite unit. Based upon cathodo- contained a large proportion of craton-derived detritus (Fig. 4A).
luminescence imaging and other criteria, we regard a Th/U = 0.1 as a Roughly 50% of the analyses were from Middle Proterozoic
meaningful cutoff to distinguish analyses in which the sputter pit over- zircon (Fig. 4G). Unfortunately, many of these grains exhibited
lapped metamorphic zircon growth (open symbols). The filled sym- variable Pb loss and high degrees of discordance. This condition
bols represent detrital grains that we consider to be largely unaffected
by metamorphic zircon growth (see also Kröner et al., 1994; Rubatto, complicates characterization of the Middle Proterozoic age dis-
2002). Zircon results with Th/U <0.1 have been excluded from all sub- tribution. Although 207Pb/206Pb ages more accurately approximate
sequent plots and calculations. the crystallization age than the highly discordant U-Pb ages, they
Figure 4. Relative probability plots of detrital zircon U-Pb age distributions from major units of the Catalina Schist
on Catalina Island. (A) Amphibolite. (B) Epidote-amphibolite. (C) Epidote-blueschist. (D) Lawsonite-blueschist.
(E) Actinolite-albite. (F) Lawsonite-albite. Note that we use a split horizontal axis at 300 Ma and that relative probability
plots between 300 and 3000 Ma have a 2× scaling factor to improve resolution of the overall age distribution. (G) True
scale cumulative probability spectra for all units.
344 Grove et al.

are insufficiently precise to reveal strong clustering of ages. Nev- Lawsonite-Blueschist Facies and Lower Grade Rocks
ertheless, the subdued maxima that occur at 1.15, 1.40, and 1.65–
1.80 Ga are compatible with a southwestern North American Data bearing upon the sedimentary protoliths of the
provenance. Age peaks also occur at 126, 145, and 162 Ma. The lawsonite-blueschist and lawsonite-albite, and albite-actinolite
five youngest zircons measured from the amphibolite unit meta- facies units are discussed together because of the highly similar
graywackes yielded a weighted mean age of 122 ± 3 Ma. These results produced (Fig. 4D–G). Overall, we obtained 164 analyses
analyses yielded a mean Th/U value of 0.30 ± 0.20 and were from from 11 lawsonite-blueschist facies rocks, 59 analyses from 3
oscillatory zoned regions in Cl− imagery that likely reflect igne- albite-actinolite facies rocks, and 82 analyses from 7 lawsonite-
ous crystallization. Accordingly, we regard 122 ± 3 Ma as a geo- albite facies metagraywackes. No metamorphic zircon growth
logically meaningful upper bound upon the depositional age of was detected; Th/U values were all >0.1. The Ca-rich metagray-
the sedimentary protolith. wackes prevalent in low-grade parts of the Catalina Schist over-
whelmingly contain clear subhedral to euhedral grains. Most of
Epidote-Amphibolite Facies these grains yield U-Pb ages of 95–130 Ma (Fig. 4D–F). Late
Jurassic–Early Cretaceous zircon was greatly subordinate, and
Five muscovite ± biotite ± garnet epidote-amphibolite facies only trace quantities of Proterozoic zircon were present. Because
metagraywackes yielded zircon that was generally more subhe- the youngest zircon U-Pb ages measured from the three lowest
dral to euhedral grains than the well-rounded grains common grade units all agreed within error (96 ± 3 Ma, 97 ± 3 Ma, and 98
in the amphibolite unit metagraywackes. CL imaging and Th/U ± 3 Ma for lawsonite-blueschist, actinolite-albite, and lawsonite-
indicate that metamorphic recrystallization of detrital grains was albite, respectively), they define a 97 ± 3 Ma upper bound for the
less common than in the amphibolite unit. Only 13 of 135 or depositional age of the sedimentary protolith for all of the low-
10% of analyses of the zircons from epidote-amphibolite facies grade Catalina Schist units.
metagraywackes had Th/U <0.1. The U-Pb age distribution indi-
cates a diminished cratonal provenance relative to amphibolite Garnet-Bearing Blueschist Block in Lawsonite-Blueschist
facies metagraywackes; only ~20% of the zircons yielded Middle Mélange
Proterozoic ages (Fig. 4G). Although fewer analyses are avail-
able, the overall age distribution of Middle Proterozoic zircons Garnet-bearing blueschist blocks occur within mélange
in epidote-amphibolite metagraywackes is broadly similar to that zones in the lawsonite-blueschist unit. Phengite 40Ar/39Ar results
of the amphibolite facies rocks (Fig. 4B). The relative lack of for two such blocks indicate that they formed prior to peak-grade
Proterozoic grains is accompanied by a proportionate increase of recrystallization in the amphibolite unit (Grove and Bebout,
<200 Ma grains. Distinct peaks are present at 115 and 126 Ma, 1995). To confirm this relationship we undertook Rb-Sr mea-
with a strong peak at 150 Ma (Fig. 4B). The five youngest detrital surements of phengite, Na-amphibole, and whole rock for one of
zircons from the epidote-amphibolite metagraywackes yielded the samples (330–4B; Fig. 5A). Results are shown in Figure 5B.
an age of 113 ± 3 Ma, with a mean Th/U value of 0.42 ± 0.05. Phengite, Na-amphibole, and whole-rock results do not define
a statistically meaningful isochron. However, the model Rb-Sr
Epidote-Blueschist Facies age defined by phengite + Na-amphibole, 135 Ma, is identical
to the total gas age calculated from 40Ar/39Ar step-heating results
Thirty analyzable zircons were recovered from a single obtained from 330 to 4B phengite.
metagraywacke sample intercalated with Na-amphibole + clino-
zoisite + albite-bearing mafic rocks (glaucophanic greenschists DISCUSSION
of Platt, 1975; Sorensen, 1986). Six similar-appearing samples
from equivalent field settings failed to yield zircon. This poor Evaluation of the Nascent Subduction Model for the
zircon yield probably reflects abundant volcanic detritus in the Catalina Schist
protolith. Nearly all of the zircons recovered consisted of clear
subhedral to euhedral grains. There was no evidence for meta- The origin of the amphibolite facies rocks of the Catalina
morphic zircon growth either petrographically or in terms of Schist is central to assessing the tectonic significance of this ter-
measured Th/U. Two distinct maxima occur at 103 and 142 Ma rane. Amphibolite facies metamorphism and anatexis are atypi-
(Fig. 4C). Although results from the epidote-blueschist facies cal features of subduction complexes and are thought to reflect
unit are transitional between those obtained from the amphibo- unusual processes along a convergent margin. High-temperature
lite and epidote-amphibolite facies units and lower grade parts of conditions accompany the earliest developmental stages of sub-
the Catalina Schist (Fig. 4G), the epidote-blueschist facies unit duction before the hanging wall is refrigerated (Platt, 1975;
appears more closely allied with the lower grade units, based Cloos, 1985; Peacock, 1987). Although ridge subduction and/or
upon the age of the youngest zircons detected in each population. slow underflow of very young oceanic crust (e.g., Peacock, 1987,
The five youngest zircons measured yielded an average U-Pb age 1992; Hacker, 1990a, 1990b; Peacock et al., 1994) can also pro-
of 101 ± 3 Ma with a mean Th/U of 0.63 ± 0.23. duce high-temperature metamorphism in subduction zones, there
The Catalina Schist 345

Figure 5. (A) Garnet-bearing blueschist


block in mélange matrix of the lawson-
ite blueschist unit. (B) Rb-Sr results for
phengite, Na-amphibole, and whole
rock. A statistically meaningful isochron
is not defined. Phengite and Na-amphi-
bole define a model isochron (135 Ma)
that is identical to the total gas 40Ar/39Ar
age yielded by this sample.

is no evidence that either process affected the middle Cretaceous Catalina Schist over a ~20 m.y. interval is required. As indicated in
southwestern margin of North America. Figure 6A, the protolith of the amphibolite unit metagraywackes
The detrital zircon results (Fig. 6) require substantial modi- was deposited between 122 Ma (youngest detrital zircon ages)
fication of Platt’s (1975) hypothesis that amphibolite through and 115 Ma (the time of peak-grade recrystallization). In contrast,
blueschist facies rocks formed in an inverted metamorphic aure- the sedimentary protolith of the lawsonite-blueschist and lower
ole in response to transient heating during nascent subduction at grade units was deposited after 97 Ma or at least 18 m.y. after
ca. 115 Ma. Sequential accretion of the major tectonic units of the peak-grade recrystallization of the amphibolite unit (Fig. 6C, D).
346 Grove et al.

Figure 6. Estimated temperature-time histories for major units of the Catalina Schist. (A) Amphibolite. (B) Epidote-
amphibolite. (C) Epidote-blueschist. (D) Lawsonite-blueschist and lower grade units of the Catalina Schist. Yellow circles
represent measured detrital zircon U-Pb ages. Maximum bounds upon the depositional age of the sedimentary protolith
represent average and standard deviation of U-Pb ages of the five youngest zircons with Th/U >0.1 (15 youngest zircons
for composite of the three lowest grade units). Data sources for other thermochronometry include Suppe and Armstrong
(1972), Mattinson (1986), Grove and Bebout (1995), and Anczkiewicz et al. (2004). Note that we have varied the bulk
closure temperatures assigned to micas from 400 °C for coarse-grained muscovites within the high-grade units to 350 °C
for finer grained, less retentive white mica within lawsonite-albite rocks.

Available data for the Catalina Schist amphibolite unit Within the Catalina Schist, only the lawsonite-blueschist and
indicate that it cooled slowly (ca. 25 °C/m.y.) from peak-grade lower grade rocks record metamorphic conditions that require
conditions (650–700 °C) at 115 Ma to K-Ar muscovite closure formation within a low-T, high-PT subduction zone. The amphib-
temperatures (350–400 °C) at 105–100 Ma (Fig. 6A; Grove olite and epidote-amphibolite units were retrograded under
and Bebout, 1995). The Catalina epidote-amphibolite unit also greenschist facies conditions, as was the epidote-blueschist unit
cooled slowly (Fig. 6B): a 7–12 m.y. interval separated peak- (Platt, 1976; Sorensen, 1986; Bebout and Barton, 1989; Grove
grade recrystallization of epidote-amphibolite mafic gneisses and and Bebout, 1995). Given their present association with lawson-
cooling through phengite K-Ar closure. Slow cooling exhibited ite-blueschist and lower grade rocks, we find it remarkable that
by the amphibolite unit is compatible with very slow subduction the amphibolite and epidote-amphibolite units exhibit such little
or subduction of very young oceanic lithosphere. However, envi- evidence for equilibration at blueschist facies conditions (Platt,
ronments such as these are far too warm to permit formation of 1976; Sorensen, 1986; Bebout and Barton, 1989; Grove and
lawsonite-blueschist and lower temperature metamorphic rocks. Bebout, 1995). Blueschist facies overprinting of higher grade
The Catalina Schist 347

rocks is the rule for subduction complexes of the North American


Cordillera (Ernst, 1988). For example, tectonic blocks of garnet
amphibolite in the Franciscan and Shuksan subduction com-
plexes are heavily overprinted by blueschist facies assemblages
(e.g., Brown et al., 1982; Cloos, 1985; Wakabayashi, 1990). In
contrast, evidence for blueschist facies overprinting within the
Catalina Schist amphibolite unit is obscure and limited to sparse
veining by pumpellyite and trace amounts of lawsonite in white
mica + zoisite + albite assemblages that replace oligoclase in
amphibolite facies rocks (Grove and Bebout, 1995). We interpret
the available thermochronology and greenschist facies overprint-
ing of the high-grade units of the Catalina Schist to indicate that
they occupied a subcrustal position characterized by 400–500 °C
temperatures prior to 100–95 Ma. The lack of blueschist facies
overprinting within the amphibolite and epidote-amphibolite
units indicates to us that these rocks were sufficiently distant
from an active subduction zone to be largely unaffected by the
low geothermal gradient (<5–10 °C/km) recorded by the Fran-
ciscan and other Cordilleran subduction complexes during the
middle Cretaceous.

Provenance Ties Linking the Catalina Schist to the


Peninsular Ranges Batholith

King et al. (2007) measured whole-rock Pb isotopes from


mélange matrix sampled from the amphibolite, lawsonite-
blueschist, and lawsonite-albite units of the Catalina Schist.
Although details of the U-Th-Pb systematics indicate the pos-
sible elemental redistributions during the devolatization history
of the schist (see King et al., 2007), it is remarkable how well the
measured Pb isotopic compositions of the sediment-dominated
lawsonite-blueschist and lawsonite-albite mélange matrix agree
with those measured from granitoids of the adjacent Peninsular
Ranges batholith (Fig. 7A, B; Kistler et al., 2003). The whole-
rock Pb isotopic compositions from the mafic- and ultramafic-
dominated mélange of the Catalina Schist amphibolite unit also
overlap strongly with the Peninsular Ranges batholith (Fig. 7C)
but tend to more radiogenic values than the lower grade units
(Fig. 7A, B). The amphibolite unit mélange results are more dif-
ficult to interpret than those from the lower grade units, because
the sources of Pb are far less uncertain owing to the scarcity of
sedimentary material within the mélange (see Bebout and Bar-
ton, 1993, 2002). Nevertheless, these results are consistent with
a cratonal provenance similar to that exhibited by southeastern

Figure 7. Whole-rock Pb isotopic data of King et al. (2007) for:


(A) Lawsonite-albite. (B) Lawsonite-blueschist. (C) Amphibolite
unit mélange matrix of the Catalina Schist. Also shown are whole-
rock Pb isotopic data for northern Peninsular Ranges batholith plutons
(red circles; Kistler et al., 2003) and equivalent data for southeastern
Arizona (orange circles; Wooden and Miller, 1990). The yellow field
represents the generalized distribution of whole-rock Pb yielded by the
Mojave–Transverse Ranges province (based upon Barth et al., 1995,
and personal observ. from J. Wooden, D. Coleman, and A. Barth).
348 Grove et al.

Arizona (Wooden and Miller, 1990) and can be differentiated readily understood in terms of the Cenomanian age of the proto-
from the Mojave–Transverse Ranges province (Fig. 7C; Barth et lith. Cenomanian-age forearc strata from the Peninsular Ranges
al., 1995; J. Wooden, D. Coleman, and A. Barth, personal observ.). batholith are also skewed to older ages because they were depos-
Because the early Mesozoic wall rocks of the Peninsular Ranges ited prior to massive exhumation of the ca. 95 Ma La Posta plu-
batholith are dominated by craton-derived Proterozoic zircon tonic suite within the eastern batholith (Mahoney et al., 2005).
that indicates a similar source (see below), we speculate that the Most of the results from the Upper Cretaceous forearc rocks are
Proterozoic zircon contributed from the cratonally derived sedi- Campanian or Maastrichtian and hence are enriched in detritus
ment eroded from the batholith’s wall rocks recrystallized dur- from the eastern plutonic zone.
ing mélange formation, liberating the highly radiogenic Pb that In order to further evaluate the strength of the provenance
was dispersed throughout the amphibolite unit mélange as it was tie between the Catalina Schist and the Peninsular Ranges, we
recrystallized and metasomatically altered (Bebout and Barton, have carried out ternary mixing calculations. We have identified
1993, 2002; King et al., 2007). three distinctive components: (1) lower Mesozoic wall rocks, rep-
In the same manner as the Pb isotope data discussed above, resented by the Bedford Canyon, French Valley, Julian Schist,
the detrital zircon results from the Catalina Schist also indicate and Vallecitos flysch localities (Fig. 9A; Morgan et al., 2005);
a close genetic relationship between the Catalina Schist and the (2) Early Cretaceous volcanics, represented by results from
Peninsular Ranges batholith: (1) early Mesozoic, Paleozoic, and Lower Cretaceous sandstones and volcanics within the volcanic
Proterozoic detrital zircon from the two terranes are similar; and arc (Fig. 9B; Alsleben et al., 2005; D.L. Kimbrough, personal
(2) middle Cretaceous detrital zircon age distributions from the observ.); and (3) Upper Cretaceous forearc, represented by
lawsonite-blueschist and lower grade Catalina Schist are virtu- Cenomanian–Maastrichtian strata distributed along the western
ally identical to the distribution of crystallization ages from the margin of the Peninsular Ranges batholith (Fig. 9C; Mahoney
northern Peninsular Ranges batholith and the detrital age distri- et al., 2005). Cumulative probability plots for these three com-
bution from the adjacent Upper Cretaceous forearc rocks. ponents are shown in Figure 9D, along with equivalent curves
Lower Mesozoic flysch wall rocks of the Peninsular Ranges representing the major units of the Catalina Schist.
batholith crop out within the deeply denuded axial zone of the We have linearly mixed these three components to obtain
northern batholith (Fig. 2) and represent a major potential source the best fit to the detrital zircon age distributions from the Cata-
of early Mesozoic, Paleozoic, and Proterozoic zircon. A com- lina Schist. Results for the amphibolite unit, epidote-amphibolite
posite detrital zircon distribution based upon results obtained by unit, epidote-blueschist unit, and lawsonite-blueschist and lower-
Morgan et al. (2005) from three different localities in southern grade units are shown in Figure 9E through H, respectively. The
California (Bedford Canyon Formation, French Valley area, and relative proportions of the three end members required to pro-
Julian Schist; localities 1–3 in Fig. 2) and northern Baja Califor- duce these fits are shown in Figure 9I. As indicated in Figure 9I,
nia (Vallecitos area; locality 4 in Fig. 2) is shown in Figure 8. the detrital zircon provenance signature of the earliest accreted
As indicated, the age distribution defined by >200 Ma detrital material within the amphibolite unit of the Catalina Schist is best
zircon from the Catalina Schist (Fig. 8A) is consistent with a approximated by a 68:32 mixture of sediment derived predomi-
source region similar to the Peninsular Ranges batholith wall nantly from lower Mesozoic wall rocks and the Early Cretaceous
rocks (Fig. 8B). The expected age distribution for detritus that volcanic arc, respectively. Such a high proportion of cratonally
originated from the southwest North American craton is repre- derived material within the amphibolite unit metasediments
sented by late Miocene to Holocene sands from the Colorado makes sense, given the abundance of Middle Proterozoic zircon,
River system (Fig. 8C). Well-defined peaks at 1.0–1.2, 1.45, the aluminous and quartz-rich nature of the protolith, and the fact
and 1.65–1.75 Ga are distinctive of the basement assemblage that pegmatites, metasediments, and mélange sampled from the
(Gehrels and Stewart, 1998). Early Paleozoic and latest Neopro- amphibolite unit tend to yield a relatively radiogenic Sr, Nd, and
terozoic (400–650 Ma) zircon was contributed by supracrustal Pb isotopic signature (i.e., similar to evolved arc crust; Bebout
cover rocks that had their source in the Appalachian and Oua- and Barton, 1993, 2002; King et al., 2006, 2007).
chita orogenic belts (Dickinson and Gehrels, 2003). Finally, The provenance signature of sediment shed from the Penin-
Permian–Triassic zircon is derived from the earliest Phanero- sular Ranges batholith shifted dramatically between early Aptian
zoic magmatic arcs established along the southwestern North to early Cenomanian time (Mahoney et al., 2005). A parallel
American margin (Barth and Wooden, 2006; González-León et shift is recorded by the Catalina Schist (Fig. 4). In the case of
al., 2006). the epidote-amphibolite unit, the relative proportions of lower
The lawsonite-blueschist and lower grade rocks of the Cata- Mesozoic wall rock to Early Cretaceous volcanic arc sediment
lina Schist that contain Upper Cretaceous forearc strata of the diminishes to 44:56 (Fig. 9I). Compositionally, there is still a
Peninsular Ranges are very similar (Fig. 8E), as is the distribu- resolvable cratonal contribution in the epidote-amphibolite meta-
tion of U-Pb zircon crystallization ages from the adjacent batho- graywackes. A further shift is exhibited by the epidote-blueschist
lith (Fig. 8F; Silver and Chappell, 1988; Walawender et al., 1990; metagraywackes. Their age distribution is well modeled by a
Kistler et al., 2003; D.L. Kimbrough, personal observ.). The fact 20:58:22 mixture of lower Mesozoic wall rock, Lower Creta-
that the Catalina Schist results are skewed to slightly older ages is ceous volcanic arc, and Upper Cretaceous forearc sediment,
The Catalina Schist 349

Figure 8. Relative probability plots (200–2200 Ma) of detrital zircon U-Pb ages. (A) All results from the Catalina Schist.
(B) Representative early Mesozoic flysch wall rocks of the northern Peninsular Ranges batholith (Morgan et al., 2005);
data originate from localities 1–4 in Figure 2. (C) Representation of the southwestern North American detrital zircon
provenance signature as represented by late Miocene–Holocene sediments of the Colorado River system (D. Kimbrough
and M. Grove, personal observ.). Yellow bands denote the primary Middle Proterozoic crystallization age maxima con-
tributed by cratonal basement and supracrustal rocks of southwestern North America. (D) All detrital zircon U-Pb age
results from the lawsonite-blueschist and lower grade units of the Catalina Schist. (E) Representative Late Cretaceous
(K) detrital zircon data from the forearc of the northern Peninsular Ranges batholith (PRB) (Mahoney et al., 2005).
(F) Pluton U-Pb zircon crystallization ages from the northern Peninsular Ranges batholith (Silver and Chappell, 1988;
Walawender et al., 1990; Kistler et al., 2003; D. Kimbrough, personal observ.).

respectively (Fig. 9I). The provenance shift was completed by 115 Ma peak-grade metamorphism within the amphibolite unit.
the time the protolith of the lawsonite-blueschist and lower grade Because the pronounced age gradients yielded by both samples
metagraywackes was deposited. The detrital zircon age distribu- reached 150–160 Ma at the highest temperatures of gas release,
tion of the latter is well described by a 49:51 mixture of Lower Grove and Bebout (1995) speculated that the garnet-bearing
Cretaceous volcanic arc and Upper Cretaceous forearc sediment, blueschist blocks formed during the Middle Jurassic and had the
respectively (Fig. 9I). same tectonic significance as similar high-grade blocks present
within the Franciscan Complex along the margin to the north
Implications of Pre-115 Ma High-PT Tectonic Blocks (Wakabayashi, and Dumitru, 2007, and references cited therein)
within Low-Grade Catalina Schist and within the western Baja terrane along the margin to the
south (Baldwin and Harrison, 1989, 1992). The concordance of
Grove and Bebout (1995) reported 40Ar/39Ar results for two the 135 Ma Rb-Sr and K-Ar model ages from sample 330–4B
garnet-bearing, blueschist blocks within the lawsonite-blueschist (Fig. 5) indicate that 40Ar/39Ar results from it cannot be explained
unit of the Catalina Schist that indicated that the blocks predated by excess 40Ar contamination. Petrographic observations of
Figure 9. Relative probability plots (0–2000 Ma) of detrital zircon U-Pb age results used as end members in ternary
mixing calculations: (A) Early Mesozoic flysch wall rocks of the northern Peninsular Ranges batholith (PRB) from
Morgan et al. (2005). (B) Early Cretaceous volcanic sandstones (Alsleben et al., 2005) and volcanic rocks (D. Kim-
brough, personal observ.) of the northern Peninsular Ranges batholith. (C) Late Cretaceous forearc strata of the northern
Peninsular Ranges batholith (Mahoney et al., 2005). (D) Cumulative probability plots of the three end members. Best-fit
ternary-mixing results: (E) Amphibolite unit. (F) Epidote-amphibolite unit. (G) Epidote-blueschist unit. (H) Lawsonite-
blueschist (and lower grade) units of the Catalina Schist. (I) Ternary diagram showing best-fit solutions. Note the general
trend away from Peninsular Ranges wall rock and volcanic zircon signature exhibited by the oldest accreted units toward
a pluton-dominated zircon provenance that characterizes the youngest rocks within the Catalina Schist. Mz—Mesozoic;
K—Cretaceous.
The Catalina Schist 351

sample 330–4B indicate that the K-Ar and Rb-Sr model ages are (Fig. 10B) explains the heat source required to produce amphibo-
most sensibly interpreted as mixed ages that resulted from partial lite facies metamorphism and anatexis within the Catalina amphib-
replacement of 0.1–1 mm diameter Middle Jurassic phengite by olite unit. It also explains why the amphibolite unit underwent
much finer grained (<50 µm) middle Cretaceous phengite. This protracted greenschist facies metamorphic conditions after peak
evidence for prior (Middle Jurassic?) subduction metamorphism metamorphism occurred at 115 Ma. Greenschist facies, rather
along the margin is in good agreement with the geologic history than blueschist facies overprinting, is expected if the amphibo-
of the Peninsular Ranges batholith. Conversely, pre–115 Ma sub- lite unit was metamorphosed close to the magmatic arc and far
duction along the segment of the margin underplated by the Cata- removed from the subduction zone.
lina Schist makes it difficult to explain how mantle lithosphere Based upon detrital zircon results, the epidote-amphibolite
composing the overriding hanging wall could have retained suf- unit was accreted after 113 ± 3 Ma or several million years
ficient heat to produce amphibolite facies metamorphism within after peak grade recrystallization in the overlying amphibolite
a newly formed subduction zone at 115 Ma. unit (Fig. 6B). Accretion of the epidote-amphibolite unit cre-
ated an imbricate thrust stack beneath the western margin of the
Subduction Erosion Model for the Formation of the Peninsular Ranges batholith (Fig. 10C). Peak grade recrystalliza-
Catalina Schist tion of the epidote-amphibolite unit occurred at 110–107 Ma on
the basis of 40Ar/39Ar hornblende results from the mafic gneiss
The results and geologic relationships described above directly underlying the amphibolite unit (Fig. 6B). Just as in the
lead us to conclude that the amphibolite, epidote-amphibolite, case of the amphibolite unit, greenschist facies metamorphic
and possibly the epidote-blueschist units of the Catalina Schist conditions likely persisted within the epidote-amphibolite unit
are unlikely to have formed within the same subduction zone for up to 10 m.y. on the basis of 40Ar/30Ar phengite results from
in which the lawsonite-blueschist and lower grade rocks of the epidote-amphibolite unit metagraywackes that indicate that cool-
Catalina Schist were also accreted and metamorphosed. Whereas ing below 400–350 °C was delayed until ca. 97 Ma (Fig. 6B).
studies of plate kinematics involving the Farallon, Kula, and Only fragmentary constraints are available for the meta-
Pacific plates are not definitive, they indicate that continuous sub- morphic evolution of the epidote-blueschist unit. Previous work
duction of comparatively old oceanic crust likely prevailed along had indicated that the rocks of the epidote-amphibolite unit and
the southwestern North American margin throughout the middle epidote-blueschist unit were contiguous (the Catalina greenschist
Cretaceous (Engebretsen et al., 1986; Stock and Molnar, 1988). unit of Platt, 1976) and shared a common greenschist facies over-
Accordingly, we favor a model in which the Catalina amphibolite print (Platt, 1976; Sorensen, 1986). Our detrital zircon results
and epidote-amphibolite units represent Early Cretaceous forearc from a single sample of epidote-blueschist metasediment inter-
strata and basement (Fig. 10A) that were underthrust and meta- calated with clinozoisite-albite–bearing blueschists cast doubt
morphosed beneath a forearc thrust that was well separated from upon the likelihood that the epidote-amphibolite and epidote-
the deeper subduction megathrust (Fig. 10B). A modern analogue blueschist units were closely related prior to ca. 101 Ma. Results
for paired megathrusts depicted in Figure 10B is imaged by deep from this single sample indicate that the epidote-blueschist rocks
seismic-reflection data from the northern Cascadia subduction were accreted after 101 ± 3 Ma or roughly 6 m.y. after peak-grade
zone (Calvert, 2004). Forearc thrusting on this scale is capable recrystallization of the epidote-amphibolite unit had occurred. The
of displacing forearc rocks to subcrustal positions and is conse- epidote-blueschist assemblages could reflect subcrustal accretion
quently an important manifestation of subduction erosion. beneath the collapsing forearc at an intermediate position between
In the Catalina Schist, high-PT amphibolite facies meta- the subduction zone and the arc. Greenschist facies overprinting
morphism and partial melting may have occurred when a part of the epidote-blueschist assemblages resulted as the rocks were
of the Early Cretaceous forearc (Fig. 10A) was underthrust to a underthrust beneath the epidote-amphibolite and higher grade
position beneath the western margin of the northern Peninsular rocks of the Catalina Schist between 101 and 97 Ma.
Ranges batholith between 122 and 115 Ma (Fig. 10B). This event The lawsonite-blueschist and lower grade rocks of the Cata-
coincided broadly with previously documented intra-arc thrust- lina Schist were accreted after 97 ± 3 Ma during a major Ceno-
ing within the adjacent Peninsular Ranges batholith (Gastil et al., manian pulse of synbatholith erosion (Fig. 10D; Kimbrough et
1981; Todd et al., 1988; Silver and Chappell, 1988; Thomson and al., 2001). Evidence that accretion took place within the sub-
Girty, 1994; Busby et al., 1998; Johnson et al., 1999: Schmidt duction zone is provided by the lawsonite-blueschist and lower
et al., 2002; Wetmore et al., 2002; Schmidt and Paterson, 2002; grade mineralogy and the fact that phengite 40Ar/39Ar ages and
Wetmore et al., 2003; Busby, 2004). Shortening of the forearc the youngest detrital zircon ages coincide, indicating that accre-
could have been triggered by lateral expansion related to batho- tion, metamorphism, and cooling took place rapidly (Fig. 6D).
lith emplacement coupled with collapse of thin, hot backarc Merging of the higher grade units of the Catalina Schist with the
crust that had separated the Peninsular Ranges batholith from the subduction complex likely occurred by ca. 95 Ma and triggered
southwestern craton margin in the Early Cretaceous. tectonic denudation via low-angle extensional faulting that atten-
Underthrusting of the Early Cretaceous forearc rocks to a uated the crust and established the presently observed contact
stalled position beneath the western margin of the magmatic arc relationships within the Catalina Schist (e.g., Platt, 1986).
(A)

(B)

(C)

(D)

(E)
The Catalina Schist 353

Figure 10. Subduction erosion model for the formation of the Catalina Schist. (A) Early Cretaceous geometry of the convergent margin of the
northern Peninsular Ranges batholith. The backarc extension and sedimentation shown are inferred from geologic relationships within east-
central Peninsular Ranges batholith and formerly adjacent mainland Mexico. Detrital zircon results from Early Cretaceous volcanic sediments
(Alsleben et al., 2005) require close proximity to southwestern North America. (B) Initial underthrusting of Early Cretaceous forearc beneath
northwestern Peninsular Ranges and amphibolite facies metamorphism and anatexis at 115 Ma. Amphibolite unit stalls beneath western batholith
at a position well separated from active subduction zone. Evidence for intra-arc shortening at this time is detailed within the text. (C) Under-
thrusting of epidote-amphibolite unit during progressive subduction erosion of the Early Cretaceous forearc beneath northwestern Peninsular
Ranges batholith. Shortening in the backarc region preconditions the crust for deep melting (>40 km) to generate La Posta magmatism at 95 Ma.
(D) Accretion of the lawsonite-blueschist and lower grade units of the Catalina Schist at 95 Ma occurs concomitantly with a major pulse of de-
nudation and erosion within the Peninsular Ranges batholith (e.g., Kimbrough et al., 2001). Deep, garnet-involved melting to generate La Posta
magmatism is facilitated by overthickened arc-ophiolite basement in the former backarc, focused asthenospheric counterflow caused by initiation
of flat subduction and eastern relocation of magmatic arc, and massive devolatilization of deep underplated Catalina Schist within the subduc-
tion channel. (E) Laramide flat subduction tectonically erodes deep crustal root beneath eastern Peninsular Ranges batholith between 80 and
65 Ma, causing renewed deep exhumation of the eastern part of this batholith (Grove et al., 2003b). Tr—Triassic; J—Jurassic; K—Cretaceous;
Pz—Paleozoic; TTG—tonalite, trondjhemite, granodiorite.

Devolatilization and Fluid Flow within the Catalina Schist the forearc. Pervasive hydration of mantle lithosphere underly-
ing the forearc region is consistent with observations, in mod-
Recent studies (Bebout et al., 1999, 2007; Bebout, 2007) ern forearcs, of extensive hydration in the hanging wall leading
have discussed varying extents of devolatilization in the tectono- to distinctive seismic-velocity signatures (Bostock et al., 2002;
metamorphic units of the Catalina Schist within the context of Brocher et al., 2003) and forearc serpentinite diapirs (Fryer et
varying prograde P-T paths that took place in a subduction zone al., 1995). As continued subduction erosion brought the loci of
setting (also see discussion in Grove and Bebout, 1995). Whereas forearc thrusting nearer to the active subduction thrust system
the subduction erosion model (Fig. 10) does not affect the general (see Fig. 10C, D, E), progressively lower T rocks, such as those
conclusions of this previous work as they pertain to devolatiliza- in the epidote-amphibolite and epidote-blueschist rocks, would
tion histories, the sources and pathways for infiltrating fluids that similarly have been infiltrated by these fluids emanating from
metasomatized the higher grade rocks of the Catalina Schist do the still lower T (lawsonite-blueschist) domains produced in the
require further consideration within the context of the subduction active subduction zone thrust.
erosional model presented here. Bebout and Barton (1989) and
Bebout (1991a, b) proposed that, in the amphibolite unit, the O Relationship to Emplacement of the La Posta Tonalite-
isotope compositions in mélange and other more permeable zones Trondjhemite-Granodiorite Suite
reflect infiltration by externally derived aqueous fluids with δ O
18

inherited from prior equilibration with similar metasedimentary At ca. 100 Ma the locus of Peninsular Ranges batholith mag-
rocks but at lower temperatures (~300–600 °C). The source of the matism shifted abruptly eastward in conjunction with intrusion
fluids that ascended into the underplated (but still hot) amphibo- of the voluminous La Posta tonalite-trondjhemite-granodiorite
lite unit was primarily related to 400–550 °C chlorite breakdown plutonic suite (Fig. 2; Gastil et al., 1975; Silver and Chappell,
reactions in rocks equivalent to those within the lower grade units 1988; Walawender et al., 1990; Tulloch and Kimbrough, 2003;
of the Catalina Schist (Bebout, 1991a, b; also see discussion in Kimbrough and Grove, 2006). The La Posta suite emplacement
Grove and Bebout, 1995). involved a sustained (ca. 98–92 Ma) magmatic flux of >100 km3/
In the subduction erosion model (Fig. 10), fluids with appro- m.y./km strike length over >1200 km (Kimbrough and Grove,
priate lower T metasedimentary signatures derived from the 2006). The high Sr, Ba, Sr/Y, Na2O, Al2O3, and highly fraction-
subducting slab are required to traverse a significant thickness ated rare-earth-element patterns exhibited by these plutons indi-
(tens of kilometers) of mantle lithosphere in order to infiltrate cate deep, garnet-involved melting of a fundamentally mafic
the amphibolite unit at near peak-grade metamorphic conditions. source region (Gromet and Silver, 1987). Oxygen and Rb-Sr
We suggest that this could have occurred without substantial isotopic measurements reported by Taylor (1986), Gromet and
modification of the δ O of the fluids. Devolatilization of sub-
18
Silver (1987), Silver and Chappell (1988), Hill and Silver (1988),
ducting oceanic crust and trench fill is considered capable of and Kistler et al. (2003) have revealed that the northern La Posta
hydrating mantle lithosphere over broad regions (Peacock, 1993; belt features elevated δ18O (9–12‰ whole-rock equivalent) at
Humphreys et al., 2003). Infiltration of aqueous fluids derived intermediate initial 87Sr/86Sr values (typically 0.704–0.708). The
from the subduction zone likely began during, or prior to, the supracrustal contribution implied by these combined isotopic
Middle Jurassic. This previous history was likely sufficient to attributes cannot be accounted for by high-level assimilation of
thoroughly hydrate and isotopically re-equilibrate the interven- highly radiogenic, early Mesozoic, cratonally derived flysch host
ing lithospheric mantle with lower temperature, slab-derived flu- rocks (Shaw et al., 2003). These characteristics are more readily
ids prior to the onset of Early Cretaceous subduction erosion of explained by partial melting and/or devolatilization of isotopically
354 Grove et al.

primitive low-grade metasedimentary and metavolcanic rocks time (i.e., between 97 and 95 Ma) beneath the northern Peninsular
and altered oceanic crust within the deep source region (Taylor, Ranges (Fig. 6B). Thermochronology and detrital zircon results
1986; Gromet and Silver, 1987). obtained from the NW-SE–trending belt of schist exposures have
Emplacement of the La Posta belt plutons involved an east- revealed that schist underplating associated with the Laramide
ern relocation of the locus of magmatism within the Peninsular event was widespread by 80–70 Ma beneath the middle Creta-
Ranges over a strike length of 1200 km (Fig. 10D; Kimbrough ceous arc (Jacobson, 1990; Jacobson et al., 2000; Barth et al.,
and Grove, 2006). This event was the initial stage of an eastern 2003; Grove et al., 2003a). By 70–60 Ma, cratonal rocks were
sweep of magmatism into northern Mexico that occurred dur- being underplated by schist at positions well east of the medial
ing the Late Cretaceous–early Tertiary (e.g., Silver and Chappell, Cretaceous arc (Grove et al., 2003a; Usui et al., 2003; Jacobson
1988; McDowell et al., 2001; Staude and Barton, 2001; Henry et et al., 2007).
al., 2003). Subduction erosion (Fig. 10D) could have delivered Laramide shallow subduction processes tectonically removed
the Catalina Schist into close proximity to the La Posta source the deep lithospheric mantle roots of the La Posta belt within
region between 100 and 95 Ma. The coincidence of crustal thick- the northern Peninsular Ranges batholith between 80 and 65 Ma
ening of the former backarc extensional basin (compare Fig. 10A (Fig. 10E). Receiver function seismic studies indicate that the
with 10D) focused asthenospheric corner flow (Fig. 10D), and deep crust and lithospheric mantle roots no longer exist beneath
massive devolatilization of the Catalina Schist (Fig. 10D) may or within the northeastern Peninsular Ranges batholith (Ichinose
have set up the optimal conditions required to trigger the La Posta et al., 1996; Lewis et al., 2001). Whereas this batholith exhib-
TTG flare-up (Kimbrough and Grove, 2006). Recent δ18O mea- ited predominantly syn- to late-batholithic cooling in its southern
surements performed by Kimbrough and Grove (2006) indicate extent (>85 Ma; e.g., Ortega Rivera, 2003), the northeastern seg-
that the La Posta plutons have high δ18O only in the northern ment of the batholith was further characterized by a delayed and
segment of the batholith, where it might have been underplated very significant pulse of rapid cooling between 80 and 65 Ma
by the Catalina Schist (Fig. 2). This is a good indication that (Krummenacher et al., 1975; George and Dokka, 1994; Lovera
Catalina Schist subduction erosion was sufficiently important to et al., 1999; Grove et al., 2003b). Lovera et al. (1999) and Grove
modify the source region characteristics of the La Posta belt mag- et al. (2003b) attributed this cooling to denudation related to the
matism. An important implication is that substantial fractions removal of lithospheric mantle and the underplating of schist dur-
of subducted forearc supracrustal material can be recycled via ing Laramide shallow subduction. Holk et al. (2006) reported that
subduction erosion processes and incorporated into arc batholith Laramide-age deformation of the northern Peninsular Ranges
magmas over a short (<10 Ma) time interval. was associated with infiltration of high δD and δ18O fluids that
are most readily explained by devolatilization of subducted oce-
Relationship to Laramide Underthrusting anic crust or underplated volcanogenic sediments. The impact of
the Laramide orogeny on the northern Peninsular Ranges appears
The accretion of the lowest grade units of the Catalina Schist to have contrasted significantly with its effect upon the south-
beneath the western Peninsular Ranges and the eastern relocation ern Sierra Nevada batholith, where an inflection in the Laramide
of the La Posta plutonic belt to the eastern Peninsular Ranges at subduction zone (Pickett and Saleeby, 1993; Malin et al., 1995;
95 Ma can be viewed as important precursors to the Laramide Saleeby, 2003) apparently allowed the deep crust and upper man-
craton-ward shift of arc magmatism and contractional deforma- tle to be preserved until much more recently (Ducea and Saleeby,
tion (Coney and Reynolds, 1977; Dickinson and Snyder, 1978). 1996, 1998; Zandt et al., 2004).
During the Laramide episode of Late Cretaceous–early Tertiary
deformation, large tracts of southern California and southwest- CONCLUSIONS
ern Arizona were underplated by the high-PT Pelona Schist and
related schists, with no intervening lithospheric mantle preserved 1. Major tectonometamorphic units of the Catalina Schist
(Fig. 10E; Ehlig, 1968, 1981; Crowell, 1968, 1981; Yeats, 1968; were successively accreted over a 15–20 m.y. interval,
Haxel and Dillon, 1978; Haxel et al., 2002; Burchfiel and Davis, beginning with the amphibolite unit at ca. 120–115 Ma
1981; Jacobson, 1983, 1990; Jacobson et al., 1988, 2002, 2007; and concluding with the lawsonite-blueschist and lower
Dillon et al., 1990; Malin et al., 1995; Wood and Saleeby, 1997; grade lithologies by 97–95 Ma.
Saleeby, 2003; Grove et al., 2003a). Late Cretaceous eclogitic 2. The amphibolite unit resided at high temperatures for
xenoliths have been recovered from kimberlite pipes as far east as a protracted period of time (10–15 m.y.), whereas the
northeastern Arizona (Usui et al., 2003). lawsonite-blueschist and lower grade units that most
The earliest recognized accretion of the distinctive Pelona likely formed near or within the subduction zone were
and related schists occurred at 91 ± 1 Ma along the southwest- deposited, accreted, metamorphosed, and cooled over
ernmost tip of the Sierra Nevada batholith (Rand Schist within a time interval too brief to be resolved by the methods
the San Emigdio Mountains; Saleeby, 2003; Grove et al., 2003a). employed in this study (<3 m.y.).
Lawsonite-blueschist and the lower grade Catalina Schist rep- 3. The provenance of the Catalina Schist metasedimentary
resent nearly equivalent material underplated at a slightly older rocks shifted as a function of time of accretion (and now,
The Catalina Schist 355

metamorphic grade). Metagraywackes from the earliest Mojave–Transverse Ranges province. Field support on Catalina
accreted amphibolite unit were derived from an early Island was provided by the Catalina Island Conservancy, with
Aptian sediment that apparently originated from ero- special thanks to Frank Starkey for providing logistical support
sion of Late Triassic–Jurassic flysch wall rocks and Early and assistance in the field. Michelle Hopkins prepared samples
Cretaceous volcanics of the Peninsular Ranges batholith. and performed cathodoluminescence imaging of zircon grains.
Successively younger accreted materials became enriched J.R. Morgan assisted with ion microprobe analysis. The ion
with Early Cretaceous plutonic zircon from the Peninsular microprobe facility at UCLA and the Laserchron Center at the
Ranges. The last accreted lawsonite-blueschist and lower University of Arizona are partly supported by grants from the
grade rocks were derived from Turonian sediment with a Instrumentation and Facilities Program, Division of Earth Sci-
detrital zircon provenance virtually identical to similarly ences, U.S. National Science Foundation.
aged sediment within the Peninsular Ranges forearc.
4. The Catalina Schist likely does not represent a synchro- REFERENCES CITED
nously formed, inverted metamorphic aureole arising
from nascent subduction. The highest grade portions of Ague, J., and Brimhall, G.H., 1988, Magmatic arc asymmetry and distribu-
the complex appear to have formed by a subduction ero- tion of anomalous plutonic belts in the batholiths of California: Effects
sion process in which parts of the forearc were under- of assimilation, crustal thickness, and depth of crystallization: Geologi-
cal Society of America Bulletin, v. 100, p. 912–927, doi: 10.1130/0016-
thrust beneath the western margin of the Peninsular 7606(1988)100<0912:MAAADO>2.3.CO;2.
Ranges batholith at ca. 122–115 Ma. Progressive subduc- Alsleben, H., Wetmore, P., Gehrels, G., and Paterson, S., 2005, Using detrital
tion erosion of the forearc by continued underthrusting in zircon geochronology to track island arc collision: Geological Society of
America Abstracts with Programs, v. 37, no., 7, p. 551–552.
the forearc region ultimately juxtaposed the higher grade Altheim, B.K., Christoffel, C.A., Bebout, G.E., and Grove, M., 1997, Detec-
units of the Catalina Schist with the subduction complex tion of subtle variations in low-T, high-P/T metamorphism in the Cata-
by 97–95 Ma. Only the lawsonite-blueschist and lower lina Schist; tectonic and kinetic implications: Program with Abstracts,
Geological Association of Canada, Mineralogical Association of Canada,
grade rocks are considered to have originated along the Canadian Geophysical Union, Joint Annual Meeting, v. 22, p. 2.
dominant subduction zone thrust within a high-PT ther- Anczkiewicz, R., Platt, J.P., Thirlwall, M.F., and Wakabayashi, J., 2004, Fran-
mal regime characteristic of such settings. ciscan subduction off to a slow start: Evidence from high-precision Lu-Hf
garnet ages on high-grade blocks: Earth and Planetary Science Letters,
5. The accretion of the Catalina Schist marked the initial v. 225, p. 147–161, doi: 10.1016/j.epsl.2004.06.003.
stage of shallowing subduction and inboard migration of Atwater, T., 1970, Implications of plate tectonics for the Cenozoic tectonic
magmatism and sediment underplating that culminated evolution of western North America: Geological Society of America
Bulletin, v. 81, p. 3513–3536, doi: 10.1130/0016-7606(1970)81[3513:
in the Late Cretaceous–early Tertiary Laramide orogeny. IOPTFT]2.0.CO;2.
The sedimentary protolith of the youngest Catalina schist Bailey, E.H., 1941, Mineralogy, petrology, and geology of Catalina Island, Cal-
units are nearly identical in age and provenance to the ifornia [Ph.D. thesis]: Stanford, California, Stanford University, 193 p.
Baird, A.K., and Miesch, A.T., 1984, Batholithic Rocks of Southern Califor-
oldest representative of the eugeoclinal schists emplaced nia—A Model for the Petrochemical Nature of Their Source Materials:
beneath the southwesternmost Sierra Nevada batholith at U.S. Geological Survey Professional Paper 1284, 24 p.
ca. 92 Ma. Baldwin, S.L., and Harrison, T.M., 1989, Geochronology of blueschists from
west-central Baja California and the timing of uplift in subduction com-
plexes: Journal of Geology, v. 97, p. 149–163.
ACKNOWLEDGMENTS Baldwin, S.L., and Harrison, T.M., 1992, The P-T-T history of blocks in
serpeninite matrix melange, west-central Baja California: Geologi-
cal Society of America Bulletin, v. 104, p. 18–31, doi: 10.1130/0016-
We gratefully acknowledge discussions with Pat Abbott, 7606(1992)104<0018:TPTTHO>2.3.CO;2.
Suzanne Baldwin, Andy Barth, Mark Barton, Ned Brown, Bannon, J.L., Bottjer, D.J., Lund, S.P., and Saul, L.R., 1989, Campanian/Maas-
Cathy Busby, Mark Cloos, Mihai Ducea, Ray Ingersoll, Peter trichtian stage boundary in southern California: Resolution and implica-
tions for large-scale depositional patterns: Geology, v. 17, p. 80–83, doi:
Lonsdale, John Platt, Jason Saleeby, Dave Scholl, Sorena 10.1130/0091-7613(1989)017<0080:CMSBIS>2.3.CO;2.
Sorensen, John Wakabayashi, and Jim Wright that helped shape Barth, A.P., and Wooden, J.L., 2006, Timing of magmatism following initial
ideas presented in this paper. We particularly thank Bob Bohan- convergence at a passive margin, southwestern U.S. Cordillera, and ages
of lower crustal magma sources: Journal of Geology, v. 114, p. 231–245,
non, Mark Cloos, John Platt, Sorena Sorensen, and John Wak- doi: 10.1086/499573.
abayashi, whose reviews of previous versions of the manuscript Barth, A.P., Wooden, J.L., Tosdal, R.M., Morrison, J., Dawson, D.L., and Hernly,
led to significant improvements. Work carried out by Brian B.M., 1995, Origin of gneisses in the aureole of the San Gabriel anorthosite
complex and implications for the Proterozoic crustal evolution of south-
Altheim and Cathy Christoffel, as parts of their M.S. research ern California: Tectonics, v. 14, p. 736–752, doi: 10.1029/94TC02901.
at Lehigh University, provided important information on low- Barth, A.P., Wooden, J.L., Grove, M., Jacobson, C.E., and Pedrick, J.N., 2003,
grade metamorphism within the Catalina Schist that helped U-Pb zircon geochronology of rocks in the Salinas Valley region of Cali-
fornia: A reevaluation of the crustal structure and origin of the Salinian
guide our sampling. Helge Alsleben, Brian Mahoney, and J.R. block: Geology, v. 31, no. 6, p. 517–520.
Morgan provided permission to use detrital zircon U-Pb age Barton, M.D., Battles, D.A., Bebout, G.E., Capo, R.C., Christensen, J.N., Davis,
data from the Peninsular Ranges batholith to compare with our S.R., Hanson, R.B., Michelsen, C.J., and Trim, H.E., 1988, Mesozoic
contact metamorphism in the western United States, in Ernst, W.G., ed.,
results from the Catalina Schist. Joe Wooden and Drew Cole- Metamorphism and Crustal Evolution of the Western United States (Rubey
man shared unpublished whole-rock Pb isotopic data from the Volume 7): Englewood Cliffs, New Jersey, Prentice Hall, p. 110–178.
356 Grove et al.

Bebout, G.E., 1991a, Field-based evidence for devolatilization in subduction Brocher, T.M., Parsons, T., Tréhu, A.M., Snelson, C.M., and Fisher, M.A., 2003,
zones: Implications for arc magmatism: Science, v. 251, p. 413–416, doi: Seismic evidence for widespread serpentinized forearc upper mantle along
10.1126/science.251.4992.413. the Cascadia margin: Geology, v. 31, p. 267–270, doi: 10.1130/0091-
Bebout, G.E., 1991b, Geometry and mechanisms of fluid flow at 15 to 45 kilo- 7613(2003)031<0267:SEFWSF>2.0.CO;2.
meter depths in an early Cretaceous accretionary complex: Geophysical Brown, E.H., Wilson, D.L., Armstrong, R.L., and Harakal, J.E., 1982, Petrologic,
Research Letters, v. 18, p. 923–926. structural, and age relations of serpentinite, amphibolite, and blueschist in
Bebout, G.E., 2007, Trace element and isotopic fluxes/subducted slab, in Trea- the Shuksan Suite of the Iron Mountain–Gee Point area, North Cascades,
tise on Geochemistry, v. 3: New York, Elsevier, p. 1–50, doi: 10.1016/ Washington: Geological Society of America Bulletin, v. 93, p. 1087–1098,
B978-008043751-4/00231-5. doi: 10.1130/0016-7606(1982)93<1087:PSAARO>2.0.CO;2.
Bebout, G.E., and Barton, M.D., 1989, Fluid flow and metasomatism in a sub- Burchfiel, B.C., and Davis, G.A., 1981, Mojave Desert and environs, in Ernst,
duction zone hydrothermal system: Catalina Schist terrane, California: W.G., ed., The Geotectonic Development of California (Rubey Volume
Geology, v. 17, p. 976–980, doi: 10.1130/0091-7613(1989)017<0976: 1): Englewood Cliffs, New Jersey, Prentice Hall, p. 217–252.
FFAMIA>2.3.CO;2. Busby, C.J., 2004, Continental growth at convergent margins facing large ocean
Bebout, G.E., and Barton, M.D., 1993, Metasomatism during subduction: Prod- basins: A case study from Mesozoic Baja California, Mexico: Tectono-
ucts and possible paths in the Catalina Schist, California: Chemical Geol- physics, v. 392, p. 241–277, doi: 10.1016/j.tecto.2004.04.017.
ogy, v. 108, p. 61–92, doi: 10.1016/0009-2541(93)90318-D. Busby, C.J., Smith, D.P., Morris, W.R., and Adams, B., 1998, Evolutionary
Bebout, G.E., and Barton, M.D., 2002, Tectonic and metasomatic mixing in a model for convergent margins facing large ocean basins: Mesozoic Baja
high-T, subduction zone mélange—Insights into the geochemical evolu- California (Mexico): Geology, v. 26, p. 227–230, doi: 10.1130/0091-
tion of the slab-mantle interface: Chemical Geology, v. 187, p. 79–106, 7613(1998)026<0227:EMFCMF>2.3.CO;2.
doi: 10.1016/S0009-2541(02)00019-0. Calvert, A.J., 2004, Seismic reflection imaging of two megathrust shear zones
Bebout, G.E., Ryan, J.G., Leeman, W.P., and Bebout, A.E., 1999, Fractionation in the northern Cascadia subduction zone: Nature, v. 428, p. 163–167,
of trace elements during subduction-zone metamorphism: Impact of con- doi: 10.1038/nature02372.
vergent margin thermal evolution: Earth and Planetary Science Letters, Chen, J.H., and Moore, J.G., 1982, Uranium-lead isotopic ages from the Sierra
v. 171, p. 63–81, doi: 10.1016/S0012-821X(99)00135-1. Nevada batholith: Journal of Geophysical Research, v. 87, p. 4761–4784.
Bebout, G.E., Bebout, A.E., and Graham, C.M., 2007, Cycling of B, Li, and Clift, P., and Vannucchi, P., 2004, Controls on tectonic accretion versus erosion in
LILE (K, Cs, Rb, Ba, Sr) into subduction zones: SIMS evidence from subduction zones: Implications for the origin and recycling of the continen-
micas in high P/T metasedimentary rocks: Chemical Geology, vol. 239, p. tal crust: Reviews of Geophysics, v. 42, doi: 10.1029/2003RG000127.
284–304, doi: 10.1016/j.chemgeo.2006.10.016. Cloos, M., 1985, Thermal evolution of convergent plate margins—Thermal
Blake, M.C., Jr., Jayko, A.S., Moore, T.E., Chavez, V., Saleeby, J.B., and Seel, modeling and reevaluation of isotopic Ar-ages for blueschists in the Fran-
K., 1984, Tectonostratigraphic terranes of Magdalena Island, Baja Cali- ciscan complex of California: Tectonics, v. 4, p. 421–433.
fornia Sur, in Frizzel, V.A., Jr., ed., Geology of the Baja California Pen- Coney, P.J., and Reynolds, S.J., 1977, Cordilleran Benioff zones: Nature, v. 270,
insula: Society of Economic Paleontologists and Mineralogists, Pacific p. 403–406, doi: 10.1038/270403a0.
Section, v. 39, p. 183–191. Critelli, S., Marsaglia, K.M., and Busby, C.J., 2002, Tectonic history of a Juras-
Bohannon, R.G., and Geist, E.L., 1998, Upper crustal structure and Neogene sic backarc-basin sequence (the Gran Cañon Formation, Cedros Island,
tectonic development of the California continental borderland: Geologi- Mexico), based on compositional modes of tuffaceous deposits: Geologi-
cal Society of America Bulletin, v. 110, p. 779–800, doi: 10.1130/0016- cal Society of America Bulletin, v. 114, p. 515–527, doi: 10.1130/0016-
7606(1998)110<0779:UCSANT>2.3.CO;2. 7606(2002)114<0515:THOAJB>2.0.CO;2.
Bohannon, R.G., and Parsons, T., 1995, Tectonic implications of post-30 Ma Crouch, J.K., and Suppe, J., 1993, Late Cenozoic tectonic evolution of the
Pacific and North American relative plate motions: Geological Society of Los Angeles Basin and inner California borderland; a model for core
America Bulletin, v. 107, no. 8, p. 937–959. complex–like crustal extension: Geological Society of America Bul-
Böhnel, H., Delgado-Argote, L.A., and Kimbrough, D.L., 2002, Discordant letin, v. 105, p. 1415–1434, doi: 10.1130/0016-7606(1993)105<1415:
paleomagnetic data for middle Cretaceous intrusive rocks from northern LCTEOT>2.3.CO;2.
Baja California: Latitude displacement, tilt, or vertical axis rotation?: Tec- Crowell, J.C., 1968, Movement histories of faults in the Transverse Ranges and
tonics, v. 21, doi: 10.1029/2001TC001298. speculations on the tectonic history of California, in Dickinson, W.R., and
Boles, J.R., 1978, Basin analysis of the Eugenia Formation (Late Jurassic), Grantz, A., eds., Proceedings of Conference on Geologic Problems of San
Punta Eugenia area, Baja, California, in Howell, D.G., and McDougall, Andreas Fault System: Stanford University Publications in the Geological
K.A., eds., Mesozoic Paleogeography of the Western United States, Sciences, v. 11, p. 323–341.
Pacific Coast Paleogeography Symposium 2: Sacramento, California, Crowell, J.C., 1981, An outline of the tectonic history of southeastern California,
Society of Economic Paleontologists and Mineralogists, p. 493–498. in Ernst, W.G., ed., The Geotectonic Development of California (Rubey
Boles, J.R., and Landis, C.A., 1984, Jurassic sedimentary mélange and asso- Volume 1): Englewood Cliffs, New Jersey, Prentice Hall, p. 583–600.
ciated facies, Baja California, Mexico: Geological Society of America Debiche, M.G., Cox, A., and Engebretson, D., 1987, The Motion of Alloch-
Bulletin, v. 95, p. 513–521, doi: 10.1130/0016-7606(1984)95<513: thonous Terranes across the North Pacific Basin: Geological Society of
JSMAAF>2.0.CO;2. America Special Paper 207, 49 p.
Bonini, J.A., and Baldwin, S.L., 1998, Mesozoic metamorphic and middle to Dibblee, T.W., Jr., 1999, Geologic map of the Palos Verdes Peninsula and
late Tertiary magmatic events on Magdalena and Santa Margarita islands, vicinity (Redondo Beach, Torrance, and San Pedro Quadrangles) Los
Baja California Sur, Mexico: Implications for the tectonic evolution of the Angeles County, California: Dibblee Geological Foundation Map DF-70
Baja California continental borderland: Geological Society of America (Ehrenspeck, H.E., Ehlig, P.L., and Bartlett, W.L., eds.), scale 1:24,000,
Bulletin, v. 110, p. 1094–1104, doi: 10.1130/0016-7606(1998)110<1094: colored, 4 cross-sections.
MMAMTL>2.3.CO;2. Dickinson, W.R., 1983, Cretaceous sinistral strike slip along Nacimiento fault
Bostock, M.G., Hyndman, R.D., Rondenay, S., and Peacock, S.M., 2002, An in coastal California: American Association of Petroleum Geologists Bul-
inverted continental Moho and the serpentinization of the forearc mantle: letin, v. 67, p. 624–645.
Nature, v. 417, p. 536–538, doi: 10.1038/417536a. Dickinson, W.R., and Gehrels, G.E., U-Pb ages of detrital zircons from Permian
Bottjer, D.J., and Link, M.H., 1984, A synthesis of Late Cretaceous southern and Jurassic eolian sandstones from the Colorado Plateau, USA: paleo-
California and northern Baja California paleogeography, in Crouch, J.K., geographic implications: Sedimentary Geology, v. 163, p. 29–66.
and Bachman, S.B., eds., Tectonics and Sedimentation along the Califor- Dickinson, W.R., and Snyder, W.S., 1978, Plate tectonics of the Laramide orog-
nia Margin: Fullerton, California, Society of Economic Paleontologists eny, in Matthews, V., III, ed., Laramide Folding Associated with Base-
and Mineralogists, Pacific Section, p. 171–188. ment Block Faulting in the Western United States: Geological Society of
Bottjer, D.J., Colburn, I.P., and Cooper, J.D., 1982, Late Cretaceous Deposi- America Memoir 151, p. 355–366.
tional Environments and Paleogeography, Santa Ana Mountains, South- Dickinson, W.R., Hopson, C.A., and Saleeby, J.B., 1996, Alternate origins of
ern California: Los Angeles, Society of Economic Paleontologists and the Coast Range ophiolite (California): Introduction and implications:
Mineralogists, Pacific Section, 121 p. GSA Today, v. 6, no. 2, p. 1–10.
The Catalina Schist 357

Dickinson, W.R., Ducea, M., Rosenberg, L.I., Greene, H.G., Graham, S.A., González-León, C.M., Stanley, G.D., Jr., Gehrels, G.E., and Centeno-García,
Clark, J.C., Weber, G.E., Kidder, S., Ernst, W.G., and Brabb, E.E., 2005, E., 2006, New data on the lithostratigraphy, detrital zircon and Nd iso-
Net Dextral Slip, Neogene San Gregorio–Hosgri Fault Zone, Coastal tope provenance, and paleogeographic setting of the El Antimonio Group,
California: Geologic Evidence and Tectonic Implications: Geological Sonora, Mexico, in Anderson, T.H., et al., eds., The Mojave-Sonora
Society of America Special Paper 391, 43 p. Megashear Hypothesis: Development, Assessment, and Alternatives:
Dillon, J.T., Haxel, G.B., and Tosdal, R.M., 1990, Structural evidence for north- Geological Society of America Special Paper 393, p. 259–282.
eastward movement on the Chocolate Mountains thrust, southeasternmost Gromet, L.P., and Silver, L.T., 1987, REE variations across the Peninsular
California: Journal of Geophysical Research, v. 95, p. 19,953–19,971. Ranges batholith: Implications for batholithic petrogenesis and crustal
Ducea, M.N., and Saleeby, J.B., 1996, Buoyancy sources for a large, unrooted growth in magmatic arcs: Journal of Petrology, v. 28, p. 75–125.
mountain range, the Sierra Nevada, California: Evidence from xenolith Grove, M., and Bebout, G.E., 1995, Cretaceous tectonic evolution of coastal
thermobarometry: Journal of Geophysical Research, v. 101, p. 8229– southern California: Insights from the Catalina Schist: Tectonics, v. 14,
8244, doi: 10.1029/95JB03452. p. 1290–1308, doi: 10.1029/95TC01931.
Ducea, M.N., and Saleeby, J.B., 1998, The age and origin of a thick mafic- Grove, M., Jacobson, C.E., Barth, A.P., and Vucic, A., 2003a, Temporal and
ultramafic keel from beneath the Sierra Nevada batholith: Contribu- spatial trends of Late Cretaceous–Early Tertiary underplating of Pelona
tions to Mineralogy and Petrology, v. 133, p. 169–185, doi: 10.1007/ and related schists beneath southern California and southwestern Arizona,
s004100050445. in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
Ehlig, P.L., 1968, Causes of distribution of Pelona, Rand, and Orocopia Schists and the Southwestern USA: Geological Society of America Special Paper
along the San Andreas and Garlock faults, in Dickinson, W.R., and 374, p. 381–406.
Grantz, A., eds., Proceedings of Conference on Geologic Problems of San Grove, M., Lovera, O.M., and Harrison, T.M., 2003b, Late Cretaceous cooling
Andreas Fault System: Stanford University Publications in the Geological of the east-central Peninsular Ranges Batholith (33 degrees N): Relation-
Sciences, v. 11, p. 294–306. ship to La Posta Pluton emplacement, Laramide shallow subduction, and
Ehlig, P.L., 1981, Origin and tectonic history of the basement terrane of the San forearc, in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern
Gabriel Mountains, central Transverse Ranges, in Ernst, W.G., ed., The Mexico and the Southwestern USA: Geological Society of America Spe-
Geotectonic Development of California (Rubey Volume 1): Englewood cial Paper 374, p. 355–379.
Cliffs, New Jersey, Prentice Hall, p. 253–283. Hacker, B.R., 1990a, Simulation of the metamorphic and deformational history
Engebretsen, D.C., Cox, A., and Gordon, R.G., 1986, Relative motions between of the metamorphic sole of the Oman Ophiolite: Journal of Geophysical
oceanic and continental plates in the Pacific Basin: Geological Society of Research, v. 95, p. 4895–4907.
America Special Paper 206, 59 p. Hacker, B.R., 1990b, The role of deformation in the formation of metamor-
Ernst, W.G., 1988, Tectonic history of subduction zones inferred from retrograde phic gradients: Ridge subduction beneath the Oman Ophiolite: Tectonics,
blueschist P-T paths: Geology, v. 16, p. 1081–1085, doi: 10.1130/0091- v. 10, p. 455–473.
7613(1988)016<1081:THOSZI>2.3.CO;2. Hall, C.A., 1991, Geology of the Point Sur–Lopez Point Region, Coast Ranges,
Evernden, J.F., and Kistler, R.W., 1970, Chronology of Emplacement of Meso- California: A Part of the Southern California Allochthon: Geological
zoic Batholithic Complexes in California and Western Nevada: U.S. Geo- Society of America Special Paper 266, 40 p.
logical Survey Professional Paper 623, 42 p. Haxel, G.B., and Dillon, J.T., 1978, The Pelona-Orocopia Schist and Vin-
Fletcher, J.M., Grove, M., Kimbrough, D.L., Lovera, O.M., and Gehrels, G.E., cent–Chocolate Mountain thrust system, southern California, in How-
2007, Ridge-trench interactions and the Neogene tectonic evolution of the ell, D.G., and McDougall, K.A., eds., Mesozoic Paleogeography of the
Magdalena Shelf: Insights from detrital zircon U-Pb ages from the Mag- Western United States: Society of Economic Paleontologists and Min-
dalena fan and adjacent areas: Geological Society of America Bulletin, v. eralogists, Pacific Section, Pacific Coast Paleogeography Symposium 2,
119, no. 11-12, p. 1313–1336. p. 453–469.
Flynn, C.J., 1970, Post-batholithic geology of La Gloria–Presa Rodriquez Haxel, G.B., Jacobson, C.E., Richard, S.M., Tosdal, R.M., and Grubensky, M.J.,
area, Baja California, Mexico: Geological Society of America Bul- 2002, The Orocopia Schist in southwest Arizona: Early Tertiary oceanic
letin, v. 81, p. 1789–1806, doi: 10.1130/0016-7606(1970)81[1789: rocks trapped or transported far inland, in Barth, A.P., ed., Crustal Evolu-
PGOTLG]2.0.CO;2. tion of the Southwestern United States: Geological Society of America
Fry, J.G., Bottjer, D.J., and Lund, S.P., 1985, Magnetostratigraphy of displaced Special Paper 365, p. 99–128.
Upper Cretaceous strata in southern California: Geology, v. 13, p. 648– Henry, C.D., McDowell, F.W., and Silver, L.T., 2003, Geology and geochronol-
651, doi: 10.1130/0091-7613(1985)13<648:MODUCS>2.0.CO;2. ogy of granitic batholith complex, Sinaloa, México: Implications for Cor-
Fryer, P., Mottl, M., Johnson, L., Haggerty, J., Phipps, S., and Maekawa, H., dilleran magmatism and tectonics, in Johnson, S.E., et al., eds., Tectonic
1995, Serpentine bodies in the forearcs of Western Pacific convergent Evolution of Northwestern Mexico and the Southwestern USA: Geologi-
margins: Origin and associated fluids, in Taylor, B., and Natland, J., eds., cal Society of America Special Paper 374, p. 237–273.
Active Margins and Marginal Basins of the Western Pacific: American Hill, D.J., 1976, Geology of the Jurassic basement rocks, Santa Cruz Island,
Geophysical Union Geophysical Monograph 88, p. 259–279. California, and correlation with other Mesozoic basement terranes in
Gastil, R.G., 1993, Prebatholithic history of Peninsular California, in Gastil, California, in Howell, D.G., ed., Aspects of the Geologic History of the
R.G., and Miller, R.H., eds., The Prebatholithic Stratigraphy of Peninsular California Continental Borderland: American Association of Petroleum
California: Geological Society of America Special Paper 279, p. 145–156. Geologists, Pacific Section, Miscellaneous Publications, v. 24, p. 16–56.
Gastil, R.G., Phillips, R.P., and Allison, E.C., 1975, Reconnaissance Geology Hill, M.L., and Dibblee, T.W., 1953, San Andreas, Garlock, and Big Pine faults,
of the State of Baja California: Geological Society of America Memoir California—A study of the character, history, and tectonic significance
140, 170 p. of their displacements: Geological Society of America Bulletin, v. 64,
Gastil, R.G., Morgan, G.J., and Krummenacher, D., 1981, The tectonic history p. 443–458, doi: 10.1130/0016-7606(1953)64[443:SAGABP]2.0.CO;2.
of Peninsular and adjacent Mexico, in Ernst, W.G., ed., The Geotectonic Hill, R.I., and Silver, L.T., 1988, San Jacinto intrusive complex: Constraints on
Development of California (Rubey Volume 1): Englewood Cliffs, New crustal magma chamber processes from strontium isotope heterogeneity:
Jersey, Prentice Hall, p. 284–306. Journal of Geophysical Research, v. 93, p. 10,373–10,388.
Gehrels, G.E., and Stewart, J.H., 1998, Detrital zircon geochronology of Holk, G.J., Taylor, H.P., and Gromet, P.L., 2006, Stable isotope evidence for
Cambrian to Triassic miogeoclinal and eugeoclinal strata of Sonora, large-scale infiltration of metamorphic fluids generated during shallow
Mexico: Journal of Geophysical Research, v. 103, p. 2471–2487, doi: subduction into the eastern Peninsular Ranges mylonite zone (EPRMZ),
10.1029/97JB03251. Southern California: International Geology Review, v. 48, p. 209–222.
George, P.G. and Dokka, R.K., 1994, Major Late Cretaceous cooling events in the Hopson, C.A., Mattinson, J.M., and Pessagno, E.A., Jr., 1981, Coast Range
eastern Peninsular Ranges, California and their implications for Cordille- Ophiolite, western California, in Ernst, W.G., ed., The Geotectonic Devel-
ran tectonics: Geological Society of America Bulletin, v. 106, p. 903–914. opment of California (Rubey Volume 1): Englewood Cliffs, New Jersey,
Girty, G.H., 1987, Sandstone provenance, Point Loma formation, San Diego, Prentice Hall, p. 418–510.
California: Evidence for uplift of the Peninsular Ranges during the Howell, D.G., and Vedder, J.G., 1981, Structural implications of stratigraphic
Laramide orogeny: Journal of Sedimentary Petrology, v. 57, p. 839–844. discontinuities across the southern California continental borderland, in
358 Grove et al.

Ernst, W.G., ed., The Geotectonic Development of California (Rubey Vol- Diego State University (field trip guidebook for the 1979 Geological Soci-
ume 1): Englewood Cliffs, New Jersey, Prentice Hall, p. 535–559. ety of America Annual Meeting, San Diego, California), p. 1–28.
Humphreys, E., Hessler, E., Dueker, K., Farmer, G.L., Erslev, E., and Atwater, Kimbrough, D.L., 1985, Tectonostratigraphic terranes of the Vizcaino Pen-
T., 2003, How Laramide-age hydration of North American lithosphere insula and Cedros and San Benitos islands, Baja California, Mexico, in
by the Farallon slab controlled subsequent activity in the Western United Howell, D.G., ed., Tectonostratigraphic Terranes of the Circum-Pacific
States: International Geology Review, v. 45, p. 575–595. Region: Houston, Texas, Circum-Pacific Council for Energy and Mineral
Ichinose, G.A., Day, S., Magistrale, H., Prush, T., Vernon, F., and Edelman, A., Resources, Earth Science Series, v. 1, p. 285–298.
1996, Crustal thickness variations beneath the Peninsular Ranges, south- Kimbrough, D.L., and Grove, M., 2006, The eastern Peninsular Ranges batho-
ern California: Geophysical Research Letters, v. 23, p. 3095–3098, doi: lith magmatic flare-up: Insight from zircon U-Pb ages and oxygen iso-
10.1029/96GL03020. tope ratios: Backbone of the Americas, Patagonia to Alaska, Geological
Ingersoll, R.V., and Rumelhart, P.E., 1999, Three-stage evolution of the Los Society of America Specialty Meetings Abstracts with Programs, no. 2,
Angeles Basin, Southern California: Geology, v. 27, p. 593–596, doi: p. 107.
10.1130/0091-7613(1999)027<0593:TSEOTL>2.3.CO;2. Kimbrough, D.L., and Moore, T.E., 2003, Ophiolite and volcanic arc assem-
Irwin, W.P., 2002, Correlation of the Klamath Mountains and Sierra Nevada: blages on the Vizcaino Peninsula and Cedros Island, Baja California Sur,
U.S. Geological Survey Open-File Report 02–490. México: Mesozoic forearc lithosphere of the Cordilleran magmatic arc,
Jacobson, C.E., 1983, Structural geology of the Pelona Schist and Vincent in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
thrust, San Gabriel Mountains, California: Geological Society of Amer- and the Southwestern USA: Geological Society of America Special Paper
ica Bulletin, v. 94, p. 753–767, doi: 10.1130/0016-7606(1983)94<753: 374, p. 43–71.
SGOTPS>2.0.CO;2. Kimbrough, D.L., Smith, D.P., Mahoney, B.J., Moore, T.E., Grove, M., Gastil,
Jacobson, C.E., 1990, The 40Ar/39Ar geochronology of the Pelona Schist and R.G., Ortega-Rivera, A., and Fanning, M.C., 2001, Forearc-basin sedimen-
related rocks, southern California: Journal of Geophysical Research, tary response to rapid Late Cretaceous batholith emplacement in the Pen-
v. 95, p. 509–528. insular Ranges of southern and Baja California: Geology, v. 29, p. 491–
Jacobson, C.E., Dawson, M.R., and Postlethwaite, C.E., 1988, Structure, meta- 494, doi: 10.1130/0091-7613(2001)029<0491:FBSRTR>2.0.CO;2.
morphism, and tectonic significance of the Pelona, Orocopia, and Rand King, R.L., Bebout, G.E., Moriguti, T., and Nakamura, E., 2006, Elemental
Schists, southern California, in Ernst, W.G., ed., Metamorphism and mixing systematics and Sr-Nd isotope geochemistry of mélange for-
Crustal Evolution of the Western United States (Rubey Volume 7): Engle- mation: Obstacles to identification of fluid sources to arc volcanics:
wood Cliffs, New Jersey, Prentice Hall, p. 976–997. Earth and Planetary Science Letters, v. 246, p. 288–304, doi: 10.1016/
Jacobson, C.E., Barth, A.P., and Grove, M., 2000, Late Cretaceous protolith j.epsl.2006.03.053.
age and provenance of the Pelona and Orocopia Schists, southern Califor- King, R.L., Bebout, G.E., Grove, M., Moriguti, T., Nakamura, E., 2007, Boron
nia: implications for evolution of the Cordilleran margin: Geology, v. 28, and lead isotope signatures of subduction-zone mélange formation:
p. 219–222. Hybridization and fractionation along the slab-mantle interface beneath
Jacobson, C.E., Grove, M., Stamp, M.M., Vucic, A., Oyarzabal, F.R., Haxel, volcanic arcs: Chemical Geology, doi: 1016/j.chemgeo.2007.01.009.
G.B., Tosdal, R.M., and Sherrod, D.R., 2002, Exhumation history of the Kistler, R.W., Wooden, J.L., and Morton, D.M., 2003, Isotopes and Ages in the
Orocopia Schist and related rocks in the Gavilan Hills area of southeast- Northern Peninsular Ranges Batholith, Southern California: U.S. Geo-
ernmost California, in Barth, A.P., ed., Crustal Evolution of the South- logical Survey Open-File Report 03–489, 45 p.
western United States: Geological Society of America Special Paper 365, Kröner, A., Jaeckel, P., and Williams, I.S., 1994, Pb-loss patterns in zircons
p. 129–154. from a high-grade metamorphic terrain as revealed by different dating
Jacobson, C.E., Grove, M., Vucic, A., Pedrick, J., and Ebert, K.A., 2007, methods: U–Pb and Pb–Pb ages for igneous and metamorphic zircons
Exhumation of the Orocopia Schist and associated rocks of southeastern from northern Sri Lanka: Precambrian Research, v. 66, p. 151–181, doi:
California: Relative roles of erosion, synsubduction tectonic denudation, 10.1016/0301-9268(94)90049-3.
and middle Cenozoic extension, in Cloos, M., Carlson, W.D., Gilbert, Krummenacher, D., Gastil, R.G., Bushee, J., and Dupont, J., 1975, K-Ar appar-
M.C., Lious, J.G., and Sorensen, S., eds., Convergent Margin Terranes ent ages, Peninsular Ranges batholith, southern California and Baja Cali-
and Associated Regions: A Tribute to W.G. Ernst: Geological Society of fornia: Geological Society of America Bulletin, v. 86, p. 760–768, doi:
America Special Paper 419, p. 1–37. 10.1130/0016-7606(1975)86<760:KAAPRB>2.0.CO;2.
Jahns, R.H., 1954, Geology of the Peninsular Ranges Province, southern Cali- Langenheim, V.E., and Jachens, R.C., 2003, Crustal structure of the Penin-
fornia and Baja California, in Jahns, R.H., ed., Geology of Southern Cali- sular Ranges batholith from magnetic data: Implications for Gulf of
fornia: California Division of Mines Bulletin, v. 170, p. 29–52. California rifting: Geophysical Research Letters, vol. 30, no. 11, doi:
Jamieson, R.A., 1986, PT paths from high temperature shear zones beneath 10.1029/2003GL017159.
ophiolites: Journal of Metamorphic Geology, v. 4, p. 3–22, doi: 10.1111/ Larsen, E.S., 1948, Batholith and Associated Rocks of Corona, Elsinore, and
j.1525-1314.1986.tb00335.x. San Luis Rey Quadrangles, Southern California: Geological Society of
Johnson, S.E., Tate, M.C., and Fanning, C.M., 1999, New geologic mapping America Memoir 29, 182 p.
and SHRIMP U-Pb zircon data in the Peninsular Ranges batholith, Baja Lewis, J.L., Day, S.M., Magistrale, H., Castro, R.R., Astiz, L., Rebollar, C.,
California, Mexico: Evidence for a suture?: Geology, v. 27, p. 743–746, Eakins, J., Vernon, F.L., and Brune, J.N., 2001, Crustal thickness of the
doi: 10.1130/0091-7613(1999)027<0743:NGMASU>2.3.CO;2. Peninsular Ranges and Gulf Extensional Province in the Californias: Jour-
Jones, D.L., Blake, M.C., Jr., and Rangin, C., 1976, The four Jurassic belts of nal of Geophysical Research, p. 13,599–13,612, doi: 2001JB000178.
northern California and their significance to the geology of the south- Lovera, O.M., Grove, M., Kimbrough, D.L., and Abbott, P.L., 1999, Magni-
ern California Borderland, in Howell, D.G., ed., Aspects of the Geologic tude and time scales in the exhumation of arc crust: An approach based
History of the California Continental Borderland: American Association upon analysis of detrital closure age distributions: Journal of Geophysical
of Petroleum Geologists, Pacific Section, Miscellaneous Publication 24, Research, v. 104, p. 29,419–29,438, doi: 10.1029/1999JB900082.
p. 343–362. Ludwig, K.R., 2003, ISOPLOT/EX Version 3: Berkeley, California, Berkeley
Karig, D.E., 1980, Material transport within accretionary prisms and the Geochronology Center Special Publication 4.
‘knocker’ problem: Journal of Geology, v. 88, p. 27–39. Mahoney, J.B., Kimbrough, D.L., Grove, M., Jacobson, C.E., and Mustard,
Kennedy, M.P., and Moore, G.W., 1971, Stratigraphic relations of Upper Creta- P.S., 2005, Variations in Late Cretaceous batholith exhumation patterns
ceous and Eocene formations, San Diego coastal area, California: Ameri- along the western North American margin: Geological Society of Amer-
can Association of Petroleum Geologists Bulletin, v. 55, p. 709–722. ica Abstracts with Programs, v. 37, no. 7, p. 482.
Kienast, J.R., and Rangin, C., 1982, Mesozoic blueschists and melanges Malin, P.E., Goodman, E.D., Henyey, T.L., Li, Y.G., Okaya, D.A., and Saleeby,
of Cedros Island (Baja California, Mexico): A consequence of nappe J.B., 1995, Significance of seismic reflections beneath a tilted exposure of
emplacement or subduction?: Earth and Planetary Science Letters, v. 59, deep continental crust, Tehachapi Mountains, California: Journal of Geo-
p. 119–138, doi: 10.1016/0012-821X(82)90121-2. physical Research, v. 100, p. 2069–2087, doi: 10.1029/94JB02127.
Kilmer, F.J., 1979, A geological sketch of Cedros Island, Baja California, Mex- Mattinson, J.M., 1986, Geochronology of high-pressure–low-temperature
ico, in Gastil, R.G., and Abbott, P.L., eds., Baja California Geology: San Franciscan metabasites: A new approach using the U-Pb system, in Evans,
The Catalina Schist 359

B.W., and Brown, E.H., eds., Blueschists and Eclogites: Geological Soci- deep levels of the Sierra Nevada batholith, Tehachapi Mountains, Califor-
ety of America Memoir 164, p. 95–106. nia: Journal of Geophysical Research, v. 98, p. 609–629.
Mattinson, J.M., and Hill, D.J., 1976, Age of plutonic basement rocks, Santa Platt, J.P., 1975, Metamorphic and deformational processes in the Francis-
Cruz Island, in Howell, D.G., ed., Aspects of the Geologic History of the can Complex, California: Some insights from the Catalina Schist ter-
California Continental Borderland: American Association of Petroleum rane: Geological Society of America Bulletin, v. 86, p. 1337–1347, doi:
Geologists, Pacific Section, Miscellaneous Publications, v. 24, p. 53–58. 10.1130/0016-7606(1975)86<1337:MADPIT>2.0.CO;2.
McDowell, F.W., Roldán-Quintana, J., and Connelly, J.N., 2001, Duration Platt, J.P., 1976, The petrology, structure, and geologic history of the Catalina
of Late Cretaceous–early Tertiary magmatism in east-central Sonora, Schist terrain, southern California: University of California Publications
Mexico: Geological Society of America Bulletin, v. 113, p. 521–531, doi: in Geological Sciences, v. 112, 111 p.
10.1130/0016-7606(2001)113<0521:DOLCET>2.0.CO;2. Platt, J.P., 1986, Dynamics of orogenic wedges and the uplift of high-pressure
McWilliams, M.O., and Howell, D.G., 1982, Exotic terranes of western Cali- metamorphic rocks: Geological Society of America Bulletin, v. 97, p. 1037–
fornia: Nature, v. 297, p. 215–217, doi: 10.1038/297215a0. 1053, doi: 10.1130/0016-7606(1986)97<1037:DOOWAT>2.0.CO;2.
Moore, T.E., 1985, Stratigraphic and tectonic significance of the Mesozoic Rothstein, D.A., and Manning, C.E., 2003, Geothermal gradients in continental
tectonostratigraphic terranes of the Vizcaíno peninsula, Baja California magmatic arcs; constraints from the eastern Peninsular Ranges Batholith,
Sur, Mexico, in Howell, D.G., ed., Tectonostratigraphic Terranes of the Baja California, Mexico, in Johnson, S.E., et al., eds., Tectonic Evolution
Circum-Pacific Region: Houston, Circum-Pacific Council for Energy and of Northwestern Mexico and the Southwestern USA: Geological Society
Mineral Resources, p. 315–329. of America Special Paper 374, p. 337–354.
Moore, T.E., 1986, Petrology and tectonic implications of the blueschist-bear- Rubatto, D., 2002, Zircon trace element geochemistry, partitioning with garnet
ing Puerto Nuevo mélange complex, Vizcaino Peninsula, Baja Califor- and the link between U-Pb ages and metamorphism: Chemical Geology,
nia Sur, Mexico, in Evans, B.W., and Brown, E.H., eds., Blueschists and v. 184, p. 123–138, doi: 10.1016/S0009-2541(01)00355-2.
Eclogites: Geological Society of America Memoir 164, p. 43–58. Saleeby, J.B., 2003, Segmentation of the Laramide Slab—Evidence from the
Morgan, J.R., Kimbrough, D.L., and Grove, M., 2005, Detrital U/Pb zircon southern Sierra Nevada region: Geological Society of America Bul-
ages from the Peninsular Ranges Mesozoic flysch belt of southern and letin, v. 115, p. 655–668, doi: 10.1130/0016-7606(2003)115<0655:
Baja California: Peninsula Geological Society, 7th International Meeting, SOTLSF>2.0.CO;2.
Abstracts with Programs, p. 20. Saleeby, J.B., and Sharp, W.D., 1980, Chronology of the structural and pet-
Nicholson, C., Sorlien, C.C., Atwater, T., Crowell, J.C., and Luyendyk, B.P., rologic development of the southwest Sierra Nevada foothills, Califor-
1994, Microplate capture, rotation of the western Transverse Ranges, nia: Geological Society of America Bulletin, v. 91, p. 317–320, doi:
and initiation of the San Andreas transform as a low-angle fault system: 10.1130/0016-7606(1980)91<317:COTSAP>2.0.CO;2.
Geology, v. 22, p. 491–495, doi: 10.1130/0091-7613(1994)022<0491: Schmidt, K.L., and Paterson, S.R., 2002, A doubly vergent, fan structure in the
MCROTW>2.3.CO;2. Peninsular Ranges batholith: Transpression or complex flow along a crustal-
Nilsen, T.H., and Abbott, P.L., 1981, Paleogeography and sedimentology of scale discontinuity?: Tectonics, v. 21, doi: 10.1029/2001TC001353.
Upper Cretaceous turbidites, San Diego, California: American Associa-
Schmidt, K.L., Wetmore, P.H., Johnson, S.E., and Paterson, S.R., 2002, Along-
tion of Petroleum Geologists Bulletin, v. 65, p. 1256–1284.
strike variation in the origin of an ocean-continent basement join in the
Nordstrom, C.E., 1970, Lusardi Formation—A post-batholithic Cretaceous con-
Mesozoic Peninsular Ranges batholith: To what degree does an island arc
glomerate north of San Diego, California: Geological Society of Amer-
collision drive orogeny?, in Barth, A., ed., Crustal Evolution of the South-
ica Bulletin, v. 81, p. 601–605, doi: 10.1130/0016-7606(1970)81[601:
western United States: Geological Society of America Special Paper 365,
LFAPCC]2.0.CO;2.
p. 49–72.
Ortega-Rivera, A., 2003, Geochronological constraints on the tectonic history
Schoellhamer, J.E., and Woodford, A.O., 1951, The floor of the Los Angeles
of the Peninsular Ranges Batholith of Alta and Baja California: Tectonic
basin, Los Angeles, Orange, and San Bernardino Counties, California:
implications for western México, in Johnson, S.E., et al., eds., Tectonic
U.S. Geological Survey Oil and Gas Investigations Map OM-117.
Evolution of Northwestern Mexico and the Southwestern USA: Geologi-
cal Society of America Special Paper 374, p. 297–335. Schoellhamer, J.E., Vedder, J.G., Yerkes, R.F., and Kinney, D.M., 1981, Geol-
Page, B.M., 1970, Sur-Nacimiento fault zone of California: Continental margin ogy of the Northern Santa Ana Mountains, California: U.S. Geological
tectonics: Geological Society of America Bulletin, v. 81, p. 667–690, doi: Survey Professional Paper 420-D, 109 p.
10.1130/0016-7606(1970)81[667:SFZOCC]2.0.CO;2. Scholl, D.W., von Huene, R., Vallier, T.L., and Howell, D.G., 1980, Sedimentary
Page, B.M., 1981, The southern Coast Ranges, in Ernst, W.G., ed., The Geo- masses and concepts about tectonic processes at underthrust ocean mar-
tectonic Development of California (Rubey Volume 1): Englewood Cliffs, gins: Geology, v. 8, p. 564–568, doi: 10.1130/0091-7613(1980)8<564:
New Jersey, Prentice Hall, p. 329–417. SMACAT>2.0.CO;2.
Page, B.M., 1982, Migration of Salinian composite block, California, and dis- Scholl, D.W., and von Huene, R., 2007, Crustal recycling at modern subduction
appearance of fragments: American Journal of Science, v. 282, p. 1694– zones applied to the past—Issues of growth and preservation of continen-
1734. tal basement, mantle geochemistry, and supercontinent reconstruction, in
Patterson, D.L., 1984, Los Chapunes and Valle sandstones: Cretaceous petrofa- Hatcher, R.D., Jr., Carlson, M.P., McBride, J.H., and Martínez Catalán,
cies of the Vizcaino basin, Baja California, Mexico, in Frizzell, V.A., Jr., J.R., eds., The 4D Framework of Continental Crust: Geological Society
ed., Geology of the Baja California Peninsula: Society of Economic Pale- of America Memoir 200, p. 9–32.
ontologists and Mineralogists, Pacific Section, v. 39, p. 161–172. Sedlock, R.L., 1988a, Metamorphic petrology of a high-pressure, low-tem-
Peacock, S.M., 1987, Creation and preservation of subduction-related metamor- perature subduction complex in west-central Baja California, Mexico:
phic gradients: Journal of Geophysical Research, v. 92, p. 12,763–12,781. Journal of Metamorphic Geology, v. 6, p. 205–233, doi: 10.1111/j.1525-
Peacock, S.M., 1992, Blueschist-facies metamorphism, shear heating, and P-T-t 1314.1988.tb00416.x.
paths in subduction shear zones: Journal of Geophysical Research, v. 97, Sedlock, R.L., 1988b, Tectonic setting of blueschist and island-arc terranes of
p. 17,693–17,707. west central Baja California, Mexico: Geology, v. 16, p. 623–626, doi:
Peacock, S.M., 1993, Large-scale hydration of the lithosphere above sub- 10.1130/0091-7613(1988)016<0623:TSOBAI>2.3.CO;2.
ducting slabs: Chemical Geology, v. 108, p. 49–59, doi: 10.1016/0009- Shaw, S.E., Todd, V.R., and Grove, M., 2003, Jurassic peraluminous gneissic
2541(93)90317-C. granites in the axial zone of the Peninsular Ranges, southern California,
Peacock, S.M., Rushmer, T., and Thompson, A.B., 1994, Partial melting of in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
subducting oceanic crust: Earth and Planetary Science Letters, v. 121, and the Southwestern USA: Geological Society of America Special Paper
p. 227–244, doi: 10.1016/0012-821X(94)90042-6. 374, p.157–183.
Peterson, G.L., and Nordstrom, C.E., 1970, Sub–La Jolla unconformity in Shervais, J.W., Murchey, B.L., Kimbrough, D.L., Renne, P.R., and Hanan, B.,
vicinity of San Diego, California: American Association of Petroleum 2005, Radioisotopic and biostratigraphic age relations in the Coast Range
Geologists Bulletin, v. 54, p. 265–274. Ophiolite, northern California: Implications for the tectonic evolution of
Pickett, D.A., and Saleeby, J.B., 1993, Thermobarometric constraints on the the Western Cordillera: Geological Society of America Bulletin, v. 117,
depth of the exposure and conditions of plutonism and metamorphism at p. 633–653, doi: 10.1130/B25443.1.
360 Grove et al.

Silver, L.T., and Chappell, B.W., 1988, The Peninsular Ranges batholith: An Peninsular Ranges, California: Tectonics, v. 13, p. 1108–1119, doi:
insight into the evolution of the Cordilleran batholiths of southwestern 10.1029/94TC01649.
North America: Royal Society of Edinburgh Transactions: Earth Sci- Todd, V.R., and Shaw, S.E., 1985, S-type granitoids and an I-S line in the Pen-
ences, v. 79, p. 105–121. insular Ranges batholith, southern California: Geology, v. 13, p. 231–233,
Smith, D.P., and Busby, C.J., 1993, Mid-Cretaceous crust extension recorded doi: 10.1130/0091-7613(1985)13<231:SGAAIL>2.0.CO;2.
in deep marine half-graben fill, Cedros Island, Mexico: Geological Todd, V.R., Erskine, B.G., and Morton, D.M., 1988, Metamorphic and tectonic
Society of America Bulletin, v. 105, p. 547–562, doi: 10.1130/0016- evolution of the northern Peninsular Ranges Batholith, southern Cali-
7606(1993)105<0547:MCCERI>2.3.CO;2. fornia, in Ernst, W.G., ed., Metamorphism and Crustal Evolution of the
Sorensen, S.S., 1985, Petrologic evidence for Jurassic island arc-like basement Western United States (Rubey Volume 7): Englewood Cliffs, New Jersey,
rocks in the southwestern Tranverse ranges and California continental Prentice Hall, p. 894–937.
borderland: Geological Society of America Bulletin, v. 96, p. 997–1006, Todd, V.R., Shaw, S.E., and Hammarstrom, J.M., 2003, Cretaceous plutons of
doi: 10.1130/0016-7606(1985)96<997:PEFJIB>2.0.CO;2. the Peninsular Ranges Batholith, San Diego and westernmost Imperial
Sorensen, S.S., 1986, Petrologic and geochemical comparison of the blueschist Counties, California: Intrusion across a Late Jurassic continental margin,
and greenschist units of the Catalina Schist terrane, southern California, in Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
in Evans, B.W., and Brown, E.H., eds., Blueschists and Eclogites: Geo- and the Southwestern USA: Geological Society of America Special Paper
logical Society of America Memoir 164, p. 59–75. 374, p. 185–235.
Sorensen, S.S., 1988a, Tectonometamorphic significance of the basement rocks Tulloch, A.J., and Kimbrough, D.L., 2003, Paired plutonic belts in convergent
of the Los Angeles Basin and the California Continental Borderland, in margins and the development of high Sr/Y magmatism: Peninsular Ranges
Ernst, W.G., ed., Metamorphism and Crustal Evolution of the Western Batholith of Baja California and Median Batholith of New Zealand, in
United States (Rubey Volume 7), p. 998–1022. Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico
Sorensen, S.S., 1988b, Petrology of amphibolite-facies mafic and ultramafic and the Southwestern USA: Geological Society of America Special Paper
rocks from the Catalina Schist, southern California: Metasomatism and 374, p. 275–295.
migmatization in a subduction zone metamorphic setting: Journal of Usui, T., Nakamura, E., Kobayashi, K., Maruyama, S., and Helmstaedt, H.,
Metamorphic Geology, v. 6, p. 405–435, doi: 10.1111/j.1525-1314.1988. 2003, Fate of the subducted Farallon plate inferred from eclogite xenoliths
tb00431.x. in the Colorado Plateau: Geology, v. 31, p. 589–592, doi: 10.1130/0091-
Sorensen, S.S., and Barton, M.D., 1987, Metasomatism and partial melting in a 7613(2003)031<0589:FOTSFP>2.0.CO;2.
subduction complex: Catalina Schist, southern California: Geology, v. 15, Vedder, J.G., 1987, Regional geology and petroleum potential of the southern
p. 115–118, doi: 10.1130/0091-7613(1987)15<115:MAPMIA>2.0.CO;2. California Borderland, in Scholl, D.W., et al., eds., Geology and Resource
Sorensen, S.S., and Grossman, J.N., 1989, Enrichment of trace elements in gar- Potential of the Continental Margin of Western North America and Adja-
net amphibolites from a paleo-subduction zone: Catalina Schist, southern cent Ocean Basins: Beaufort Sea to Baja California: Houston, Circum-
California: Geochimica et Cosmochimica Acta, v. 53, p. 3155–3177, doi: Pacific Council for Energy and Mineral Resources, p. 403–447.
10.1016/0016-7037(89)90096-3. Vedder, J.G., Howell, D.G., and McLean, H., 1983, Stratigraphy, sedimenta-
Staude, J.M., and Barton, M.D., 2001, Jurassic to Holocene tectonics, mag- tion, and tectonic accretion of exotic terranes, southern Coast Ranges,
matism, and metallogeny of northwestern Mexico: Geological Soci- California, in Watkins, J.S., and Drake, C.L., eds., Studies in Continental
ety of America Bulletin, v. 113, p. 1357–1374, doi: 10.1130/0016- Margin Geology: American Association of Petroleum Geologists Memoir
7606(2001)113<1357:JTHTMA>2.0.CO;2. 34, p. 471–496.
Stern, R.J., and Bloomer, S.H., 1992, Subduction zone infancy: Examples from von Huene, R., and Scholl, D.W., 1991, Observations at convergent margins
the Eocene Izu-Bonin-Mariana and Jurassic California arcs: Geological concerning sediment subduction, subduction erosion, and the growth of
Society of America Bulletin, v. 104, p. 1621–1636, doi: 10.1130/0016- continental crust: Reviews of Geophysics, v. 29, p. 279–316.
7606(1992)104<1621:SZIEFT>2.3.CO;2. von Huene, R., and Scholl, D.W., 1993, The return of sialic material to the mantle
Stern, T., Bateman, P.C., Morgan, B.A., Newell, M.F., and Peck, D.L., 1981, Iso- indicated by terrigeneous material subducted at convergent margins: Tec-
topic U-Pb Ages of Zircon from the Granitoids of the Central Sierra Nevada, tonophysics, v. 219, p. 163–175, doi: 10.1016/0040-1951(93)90294-T.
California: U.S. Geological Survey Professional Paper 1185, 17 p. Wakabayashi, J., 1990, Counterclockwise P-T-t paths from amphibolites, Fran-
Stock, J., and Molnar, P., 1988, Uncertainties and implications of the Late Cre- ciscan Complex, California: Relics from the early stages of subduction
taceous and Tertiary position of North America relative to the Farallon, zone metamorphism: Journal of Geology, v. 98, p. 657–680.
Kula, and Pacific plates: Tectonics, v. 7, p. 1339–1384. Wakabayashi, J., 1992, Nappes, tectonics of oblique plate convergence, and met-
Stuart, C.J., 1979, Middle Miocene paleogeography of coastal southern Cali- amorphic evolution related to 140 million years of continuous subduction,
fornia and the California borderland: Evidence from schist-bearing sedi- Franciscan Complex, California: Journal of Geology, v. 100, p. 19–40.
mentary rocks, in Armentrout, J.M., et al., eds., Cenozoic Paleogeography Wakabayashi, J., 1999, Subduction and the rock record: Concepts developed
of the Western United States: Society of Economic Paleontologists and in the Franciscan Complex, California, in Sloan, D., et al., eds., Clas-
Mineralogists, Pacific Section, Field Guide 3, p. 29–44. sic Cordilleran Concepts: A View from California: Geological Society of
Sundberg, F.A., and Cooper, J.D., 1978, Late Cretaceous depositional envi- America Special Paper 338, p. 123–133.
ronments, northern Santa Ana Mountains, Southern California, in How- Wakabayashi, J., and Dilek, K., 2000, Spatial and temporal relations between
ell, D.G., and McDougall, K.A., eds., Mesozoic Paleogeography of the ophiolites and their subophiolitic soles: A test of models of forearc ophio-
Western United States: Society of Economic Paleontologists and Min- lite genesis, in Dilek, Y., et al., eds., Ophiolites and Oceanic Crust: New
eralogists, Pacific Section, Pacific Coast Paleogeography Symposium 2, Insights from Field Studies and Ocean Drilling: Geological Society of
p. 535–546. America Special Paper 349, p. 53–64.
Suppe, J., 1970, Offset of late Mesozoic basement terrains by the San Andreas Wakabayashi, J., and Dumitru, T.A., 2007, 40Ar/39Ar ages from coherent high-
fault system: Geological Society of America Bulletin, v. 81, p. 3253– pressure metamorphic rocks of the Franciscan Complex, California:
3258, doi: 10.1130/0016-7606(1970)81[3253:OOLMBT]2.0.CO;2. Revisiting the timing of metamorphism of the world’s type subduction
Suppe, J., and Armstrong, R.L., 1972, Potassium-argon dating of Franciscan complex: International Geology Review, v. 49, no. 10, p. 873–906.
metamorphic rocks: American Journal of Science, v. 272, p. 217–233. Walawender, M.J., Gastil, R.G., Clinkenbeard, J.P., McCormick, W.V., East-
Taylor, H.P., 1986, Igneous rocks: Isotopic case studies of Circum-Pacific mag- man, B.G., Wernicke, R.S., Wardlaw, M.S., Gunn, S.H., and Smith, B.M.,
matism, in Valley, J.W., et al., eds., Stable Isotopes in High-Temperature 1990, Origin and evolution of the zoned La Posta–type plutons, eastern
Geological Processes: Mineralogical Society of America, p. 273–317. Peninsular Ranges batholith, southern and Baja California, in Anderson,
ten Brink, U.S., Zhang, J., Brocher, T.M., Okaya, D.A., Klitgord, K.D., and Fuis, J.L., ed., The Nature and Origin of Cordilleran Magmatism: Geological
G.S., 2000, Geophysical evidence for the evolution of the California inner Society of America Memoir 174, p. 1–18.
continental borderland as a metamorphic core complex: Journal of Geo- Weaver, D.W., and Nolf, B., 1969, Geology of Santa Cruz Island (map), in
physical Research, v. 105, p. 5835–5857, doi: 10.1029/1999JB900318. Weaver, D.W., ed., Geology of the Northern Channel Islands, Southern
Thomson, C.N., and Girty, G.H., 1994, Early Cretaceous intra-arc ductile California Borderland: American Association of Petroleum Geologists,
strain in Triassic–Jurassic and Cretaceous continental margin arc rocks, Pacific Section, Special Publication, scale 1:24,000.
The Catalina Schist 361

Wetmore, P.H., Schmidt, K.L., Paterson, S.R., and Herzig, C., 2002, Tectonic Wright, J.E., and Fahan, M.R., 1988, An expanded view of Jurassic orogenesis
implications for the along-strike variation of the Peninsular Ranges in the western United States Cordillera: Middle Jurassic (pre-Nevadan)
batholith, southern and Baja California: Geology, v. 30, p. 247–250, doi: regional metamorphism and thrust faulting within an active arc environ-
10.1130/0091-7613(2002)030<0247:TIFTAS>2.0.CO;2. ment, Klamath Mountains, California: Geological Society of America
Wetmore, P.H., Herzig, C., Schultz, P.W., Alsleben, H., Paterson, S.R., and Bulletin, v. 100, p. 859–876, doi: 10.1130/0016-7606(1988)100<0859:
Schmidt, K.L., 2003, Cretaceous tectonic evolution of the western zone AEVOJO>2.3.CO;2.
of the Peninsular Ranges batholith of southern and Baja California, in Wright, T.L., 1991, Structural geology and tectonic evolution of the Los Ange-
Johnson, S.E., et al., eds., Tectonic Evolution of Northwestern Mexico les Basin, California, in Biddle, K.T., ed., Active Margin Basins: Ameri-
and the Southwestern USA: Geological Society of America Special Paper can Association of Petroleum Geologists Memoir 52, p. 35–134.
374, p. 93–116. Yeats, R.S., 1968, Southern California structure, seafloor spreading, and history
Williams, I.S., and Claesson, S., 1987, Isotopic evidence for the Precambrian of the Pacific Basin: Geological Society of America Bulletin, v. 79, p. 1693–
provenance and Caledonian metamorphism of high grade paragneisses 1702, doi: 10.1130/0016-7606(1968)79[1693:SCSSSA]2.0.CO;2.
from the Seve Nappes, Scandinavian Caledonides: Contributions to Min- Yeats, R.S., 1973, Newport-Inglewood fault zone, Los Angeles basin, Cali-
eralogy and Petrology, v. 97, p. 205–217, doi: 10.1007/BF00371240. fornia: American Association of Petroleum Geologists Bulletin, v. 57,
Wood, D.J., and Saleeby, J.B., 1997, Late Cretaceous–Paleocene extensional p. 117–135.
collapse and disaggregation of the southernmost Sierra Nevada batholith: Yerkes, R.F., McCulloh, T.H., Schoellhamer, J.E., and Vedder, J.G., 1965, Geol-
International Geology Review, v. 39, p. 973–1009. ogy of the Los Angeles Basin, California—An Introduction: U.S. Geo-
Wooden, J.L., and Miller, D.M., 1990, Chronologic and isotopic framework for logical Survey Professional Paper 420-A, 57 p.
Early Proterozoic evolution of the eastern Mojave Desert region, SE Cali- Zandt, G., Gilbert, H., Owens, T.J., Ducea, M., Saleeby, J., and Jones, C.H.,
fornia: Journal of Geophysical Research, v. 95, p. 20133–20146. 2004, Active foundering of a continental arc root beneath the southern
Woodford, A.O., 1924, The Catalina Metamorphic Facies of the Franciscan Sierra Nevada in California: Nature, v. 431, p. 41–46, doi: 10.1038/
series: University of California Publications in Geological Sciences, nature02847.
v. 15, p. 49–68.
Woodring, W.P., and Popenoe, W.P., 1942, Upper Cretaceous formations and
faunas of southern California: American Association of Petroleum Geolo-
gists Bulletin, v. 26, p. 166–176.
Woodring, W.P., Bramlette, M.N., and Kew, W.S.W., 1946, Geology and Pale-
ontology of the Palos Verdes Hills, California: U.S. Geological Survey
Professional Paper 207, 145 p. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Sedimentary response to arc-continent collision, Permian, southern Mongolia

C.L. Johnson*
Geology and Geophysics, WBB 609, University of Utah, 135 South 1460 East, Salt Lake City, Utah 84112, USA

J.A. Amory
Salym Petroleum Development, Griboedova Street #2, 625000 Tyumen, Russia

D. Zinniker
Geology and Geophysics, Yale University, P.O. Box 2080109, New Haven, Connecticut 06520, USA

M.A. Lamb
Geology Department, OWS 153, University of St. Thomas, 2115 Summit Avenue, St. Paul, Minnesota 55105, USA

S.A. Graham
Geological and Environmental Sciences Department, Stanford University, 450 Serra Mall (Building 320) #118, Stanford,
California 94305, USA

M. Affolter
Geology and Geophysics, WBB 609, University of Utah, 135 South 1460 East, Salt Lake City, Utah 84112, USA

G. Badarch†
Institute of Geology and Mineral Resources, Mongolian Academy of Sciences, Ulaanbaatar 210351, Mongolia

ABSTRACT

The Eurasian Tien Shan–Yin Shan suture is a ~3000-km-long boundary between


Paleozoic arc and accretionary complexes (the Altaids) and Precambrian microcon-
tinental blocks (Tarim and North China block). Stratigraphic data are presented
from localities in southern Mongolia spanning >800 km along the northern margin
of the suture. Facies descriptions, climatic indicators, sandstone provenance, and
paleocurrent data help reconstruct Permian basin evolution during and following
arc-continent collision, and results are integrated with previously published data to
create a preliminary regional synthesis. Upper Permian strata of southern Mongolia
comprise fluvial successions in the southwest and marine turbidite deposits in the
southeast. Floral assemblages show mixing of Siberian craton and North China block
communities, indicating their close proximity to Mongolia by Permian time. There
was a rapid transition from humid environments in the Late Permian to more arid

*cjohnson@earth.utah.edu

Deceased.

Johnson, C.L., Amory, J.A., Zinniker, D., Lamb, M.A., Graham, S.A., Affolter, M., and Badarch, G., 2008, Sedimentary response to arc-continent collision,
Permian, southern Mongolia, in Draut, A.E., Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record in Arc Collision Zones:
Geological Society of America Special Paper 436, p. 363–390, doi: 10.1130/2008.2436(16). For permission to copy, contact editing@geosociety.org. ©2008 The
Geological Society of America. All rights reserved.

363
364 Johnson et al.

conditions in the Early Triassic, corresponding to the global Permian–Triassic bound-


ary event, but which may also reflect more local driving mechanisms such as rain
shadow effects. Permian sandstones from Mongolia have undissected to dissected arc
provenance, with little contribution from continental or recycled orogen sources. The
timing of the nonmarine-marine facies transition and cessation of arc magmatism
broadly supports earlier collision along the western part of the suture zone than along
the eastern part (e.g., Late Carboniferous–Late Permian). However, when regional
geologic constraints are integrated, a more complex model involving differential rota-
tion of Tarim and the North China block is preferred. Late Paleozoic rocks of southern
Mongolia have been subsequently dismembered along Mesozoic–Cenozoic strike-slip
faults, and thus also represent the long-term record of intracontinental deformation
within accreted, heterogeneous crust.

Keywords: collisional tectonics, arc accretion, palynology, sedimentary basins,


Mongolia.

INTRODUCTION the timing, mechanisms, and sedimentary record of Mesozoic and


younger collisional events in this region (Graham et al., 1993;
Continental growth via accretion is well illustrated by the Graham, 1996; Yin and Nie, 1996). However, the suture zone
Cenozoic India-Eurasia collision (Harrison et al., 1992; Rowley, that formed the main nucleation point for Mesozoic accretion
1996). This phenomenon broadly represents the entire Phanero- remains poorly understood; specifically, the boundary between
zoic history of the region, as demonstrated by a series of generally Paleozoic arc terranes that lie south of the Siberian craton and
southward-younging suture zones extending from the Siberian tectonic blocks in China that are floored mainly by Precambrian
craton through Mongolia and China (Zhang et al., 1984; Fig. 1). continental crust (e.g., North China block and Tarim–Central Tien
Such rapid continental growth not only highlights an exceptional Shan blocks; Watson et al., 1987). This accretionary complex (the
protracted record of terrane accretion but also illustrates sub- Altaid tectonic collage or “Altaids” after Şengör and Natal’in,
sequent intracontinental deformation along inherited zones of 1996, Fig. 1; cf. the Central Asian Orogenic Belt of Jahn, 1999) is
crustal weakness, even far inboard of the active margins (Tappon- well exposed in the study area of southern Mongolia. As such, this
nier et al., 1982; Burchfiel et al., 1989). Multiple studies highlight region constitutes a significant portion of one of the largest zones

Figure 1. Map of major tectonic units,


boundaries, and fault zones in east-
central Asia. ATF—Altyn Tagh fault;
BS—Bogda Shan; CTS—Central Tien
Shan; EGFZ—East Gobi fault zone;
JH—Junggar-Hegen suture; NTS—
North Tien Shan; QFS—Qilian fault sys-
tem; SIB—Siberian craton; THB—Tur-
pan-Hami basin; TLF—Tan Lu fault;
UB—Ulaanbaatar (capital of Mongolia).
Sedimentary response to arc-continent collision 365

of arc accretion on earth (Şengör et al., 1993), so there is much to highlight outstanding questions related to regional tectonic evo-
learn about arc-arc collision as well as terminal arc-continent col- lution. Complementary data are then presented from seven main
lision, the subject of this paper. Few integrated studies of Permian localities spanning >800 km from southwestern to southeastern
rocks in southern Mongolia are available, so this study is a first Mongolia (Table 1; Fig. 2). New data and observations are derived
step toward documenting that record. primarily from Upper Permian accretionary and postaccretionary
The boundary between the Altaids and Precambrian micro- sedimentary complexes. Considered in the regional context, these
continental blocks of north China has been variously named the data help constrain the timing, mechanisms, and implications of
Suolon, Solonker, Junggar-Hegen, South Mongolia–Hinggan, this early phase of continental growth in Asia. Similarly, the arc
and Tien Shan–Yin Shan sutures (Şengör et al., 1993; Wang and setting of southern Mongolia in the middle to late Paleozoic has
Mo, 1995; Badarch et al., 2002; Cope et al., 2005; Shi, 2006). been compared to the modern southwest Pacific Ocean (Lamb
The term Tien Shan–Yin Shan suture is favored because it implies and Badarch, 2001), and additional field constraints permit eval-
a regional continuation of the collision zone from northwest to uating this analogy and various models that describe the strati-
northeast China (Fig. 1; Yin and Nie, 1996; Xiao et al., 2004), graphic record of interarc and arc-continent collisions (Ingersoll
a correlation that is supported by this study. The geology of this et al., 1995; Huang et al., 2006).
region is poorly defined on many levels beyond these problem-
atic naming conventions. The latitude and extent of the suture LATE PALEOZOIC SEDIMENTATION AND
zone relative to the China-Mongolia political border is debated TECTONICS OF NORTH CHINA
(Xiao et al., 2003). Assuming that collision involved a combined
Tarim–North China block moving north relative to the Altaids The southern margin of the Altaids is inferred to follow
(this assumption is discussed more fully below; Enkin et al., either a strongly arcuate shape roughly parallel to the modern
1992; Ziegler et al., 1996), the Tien Shan–Yin Shan suture spans political border, or alternatively a more continuous east-west–
>3000 km from northwest to northeast China, making it one of trending boundary across northern China (Fig. 1, Junggar-
the longest and oldest tectonic boundaries within the amalgam- Hegen versus Tien Shan–Yin Shan sutures; Zhang et al., 1984;
ated crustal material south of the Siberian craton. Yin and Nie, 1996). Tarim and the North China block represent
Formation of this suture during closure of the Paleo-Asian continental margins with Precambrian crust, forming distinctive
ocean (Dobretsov et al., 1994) is generally thought to have been southern boundaries to the accretionary complex. In contrast,
complete by Late Permian to Early Triassic time, but alternate the Altaid tectonic collage primarily comprises late Paleozoic
interpretations range from Devonian to Cretaceous (reviewed in subduction-accretion complexes with probably only small and
Xiao et al., 2003). Timing relationships across the region have isolated microcontinent blocks (Şengör et al., 1993). Confusion
also been interpreted as either broadly synchronous or highly dia- over the location of the suture that marks final closure of the
chronous collision from west to east (Amory, 1996; Yin and Nie, Paleo-Asian ocean is exacerbated by this rather diffuse northern
1996). Although many studies provide important constraints for margin, which consists of several magmatic arc and ophiolite
parts of the suture zone (particularly in northwest China; Cole- zones, and thus potentially multiple terrane boundaries.
man, 1989; Windley et al., 1990; Laurent-Charvet et al., 2003),
in general the paleogeography of the entire closing ocean basin Northwest China Tectonostratigraphy
is poorly understood in terms of polarity of subduction, and the
number and nature of magmatic arcs that may have been active Tarim and the North China block bound the Altaids south
(Lamb and Badarch, 2001; Zhou et al., 2001; Xiao et al., 2004). of Mongolia (Fig. 1). These two microcontinents are inferred
Without a better understanding of the location, timing, and by many to have amalgamated prior to the Ordovician (Yin and
mechanisms associated with this event, it is difficult to evaluate Nie, 1996; Heubeck, 2001). Similar middle to upper Paleozoic
implications such as whether postorogenic collapse could have
been active in the Permian (Tang, 1990) or later into the Trias-
sic (Davis et al., 2004), whether later intracontinental deforma- TABLE 1. LATITUDE AND LONGITUDE COORDINATES FOR
tion has been concentrated along major crustal boundaries, and KEY STUDY AREAS*
whether regional climate change in the Permian of China can be Shin Jinst N 44° 30', E 99° 18'
Noyon Uul N 43° 12', E 101° 54'
attributed to a rain shadow effect created by collisional orogens
N 43° 09', E 102° 04'
(Cope et al., 2005).
Tavan Tolgoy N 43° 33', E 105° 31'
This study reviews known, inferred, and newly constrained Naftgar Uul N 43° 14', E 105° 12'
relationships relevant to the tectonic and stratigraphic history on Tsaagan Tolgoy N 42° 53', E 105° 27'
either side of the nominal collision zone in order to extract a new Nomgon N 42° 57', E 108° 20'
model regarding its formation. Reconstruction of the Tien Shan– N 42° 46', E 108° 29'
Yin Shan suture is aided by decades of research in northern China. Hovsgol N 43° 35', E 109° 31'
We first review these studies, with an emphasis on available basin N 43° 37', E 109° 47'
analysis constraints, to provide a regional tectonic context and to *See Figure 2.
366 Johnson et al.

Shin Jinst Faults associated with the Permian granite Ordovician-Silurian sequences,
East Gobi fault zone undifferentiated
Mesozoic-Cenozoic rocks, Permian sedimentary
undifferentiated and volcanic rocks Rocks mapped as Precambrian,
Devonian-Carboniferous arc including carbonate klippen and
44º N Triassic sedimentary rocks sequences, undifferentiated possible Mesozoic tectonites

in
b as
o bi Hovsgol
Tavan Tolgoy tG
s
Noyon Uul Ea
Naftgar Uul
Nomgon
Tsaagan
China Tolgoy
border-Mongolia

42º N Onch
Hayrhan Bulgan Uul N
99º E

108º E
0 50 100 km
102º E

105º E
Southwest localities South-central localities Southeast localities

Figure 2. Geologic map of southern Mongolia, showing main study areas. This “timeslice” map is generalized and modified from Tomurtogoo
(1999) and Yanshin (1989) and highlights Triassic and older rocks most relevant to this study. See Table 1 for latitude and longitude coordinate
locations of the study areas.

stratigraphy on either side of the Qilian Shan supports this early 1984; Watson et al., 1987). Although crustal thickness throughout
assembly (Yang et al., 2001; Ritts et al., 2004), and no clear northwest China ranges up to 50 km, this is mainly attributed to
Phanerozoic suture has been found between Tarim and the North post-Paleozoic sedimentary basin deposits and Cenozoic crustal
China block. On the contrary, paleomagnetic data indicate sepa- thickening owing to India-Asia collision (Wang et al., 2003). No
rate apparent polar wander paths throughout the late Paleozoic, Precambrian rocks have been found surrounding the Junggar
and such reconstructions imply that the North Tien Shan and Yin basin, in contrast to widespread Paleozoic ophiolite complexes
Shan segments of the suture zone formed at about the same time and mafic igneous rocks that are present (Ren et al., 1987; Cole-
(Carboniferous–Permian) but at very different latitudes (McEl- man, 1989; Wang et al., 2003). Thus the Junggar basin does not
hinny et al., 1981; Li et al., 1988; Enkin et al., 1992; Zhao et al., seem to be underpinned by Precambrian continental crust on the
1996). Floral distributions generally contradict this interpreta- basis of multiple lines of evidence, and the area is better grouped
tion, instead indicating that Tarim and the North China block into Altaid arc affinities (Şengör et al., 1993).
were at similar latitudes in the Late Permian (Ziegler et al., Collision between the combined Tarim–Central Tien Shan
1996). Paleozoic paleomagnetic poles from Tarim may be over- and the North Tien Shan–Bogda Shan Arcs occurred between
printed, not fully reconciled with Mesozoic–Cenozoic shorten- Late Carboniferous and Early Permian time (ca. 300–280 Ma),
ing and vertical axis rotations, and/or compared with problem- based on radiometric dating of arc-related igneous rocks (Fig. 1;
atic Eurasian reference poles (Gilder et al., 1996; Roger et al., Yin and Nie, 1996; Dumitru et al., 2001; Zhou et al., 2001). Sedi-
2003). With respect to late Paleozoic collision with the Altaids, mentary basin analyses provide further constraints on formation
it seems that Tarim and the North China block can be treated as of the North Tien Shan. The youngest marine rocks directly north
constituting a northward-migrating Andean-style margin, but it of Tarim (in the Junggar basin) are Upper Carboniferous–Lower
is important to note that the paleogeography is poorly known Permian turbidites and shallow-marine sandstone, mudstone, and
and controversial. carbonates (Carroll et al., 1995). Overlying Permian deposits are
Amalgamation of northwest China occurred in two main nonmarine, with at least 3000 m of upward-coarsening succes-
phases, beginning with collision of Tarim and the Central Tien sions ranging from lacustrine mudstone to coarse alluvial-fluvial
Shan by Early Carboniferous time (Fig. 1; Carroll et al., 1990; conglomerate (Carroll et al., 1990).
Windley et al., 1990; Allen et al., 1993; Zhou et al., 2001; Wang Apparent climatic shifts accompany and postdate collision
et al., 2005). The second phase of amalgamation in northwest in the Junggar and Turpan basins (connected in the Permian–
China is inferred to represent the southern boundary of the Altaids Triassic, Hendrix et al., 1992; Carroll et al., 1995). There is a
in northwest China and thus is of primary interest for this study. Late Permian transition from organic-rich, wood- and coal-bear-
Alternate interpretations place the boundary to the north of the ing fluvial and lacustrine rocks associated with relatively humid
Junggar basin (e.g., Junggar-Hegen suture, Fig. 1; Zhang et al., climates, to braided fluvial red beds with common desiccation
Sedimentary response to arc-continent collision 367

features (Carroll, 1991; Wartes et al., 2002). Arid conditions Solonker (e.g., Yin Shan) suture in northeast China, with dual-
persisted from the latest Permian to the Early Triassic, whereas polarity subduction beneath two continental margins (the North
coal-bearing, wet-climate rocks were once again widespread China block to the south and a Devonian continental margin to
in the Middle to Late Triassic (Hendrix et al., 1992; Greene et the north), followed by Himalayan-style, final-stage continental
al., 2001). Along with paleocurrent and provenance data, these collision of these margins at the end of the Permian. Although
lithologic and climatic changes have been attributed to syndepo- few paleomagnetic data are available for Carboniferous–Permian
sitional deformation during uplift of the North Tien Shan in the rocks of the North China block (cf. Huang et al., 2001), Mesozoic
Early Permian and continued deformation and partitioning of the paleogeographic reconstructions suggest that Mongolian arc ter-
flexural Junggar-Turpan-Hami basin from the Late Permian into ranes had accreted to the North China block by the Late Permian,
the Mesozoic (Hendrix, 2000; Greene et al., 2001; Vincent and and as a combined block they rotated counterclockwise relative
Allen, 2001). Carroll et al. (1992) also linked evidence for saline to the Siberian craton from the Triassic to the Cretaceous during
versus fresh-water depositional conditions in Permian lacustrine closure of the Mongol-Okhotsk ocean (Fig. 1; Enkin et al., 1992;
oil shale to an orographic effect, driven by increased precipitation Heubeck, 2001). As noted for paleomagnetic data from northwest
runoff adjacent to the North Tien Shan. China, it seems possible that this rotation began during or prior to
The Junggar basin likely originated in an arc setting (back-, Permian collision.
intra- or interarc), and by the late Paleozoic it was a trapped ocean Along the North China block boundary in the Yin Shan
basin, evolving into a flexural foreland with ongoing collisional Range (Fig. 1), >1000 m of Upper Carboniferous–Upper(?)
deformation in the Mesozoic (Allen et al., 1991; Carroll et al., Permian sedimentary rocks were deposited in braided fluvial
1990; Vincent and Allen, 2001). Sediment transport in the Perm- to meandering fluvial and floodplain environments (Mueller et
ian Junggar trapped ocean basin is inferred to have been toward al., 1991; Cope et al., 2005). The youngest marine rocks in this
the west-northwest, as indicated by regional paleocurrent data region are Ordovician shallow-marine carbonates that uncon-
(Carroll et al., 1995). This may reflect closure of a wedge-shaped formably underlie the Carboniferous and a younger nonmarine
ocean that opened toward the west, perhaps accommodated by sequence. Significant changes occur within the upper Paleozoic
clockwise rotation of Tarim and the Central Tien Shan during stratigraphy in northeast China. Coeval with a shift from braided
this time. This interpretation is consistent with published remnant fluvial to meandering fluvial deposition, paleocurrents reversed
ocean-basin models, where sediment is dispersed outboard of the from north-directed dispersal in Carboniferous–Lower Permian
migrating suture zone and mainly parallel to colliding margins strata to south-directed dispersal in Upper Permian–Lower Tri-
(Ingersoll et al., 1995). However, it may contradict the broader assic(?) rocks (Cope et al., 2005). Sandstone petrography indi-
interpretation of the entire Tien Shan–Yin Shan suture having cates a recycled orogen to dissected arc provenance, representing
closed progressively from west to east (Nie et al., 1990; Muel- unroofing of a thin sedimentary cover into continental arc plutons
ler et al., 1991; Amory et al., 1994). Other geologic evidence during the Late Permian (Cope et al., 2005).
for Permian block rotation in northwest China includes dextral The late Paleozoic succession in northeast China is also
strike-slip shear zones in the North Tien Shan with synkinematic marked by a climatic shift from humid-environment, coal-bearing
biotite 40Ar/39Ar ages of 290–245 Ma (Laurent-Charvet et al., deposits in the Carboniferous–Early Permian to widespread arid-
2003). There is also paleomagnetic evidence for continued move- climate, red bed deposition by Late Permian time (Mueller et al.,
ment of Tarim with respect to Siberia since the Late Permian (Li 1991). Cope et al. (2005) suggest that the climatic shift represents
et al., 1988; Gilder et al., 1996), so it seems permissible that rota- a rain shadow effect created during uplift of the continental mar-
tion could have begun during late Paleozoic collision, coincident gin owing to collision, and/or northward movement of northern
with dextral shear in the North Tien Shan. China to arid subtropical latitudes during this time. Uplift and sig-
nificant crustal thickening along this margin during the Permian
Northeast China Tectonostratigraphy may have been further supported by Late Triassic–Early Jurassic
(ca. 220–190 Ma) collapse of the orogen on the basis of radiomet-
The broader interpretation of west-to-east closure across the ric ages of rocks involved in metamorphic core complex forma-
entire Tien Shan–Yin Shan suture zone is based mainly on evi- tion in northeast China (Davis et al., 2004).
dence for younger collision along the Yin Shan belt (northeast
China) ~2000 km east of the North Tien Shan (Fig. 1; Amory et al., Summary
1994). Cope et al. (2005) interpreted active arc magmatism from
ca. 400–275 Ma via detrital zircon ages from Permian sandstones. Multiple lines of evidence indicate that collision between the
This is further constrained by a geochemical shift to inferred post- southern Altaids and microcontinent blocks of north China (Tarim
collisional granites younger than ca. 257 Ma (Chen et al., 2000). and the North China block) occurred during Carboniferous to
Dissected continental arcs are found south of the Mongolian bor- Permian time. More specifically, suturing appears to have occurred
der in the Yin Shan highlands, indicating south-directed subduc- earlier in the Tien Shan than in the Yin Shan Ranges, suggesting
tion in the region through Late Permian time. Alternatively, Xiao diachronous west-to-east closure. However, available radiometric
et al. (2003) define both a northern and southern boundary to the data indicate that arc activity was at least partly overlapping in
368 Johnson et al.

northeast and northwest China, and the transition from nonmarine likely close to a continental margin or highly dissected sediment
to marine deposition is not well preserved across the length of the source to the north (Badarch et al., 2002). The Carboniferous arc
suture zone. Paleomagnetic data generally support the possibility shifted slightly south, closer to the present-day China-Mongolia
of differential rotation of Tarim and the North China block during border, and is at least partly built on Devonian strata. Carbon-
the late Paleozoic, and stratigraphic studies suggest that the Jung- iferous sandstones suggest a possible granitic source, although
gar trapped ocean basin may have opened toward the west rather this may represent uplift and dissection of the Devonian arcs
than having been contiguous to the east. Thus the prospect of a rather than proximity to a microcontinent (Lamb, 1998). Both
more complex collisional event, with Tarim and the North China systems have analogous components across the border and likely
block acting somewhat independently and perhaps separated by extended west into the Bogda Shan–North Tien Shan arc sys-
an indenter (Lamb and Badarch, 2001), is intriguing and perhaps tem (Fig. 1; Lamb and Badarch, 2001), albeit with some changes
supported by our new data from Mongolia. Another important but toward more oceanic arc affinities to the southwest. Similarly, the
poorly constrained aspect of this collision is what, if any, climate system is tentatively linked through southeast Mongolia and into
response occurred during and after arc-continent collision, and northeast China (Xiao et al., 2003). These arcs were likely more
whether this might have been driven by orographic relief along complicated regionally, but the simpler interpretation reflects the
the suture zone. Rain shadow and other topographic influences resolution afforded by available data (Lamb and Badarch, 2001).
driven by collisional and postcollisional tectonics have been cited This interpretation is also more representative of continuous arc
to explain observed stratigraphic changes in Permian basins of systems, such as the modern Philippine Arc or other parts of the
northwest and northeast China, but this is only one of multiple southwestern Pacific Ocean, where arc type, geometry, and sub-
possible driving factors. duction polarity are variable along strike but nevertheless are part
Outstanding questions that we now seek to address include of the same geodynamic setting.
whether the Permian of Mongolia supports or refutes the tim- Regional geologic maps of southern Mongolia show mul-
ing and location of this suture zone, as suggested by data from tiple Precambrian units, implying the existence of a microconti-
the continental margin to the south. The potential climatic effects nent along this part of the Tien Shan–Yin Shan suture, e.g., the
of this event are also investigated. Additional data from the South Gobi microcontinent (Fig. 2; Şengör et al., 1993; Xiao et
southern Altaids help to evaluate how well various modern ana- al., 2003). This interpretation is supported by a single U-Pb zir-
logues, including the southwest Pacific Ocean, may apply to this con age (916 ± 156 Ma) from gneiss on the Chinese side of the
setting. This ancient record addresses existing models of basin Yagan-Onch Hayrhan metamorphic core complex (Fig. 2; Wang
evolution, evolving sedimentary systems and source areas, and et al., 2001); however, protolith ages have not been tested rigor-
issues of preservation in comparison with modern analogues. ously across the region. High-grade gneiss-schist-mylonite com-
Finally, interpreting subsequent Mesozoic–Cenozoic deforma- plexes in southeastern Mongolia are mainly Mesozoic tectonites
tion requires understanding the pre- and syn-assembly record of (Webb et al., 1999; Webb and Johnson, 2006), and in many areas
southern Mongolia provided by this study. inferred protolith lithologies are more consistent with Paleozoic
arc successions than with continental crust (Johnson et al., 2001;
LATE PALEOZOIC SEDIMENTATION AND Webb and Johnson, 2006). In addition to metamorphic rocks,
TECTONICS OF SOUTHERN MONGOLIA Proterozoic marble-quartzite units are mapped throughout the
region. These are typically in flat, thrust-fault contact over Perm-
Overview ian and younger rocks and are inferred to be klippen of alloch-
thonous lower Mesozoic thrust sheets (Webb et al., 1999). Thus
The broad swath of land from southwestern to southeastern the existence of the South Gobi microcontinent is uncertain, and
Mongolia encompasses some 1000 km along the northern edge of in contrast to the Tarim and North China blocks the Precambrian
the Tien Shan–Yin Shan suture zone. Arc and accretionary com- crust likely forms a very minor component of the southern Altaids
plexes that define the Altaids are well exposed along basement in Mongolia, if at all.
uplifts in southern Mongolia (Fig. 2). Nevertheless, the number Although these rocks have not been described extensively
and nature of these intraoceanic arc systems is debated; the geo- in the English-language literature, Permian sequences in south-
logic complexity of the region is often represented by dozens of ern Mongolia are inferred to represent final closure of the Paleo-
individual tectonic “terranes” and intricate models for their colli- Asian ocean between Mongolian arc terranes and continental
sion and evolution (Zonenshain et al., 1990; Şengör et al., 1993; block(s) to the south (Amory, 1996). The timing of this terminal
Şengör and Natal’in, 1996; Badarch et al., 2002). collision is perhaps best constrained by Permian floral assem-
Based on geochemical data and sandstone provenance data blages that indicate mixing between polar-temperate, Siberian
from >20 sites in southern Mongolia, Lamb and Badarch (2001) (Angaran) communities and tropical, north China (Cathaysian)
proposed a simpler interpretation of two main evolving arc sys- communities (Durante, 1971, 1983; Shen et al., 2006). Similarly,
tems, Devonian and Carboniferous, followed by Permian closure interpreted syn- or postcollisional granites in northeast China
of the Paleo-Asian ocean. Devonian units are generally assigned have ca. 250–230 Ma U-Pb and Rb-Sr ages, supporting Late
to an island arc setting above a north-dipping subduction zone, Permian collision (Chen et al., 2000).
Sedimentary response to arc-continent collision 369

The ages of Permian rocks in Mongolia are generally only purposes of this study we retain the use of Upper-Late Permian
poorly known. Volcanic lithologies are uncommon except for the (Cisuralian, 299–270.6 Ma) versus Lower-Early Permian (Guada-
Early Permian (Fig. 3), and even for this interval have not been lupian–Lopingian, 270.6–253 Ma) terminology (Fig. 3; cf. Grad-
extensively dated by modern radiometric techniques. Invertebrate stein et al., 2004), as this is still widely used in regional literature
biostratigraphy for marine sections in southeastern Mongolia and provides the best means for suggesting regional correlations.
relies mainly on brachiopod zonations, some with identifica- Generalized stratigraphic correlations (Fig. 3) show broad
tion of bryozoans, crinoids, and foraminifers (Manankov, 1988, subdivisions for the Permian of southern Mongolia. Lower Perm-
2004). Nonmarine strata are constrained by plant fossils, inverte- ian red bed and rhyolite-andesite volcanic sequences are observed
brates, and spore-palynology where available. These forms tend across the region, typically in fault contact with Carboniferous–
to be low-resolution age constraints (>1 m.y.), and the correlation Devonian arc sequences. Overlying middle Permian strata include
from Russian “horizons” to the most recent global absolute time shallow-marine carbonate and clastic successions, which con-
scales (Gradstein et al., 2004) is just currently under way (Shen et formably underlie turbidite deposits of the Lugyn Gol Formation
al., 2006; Shi, 2006; Manankov et al., 2006). Thus it is difficult to (Fig. 3). A regional unconformity divides these flysch deposits
provide stage or even epoch assignments for many of the studied from the nonmarine overlap succession, which includes coal-bear-
intervals. Improving age resolution and correlations to a global ing fluvial strata that graded into braided fluvial successions in the
time scale are a focus of our ongoing work in Mongolia. For the Early Triassic (Fig. 3).

Standard Southwest South-central Southeast


Chronostratigraphy Mongolia Mongolia Mongolia
Triassic
250 and younger
Tavan Tolgoy and ?
Tsaagan Tolgoy Nonmarine Upper
~1600 m meandering Formations, ~100-300 m Permian units not
fluvial deposits, Noyon thick. Coal bearing fluvial Uncon- exposed.
Lopingian

255 Uul (Figs. 4-5). formity


deposits (Figs. 6-10).
?
Late Permian

Uncon-
260 formity
Lugyn Gol Formation,
Lugyn Gol Formation,
Guadalupian

? >500 m thick. Distal


>100 m thick (partially turbidite deposits (Figs.
265 preserved in 11, 12).
south-central Mongolia).
Distal turbidite deposits.
Uncon-
formity
270
Age (Ma)

Tavan Har and Har


? Agui Uul and Bulgan Uul Erdene Formations,
275 Formations, >3000 m >300 m thick. Shallow
thick. Shallow marine marine clastics, minor
Early Permian

carbonate and clastic carbonate and volcanic


deposits. units (Figs. 11, 12).
Ulaan Nur Formation,
280 ~300 m thick.
Rhyolitic-andesitic
volcanic rocks and
nonmarine clastic
Cisuralian

285 deposits (Figs. 4-5).


? Uncon-
contact formity
not
? exposed
290
Argalant Formation,
>100 m thick, red beds
and volcanic rocks. Argalant Formation,
>100 m thick, red beds
295 Carboniferous ? Unconformable and fault and volcanic rocks.
? sedimentary and volcanic contact with
C2 rocks thrust over Permian Carboniferous and older.
(Noyon Uul, Fig. 5E) or ? Unconformable and
P1 fault contact with
Carboniferous angular unconformity with
Carboniferous and older. Devonian.
and older

Figure 3. General chronostratigraphic correlation of Permian sedimentary and volcanic rocks from southwest to southeast Mongolia (cf. locali-
ties shown in Fig. 2). Lithologies are indicated by patterns as shown in Figure 4 (key). Formation names, approximate thicknesses, lithologies,
and contact relationships are compiled from our own work and published interpretations (Berkey and Morris, 1927; Grabau, 1931; Durante,
1971; Mossakovsky and Tomurtogoo, 1976; Zaitsev, 1974; Nikolov et al., 1981; Manankov et al., 2006). Dashed lines between sections represent
inferred lithostratigraphic correlations, which in some cases are interpreted to be time-transgressive. Time scale is from Gradstein et al. (2004),
although we have retained the use of Early versus Late Permian epochs, as discussed in the text. Age assignments are only approximate, given
the low resolution of available age constraints from Mongolia.
370 Johnson et al.

To fill in such broad correlations with more detailed field- Approximately 250 km southeast of Shin Jinst at Noyon
based constraints, we now present new data from a regional Uul (Fig. 2), ~1600 m of Upper Permian strata include inter-
southwest to southeast transect, highlighting localities where bedded sandstone and mudstone intervals that are also inter-
Permian strata are best preserved. The sections were visited preted to represent meandering fluvial environments (Fig. 4;
from 1992 to 2005, and so this is a long-term compendium of full section shown in Hendrix et al., 1996). Sandstone beds are
our studies. We present these observations at a reconnaissance lenticular over ~10 m (Fig. 5F) with minor erosional scour-
scale in order to emphasize regional relations; much of this inter- ing at their bases and are upward fining, with common trough
pretation is derived from more detailed descriptions in Amory and planar cross-bedding. Reworked terrestrial organic matter
(1996) and Lamb et al. (1999). This work focuses on the iden- is present throughout. The section is dominated by floodplain
tification of sedimentary facies to highlight basin setting and deposits of variegated silt and mudstone; carbonate nodules and
paleogeography, to constrain marine to nonmarine transitions, minor paleosols (argillosols) are also present (Fig. 5G, H). Age
and to identify shifting climatic and tectonic controls on sedi- control in this part of the section is based on well-preserved
mentation. Stratigraphic relations were identified on the basis of Late Permian flora and fauna (Zaitsev et al., 1973; Durante,
measuring and describing sections and, where possible, collect- 1976; other references in Hendrix et al., 1996). However, the
ing paleocurrent data. Climatic indicators are mainly inferred uppermost portion of the meandering fluvial succession has
from study of nonmarine sedimentary facies, particularly the been assigned an Early Triassic age, based upon tetrapod fossils
distinction between highly organic and coal-bearing strata that (Gubin and Sinitza, 1993). This generally fine-grained sequence
formed in wet to seasonally humid environments versus red is sharply overlain by thick Middle(?) Triassic conglomerates
beds, with common exposure and desiccation features indicative (Zaitsev et al., 1973; Hendrix et al., 2001). The coarse-grained
of aridity (Parrish et al., 1982; Wright, 1992; Dubiel and Smoot, Triassic–Jurassic units at Noyon Uul were deposited in an intra-
1994). Spores and other palynomorphs from the Tsaagan Tol- plate setting, likely within a flexural foreland basin (Hendrix
goy locality (Fig. 2) support these interpretations. Sedimentary et al., 1996). They clearly postdate the main collisional event
provenance, mainly derived from sandstone petrofacies, is also along the Tien Shan–Yin Shan suture but likely record ongoing
discussed as an additional constraint on basin tectonic setting intraplate contractile deformation.
and evolving source areas. The Late Permian floral assemblage at Noyon Uul is also
notable because it contains a mixture of Cathaysian (north China
Southwest Localities affinity) and Angaran (Siberian affinity) flora (Durante, 1971;
Amory, 1996). This is one of the early phases of population
Permian strata in southwest Mongolia (Fig. 2) lie uncon- mixing prior to complete homogenization in the Triassic, and is
formably on Lower Carboniferous and older marine sedimentary detected at localities in south-central Mongolia as well. In particu-
and volcanic units. Although Devonian–Carboniferous succes- lar, an unusual cycad genus (Guramsania) has been found at both
sions have distinct arc-related geochemistry, Permian strata are Noyon Uul and Tsaagan Tolgoy (Figs. 2, 6), providing further
nonmarine and are interbedded with basalts that have intraplate evidence for a connection between these localities (Vakrameyev
geochemical signatures (Lamb and Badarch, 2001). These rela- et al., 1986). Such floral mixing in the transition zone along the
tions support a postcollisional setting by Permian time. China-Mongolia border provides compelling evidence of close
At Shin Jinst (Fig. 4), Zonenshain et al. (1974) reported approach and/or collision with the North China block (Li, 2006;
Late(?) Permian plant fossils in a section of ~100 m of reddish Shen et al., 2006).
argillite, sandstone, and conglomerate (Fig. 5A). Although beds Although Upper Permian fluvial strata measured within the
are cleaved and deformed in places, sedimentary structures are Noyon Uul syncline are generally consistent with at least peri-
relatively well preserved. Organic material is abundant through- odically wet-humid conditions (Hendrix et al., 1996), Permian
out, including in-place petrified wood stumps. Sandstone beds strata unconformably underlying this section are quite different.
are typically lenticular over several meters, have sharp and Anatoleva (1974) describes ~315 m of distinctively red extrusive
erosional bases, and are trough to low-angle cross-bedded and volcanic units interbedded with sandstone and mudstone exposed
current-rippled (Fig. 5B–D). These features are interpreted southeast of the Noyon Uul syncline (Figs. 2, 4). Our observa-
as fluvial channels, and given the abundance of red and gray tions of this section are generally consistent with Anatoleva’s
mudstone, representing floodplain deposits, we suggest mainly (1974) descriptions. The lowermost unit is in thrust-fault contact
meandering fluvial environments. The presence of soft sedimen- with middle–Upper Carboniferous arc sequences (Fig. 5E). It
tary deformation features, climbing ripples, and syndepositional consists of ~55 m of poorly sorted cobble-boulder conglomerate
clastic dikes may also be consistent with a delta-plain environ- interbedded with andesitic-rhyolitic crystalline tuffs ranging from
ment (Battacharya and Walker, 1992), even though distal marine 0.5 to 2.5 m thick. This sequence fines upward into a distinctive
or lacustrine successions are not preserved. There is an over- red and reddish-brown succession (~140 m total) of interbedded
all upward-coarsening trend, possibly representing a shift from siltstone, sandstone, and conglomerate interbedded with tuff,
meandering to braided fluvial deposition from Late Permian to tuff breccias, and rhyolitic ash-fall deposits and rare trachyte-
Early Triassic time. basalt flows. The uppermost ~120 m of section grades to mainly
Key for all stratigraphic sections
siderite horizon current ripples Upper Permian,
root marks low angle Shin Jinst
cross beds (m)
mud chips Upper Permian (?) - lowermost 100
low angle /
carbonate wavy bedding Triassic, Noyon Uul syncline
nodules planar laminae (Western Transect,
soft sediment Hendrix et al., 1996) trough
deformation trough cross beds cross beds
cover n=58
coal (m)
mudstone-siltstone
fine-medium grained sandstone
volcanic and volcaniclastic
sandstone
gravel-pebble conglomerate 40
limestone or dolomite
cobble-boulder conglomerate
75

Lower Permian,
SE of Noyon Uul syncline
(m) angular unconformity

300 30

250 50

200 20

150

25

100 10

50

0m 0m 0m
r t. t. l. r t. t. l. r t. t. l.
ve lts ss cg ve lts ss cg ve lts ss cg
co d/si d. ble co d/si d. ble co d/si d. ble
u e b u e b u e b
m m pe m m pe m m pe

Figure 4. Measured sections from southwest localities (see Fig. 2 for locations). Key applies to all measured sections
(Figs. 4, 6, 9, 11). The left-hand section (Lower Permian, southeast of Noyon Uul syncline) is a schematic representation
based on Anatoleva (1974) and our own field observations. The Noyon Uul syncline detailed section was measured in
1992 by M. Hendrix and E. Sobel (Western Transect of Hendrix et al., 1996; ~1500 m from base), and is redrafted with
permission. This section was measured close to the Permian–Triassic boundary, and its exact age is uncertain; however,
the depositional style is representative of the underlying, continuous Permian section. Paleocurrent measurements are
from the same approximate stratigraphic interval, measured ~10 km along strike (Sain Sar Bulag transect; Hendrix et al.,
2001). The Shin Jinst section was measured as part of this study, as described in the text.
A B

~ 30 meters

C D

E F

~10 m

G H

Figure 5. Photographs from Permian strata at the southwest localities (F–H from M. Hendrix). (A) Overview of Shin
Jinst section. (B, C) Channel cut-and-fill geometries at Shin Jinst. (D) Soft sedimentary deformation within a channel
sandstone body, Shin Jinst. (E) Lower Permian rocks southeast of Noyon Uul syncline, showing red beds and volcani-
clastic rocks in the foreground. High ridges in the background are Carboniferous sedimentary and volcanic rocks thrust
over Permian (teeth on overriding plate). (F) Overview of detailed Upper Permian section at Noyon Uul. (G) Carbonate
nodules within interdistributary mudstone-siltstone deposits, Upper Permian, Noyon Uul. Stratigraphic up is to the upper
left. (H) Variegated red-green argillosol and resistant carbonate zone, Upper Permian, Noyon Uul (hammer for scale).
U. Permian-L.Triassic
Tsaagan Tolgoy, detailed section
U. Permian-L.Triassic (m)
Tsaagan Tolgoy, long section 70
last petrified wood fragment
(m) (at about 80 meters)
400
X

60
last siderite horizon

major shift in pollen assemblage

300 50
X Lower Triassic
(measured along strike)
detailed section at right

first nodular carbonate


X Upper Permian
color change to dark gray
to greenish-light gray mudstones

40

200
last coal bed
last Cordaites leaf impression
30

20

100

X
10

X
0m 0m
l al
r oa t. t. l. er coltst. sst. cgl.
o ve c ilts . ss cg v i
c d/s d ble
e b co d/s d. ble
u u e b
m m pe m m pe

Figure 6. Long (at left) and detailed (right) measured sections across the Permian–Triassic boundary at south-central local-
ity Tsaagan Tolgoy. Xs mark palynological samples used in Figure 7. See Figure 4 for key to lithologies and symbols.
374 Johnson et al.

cross-bedded sandstone and siltstone, which are variegated with coarsens upward abruptly, with the first appearance of pebble-
yellow and greenish gray more commonly than in the underly- cobble conglomerate at ~400 m. Conglomeratic and sandy inter-
ing sections (Fig. 4). In general the section is coarse grained and vals contain large-scale trough cross-beds and sharp to erosive
poorly sorted, with a trend to better sorting and rounding starting bases (Fig. 8E). Above 280 m, mudstone units grade abruptly
at ~100 m from the base. Organic matter is rare, probably owing from organic-rich, dark-gray units to lighter colored tan-green
to volcanic contributions and poor preservation under oxidizing and organic-poor beds. Carbonate nodules also appear in the
conditions. These observations are consistent with an alluvial fan upper section, and siderite is comparatively rare, indicating a
setting, close to an active volcanic source area, possibly grading shift to relatively well-drained soils. Overall, this vertical trend
to braided fluvial deposition through time. is interpreted as a change to higher energy fluvial environments,
A Late Permian age for this volcanic and clastic succession possibly including braided-stream deposition, with more oxi-
southeast of the Noyon Uul syncline has been suggested (Ana- dized and less organic-rich floodplain deposition.
toleva, 1974); however, this lower section appears to have an A distinctive shift in floral assemblages also occurs at Tsaa-
angular relationship with the Upper Permian strata described by gan Tolgoy (Fig. 7). Late Permian flora are dominated by Cor-
Hendrix et al. (1996) and is also quite distinctive lithologically daitales (pre-gymnosperms), ferns, and bottryococcus algae,
(with its red color, abundant volcanic rocks, and lack of coal mea- whereas the Triassic assemblage includes taeniate-striate bisac-
sures). The age of the lower section is bracketed between Late cate pollen typically associated with evaporites and more arid
Carboniferous, which is in fault contact with this section, and conditions. Cavate microspores (Lundbladispora sp.) and mega-
Late Permian; thus we favor an Early Permian age for these clas- spores characteristic of Early Triassic heterosporous lycopods
tic and volcanic units. (e.g., the desert xerophyte Pleuromeia) were also found in the
upper part of the section. An observed 4–6‰ decrease in the δ13C
South-Central Localities ratio of organic carbon is coincident with changes in lithology
and pollen assemblage (Fig. 7A). Similar shifts across the Perm-
One of the better known nonmarine Permian successions in ian–Triassic boundary in marine and nonmarine sections around
south-central Mongolia is found at Tsaagan Tolgoy (also known the world are believed to reflect a large, globally synchronous
as Ih Uvgon, Zinniker and Badarch, 1997; Figs. 2, 6). Age control contribution of isotopically light carbon into the atmosphere
here is excellent, based on our own and previous paleontological (Baud et al., 1989; Krull and Retallack, 2000).
analyses (Durante, 1976; Vakrameyev et al., 1986; Zinniker and Along with the lithologic changes noted previously, the
Badarch, 1997; Uranbileg, 2003; Aristov, 2005), which document rapid floral and isotopic shift is thought to represent the demise
Upper Permian through Lower Triassic strata (Fig. 6). Character- of Late Permian coal-bearing swamp facies and replacement with
istic Permian pollen genera found at Tsaagan Tolgoy (identified primarily herbaceous flora tolerant of seasonal droughts. The
by D. Zinniker, Fig. 7) include Cordaitina, Florintes, Potonie- change may reflect the global record of warming across the Perm-
isporites, and Neoraistrickia. Triassic and transition zone species ian–Triassic boundary (Faucett et al., 1994; Retallack et al., 1996;
include Lueckisporites virkiae, Lundbladispora sp., and Taeni- Berner, 2002). However, a similar aridification trend observed
sporites sp. The Permian–Triassic palynological shift is similar within Permian rocks of northeast China has also been attributed
to that documented along the southern margin of the Junggar to a possible rain shadow effect, reflecting ongoing collision and
basin (Ouyang and Norris, 1999). uplift along the Tien Shan–Yin Shan suture (Cope et al., 2005).
The Upper Permian section at Tsaagan Tolgoy (Fig. 6) is Supporting evidence for this climate shift is found at the Tavan
~300 m thick and consists primarily of dark, organic-rich mud- Tolgoy and Naftgar Uul localities, described below. Although
stone, coaly zones, and thin sandstone beds (Figs. 7, 8A). Pet- neither of these sections records the actual Permian–Triassic tran-
rified wood is common, including paleohorizontal and growth- sition, as at Tsaagan Tolgoy, individually they provide supporting
position logs encased in tabular sandstone beds (Fig. 8B). The evidence of distinctive and shifting climate regimes.
section is generally fine grained, with multiple concentrated Additional Upper Permian coal-bearing fluvial deposits are
zones of centimeter-scale siderite nodules, indicating swampy, preserved within coal quarries at Tavan Tolgoy (Durante, 1971;
reducing conditions (Landuydt, 1990). Sandstone units range Fig. 9). Coal seams are individually up to 14–33 m thick and
from thin beds to 1–4-m-scale individual and amalgamated units interbedded with coarser units, including rare pebble conglom-
and commonly exhibit convolute bedding and slumping and erate and medium- to fine-grained sandstone. Sandstones are
upward-fining trends (Fig. 8C). Trough cross-beds and asym- trough cross-bedded with sharp, erosional bases, and are lenticu-
metric current ripples are also found within the discontinuous lar, although amalgamated units are laterally continuous over tens
sandstone beds. Overall, this section is consistent with poorly of meters (Fig. 10). Sandstone beds exhibit normal grading, and
drained interdistributary or backswamp deposition, with sandier fine upward to rippled siltstone and mudstone. Paleocurrent mea-
units representing mainly crevasse splay deposition or meander- surements at Tavan Tolgoy are primarily south-southwest directed
ing of minor channels across the floodplain. (Fig. 9). Overall, outcrops at Tavan Tolgoy are broadly analogous
Above the Permian–Triassic boundary (~280 m above the to the Upper Permian at Tsaagan Tolgoy and are consistent with a
base of the section; Figs. 7, 8D), the section at Tsaagan Tolgoy meandering fluvial system in a relatively humid environment.
A Relative abundance of
palynomorph groups: present common abundant

smooth or small sculptured


cavate spores (7E)
characteristic

megaspore wall fragments


Triassic taxa

spore tetrad

taeniate bisaccate (7D)

striate bisaccate
transitional

unsculptured acavate
spores
normal bisaccate

monosaccate
characteristic
Permian taxa

sculptured acavate
spores (7B and 7C)
large sculptured cavate
spores
bottryococcus

cycadopites
-23
d 13 C (‰)

-25

-27
Permian Triassic
-29
0 40 80 120 160 200 240 280 320 360 400
distance in meters from base of section

B C D E

10 µ
µmm 10 µm 10 µm 10 µm

Figure 7. (A) Spore and pollen abundance and δ13C isotope data from Tsaagan Tolgoy, plotted against distance above the base of the long sec-
tion shown in Figure 6. (B–E) Photomicrographs of select spores including Neoraistrickia (B), Lycopodiumsporites (C), Leuckisporites (D), and
Lundbladispora (E), keyed to the palynomorph groups shown above. Carbon isotopic data were collected on acid-resistant kerogen isolated for
palynological analysis, conducted at Petrobras stable isotope laboratory in 1997 following industry-standard methods.
A B

10 cm

Figure 8. Photographs from Tsaagan Tolgoy. (A) Coaly intervals at the base of the section. (B) Petrified log in growth
position. (C) Convolute bedding in overbank deposits. (D) Transition from dark, organic-rich mudstone to tan, organic-poor
mudstone at the Permian–Triassic boundary (~290 m). (E) Trough cross-beds in a lenticular, channelized sandstone.
Figure 9. Measured sections from south-central localities Naftgar Uul and Tavan Tolgoy (see Fig. 2 for locations). Paleo-
current data from Naftgar Uul include data from possible Triassic strata overlying this section. See Figure 4 for key to
symbols and lithologies.
378 Johnson et al.

A The lower part of the section is more consistent with sedimentary


facies found in the Triassic across the region, and thus it is likely
that these strata are younger than mapped. The section at Naft-
gar Uul includes more than 300 m of interbedded sandstone, silt-
stone, and rare granular to pebble conglomerate. Sandstone beds
contain planar laminations, low-angle cross-stratification, and
trough cross-beds with >1 m of relief. These beds typically have
sharp, slightly erosive bases, and amalgamated units are laterally
continuous over >50 m. Paleoflow directions measured in trough
cross-beds are variable but are generally west-directed (Fig. 9).
Coal is absent in the section, and only rare, poorly preserved and
oxidized plant fossils are present. Lower Triassic strata here and
at Tsaagan Tolgoy are consistent with sand-rich meandering rivers
to braided-stream environments, with oxidizing environments and
likely periodic arid-drought conditions.

B Southeast Localities

The final phase of marine deposition in southern Mongo-


lia is preserved in outcrops of the Permian Lugyn Gol Forma-
tion, exposed at Nomgon and Hovsgol (Figs. 2, 11). Devonian
turbidite deposits are also reported in the area, which has led to
some confusion in mapping, where these widespread outcrops
are alternately assigned either to the Early–Middle Devonian or
the Late Permian (cf. Yanshin, 1989, and Tomurtogoo, 1999).
However, the Late Permian age of these deposits is well estab-
lished in multiple areas (Pavlova et al., 1991) and is confirmed
by our own findings in 1995 and 2005 of Late Permian crinoids
and bryozoans (J. Undariya, 2005, personal commun.). It is pos-
sible that Devonian crinoids and corals found in the section are
Figure 10. Photographs of the Tavan Tolgoy coal mine section shown reworked and transported. The Lugyn Gol Formation crops out
in Figure 9. (A) Lenticular sandstone bodies and interbedded coals.
Person (circled) for scale. (B) Detailed photo of finely laminated silt- along northeast-trending ridges at Nomgon (Fig. 2). Beds are
coal beds in interdistributary deposits. highly deformed and cleaved with large-scale isoclinal folding
and overturned units. Thus true thickness is difficult to deter-
mine, but we estimate a minimum thickness of ~800 m, based on
As at Noyon Uul, Upper Permian strata at Tavan Tolgoy and individually continuous, non-repeated sections (Fig. 11).
Tsaagan Tolgoy contain uniquely mixed Angaran-Cathaysian Permian flysch units mainly lie unconformably or in fault
assemblages (Vakrameyev et al., 1986) and thus indicate that contact with Carboniferous and older metasedimentary, volcanic,
south Mongolia and north China were in close proximity at that and intrusive rocks, including ophiolites, carbonate klippen, and
time. Widespread terrestrial deposition in this area likely occurred metamorphic rocks mapped as Precambrian (Fig. 3; Tomurtogoo,
in syn- and postcollisional flexural basins (Hendrix et al., 2001). 1999). The lowest part of the section includes a possible shallow-
Another similarity to the Noyon Uul section is that Upper Perm- marine succession that is marked by 10–50-m-scale upward-
ian strata near Tavan Tolgoy overlie extrusive intermediate-felsic coarsening and -thickening sandstone packages (Figs. 11, 12A).
volcanic rocks and interbedded reddish conglomerate, sandstone, Dolomitic units include crinoid, brachiopod, and bivalve coquina
and siltstone that are thought to be Lower Permian (Ulaan Nur and grainstone. Low-angle cross-bedding includes trough and
Formation, Fig. 3; Amory, 1996). Thus there is apparently an possible hummocky cross-stratification, and ripple marks are
older climatic shift from generally arid to more humid climates also present (Figs. 12B, C). Beds are generally tabular rather
in the middle Permian, followed by a return to aridification at or than strongly lenticular, but they do pinch out over hundreds of
near the Permian–Triassic boundary. meters along strike. Sandstone beds have sharp bases but gener-
In comparison with Tsaagan Tolgoy, Triassic strata are bet- ally lack channel-like scour features or bar forms, and no sub-
ter exposed nearby at Naftgar Uul (Figs. 2, 9). Although shown aerial exposure surfaces were identified. This lowest section is
as Permian on regional maps (Tomurtogoo, 1999), one of us (D. interpreted as a shallow-marine succession, possibly along a flu-
Zinniker) found Late Triassic to Middle Jurassic pollen assem- vial shoreline with upward-shallowing prodeltaic parasequence
blages (Classopollis) in lacustrine facies near the top of the section. packages (Fig. 11; Posamentier and Allen, 1999). Similar units
Figure 11. Measured sections through the Upper Permian Lugyn Gol Formation at Nomgon, southeast Mon-
golia (see Fig. 2 for location). Sections are noncontinuous but are shown in stratigraphic order from oldest to
youngest (left to right). Paleocurrent data are all from uppermost stratigraphic section. See Figure 4 for key to
lithologies and symbols.
380 Johnson et al.

A B

C D

~25 m

E F

Figure 12. Photographs of the Lugyn Gol Formation exposed at Nomgon. (A) Upward-coarsening and thickening pack-
ages (lowermost stratigraphic section, Fig. 11). (B) Trough cross-beds (lowermost section). (C) Ripple marks at top of
bed, middle section. (D) Overview of turbidite sequence. (E) Single upward-fining and -thinning unit (overturned beds,
from uppermost section, Fig. 11). (F) Flute casts at the base of sandy turbidite bed.

are found to the east at Hovsgol (Fig. 2), where crinoid and bra- Gol Formation includes hundreds of meters of tabular, continuous
chiopod grainstones are present with sporadic plant fossils, bed- sandstone and shale beds (Fig. 12D). Sandstone units are centime-
ding-plane parallel and perpendicular bioturbation, and ripples ter to decimeter scale in thickness and are typically upward fining
and convolute bedding, which may also support shallow-marine (Fig. 12E). Basal contacts are sharp to slightly erosive and com-
and/or prodeltaic deposition. monly preserve tool marks and flute casts, which generally show
There is a rapid transition from the shallow-marine packages south-southeast transport directions (Figs. 11, 12F). Structures
into regular turbidite sequences (Fig. 11, uppermost section) at include normal grading, sporadic load structures, planar laminae,
Nomgon. The classic turbidite flysch architecture of the Lugyn and some current ripples at the tops of beds, corresponding to Ta,
Sedimentary response to arc-continent collision 381

Tb, and Tc/Te petrofacies (respectively) of Mutti and Ricci Lucchi Qm


(1978). Local zones of intense cleavage preclude identification of
small-scale sedimentary structures, but the rhythmic sand-shale
interbedding and tabular, laterally continuous units (Fig. 11) are
consistent with turbidite deposition, likely within Mutti and Ricci
Lucchi’s (1978) mixed sand-mud facies (facies C).

rec
These turbidite deposits are tentatively assigned to an outer

yc
submarine fan setting that is intermediate between proximal and

led
k
loc
distal fan deposition (Walker, 1992). Coarse-grained facies, high-

oro
mixed

lb
density turbidite flows, slumping and channel-levee systems were

ge
nta
not observed, which argues against a more proximal inner fan

n
ne
setting. Similarly, the system is quite sandy, and bioturbation is
Dissected

nti
rare, arguing against a distal basin-floor fan environment. How- arc

co
ever, overall packaging is variable between more shaly and more
sandy intervals, and this likely indicates fluctuations in relative
sea level through time or shifting feeder channels within the fan
(Amory, 1996). The distal submarine environments observed at Transitional
Nomgon and Hovsgol are consistent with depositional settings arc
observed in analogous remnant ocean-basin fan systems, where F Lt
basin geometry favors trough-like axial deposition, and turbidite Undissected arc
pulses are funneled outboard over long distances (Graham et al.,
1975). We suggest that the underlying shallow-marine successions Shin Jinst (n=11) Southwest localities
formed along the margin of an extinct Carboniferous–Devonian Noyon Uul (n=21) (Upper Permian)
arc, immediately prior to final collision stage marked by rapid South-central
Naftgar Uul (n=18)
subsidence and deposition of overlying turbidite successions. locality (Triassic?)
Nomgon (n=21)
Southeast localities
Sandstone Provenance Data Hovsgol (n=10) (Upper Permian)
Effect of dynamic metamorphism
Sandstone samples were collected for petrographic analysis
(Nomgon and Shin Jinst samples)
and point counting (Fig. 13; Table 2) to complement facies anal-
yses. Sandstone provenance studies permit evaluation of the tec- Figure 13. Ternary diagram of sandstone modal compositions, using
tonic setting in which sedimentary sequences are deposited and monocrystalline quartz (Qm), potassium plus plagioclase feldspar (F),
can demonstrate evolution of source areas through time (Dickin- and total lithic fragments (Lt). Mean values are shown with a field
son and Suczek, 1979; Dickinson et al., 1983). In other arc colli- of one standard mean deviation (Table 2). Provenance fields are from
Dickinson et al. (1983), following methods described in the text.
sional settings, sandstone compositional trends have been attrib-
uted to erosion of magmatic arcs and unroofing of collisional
orogens (Dorsey, 1988; Graham et al., 1993) and thus record tim-
ing and evolution of suturing and postcollisional uplift. F = plagioclase + potassium feldspar, Lt = sedimentary, meta-
Sandstone provenance studies are best conducted on medium- morphic, and volcanic lithic fragments + carbonate and chert) is
grained sandstone samples that are relatively unaltered. Given the the most reliable comparison between all localities. Recent stud-
level of deformation and burial in many Permian outcrops of the ies of modern arc-derived sands also use the QtFL ternary plot
region, not all localities are represented. Nevertheless, 82 Perm- (Marsaglia and Ingersoll, 1992), which combines monocrystal-
ian samples from 6 localities across southwestern to southeastern line, polycrystalline, and chert into the Qt (total quartz) field.
Mongolia were point-counted using a modified Gazzi-Dickinson However, Qp (polycrystalline quartz) is a minor component in
method (Ingersoll et al., 1984), which compensates for the poten- most modern arc–derived sands (Marsaglia, 1992), so use of
tial adverse effect of grain-size variations. Principal framework QmFLt modal compositions provides a reasonable comparison
grains were counted by standard categories (Dickinson, 1970) and for the southern Mongolia Permian sample suite.
then normalized as detrital modes on standard ternary diagrams Upper Permian sandstones from southern Mongolia range
for comparison with global modal distribution fields (Table 2; from undissected to transitional to dissected arc provenance fields
Dickinson et al., 1983; Marsaglia and Ingersoll, 1992). In some (Fig. 13), indicating recycling or inheritance of late Paleozoic arc
cases, sandstones were extensively altered, such that distinguish- sources (Graham et al., 1993). None of the samples has a recycled
ing polycrystalline quartz and chert from altered and devitrified orogen or continental block signature. There does not appear to be
zeolites or volcanic lithic fragments was impossible. Thus the a consistent spatial trend in samples from across southern Mongo-
QmFLt ternary diagram (Fig. 13; Qm = monocrystalline quartz, lia: i.e., dissected versus undissected arc provenance fields are not
TABLE 2. POINT COUNT DATA FROM PERMIAN SANDSTONES OF SOUTHERN MONGOLIA
Naftgar Uul
Shin Jinst Qm F Lt (continued) Qm F Lt
94-SJ-181 17.66 29.58 52.76 93-NG-14a 18.37 38.86 42.77
94-SJ-182 16.36 43.18 40.45 93-NG-15 10.83 35.04 54.13
94-SJ-230 41.11 42.78 16.11 93-NG-15a 14.46 28.19 57.35
94-SJ-231 26.98 43.25 29.76 93-NG-24 20.60 36.54 42.86
94-SJ-240 24.24 34.34 41.41 93-NG-25 17.49 28.72 53.79
94-SJ-241 25.87 11.89 62.24 93-NG-26 16.84 31.02 52.14
94-SJ-243 30.70 28.73 40.57 93-NG-27 25.27 22.87 51.86
94-SJ-245 32.75 21.62 45.63 93-NG-27a 17.36 38.14 44.50
94-SJ-246 34.67 21.07 44.27 NG Mean 16.52 33.52 49.96
94-SJ-247 31.99 18.01 50.00 NG St Dev 4.24 5.08 5.71
94-SJ-251 23.48 30.19 46.33
SJ Mean 27.80 29.51 42.69 Nomgon (Lugyn Gol Fm.)
SJ St Dev 7.03 10.24 11.47 93-LG-201 30.80 22.90 46.20
93-LG-202 38.70 22.10 39.10
Noyon Uul (TU = Tost Uul, near Noyon Uul) 93-LG-204 26.10 25.00 48.90
95-NU-2 0.00 4.55 95.45 93-LG-206 39.10 19.00 41.90
95-NU-3 0.76 15.78 83.46 93-LG-207 19.30 28.90 51.90
95-NU-4 3.57 13.74 82.69 93-LG-208 24.60 23.50 51.80
95-NU-8 5.79 3.22 91.00 93-LG-209 23.40 27.10 49.50
92-NU-8 1.43 9.07 89.50 93-LG-210 26.40 23.40 50.20
92-NU-10 1.20 19.60 79.20 93-LG-211 25.60 25.90 48.50
92-NU-13 6.20 19.90 73.90 93-LG-301 26.30 29.30 44.40
92-NU-14 6.80 22.20 71.00 93-LG-302 31.00 27.00 42.00
92-NU-15 10.59 9.85 79.56 93-LG-303 32.90 25.20 41.90
92-NU-39 5.52 4.94 89.53 93-LG-304 30.90 18.00 51.10
92-NU-40 6.10 18.30 73.30 93-LG-305 20.60 29.40 50.00
92-NU-41 8.25 18.45 73.30 93-LG-306 27.40 31.80 40.80
97-TU-1 6.40 18.50 75.10 93-LG-307 30.90 21.10 48.00
97-TU-4 8.80 12.50 78.70 93-LG-308 26.60 29.60 43.80
97-TU-7 5.40 16.40 78.30 93-LG-309 27.50 30.50 41.90
97-TU-8 5.90 10.80 83.30 93-LG-401 32.00 24.70 43.30
97-TU-13 4.80 18.30 76.80 93-LG-402 43.20 24.80 32.00
97-TU-14 3.70 15.50 80.80 93-LG-403 26.60 15.70 57.60
97-TU-15 5.90 11.80 82.20 LG Mean 29.04 25.00 45.94
97-TU-16 10.20 18.50 71.30 LG St Dev 5.78 4.17 5.54
97-TU-19 2.00 8.90 89.10
NU Mean 5.21 13.85 80.83 Hovsgol (Lugyn Gol Fm.)
NU St Dev 2.89 5.41 6.79 XO-1A 4.14 25.30
XO-1B 5.08 28.43 66.50
Naftgar Uul XO-1C 1.52 33.59 64.90
93-NG-1 15.20 41.52 43.27 XO-1D 6.57 28.71 64.72
93-NG-2 15.01 29.01 55.98 93-XO-3A 1.63 12.23 86.14
93-NG-3 18.25 31.22 50.53 93-XO-3B 1.70 14.16 84.14
93-NG-5 18.69 26.11 55.19 93-XO-3C 0.59 24.56 74.85
93-NG-6 12.00 39.50 48.50 93-XO-3E 5.23 18.90 75.87
93-NG-7 5.43 33.33 61.23 93-XO-3F 5.85 22.15 72.00
93-NG-9 15.79 34.59 49.62 93-XO-3G 1.56 17.19 81.25
93-NG-12 20.59 32.94 46.47 XO Mean 3.39 22.52 74.09
93-NG-13 20.17 40.34 39.49 XO St Dev 2.09 6.52 7.40
93-NG-14 14.91 35.50 49.59
Note: QmFLt normalized to 100%. Qm—monocrystalline quartz; F—total feldspar, Lt—total lithic fragments (see text for
explanation of methods); St Dev—standard deviation.
Sedimentary response to arc-continent collision 383

linked to the southwestern, south-central, or southeastern study amalgamated middle to late Paleozoic oceanic arc systems and
areas, respectively. The dissected arc provenance indicated by microcontinental blocks of north China. Upper Permian sedi-
sandstone samples from Shin Jinst and Nomgon may in fact be an mentary successions in southwestern Mongolia (e.g., Shin Jinst
artifact of alteration level. Dynamic, greenschist-grade metamor- and Noyon Uul localities) comprise entirely nonmarine deposits,
phism causes notable recrystallization and breakdown of labile including meandering to braided fluvial environments. Similar
grains, including chert, calcite, and fine-grained lithic volcanic successions are found in south-central Mongolia (e.g., Naft-
fragments, into a slaty fabric matrix. This causes a relative com- gar Uul, Tavan Tolgoy, and Tsaagan Tolgoy), albeit with more
positional increase in the quartzofeldspathic fraction, which might extensive coal measures. These deposits likely formed in sepa-
indicate a more transitional-arc origin for the original composi- rate intramontane, flexural basins formed in response to collision
tion of these sandstones (gray arrow in Fig. 13). This may explain and continued crustal thickening along the western end of the
the difference between Hovsgol (undissected arc) and Lugyn Gol Tien Shan–Yin Shan suture (cf., Hendrix et al., 2001). In con-
sand compositions, as the Lugyn Gol turbidite sand samples have trast, Upper Permian strata in southeast Mongolia (e.g., Nomgon
undergone a comparatively high degree of alteration. and Hovsgol) include distal turbidite sequences that record the
By comparison, Naftgar Uul sandstone provenance samples final phase of marine deposition in a closing, trapped ocean basin
lack any metamorphic overprint. These samples are more com- immediately prior to final collision by Early Triassic time.
positionally mature, but they are also interpreted as possibly These regional facies relationships (summarized in Fig. 14)
younger (Triassic) deposits than the other localities plotted here. broadly support diachronous collision along the Tien Shan–Yin
They are distinctly richer in monocrystalline quartz. At Naftgar Shan suture zone. This interpretation is based mainly on the pro-
Uul there is no statistical change in sandstone composition over gression of youngest marine rocks from the middle Carboniferous
a 2000-m-thick sampling interval (Amory, 1996). In contrast, in southwestern Mongolia to the Late Permian in the southeast,
sandstone composition at Noyon Uul changes drastically near the as well as evidence for earlier cessation of arc-related magmatism
Permian–Triassic boundary from lithic-rich to more quartzofeld- in northwest China (Zhou et al., 2001; Cope et al., 2005). How-
spathic (Hendrix et al., 1996; note that only the Permian sample ever, several caveats accompany this general interpretation. The
compositions from Noyon Uul are presented here). marine-to-nonmarine facies transition is not smoothly continu-
ous from west to east. The only classic flysch deposits preserved
DISCUSSION in southeastern Mongolia are part of the Lugyn Gol Formation
(Nomgon and Hovsgol localities, Fig. 14). Carboniferous or
Tectonic Setting and Stratigraphic Evolution Lower Permian flysch deposits are not present in the study areas;
indeed, distal marine deposits predating the Upper Permian are
Permian sedimentary basins of southern Mongolia formed sparse throughout southern Mongolia (Lamb and Badarch, 2001).
along the northern margin of an evolving collision zone between Furthermore, Upper Carboniferous strata across the border in the

Mongol-Okhotsk ocean:
T1 marine sedimentation
in NE Mongolia

SW localities
(NU, SJ) SC localities T1 ? SE localities
(NA, TT, TsT) (NO, HO)
T1 P2
T1 Figure 14. Schematic summary of re-
Junggar: P2 P1
W-facing P1 P2 C2
gional tectonic and stratigraphic re-
P1 trapped C2 P1 lationships discussed in this study.
ocean basin? C2 EGFZ—East Gobi fault zone. South-
FZ E-facing P2 Solonker
T1 EG trapped ocean basin west (SW) localities: NU—Noyon Uul;
P2 Bulgan Uul:
offset P2 SJ—Shin Jinst. South-central (SC) lo-
P1 North Tie Yinshan
n Shan turbidites? calities: NA—Naftgar Uul; TT—Tavan
C2
Tolgoy; TsT—Tsaagan Tolgoy. South-
lt
Tarim fau Nanshan T1 east (SE) localities: NO—Nomgon;
agh
y nT North China P2 HO—Hovsgol. Age divisions are only
Alt block P1 approximate, based on available resolu-
C2 ? tion as discussed in the text.

T1 Lower Triassic nonmarine, arid climate facies 0 250 500 km


P2 Upper Permian nonmarine, humid climate facies
P1 Lower Permian marine strata (ocean basin) and/or oceanic arc sequences
C2 Upper Carboniferous
sediment dispersal direction
Inferred Late Paleozoic rotation direction.
384 Johnson et al.

Junggar basin may indicate west-directed closure and clockwise there evidence for systematic cessation of arc activity from west
rotation of Tarim, which suggests that perhaps Tarim and the North to east in Mongolia. Sand compositions represent recycling of
China block were not colliding as a single block (Fig. 14). Thus the Carboniferous and Devonian arc systems (Graham et al., 1993;
sedimentary record appears to support Late Carboniferous–Early Lamb and Badarch, 2001) in contrast to Permian sandstones in
Permian collision between the Altaids and Tarim, versus Late northeast China, which range from dissected arc to recycled oro-
Permian amalgamation between the Altaids and the North China gen provenance (Fig. 13; Mueller et al., 1991; Cope et al., 2005).
block; however, this does not seem to have occurred in a contigu- This indicates a much stronger continental (quartzofeldspathic)
ous east-west basin with simple progressive suturing proceeding contribution to Permian basins along the southern margin of the
toward the east (Fig. 15). Tien Shan–Yin Shan suture in China and argues against exposure
Provenance data underscore that the progressive closure of significant Precambrian crust in Mongolia (i.e., the hypoth-
model should not be oversimplified. A transition from undis- esized South Gobi microcontinent). Finally, the lack of direct
sected-dissected arc to recycled orogen provenance is observed evidence for Permian arc activity in southern Mongolia supports
as an unroofing trend in many arc-collisional settings (Dorsey, south-dipping subduction beneath the North China block, at least
1988; Clift et al., 2003). If closure of the Paleo-Asian ocean was during this terminal collision phase.
accommodated by progressive counterclockwise rotation and One way of reconciling the simple regional interpretation
northward movement of Tarim and North China blocks com- (of progressive west-to-east suturing) with the complexities
bined, we might expect to see a west to east trend from dissected indicated by field data is to depart from some aspects of tra-
to undissected arc provenance in our study areas indicating ear- ditional remnant ocean-basin models. According to Ingersoll et
lier arc dissection in the west. No such trend was observed, nor is al. (1995, p. 363), “A remnant ocean basin is a shrinking ocean

closing
turbidite
fan system
ocean basin ESE
Altaids, southern Mongolia
SE C
SC
Yin Shan active
North magmatic arcs
SW B China
North (Nanshan) block
A Tien Shan
Tarim

subducting
WNW Altaid lithosphere

microcontinent crust

A Junggar basin and North Tien Shan, northwest China


- Late Carboniferous to Early Permian collision and closure of Junggar trapped ocean basin
from east to west during clockwise (?) rotation of Tarim.
B North Tien Shan--Yinshan suture zone
- Collisional zone formed during late Paleozoic closure of the Paleo-Asian ocean. The orogen
grows through bidirectional suturing, first westerly with closure of the Junggar ocean basin
(Late Carboniferous to Early Permian), then easterly with closure of the Solonker ocean (Early
to Late Permian). The orogenic belt comprises mainly uplifted Carboniferous-Devonian arc
sequences, which are the main source for sediment delivered by turbidite fan deposits in the
closing ocean basins to the west and east.

C Solonker ocean basin


- Early to Late Permian collision between the North China block and the Altaids of southeast
Mongolia. Suturing progresses eastward over time due to counterclockwise rotation of the
North China block relative to Mongolia. Sediment is mainly derived from the growing orogenic
belt to the west (including uplifted Carboniferous-Devonian arc sequences), and distributed in
the form of turbidite fan deposits within the trapped ocean basin.

Figure 15. Paleogeographic block diagram for the late Paleozoic of the China-Mongolia border zone, summarizing ob-
served tectonic and stratigraphic relationships discussed in this study. Cartoon is highly schematic, and no specific scale
is implied by the features shown. SW, SC, and SE refer to approximate locations of the southwest, south-central, and
southeast study areas in Mongolia, discussed in the text (cf. Fig. 14).
Sedimentary response to arc-continent collision 385

basin, which is flanked by at least one convergent margin and Data from Mongolia help to address these hypothesized cli-
whose floor is typically covered by turbidites derived predomi- matic forcing factors. Regarding the impact of northward plate
nantly from associated suture zone(s).” Although this definition movement, southern Mongolia already had mixed Cathaysian-
broadly applies to the Permian setting in the China-Mongolia Angaran floral assemblages by Late Permian time, underscoring
border region, many aspects of the model are derived mainly the influence of flora with Siberian and North China craton affini-
from end-member cases in which significant volumes of sediment ties. However, paleomagnetic data still indicate a wide latitudinal
are deposited and preserved as turbidite fans—i.e., the Cenozoic gap between Siberia and the North China block (and by inference
Himalayan-Bengal system and the late Paleozoic Appalachian- the amalgamated Mongolian arcs) extending into the Cretaceous
Ouachita orogen (Graham et al., 1975). In these case studies, (Enkin et al., 1992). The sedimentary record in northeast Mongo-
sequential suturing of an irregular continental margin results in lia does not support such a late collision with the Siberian craton,
transition from fine-grained marine flysch to coarse-grained ter- although Lower Triassic marine rocks are present in northeast
restrial molasse in the direction of progressive collision (Dewey Mongolia (Zonenshain et al., 1990; Badarch et al., 2002), indicat-
and Burke, 1974). This predictive aspect of the remnant ocean ing that final amalgamation with Siberia likely extended into the
basin model may be broadly supported by the fact that turbi- early Mesozoic along this remnant ocean (the Mongol-Okhotsk
dite facies in southeastern Mongolia are mainly distal, outer fan ocean, Fig. 14) to the north of the Tien Shan–Yin Shan suture. As
deposits, favoring long-distance funneling of submarine flows such, continued northward movement of both the North China
outboard of a growing orogen. However, the flysch-molasse block and southern Mongolian arcs is still possible, at least
transition is not well preserved across the entire suture zone, and into the Triassic. However, the Permian–Triassic climatic shift
this part of the Altaids may be better represented by ongoing in southern Mongolia apparently postdates the middle Permian
intra-arc and microcontinent collisions with small trapped ocean event in China, whereas we would expect an earlier response to
basins, as are found in parts of the modern southwest Pacific the north if caused by such plate movement.
Ocean (cf. Fig. 15). Although it is currently unclear how much crustal thicken-
ing and orographic relief accompanied collision in the southern
Collision and Climate Altaids, Davis et al. (2004) have documented metamorphic core
complex formation in northeast China at ca. 220–190 Ma. This
Regional sedimentary facies relationships demonstrate chang- may indicate collapse of the Permian–Early Triassic collisional
ing climate regimes that accompany collision. In Mongolia, two orogen and would support the presence of high topography
possible transitions from relatively arid climate to humid climate prior to collapse. Cope et al. (2005) suggested creation of a rain
conditions occur (Fig. 14). The older shift is poorly constrained but shadow, resulting in red bed deposition, starting in the middle to
seems to have occurred in the middle Permian at Noyon Uul and Late Permian. Interestingly, this coincides with Upper Permian
Tavan Tolgoy, where red shale and sandstone beds generally rep- coal-bearing deposits in south-central Mongolia (Fig. 14), which
resenting well-drained soils and oxidizing conditions are distinct could also indicate trapping of moisture on the north side of an
from overlying Upper Permian coal-bearing units. The younger inferred mountain range along the Tien Shan–Yin Shan suture.
climatic shift is better constrained in southwest and south-central The rain shadow interpretation seems permissible within the
Mongolia, where coal-bearing Upper Permian strata are overlain range of available age constraints, although we note that turbidite
by Lower Triassic rocks that lack extensive coal deposits. This deposition continued in the Late Permian in southeast Mongolia
is illustrated by a dramatic record of floral transition to drought- (directly north of the Yin Shan belt), suggesting slightly later col-
tolerant organisms at Tsaagan Tolgoy. As noted previously, this lision. In any case, the Early Triassic appears to have been a time
sudden shift suggests a correlation with the global warming event of arid climate across the region, which likely reflected the global
at the Permian–Triassic boundary (Retallack et al., 1996; Berner, warming event at the Permian–Triassic boundary, perhaps modi-
2002). However, orographic relief generated by local tectonic fied by local topography.
uplift may also have influenced such climatic changes.
A similar aridification trend, though inferred to be some- IMPLICATIONS
what older, is reported from northeast China, where Carbonifer-
ous and Lower Permian humid-climate facies were replaced by If, as is argued here, the Tien Shan–Yin Shan suture links the
arid-climate indicators that started in the Late Permian (Mueller Tarim-Altaids and the North China block–Altaids collision zones,
et al., 1991). The relationship is also present in northwest China, the entire suture is >3000 km long and thus rivals some of the
based on a Late Permian transition from humid to arid environ- longest accretionary zones known (e.g., Indian-Eurasia collision,
ments (Carroll, 1991; Hendrix et al., 1992; Wartes et al., 2002; Ouachita-Appalachian orogen). The diffuse northern boundary
Fig. 14). In the Yin Shan belt, the drying event is attributed either represented by the Altaids creates a particular challenge in delin-
to northward latitudinal migration of north China into subtropical eating the final suture zone: perhaps the best marker is the south-
realms or to a rain shadow effect following development of high ern boundary against known Precambrian-floored blocks, with
topography during collision with the Altaids (Carroll et al., 1992; the understanding that multiple arc systems mark the collision
Cope et al., 2005). zone to the north. This record seems to be well represented by
386 Johnson et al.

modern accretionary processes in the southwestern Pacific ocean, 1978) or from intracontinental deformation within an accretion-
including the “triple-junction” region of convergence between ary complex (Tapponnier et al., 1982). The latter seems to have
the Pacific-Philippine, Australia, and Southeast Asia plates (Hin- been the case in Mongolia, where these two phases of sinistral
schberger et al., 2005), with several magmatic arcs and micro- motion have been tentatively linked to post-Paleozoic collisions
plates occupying the Java, Molucca, and Banda Seas. Although along the southward-growing margin of Asia, including the
the geometry of this collision zone may not directly apply to the Cenozoic India-Asia collision >2000 km from the East Gobi fault
Altaids, it is notable that the oceanic collision zone is marked zone. Cenozoic strike-slip is also demonstrated in southwestern
by multiple arc and backarc basins between major plates that are Mongolia (Cunningham et al., 2003). Rather than following a
moving independently, which may be analogous to the Siberian single terrane boundary, this intracontinental fault system seems
craton, the Tarim–North Tien Shan, and the North China block to exploit multiple zones of weakness within heterogeneous crust
in the late Paleozoic. Multiple trapped ocean basins are present of the Altaids. Thus consideration of syn-and postcollisional
along the Tien Shan–Yin Shan suture, but evidence is lacking for strike-slip deformation is critical to reconstruction of these col-
continuous zipperlike regional closure (Fig. 15). This broad com- lisional zones, and, conversely, better constraints on the location
parison is supported by other modern Pacific arc-collision ana- and nature of arc-collision zones help to predict modes and loca-
logues where climatic effects of orogeny are documented (e.g., a tions of subsequent deformation.
present-day rain shadow effect in southern Papua New Guinea) as
well as the evolution to syn- or postcollisional collapse and cre- CONCLUSIONS
ation of metamorphic core complexes, as in Taiwan (Clift et al.,
2003; Huang et al., 2006). Perhaps the main strength of Permian Stratigraphic data presented here are a first step toward
studies near the China-Mongolia border is that this region pro- unraveling the history of arc-arc and ultimately arc-continent col-
vides a good ancient analogue for such modern systems and pre- lision that is dramatically represented in the southern Altaids, par-
dicts what the southwest Pacific might look like following termi- ticularly from north of the China-Mongolia border. Future studies
nal collision and closure over the next tens of millions of years. that attempt to further constrain the timing and modes of conti-
Strike-slip deformation is also an important aspect of the nental growth by accretion, both in this part of Asia and in analo-
long-term results of these accretionary complexes. The role of gous settings worldwide, should take into account several key
syncollisional strike-slip faults in accommodating rotational clo- constraints afforded by this study. Regional relations that extend
sure and oblique collisions is noted in modern analogs (Moore and some 3000 km from northwest China, through south Mongolia,
Silver, 1983; Hinschberger et al., 2005; Huang et al., 2006). Late to northeast China, argue for broad correlation of the Tien Shan–
Paleozoic strike-slip deformation has been documented in parts Yin Shan sutures. The main suture zone is best drawn along the
of the Altaids (Şengör and Natal’in, 1996; Laurent-Charvet et al., northern margin of known continental blocks; tectonic boundar-
2003; Zhou et al., 2001) but has not yet been well documented ies that extend into the Altaids represent only extensions of this
in the Permian of Mongolia. Reconstruction of syncollisional broad collision zone into a heterogeneous accretionary margin.
paleogeography is complicated by later strike-slip deformation Generally south-dipping subduction during the Permian is
that has only recently been recognized in southern Mongolia. At supported by the lack of active Permian arcs in southern Mongo-
least two phases of left-lateral movement are documented along lia, and by complementary studies of active Permian arc systems
the East Gobi fault zone (Figs. 1, 2). The first phase occurred in in north China. In very general terms, west-to-east diachronous
the latest Triassic (ca. 207–209 Ma) associated with a northeast- closure may be argued mainly on the basis of the age of the young-
trending mylonite zone that is continuous for >250 km (Lamb et est marine strata from the Tien Shan to the Yin Shan (from Upper
al., 1999). This was followed by Late Cretaceous–Cenozoic reac- Carboniferous to Upper Permian). However, paleogeographic
tivation via brittle sinistral faulting (Johnson, 2004). The amount reconstructions based on sedimentary facies and paleocurrent
of offset associated with these sinistral movements is poorly con- data, sandstone petrofacies, and dating of magmatic activity indi-
strained but likely represents ~200–400 km of total slip (Webb cate that the collisional event was not a simple progressive sutur-
and Johnson, 2006). One important implication of this postcol- ing process. There is no record of progressive arc dissection in
lisional deformation is that attempts to reconstruct the southern the Permian from southwest to southeast Mongolia, nor is there
margin of the Altaids by conducting mainly west-to-east corre- a continuously eastward-younging flysch to molasse transition.
lations may incorrectly assume that these units were originally Sandstone provenance studies argue for very little input from
formed at the same latitude. For example, turbidite deposits in the continental crust or recycled orogen sources into Permian basins
Nomgon area may have equivalents along the southernmost bor- of Mongolia, in contrast to analogous deposits in north China.
der of Mongolia at Bulgan Uul (Figs. 2, 14; Lamb et al., 1999), Final arc-continent accretion may have included both west- and
and thus the record of continued marine deposition may extend east-facing trapped oceans and differential rotation of the North
farther westward than currently realized. China block and Tarim, possibly driven by uneven colliding mar-
It is common for convergent margins to be reactivated as gins with indenters that are currently poorly constrained. Thus
transform or strike-slip settings from shifting convergence direc- Permian strata of southern Mongolia are best considered an end-
tions along a plate margin (e.g., Cenozoic of California; Graham, member representation of published models for remnant ocean
Sedimentary response to arc-continent collision 387

basin closure. This record also provides a predictive glimpse into Allen, M.B., Windley, B.F., and Zhang, C., 1993, Palaeozoic collisional tecton-
future results of arc-arc and arc-continent collisions in the present ics and magmatism of the Chinese Tien Shan, central Asia: Tectonophys-
ics, v. 220, p. 89–115, doi: 10.1016/0040-1951(93)90225-9.
southwest Pacific Ocean. Amory, J., 1996, Permian sedimentation and tectonics of southern Mongolia
The possible implications of this collisional event are sig- [M.S. thesis]: Stanford, California, Stanford University, 183 p.
nificant. Regional climatic shifts are linked to evolution of the Amory, J., Hendrix, M., Lamb, M., Keller, A., Badarch, G., and Tomurtogoo,
O., 1994, Permian sedimentation and tectonics of southern Mongolia;
accretionary belt, including the possibility of a rain shadow effect implications for a time-transgressive collision with North China: Geologi-
south of the collisional orogen, which may have subsequently cal Society of America Abstracts with Programs, v. 26, no. 7, p. 242.
collapsed during the Triassic. Similar impacts are reported in Anatoleva, A.I., 1974, The structure and composition of the Permian red vol-
canogenic-sedimentary complex in southern Mongolia: Geologiya i Geo-
modern arc-continent collisions such as Taiwan and Papua New fizika, v. 15, p. 32–43.
Guinea. Finally, strike-slip deformation is an important syncol- Aristov, D.S., 2005, New Grylloblattids (Insecta: Gryllobattida) from the Trias-
lision mechanism to consider as a way of accommodating rota- sic of Eastern Europe, Eastern Kazakstan, and Mongolia: Paleontological
Journal, v. 39, p. 173–177.
tional closure. The recognition of subsequent strike-slip defor- Badarch, G., Cunningham, W.D., and Windley, B.F., 2002, A new terrane sub-
mation is also critical to reconstruction of the suture zone, which division for Mongolia; implications for the Phanerozoic crustal growth
in southern Mongolia is now partly offset and dismembered of Central Asia: Journal of Asian Earth Sciences, v. 21, p. 87–110, doi:
10.1016/S1367-9120(02)00017-2.
along the East Gobi fault zone. It is likely that multiple phases of Battacharya, J.P., and Walker, R.G., 1992, Deltas, in Walker, R.G., and James,
post-Permian movement along this fault system were driven by N.P., eds., Facies Models: St. Johns, Newfoundland, Geological Associa-
continued collisions along the growing margin of Asia, and thus tion of Canada, p. 157–178.
the long-term record of intracontinental deformation is also tied Baud, A., Magaritz, M., and Holser, W.T., 1989, Permian–Triassic of the Tethys;
carbon isotope studies: Geologische Rundschau, v. 78, p. 649–677, doi:
to this region as an inherited zone of crustal weakness. 10.1007/BF01776196.
Berkey, C.P., and Morris, F.K., 1927, Geology of Mongolia, in Natural History
ACKNOWLEDGMENTS of Central Asia, Volume 2: New York, American Museum of Natural His-
tory, 475 p.
Berner, R.A., 2002, Examination of hypothesis for the Permo–Triassic bound-
This work was supported by U.S. National Science Founda- ary extinction by carbon cycle modeling: Proceedings of the National
tion grants to Graham (EAR-9614555 and EAR-9708207) and Academy of Sciences of the United States of America, v. 99, p. 4172–
4177, doi: 10.1073/pnas.032095199.
Johnson (EAR-0537318, Tectonics and OISE, L.E. Webb, co- Burchfiel, B.C., Deng, Q., Molnar, P., Royden, L., Wang, Y., Zhang, P.,
PI), and funding from the American Chemical Society (Petro- and Zhang, W., 1989, Intracrustal detachment within zones of conti-
leum Research Fund 40193-G8 to Johnson). Funds from Stan- nental deformation: Geology, v. 17, p. 748–752, doi: 10.1130/0091-
7613(1989)017<0448:IDWZOC>2.3.CO;2.
ford University also supported this work, including the Gradu- Carroll, A.R., 1991, Late Paleozoic tectonics, sedimentation, and petroleum
ate Fellowship program, the McGee Fund, and the Stanford potential of the Junggar and Tarim basins, northwest China [Ph.D. thesis]:
Undergraduate Research Opportunity Grant. The American Stanford, California, Stanford University, 405 p.
Carroll, A.R., Liang, Y., Graham, S.A., Xiao, X., Hendrix, M.S., Chu, J., and
Association of Petroleum Geologists and the Geological Soci- McKnight, C., 1990, Junggar basin, northwest China: Trapped Late
ety of America supported field seasons in Mongolia through Paleozoic ocean: Tectonophysics, v. 181, p. 1–14, doi: 10.1016/0040-
several student grants to the authors. We thank Ch. Minjin and 1951(90)90004-R.
Carroll, A.R., Brassell, S.C., and Graham, S.A., 1992, Upper Permian lacus-
colleagues at the Mongolian University of Science and Tech- trine oil shales, southern Junggar basin, northwest China: American Asso-
nology for their support, and M.S. Hendrix and L.E. Webb for ciation of Petroleum Geologists Bulletin, v. 76, p. 1874–1902.
many years of collaboration. D. Zinniker thanks S. Fowell and Carroll, A.R., Graham, S.A., Hendrix, M.S., Ying, D., and Zhou, D., 1995, Late
Paleozoic tectonic amalgamation of northwestern China: Sedimentary
G. Norris for assistance in paleopalynology. Countless field record of the northern Tarim, northwestern Turpan, and southern Junggar
assistants and colleagues assisted in field work, including A. basins: Geological Society of America Bulletin, v. 107, p. 571–594, doi:
Chimitsuren, J. Crider, M. Heumann, A. Keller, T. Hickson, N. 10.1130/0016-7606(1995)107<0571:LPTAON>2.3.CO;2.
Chen, B., Jahn, B.-M., Wilde, S., and Xu, B., 2000, Two contrasting Paleo-
Manchuk, G. Sersmaa, E. Sobel, and J. Undariya. We appreci- zoic magmatic belts in northern Inner Mongolia, China: Petrogenesis and
ate thoughtful and thorough reviews by A. Carroll, R. Dorsey, tectonic implications: Tectonophysics, v. 328, p. 157–182, doi: 10.1016/
M. Faure, and P. Clift, which improved the final manuscript. S0040-1951(00)00182-7.
Clift, P.D., Schouten, H., and Draut, A.E., 2003, A general model of arc-conti-
This study is dedicated to our colleague, advisor, and dear nent collision and subduction polarity reversal from Taiwan and the Irish
friend Gombosuren Badarch, who died during preparation of Caledonides, in Larter, R.D. and Leat, P.T., eds., Intra-oceanic Subduction
this manuscript. This work would not have been possible with- Systems: Tectonic and Magmatic Processes: Geological Society [London]
Special Publication 219, p. 81–98.
out Badarch’s guidance, expertise, and patience. We miss him
Coleman, R.G., 1989, Continental growth of northwest China: Tectonics, v. 8,
deeply, but his love for Gobi field work and the geology of Asia p. 621–635.
continues to inspire us. Cope, T., Ritts, B.G., Darby, B.J., Fildani, A., and Graham, S.A., 2005, Late
Paleozoic sedimentation on the northern margin of the North China
Block; implications for regional tectonics and climate change: Interna-
REFERENCES CITED tional Geology Review, v. 47, p. 270–296.
Cunningham, D., Dijkstra, A.H., Howard, J., Quarles, A., and Badarch, G.,
2003, Active intraplate strike-slip faulting and transpressional uplift in the
Allen, M.B., Windley, B.F., Zhang, C., Zhao, Z., and Wang, R.-G., 1991, Basin Mongolian Altai in Storti, F., Holdsworth, R.E., and Salvini, F., Intraplate
evolution within and adjacent to the Tien Shan Range, NW China: Geo- Strike-slip Deformation Belts: Geological Society [London] Special Pub-
logical Society [London] Journal, v. 148, p. 369–378. lication 210, p. 65–87.
388 Johnson et al.

Davis, G.A., Xu, B., Zheng, Y., and Zhang, W., 2004, Indosinian extension in composition: Geological Society of America Bulletin, v. 105, p. 323–344,
the Solonker suture zone: The Sonid Zuoqi metamorphic core complex, doi: 10.1130/0016-7606(1993)105<0323:CSBOWC>2.3.CO;2.
Inner Mongolia, China: Beijing, China University of Geosciences, Earth Greene, T.J., Carroll, A., Hendrix, M., Graham, S., Wartes, M., and Abbink, O.,
Science Frontiers, v. 11, p. 135–143. 2001, Sedimentary record of Mesozoic deformation and inception of the
Dewey, J.F., and Burke, K., 1974, Hot spots and continental break-up: Implica- Turpan-Hami Basin, Northwest China, in Hendrix, M.S., and Davis, G.A.,
tions for collisional orogeny: Geology, v. 2, p. 57–60, doi: 10.1130/0091- eds., Paleozoic and Mesozoic Tectonic Evolution of Central and Eastern
7613(1974)2<57:HSACBI>2.0.CO;2. Asia: From Continental Assembly to Intracontinental Deformation: Geo-
Dickinson, W.R., 1970, Interpreting detrital modes of graywacke and arkose: logical Society of America Memoir 194, p. 317–340.
Journal of Sedimentary Petrology, v. 40, p. 695–707. Gubin, Y.M., and Sinitza, S.M., 1993, Triassic terrestrial tetrapods of Mon-
Dickinson, W.R., and Suczek, C.A., 1979, Plate tectonics and sandstone com- golia and the geological structure of the Sain-Sar-Bulak locality: Bul-
positions: American Association of Petroleum Geologists Bulletin, v. 63, letin of the New Mexico Museum of Natural History and Science, v. 3,
p. 2164–2182. p. 169–170.
Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Harrison, T.M., Copeland, P., Kidd, W.S.F., and Yin, A., 1992, Raising Tibet:
Inman, K.F., Knepp, R.A., Lindberg, F.A., and Ryberg, P.T., 1983, Prov- Science, v. 255, p. 1663–1670, doi: 10.1126/science.255.5052.1663.
enance of North American Phanerozoic sandstones in relation to tectonic Hendrix, M., 2000, Evolution of Mesozoic sandstone compositions, southern
setting: Geological Society of America Bulletin, v. 94, p. 222–235, doi: Junggar, northern Tarim, and western Turpan basins, Northwest China;
10.1130/0016-7606(1983)94<222:PONAPS>2.0.CO;2. a detrital record of the ancestral Tian Shan: Journal of Sedimentary
Dobretsov, N.L., Coleman, R.G., and Berzin, N.A., eds., 1994, Geodynamic Research, v. 70, p. 520–532.
Evolution of the Paleoasian Ocean: Russian Geology and Geophysics, Hendrix, M., Graham, S., Carroll, A., Sobel, E., McKnight, C., Schulein, B., and
v. 35, 233 p. Wang, Z., 1992, Sedimentary record and climatic implications of recur-
Dorsey, R.J., 1988, Provenance evolution and unroofing history of a modern rent deformation in the Tian Shan; evidence from Mesozoic strata of the
arc-continent collision; evidence from petrography of Plio–Pleistocene north Tarim, south Junggar, and Turpan basins, Northwest China: Geo-
sandstones, eastern Taiwan: Journal of Sedimentary Petrology, v. 58, logical Society of America Bulletin, v. 104, p. 53–79, doi: 10.1130/0016-
p. 208–218. 7606(1992)104<0053:SRACIO>2.3.CO;2.
Dubiel, R.F., and Smoot, J.P., 1994, Criteria for interpreting paleoclimate from Hendrix, M.S., Graham, S.A., Amory, J.Y., and Badarch, G., 1996, Noyon Uul
red beds: A tool for Pangean reconstructions, in Embry, A.F., Beauchamp, Syncline, southern Mongolia; lower Mesozoic sedimentary record of the
B., and Glass, D.J., Pangea: Global Environments and Resources: Cana- tectonic amalgamation of Central Asia: Geological Society of America
dian Society of Petroleum Geology Memoir 17, p. 295–310. Bulletin, v. 108, p. 1256–1274, doi: 10.1130/0016-7606(1996)108<1256:
Dumitru, T., Zhou, D., Chang, R., Graham, S., Hendrix, M., Sobel, E., and Car- NUSSML>2.3.CO;2.
roll, A., 2001, Uplift, exhumation, and deformation in the Chinese Tian Hendrix, M., Beck, M., Badarch, G., and Graham, S., 2001, Triassic synoro-
Shan, in Hendrix, M.S., and Davis, G.A., Paleozoic and Mesozoic Tec- genic sedimentation in southern Mongolia: Early effects of intracontinen-
tonic Evolution of Central and Eastern Asia: From Continental Assembly tal deformation, in Hendrix, M.S., and Davis, G.A., eds., Paleozoic and
to Intracontinental Deformation: Geological Society of America Memoir Mesozoic Tectonic Evolution of Central and Eastern Asia: From Conti-
194, p. 71–99. nental Assembly to Intracontinental Deformation: Geological Society of
Durante, M.V., 1971, Late Permian flora of Mongolia and the southern bound- America Memoir 194, p. 199–214.
ary of the Angaran floristic region of that time: Paleontological Journal, Heubeck, C., 2001, Assembly of central Asia during the middle and late
v. 4, p. 511–522. Paleozoic, in Hendrix, M.S., and Davis, G.A., eds., Paleozoic and Meso-
Durante, M.V., 1976, The Carboniferous and Permian Stratigraphy of Mongolia zoic Tectonic Evolution of Central and Eastern Asia: From Continental
on the Basis of Paleobotanical Data: Joint Soviet-Mongolian Scientific Assembly to Intracontinental Deformation: Geological Society of Amer-
Research Geological Expedition, v. 19, 280 p. ica Memoir 194, p. 1–22.
Durante, M.V., 1983, Existence of an Upper Permian mixed Cathaysio-Angar- Hinschberger, F., Malod, J.A., Rehault, J.P., Villeneuve, M., Royer, J.Y.,
ian flora in Nanshan (North China): Geobios, v. 16, p. 241–242, doi: and Burhanuddin, S., 2005, Late Cenozoic geodynamic evolution of
10.1016/S0016-6995(83)80022-9. eastern Indonesia: Tectonophysics, v. 404, p. 91–118, doi: 10.1016/
Enkin, R.J., Yang, Z., Chen, Y., and Courtillot, V.E., 1992, Paleomagnetic con- j.tecto.2005.05.005.
straints on the geodynamic history of the major blocks of China from the Huang, B., Otofuiji, Y., Zhu, R., Shi, R., and Wang, Y., 2001, Paleomagnetism
Permian to the present: Journal of Geophysical Research, B, Solid Earth of Carboniferous sediments in the Hexi Corridor; its origin and tectonic
and Planets, v. 97, p. 13,953–13,989. implications: Earth and Planetary Science Letters, v. 194, p. 135–149,
Faucett, P.J., Barron, E.J., Robison, V.D., and Katz, B., 1994, The climatic evo- doi: 10.1016/S0012-821X(01)00557-X.
lution of India and Australia from the Late Permian to mid-Jurassic: A Huang, C.-Y., Yuan, P.B., and Tsao, S.-J., 2006, Temporal and spatial records of
comparison of climate model results within the geologic record, in Klein, active arc-continent collision in Taiwan: A synthesis: Geological Society
G.D., ed., Pangea: Paleoclimate, Tectonics, and Sedimentation during of America Bulletin, v. 118, p. 274–288, doi: 10.1130/B25527.1.
Accretion, Zenith and Breakup of a Supercontinent: Geological Society Ingersoll, R.V., Fullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., and Sares,
of America Special Paper 288, p. 139–157. S.W., 1984, The effect of grain size on detrital modes; a test of the Gazzi-
Gilder, S., Zhao, X., Coe, R., Meng, Z., Courtillot, V., and Besse, J., 1996, Dickinson point-counting method: Journal of Sedimentary Petrology,
Paleomagnetism and tectonics of the southern Tarim basin, northwestern v. 54, p. 103–116.
China: Journal of Geophysical Research, v. 101, p. 22,015–22,031, doi: Ingersoll, R.V., Graham, S.A., and Dickinson, W.R., 1995, Remnant ocean
10.1029/96JB01647. basins, in Ingersoll, R.V., and Busby, C.J., eds., Tectonics of Sedimentary
Grabau, A.W., 1931, The Permian of Mongolia, in Natural History of Central Basins: Malden, Massachusetts, Blackwell Science, p. 363–392.
Asia, Volume 4: New York, American Museum of Natural History, 665 p. Jahn, B., 1999, Introduction to IGCP-420; Continental growth in the Phanero-
Gradstein, F.M., Ogg, J.G., et al., 2004, A Geologic Time Scale 2004: Cam- zoic; evidence from the Central Asian orogenic belt (CAOB): Mémoires
bridge, UK, Cambridge University Press, 589 p. de Géosciences Rennes, Special Issue no. 2, p. 1–5.
Graham, S.A., 1978, Role of Salinian Block in evolution of San Andreas fault Johnson, C.L., 2004, Polyphase evolution of the East Gobi basin: Sedimen-
system, California: American Association of Petroleum Geologists Bul- tary and structural records of Mesozoic–Cenozoic intraplate deforma-
letin, v. 62, p. 2214–2231. tion in Mongolia: Basin Research, v. 16, p. 79–99, doi: 10.1111/j.1365-
Graham, S.A., 1996, Controls on intracontinental deformation in Central Asia: 2117.2004.00221.x.
Geological Society of America Abstracts with Programs, v. 28, p. 112. Johnson, C.L., Webb, L.E., Graham, S.A., Hendrix, M.A., and Badarch, G.,
Graham, S.A., Dickinson, W.R., and Ingersoll, R.V., 1975, Himalayan-Bengal 2001, Sedimentary and structural records of late Mesozoic high-strain
model for flysch dispersal in Appalachian-Ouachita system: Geologi- extension and strain partitioning, East Gobi basin, southern Mongolia, in
cal Society of America Bulletin, v. 86, p. 273–286, doi: 10.1130/0016- Hendrix, M.S., and Davis, G.A., eds., Paleozoic and Mesozoic Tectonic
7606(1975)86<273:HMFFDI>2.0.CO;2. Evolution of Central and Eastern Asia: From Continental Assembly to
Graham, S.A., Hendrix, M.S., Wang, L.B., and Carroll, A.R., 1993, Collisional Intracontinental Deformation: Geological Society of America Memoir
successor basins of western China; impact of tectonic inheritance on sand 194, p. 413–434.
Sedimentary response to arc-continent collision 389

Krull, E.S., and Retallack, G.J., 2000, δ13C depth profiles from paleosols across of southern Gobi in Mongolia: Spisanie na Bulgarskogo Geologichesko
the Permian–Triassic boundary: Evidence for methane release: Geologi- Druzhestvo, v. 42, p. 107–117.
cal Society of America Bulletin, v. 112, p. 1459–1472, doi: 10.1130/0016- Ouyang, S., and Norris, G., 1999, Earliest Triassic (Induan) spores and pol-
7606(2000)112<1459:CDPFPA>2.0.CO;2. len from the Junggar Basin, Xinjiang, northwestern China: Review of
Lamb, M.A., 1998, Paleozoic sedimentation, volcanism, and tectonics of south- Palaeobotany and Palynology, v. 106, p. 1–56, doi: 10.1016/S0034-
ern Mongolia [Ph.D. thesis]: Stanford, California, Stanford University, 6667(98)00078-5.
207 p. Parrish, J.T., Ziegler, A.M., and Scotese, C.R., 1982, Rainfall patterns and the
Lamb, M.A., and Badarch, G., 2001, Paleozoic sedimentary basins and vol- distribution of coals and evaporites in the Mesozoic and Cenozoic: Pal-
canic arc systems of southern Mongolia: New geochemical and petro- aeogeography, Palaeoclimatology, Palaeoecology, v. 40, p. 67–101.
graphic constraints, in Hendrix, M.S., and Davis, G.A., eds., Paleozoic Pavlova, E.E., Manankov, I.N., Morozova, I.P., Solov’yeva, M.N., Suyetenko,
and Mesozoic Tectonic Evolution of Central and Eastern Asia: From Con- O.D., and Bogoslovskaya, M.F., 1991, Permian invertebrates of southern
tinental Assembly to Intracontinental Deformation: Geological Society of Mongolia: Joint Soviet Mongolian Scientific Research Geological Expe-
America Memoir 194, p. 117–147. dition, Transactions, v. 40, 173 p.
Lamb, M.A., Hanson, A.D., Graham, S.A., Badarch, G., and Webb, L.E., 1999, Posamentier, H.W., and Allen, G.P., 1999, Siliciclastic Sequence Stratigraphy–
Left-lateral sense offset of upper Proterozoic to Paleozoic features across Concepts and Applications: SEPM (Society for Sedimentary Geology),
the Gobi Onon, Tost, and Zuunbayan faults in southern Mongolia and Concepts in Sedimentology and Paleontology no. 7, 210 p.
implications for other Central Asian faults: Earth and Planetary Science Ren, J., Jiang, C., Zhang, Z., Qin, D., and Huang, T.K., 1987, Geotectonic Evo-
Letters, v. 173, p. 183–194, doi: 10.1016/S0012-821X(99)00227-7. lution of China: Beijing, Science Press, 217 p.
Landuydt, C.J., 1990, Micromorphology of iron minerals from bog ordes of Retallack, G.J., Veevers, J.J., and Morante, R., 1996, Global coal gap between
the Beliian Camine area, in Douglas, L.A., ed., Soil Micromorphology: A Permian–Triassic extinctions and middle Triassic recovery of peat form-
Basic and Applied Science: Amsterdam, Elsevier, p. 289–294. ing plants: Geological Society of America Bulletin, v. 108, p. 195–207,
Laurent-Charvet, S., Charvet, J., Monie, P., and Shu, L., 2003, Late Paleo- doi: 10.1130/0016-7606(1996)108<0195:GCGBPT>2.3.CO;2.
zoic strike-slip shear zones in eastern Central Asia (NW China); new Ritts, B.D., Yue, Y.J., and Graham, S.A., 2004, Oligocene–Miocene tectonics
structural and geochronological data: Tectonics, v. 22, no. 2, doi: and sedimentation along the Altyn Tagh Fault, northern Tibetan Plateau:
10.1029/2001TC901047. Analysis of the Xorkol, Subei, and Aksay basins: Journal of Geology,
Li, J.-Y., 2006, Permian geodynamic setting of Northeast China and adjacent v. 112, p. 207–229, doi: 10.1086/381658.
regions: Closure of the Paleo-Asian Ocean and subduction of the Paleo- Roger, F., Arnaud, N., Gilder, S., Tapponnier, P., Jolivet, M., Brunel, M., Mala-
Pacific: Journal of Asian Earth Sciences, v. 26, p. 207–224, doi: 10.1016/ vieille, J., Xu, Z., and Yang, J., 2003, Geochronological and geochemical
j.jseaes.2005.09.001. constraints on Mesozoic suturing in east central Tibet: Tectonics, v. 22,
Li, Y., McWilliams, M., Cox, A., Sharps, R., Li, Y., Gao, Z., Zhang, Z., and p. 1037, doi: 10.1029/2002TC001466.
Zhai, Y., 1988, Late Permian paleomagnetic pole from dikes of the Rowley, D., 1996, Age of initiation and collision between India and Asia: A
Tarim craton, China: Geology, v. 16, p. 275–278, doi: 10.1130/0091- review of stratigraphic data: Earth and Planetary Science Letters, v. 145,
7613(1988)016<0275:LPPPFD>2.3.CO;2. p. 1–13, doi: 10.1016/S0012-821X(96)00201-4.
Manankov, I.N., 1988, Late Permian productida (Brachiopoda) from southeast- Şengör, A.M.C., and Natal’in, B.A., 1996, Paleotectonics of Asia: Fragments of
ern Mongolia: Paleontological Journal, v. 32, p. 486–492. a synthesis, in Yin, A., and Harrison, M., eds., The Tectonic Evolution of
Manankov, I.N., 2004, New species of Early Permian brachiopods and biostra- Asia: Cambridge, UK, Cambridge University Press, p. 486–640.
tigraphy of the boreal basin of Mongolia: Paleontological Journal, v. 38, Şengör, A.M.C., Natal’in, B.A., and Burtman, V.S., 1993, Evolution of the
p. 366–372. Altaid tectonic collage and Palaeozoic crustal growth in Eurasia: Nature,
Manankov, I.N., Shi, G.R., and Shen, S.-J., 2006, An overview of Permian v. 364, p. 299–307, doi: 10.1038/364299a0.
marine stratigraphy and biostratigraphy of Mongolia: Journal of Asian Shen, S.-Z., Zhang, H., Shang, Q.H., and Li, W.-Z., 2006, Permian stratigraphy
Earth Sciences, v. 26, p. 294–303, doi: 10.1016/j.jseaes.2005.11.008. and correlation of Northeast China: A review: Journal of Asian Earth Sci-
Marsaglia, K.M., 1992, Petrography and provenance of volcaniclastic sands ences, v. 26, p. 304–326, doi: 10.1016/j.jseaes.2005.07.007.
recovered from the Izu-Bonin arc and the Mariana arc, Leg 126: Proceed- Shi, G.R., 2006, The marine Permian of East and Northeast Asia: An over-
ings of the Ocean Drilling Program, Scientific Results, v. 126, p. 139–154. view of biostratigraphy, palaeobiogeography and palaeogeographi-
Marsaglia, K.M., and Ingersoll, R.V., 1992, Compositional trends in arc-related, cal: Journal of Asian Earth Sciences, v. 26, p. 175–206, doi: 10.1016/
deep-marine sand and sandstone; a reassessment of magmatic-arc prov- j.jseaes.2005.11.004.
enance: Geological Society of America Bulletin, v. 104, p. 1637–1649, Tang, K., 1990, Tectonic development of Paleozoic foldbelts at the north mar-
doi: 10.1130/0016-7606(1992)104<1637:CTIARD>2.3.CO;2. gin of the Sino-Korean craton: Tectonics, v. 9, p. 249–260.
McElhinny, M.W., Embleton, B.J.J., Ma, X.A., and Zhang, Z.K., 1981, Frag- Tapponnier, R., Peltzer, G., Le Dain, A.Y., Armijo, R., and Cobbold, P., 1982,
mentation of Asia in the Permian: Nature, v. 293, p. 212–215, doi: Propagating extrusion tectonics in Asia; new insights from simple experi-
10.1038/293212a0. ments with plasticine: Geology, v. 10, p. 611–616, doi: 10.1130/0091-
Moore, G.F., and Silver, E.A., 1983, Collisional processes in the northern 7613(1982)10<611:PETIAN>2.0.CO;2.
Molucca Sea, in Hayes, D.E., ed., The Geology and Tectonics of East and Tomurtogoo, O., 1999, Geological Map of Mongolia, 1:1,000,000 scale:
Southeast Asian Seas: Part 2: Geophysical Monograph, v. 27, p. 360–372. Ankara, Turkey, General Directorate of Mineral Research and Explora-
Mossakovsky, A.A., and Tomurtogoo, O., 1976, Upper Paleozoic of Mongo- tion, 1 sheet.
lia: Joint Soviet Mongolian Scientific Research Geological Expedition, Uranbileg, L., 2003, The new plants of Upper Permian coal deposits in southern
Transactions, v. 15, 126 p. Mongolia: Mongolian Geoscientist, v. 23, p. 47–50.
Mueller, F., Rogers, J.J.W., Jin, Y., Wang, H., Li, W., Chronic, J., and Mueller, Vakrameyev, V.A., Lebedev, Y.L., and Sodov, Z., 1986, A cycad(?) Guramsania
J., 1991, Late Carboniferous to Permian sedimentation in Inner Mongolia, gen. nov. from the Upper Permian of South Mongolia: Paleontological
China, and tectonic relationships between North China and Siberia: Jour- Journal, v. 20, p. 95–101.
nal of Geology, v. 99, p. 251–263. Vincent, S.J., and Allen, M.B., 2001, Sedimentary record of Mesozoic intrac-
Mutti, E., and Ricci Lucchi, F., 1978, Turbidites of the Northern Apennines; ontinental deformation in the eastern Junggar Basin, northwestern China;
Introduction to Facies Analysis: Falls Church, Virginia, American Geo- response to orogeny at the Asian margin, in Hendrix, M.S., and Davis,
logical Institute, 166 p. G.A., eds., Paleozoic and Mesozoic Tectonic Evolution of Central and
Nie, S.-Y., Rowley, D.B., and Ziegler, A.M., 1990, Constraints on the locations Eastern Asia: From Continental Assembly to Intracontinental Deforma-
of Asian microcontinents in Palaeo-Tethys during the Late Paleozoic, in tion: Geological Society of America Memoir 194, p. 341–360.
McKerrow, W.S., and Scotese, C.R., Palaeozoic Palaeogeography and Walker, R.G., 1992, Turbidites and submarine fans, in Walker, R.G., and James,
Biogeography: Geological Society [London] Memoir, 12, p. 397–409. N.P., eds., Facies Models: St. Johns, Newfoundland, Geological Associa-
Nikolov, Z., Ehehbum, C., Shishkov, G., and Sallabasheva, V., 1981, Novye tion of Canada, p. 239–265.
dannye o geologiy i petrologiy kamennougol’nogo mestorozhdeniya Wang, B., Faure, M., Cluzel, D., Shu, L., and Charvet, J., 2005, Tectonic evo-
“Tavan Tolgoy” y yuzhnoj Gobi Mongol’skoy narodnoy respubliki—New lution of the northern part of the Yili Block (Western Chinese Tianshan),
data on the geology and petrology of the “Tavan Tolgoy” coal deposit in Sklyarov, E.V., ed., Structural and Tectonic Correlation across the
390 Johnson et al.

Central Asia Orogenic Collage: North-eastern Segment: Irkutsk, IGCP- From Continental Assembly to Intracontinental Deformation: Geological
480, Guidebook and Abstract Volume of the Siberian Workshop IGCP- Society of America Memoir 194, p. 151–170.
480, p. 274–278. Yanshin, A.L., 1989, 1:1,500,000 Map of Geologic Formations of the Mongo-
Wang, H., and Mo, X., 1995, An outline of the tectonic evolution of China: lian People’s Republic: Moscow, Academia Nauka USSR, 1 sheet.
Episodes, v. 18, p. 6–16. Yin, A., and Nie, S., 1996, A Phanerozoic palinspastic reconstruction of China
Wang, T., Zheng, Y., Gehrels, G., and Zhigou, M., 2001, Geochronological evi- and its neighboring regions, in Yin, A., and Harrison, M., eds., The Tec-
dence for existence of a South Mongolian microcontinent: Chinese Sci- tonic Evolution of Asia: Cambridge, UK, Cambridge University Press,
ence Bulletin, v. 46, p. 2005–2008. p. 442–485.
Wang, Y., Mooney, W., Xuecheng, Y., and Coleman, R., 2003, The crustal Zaitsev, N.S., 1974, Stratigraphy and Tectonics of the Mongolian People’s
structure from the Altai Mountains to the Altyn Tagh fault, northwest Republic: Joint Soviet Mongolian Scientific Research Geological Expedi-
China: Journal of Geophysical Research, Solid Earth, v. 108, no. B6, doi: tion, Transactions, v. 1, 148 p.
10.1029/2001JB000552. Zaitsev, N.S., Mossakovsky, A.A., and Shishkin, M.A., 1973, Type section of
Wartes, M.A., Carroll, A.R., and Greene, T.J., 2002, Permian sedimentary Upper Paleozoic and Triassic with the first remains of Labyrinthodonts,
record of the Turnpan-Hami basin and adjacent regions, northwest Southern Mongolia: Izvestiia Akademii Nauk USSR, Seriia Khimiches-
China: Constraints on postamalgamation tectonic evolution: Geologi- kaia, v. 7, p. 133–144.
cal Society of America Bulletin, v. 114, p. 131–152, doi: 10.1130/0016- Zhang, Z.M., Liou, J.G., and Coleman, R.G., 1984, An outline of the plate tec-
7606(2002)114<0131:PSROTT>2.0.CO;2. tonics of China: Geological Society of America Bulletin, v. 95, p. 295–
Watson, M.P., Hayward, A.B., Parkinson, D.N., and Zhang, Z.M., 1987, Plate 312, doi: 10.1130/0016-7606(1984)95<295:AOOTPT>2.0.CO;2.
tectonic history, basin development and petroleum source rock deposition Zhao, X., Coe, R.S., Gilder, S., and Frost, G.M., 1996, Paleomagnetic con-
onshore China: Marine and Petroleum Geology, v. 4, p. 205–225, doi: straints on the palaeogeography of China: Implications for Gondawana-
10.1016/0264-8172(87)90045-6. land: Australian Journal of Earth Sciences, v. 43, p. 643–672.
Webb, L.E., and Johnson, C.L., 2006, Tertiary strike-slip faulting in southeast- Zhou, D., Graham, S.A., Chang, E.Z., Wang, B., and Hacker, B., 2001, Paleo-
ern Mongolia and implications for Asian tectonics: Earth and Planetary zoic tectonic amalgamation of the Chinese Tian Shan: Evidence from a
Science Letters, v. 241, p. 323–335, doi: 10.1016/j.epsl.2005.10.033. transect along the Dushanzi-Kuqa Highway, in Hendrix, M.S., and Davis,
Webb, L.E., Graham, S.A., Johnson, C.L., Badarch, G., and Hendrix, M.S., G.A., eds., Paleozoic and Mesozoic Tectonic Evolution of Central and
1999, Occurrence, age, and implications of the Yagan-Onch Hayrhan met- Eastern Asia: From Continental Assembly to Intracontinental Deforma-
amorphic core complex, southern Mongolia: Geology, v. 27, p. 143–146, tion: Geological Society of America Memoir 194, p. 23–45.
doi: 10.1130/0091-7613(1999)027<0143:OAAIOT>2.3.CO;2. Ziegler, A.M., Rees, P.M., Rowley, D.B., Bekker, A., Li, Q., and Hulver, M.L.,
Windley, B.F., Allen, M.B., Zhang, C., Zhao, Z.Y., and Wang, G.R., 1990, 1996, Mesozoic assembly of Asia: Constraints from fossil floras, tecton-
Paleozoic accretion and Cenozoic redeformation of the Chinese Tien ics, and paleomagnetism, in Yin, A., and Harrison, M., eds., The Tec-
Shan Range, central Asia: Geology, v. 18, p. 128–131, doi: 10.1130/0091- tonic Evolution of Asia: Cambridge, UK, Cambridge University Press,
7613(1990)018<0128:PAACRO>2.3.CO;2. p. 371–400.
Wright, V.P., 1992, Paleopedology: Stratigraphic relationships and empirical Zinniker, D.A., and Badarch, G., 1997, Reconnaissance palynology and evi-
methodology, in Martini, I.P., and Chesworth, W., eds., Weathering, Soils, dence of climate change across the Permian/Triassic boundary, Ih Uvgon,
and Paleosols: Amsterdam, Elsevier, p. 475–499. southern Mongolia: Geological Society of America Abstracts with Pro-
Xiao, W., Windley, B.F., Hao, J., and Zhai, M., 2003, Accretion leading to grams, v. 29, no. 6, p. 97.
collision and the Permian Solonker suture, Inner Mongolia, China: Ter- Zonenshain, L.P., Durante, M.V., Markova, N.G., Filippova, I.V., and Check-
mination of the central Asian orogenic belt: Tectonics, v. 22, no. 6, doi: hovich, M.V., 1974, The main features of the geological structure and
10.1029/2002TC001484. development of contiguous parts of the Mongolian and Gobi Altai Moun-
Xiao, W., Windley, B.F., Badarch, G., Sun, S.H., Li, J., Qin, K., and Wang, Z., tains, in Zaitsev, N.S., ed., Stratigraphy and Tectonics of the Mongolian
2004, Palaeozoic accretionary and convergent tectonics of the southern People’s Republic: Joint Soviet Mongolian Scientific Research Geologi-
Altaids: Implications for the growth of Central Asia: Geological Society cal Expedition, Transactions, v. 1, p. 114–131.
[London] Journal, v. 161, p. 339–342. Zonenshain, L.P., Kuzmin, M.I., and Natapov, L.M., 1990, Geology of the
Yang, J., Xu, Z., Zhang, J., Chu, C., Zhang, R., and Liou, J.G., 2001, Tectonic USSR; a Plate-Tectonic Synthesis: American Geophysical Union Geody-
significance of early Paleozoic high-pressure rocks in Altun-Qaidam-Qil- namics Series, v. 21, 242 p.
ian Mountains, northwest China, in Hendrix, M.S., and Davis, G.A., eds.,
Paleozoic and Mesozoic Tectonic Evolution of Central and Eastern Asia: MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA


The Geological Society of America
Special Paper 436
2008

Links among mountain building, surface erosion, and growth of an accretionary prism in
a subduction zone—An example from southwest Japan

Gaku Kimura*
Department of Earth and Planetary Science, University of Tokyo, Japan, and Institute for Frontier Research on Earth Evolution,
Japan Agency for Marine-Earth Science and Technology, Yokosuka, Japan

Yujin Kitamura
Department of Earth and Planetary Science, University of Tokyo, Japan, now at IFM-GEOMAR, Kiel, Germany

Asuka Yamaguchi
Hugues Raimbourg
Department of Earth and Planetary Science, University of Tokyo, Japan

ABSTRACT

The relationships between mountain building, surface erosion, sediment supply to


the trench, and growth of the accretionary prism are examined in southwest Japan and
the Nankai Trough. Mountain building caused by the subduction of the Philippine Sea
plate in the Nankai Trough and collision in central Japan has resulted in a rock uplift
rate of ~4 mm/yr. Surface denudation rates in the mountain regions are on the order of
3–4 mm/yr, resulting from the heavy rainfall of the Asian monsoon. This fact suggests
that mountain building is almost in an equilibrium stage in which surface erosion and
rock uplift balance each other, resulting in a constant altitude of ~2000 m.
Several drainage systems on land and in offshore submarine canyons enable
the transport of eroded sediments directly into the Nankai Trough. Most of the ter-
rigenous sediments supplied to the Nankai Trough are accreted in the subduction
zone of the Philippine Sea plate. The accretion rate of the sediments in the eastern
Nankai Trough is ~1.68 × 107 m3/yr, which is consistent with the denudation rate of
the Akaishi Mountains, contributing to the supply of 1.72 × 107 m3/yr of sediment in
central Japan.
The growth of the accretionary prism is an important controlling factor for the
onset of large earthquakes in the Nankai Trough, because the hanging wall of the
rupture area of the seismogenic zone is composed entirely of the accretionary prism.
Repeated large earthquakes with a recurrence time of ~100–200 yr, which are well
recorded in the Nankai Trough, in turn promote surface erosion through consecutive
landslides and tsunamis.

*gaku@eps.s.u-tokyo.ac.jp

Kimura, G., Kitamura, Y., Yamaguchi, A., and Raimbourg, H., 2008, Links among mountain building, surface erosion, and growth of an accretionary prism in a
subduction zone—An example from southwest Japan, in Draut, A.E, Clift, P.D., and Scholl, D.W., eds., Formation and Applications of the Sedimentary Record
in Arc Collision Zones: Geological Society of America Special Paper 436, p. 391–403, doi: 10.1130/2008.2436(17). For permission to copy, contact editing@
geosociety.org. ©2008 The Geological Society of America. All rights reserved.

391
392 Kimura et al.

Southwest Japan, with its extensive record of both erosional processes and seis-
mic events, shows the intimate long-term relationships between tectonically driven
mountain building, surface erosion under the Asian monsoon climate, growth of the
accretionary prism in the trench, and the generation of large earthquakes.

Keywords: Nakai Trough, Asian monsoon, denudation, accretionary prism, large


earthquake.

INTRODUCTION TECTONIC SETTING OF SOUTHWEST JAPAN AND


MOUNTAIN BUILDING IN CENTRAL JAPAN
The growth of accretionary prisms is controlled mainly by
the sediment supply into the trench, which depends on surface The tectonic setting of central to southwest Japan has resulted
erosion on land and on the drainage system down to the trench. from the convergence of four plates or blocks, and it is commonly
The development of the drainage system from land to trench and decomposed into three distinct layers (Fig. 1): the top layer is
active surface erosion are necessary for the development of an a conjunction of the Amurian plate of southwest Japan and the
accretionary prism. Surface erosion on land is especially intense northeast Japan block, which have collided with each other in
in mountain regions under climatic conditions that include heavy central Japan; the second layer is the Philippine Sea plate; and
rainfall, such as tropical to subtropical rainy mountains (e.g., in the lowest layer is the Pacific plate, which is subducting beneath
southeastern Asia) or high-latitude glacier-covered mountains the above two layers.
(e.g., Alaska, Aleutian Trench, or southernmost regions of South The upper plate of the Japanese islands is composed mainly
America). Mountain building in these regions results from plate of volcanic arcs and ancient accretionary complexes, whose rigid
convergence associated with continent-continent, continent-arc, parts consist only of upper brittle crust without rigid mantle litho-
or arc-arc collisions, or intracontinental shortening. sphere. As the island arcs are strongly deformed, they are sepa-
Southwest Japan is a particularly appropriate context for rated into several microblocks by large faults or tectonic disconti-
analysis of the complex interplay among active mountain build- nuities. In such cases the distinction between plate boundaries and
ing on land, active surface erosion from the Asian monsoon intraplate faults is subtle, and the definition of plates differs from
climate, and the resulting development of a huge accretionary one author to another, from North America (Nakamura, 1983),
prism in the Nankai Trough (Taira et al., 1980, 1988). Taira et al. to Okhotsk (Savostin and Karasik, 1981; Kimura and Tamaki,
(1988) and Taira and Niitsuma (1986) emphasized that effective 1986), to the northeast Japan plate or block (Seno, 1985), to the
drainage systems from the central mountain region to the Nankai northeast Japan Arc, to Eurasia (Nakamura, 1983), to the Amu-
Trough in Japan have contributed to the buildup of the Nankai rian (Savostin and Karasik, 1981; Kimura and Tamaki, 1986)
accretionary prism. The mountain building in central Japan is the plate, to the southwest Japan Arc (Fig. 1). Whether the north-
tectonic result of the collision between the Izu-Bonin Arc and the east and southwest Japan Arcs are plates or microblocks, they
main Honshu Arc, and the collision between the southwest Japan are separated by a large fault in central Japan (Fig. 1; Nakamura,
and northeast Japan Arcs (Sugimura, 1972; Huzita, 1980). 1983), and ongoing convergence between the two is responsible
The Asian monsoon brings heavy rain to Japan every year for mountain uplift of the Japan Alpine Mountains in central
during the spring rainy season, called Tsuyu, which means con- Japan. The convergence and collision of these two arcs began in
tinuous rain. In addition, typhoons attack the Japanese islands, Pliocene time (Tsunakawa and Takeuchi, 1986) and accelerated
usually at the end of summer. This climatic condition enhances at ca. 2 Ma (Kaizuka, 1975; Huzita, 1980; Nakamura and Uyeda,
the surface erosion of the mountains, and eroded debris and sedi- 1980; Tsunakawa and Takeuchi, 1986).
ments drain into the rivers down to the ocean. The southern part of central Japan is also the collision zone
Frequent earthquakes also might promote surface erosion. between the Izu-Ogasawara Arc on the Philippine Sea plate and
Earthquakes in the Nankai Trough have repeatedly taken place the northeast Japan Arc (Sugimura, 1972; Fig. 1). Since the onset
during historic time (Ando, 1975). Most of the earthquakes in the of collision at ca. 15 Ma the arc crust has continuously flaked and
Nankai Trough are larger than Mw = 8. Ruff and Kanamori (1980) accreted to the upper plate of central Japan (Takahashi, 2006).
suggested that the occurrence of large earthquakes in the subduc- The Izu Peninsula is a colliding part of the Philippine Sea plate,
tion zone might be related to the development of an accretionary and the intracrustal breakage of the peninsula has already started
prism in the trench. because of the collision. The Pacific side of the southwest Japan
Thus, mountain building, heavy surface erosion, develop- Arc (Fig. 1) is composed mainly of Jurassic, Cretaceous, and
ment of the accretionary prism, and large earthquakes in the Tertiary accretionary complexes, which are exposed parallel to
Nankai Trough might ultimately be linked to one another. In this the Nankai Trough. The complexes in central Japan are bent in a
paper, we examine the linkage in southwest Japan and discuss the northward-convex shape. This bending is the result of the contin-
relationships among these geologic processes. uous collision and accretion of the Izu-Ogasawara Arc (Matsuda
130û
130º 135û
135º 140û
140º 145û
145º
45û
45º Okhotsk 45º
Plate

40û
40º Amurian 40º
Plate NE Japan

ch
ren
10 cm/y
Arc

an T
Jap
Collision
35û
35º SW Japan 35º
Arc Saga
m
Trou i
h
gh Pacific Plate
r o ug
iT
n ka
Na
4-6 cm/y
30û
30º 30º
Figure 2,3,5 and 6
Izu-Og
ch

asawar
en

Philippine Sea
Tr
yu

Plate
uk

25û
25º 25º
a Arc
Ry

km
0 500
130º 135º 140º 145º
Figure 1. Map showing tectonic setting of the Japanese islands. Note that central Japan lies at the collision zone between
the Amurian plate of southwest Japan, the northeast Japan block, and the Izu-Ogasawara Arc on the Philippine Sea plate.
394 Kimura et al.

and Uyeda, 1971). The main part of the Philippine Sea plate is system (GPS), put in place after the 1995 Hanshin-Awaji earth-
now subducting beneath southwest Japan along the Nankai and quake, has enabled highly detailed measurements over the past
Sagami Troughs at a rate of 4–6 cm/yr (Seno, 1989). 10 yr (Sagiya, 2004). The century-scale record shows that uplift
The Pacific plate is subducting beneath the northeast Japan in central Japan is concentrated in mountain regions (Figs. 2, 3),
Arc at ~10 cm/yr along the Japan Trench. A trench-trench-trench especially in the southern part—the Akaishi Mountains—which
type of triple junction lies to the southeast of central Japan at the lie to the west of the Itoigawa-Shizuoka Tectonic Line and the
southeastern end of the Sagami Trough. Izu Peninsula.
The mountains of the Kii Peninsula, and the Shikoku and
UPLIFT RATE OF THE MOUNTAINS IN CENTRAL Kyushu island regions in southwest Japan, also have undergone
AND SOUTHWEST JAPAN some uplift, especially in their southern parts (Figs. 2, 3). The
most uplifted regions are parallel to the Nankai Trough but are
The present crustal movement of the Japanese islands has not continuous. They actually alternate with subsiding regions
been precisely recorded with modern geodetic observations characterized by flat plains, channels, and bays, resulting in
within eight distinct periods from 1883 to 1999 (Kunimi et al., the following pattern from east to west: the Akaishi Mountains
2001). In addition to these observations, the global positioning (uplift), Ise Bay (subsidence), Kii Mountain in the Kii Peninsula

Itoigaw
Figure 4

a- S
iz

h
uo
T.L.

s ka T
Akaishi

.L.
in
Akaishi
Mounta
d Sea
I nlan
Seto
Is
e
Ba
aka
Os y O
ains

h
Bay ount
roug

Ogasawara Islands
Kii M
s
tain
ga T
M oun
oku no
e
Lin ma
Shik
Suru

ic Sh S Ku in
n Ch ion s
cto an o-m Ba
Te M
dia
n Tosa ne is
l ak
Me
nnel

Bay i
hu
us
o Cha

Ky A
gh
Bung

i Trou
nka
Na

Philippine Sea Plate

Figure 2. Map showing topographic and offshore features in central and southwest Japan. Rectangle shows the area of
Figure 4. Circles with letters show the following capes: O—Omaezaki; S—Shiono; M—Muroto; A—Ashizuri.
Mountain building, surface erosion, and growth of an accretionary prism 395

130
130º 131º 132º 133º 134º 135º 136º 137º 138º 139º 140º
37º
A km
0 100 200 300
36º

35º

34º

B
36º
33º

32º
34º

31º
m 32º
-0.4 0 0.4 m km
-0.05 0.00 0.05 0 100 200
30º
130º 132º 134º 136º 138º 140º

Figure 3. Maps showing vertical movement of the Japanese islands. (A) 1883–1999, modified from Kunimi et al. (2001).
Note that the Akaishi Mountains in central Japan and the Pacific side of the Kii Peninsula, as well as the Shikoku and
Kyushu islands, are characterized by an uplift rate of ~40 cm/100 yr. The Seto Inland Sea region, in contrast, is subsid-
ing. (B) Vertical movement of southwest Japan, represented from GPS networks from 1996 to 2004 (Sagiya, 2004). Dots
represent 358 GPS stations.

(uplift), Kii channel (subsidence), eastern Shikoku Mountains (Fig. 3B) except for the southern tip of the outer arc. Mountain
(uplift), Tosa Bay (stable), and the western Shikoku Mountains regions in the outer arc are generally uplifted, but their south-
in Kyushu (uplift). This geometry is characteristic of 10-km-scale ern parts in Capes Omaezaki in Tokai, Shionomisaki in Kii, and
crustal folds in the along-arc orientation and is explained as a Muroto in Shikoku, from east to west, are undergoing subsidence
superimposed compression that resulted from two different con- (Fig. 2B). This subsidence has been explained as an interseismic
vergence directions owing to the subduction of the Philippine Sea phenomenon between the large earthquakes in 1944 in Tonankai
plate along the Nankai Trough and the collision in central Japan and in 1946 in Nankai, and a near-future one predicted for the
(Huzita, 1980). twenty-first century (Sagiya, 2004). The uplift rate in the moun-
The Seto Inland Sea to the north of the island of Shikoku tain region has reached a maximum of 4 mm/yr (Fig. 2) in the
(Fig. 2) is a subsiding region. This aspect was explained by Akaishi Mountains and other nearby areas.
the same tectonic origin of crustal-scale folding, superposition
caused by the subduction of the Philippine Sea plate, and the col- SURFACE DENUDATION RATE OF THE MOUNTAINS
lision in central Japan similar to that in the outer arc of southwest IN CENTRAL JAPAN
Japan (Kaizuka, 1975; Huzita, 1980).
The recent compilation of geodetic observations by GPS The relationship between sediment production and tectonic
(Sagiya, 2004) shows almost the same vertical crustal move- uplift in the Japanese islands, and how the climate contributes
ment as was recorded earlier by long-term geodetic observations to sediment production, were thoroughly studied by Yoshikawa
396 Kimura et al.

h2
(1974) and Ohmori (1978). Their contribution is a milestone in 1
this research field. We briefly introduce their results in this sec- t= ∫ U − 0.3031 × 1 0
h1
−9
× H 2.1894
dH
tion and discuss them in light of data acquired from geographic .
information system (GIS) analyses in central Japan.
The denudation rate of the mountain region was estimated This is a law of diminishing returns, which implies that the
from the amount of accumulated sediments in the water reser- mountains are equilibrated at some altitude at some time t. In
voirs, which were constructed by complete damming of the pri- the case of 4 mm/yr of U, ~1 m.y. is enough to reach the equilib-
mary river system (the mainstream system important to the nation rium stage at which surface erosion and uplift balance each other,
as defined by the Ministry of Land, Infrastructure and Transport resulting in a constant altitude of ~2000 m.
of the Japanese Government) that drains the mountains. As the Figure 4B shows that most of the Akaishi Mountains are
volume of small clayey particles and solution chemicals from steeper than 30°, which is the critical slope angle consistent with
erosion is likely to be smaller by several orders of magnitude internal friction of crustal rocks. Slopes steeper than 30° are not
than that of solid accumulation in the reservoirs, Ohmori (1978) stable and collapse easily by gravitational sliding.
ignored such components for a first-order approximation. By multiplying the denudation rate (DR) obtained from the
The period studied by Ohmori (1978) is also important. local stream system (Ohmori, 1978) by the surface of the land
Many small dams were constructed in upstream regions in the whose altitude is higher than 1000 m, a total solid erosion rate is
tributaries of the primary rivers since the 1970s to prevent land- obtained. We reanalyzed the surface areas of the Akaishi Moun-
slides and debris flows in the mountains. This means that sedi- tains higher than 1000 m, which amounted to 4.631 × 106 m2,
ment accumulation in the large reservoirs can no longer be eas- obtained from GIS analysis. Applying the DR from the river sys-
ily related to the denudation rate of the mountains. Thus, the tems in the Akaishi Mountains in central Japan, an erosional solid
data set collected in the 1960s is the only available quantitative volume of ~1.72 × 107 m3/yr is estimated (Table 1).
measurement of the natural erosion rate in the mountains in the Ohmori (1978) pointed out that the lithology of the moun-
Japanese islands. tains and location-controlled climatic differences (such as snowy
Ohmori (1978) clarified that an empirical relationship mountains in the north or rainy mountains in the south within the
between the mean altitude of mountains and sediment produc- Japanese islands) are not strong factors in denudation rates. This
tion is expressed by is important, because climatic conditions during the Last Glacial
Maximum were different from those of the present. Sea level was
SDR = W × H a , much lower, and the land area was much greater, than in the pres-
ent, but low-altitude mountains might have not contributed much
where SDR = the mean annual sediment delivery rate (m3/km2/ to surface erosion, as suggested by the SDR-H function described
yr), H = the mean altitude of mountains (m), W = 0.4642−3 (/yr), above. The present Japanese islands widely cover the climatic
and a = 2.1894. Thus SDR is almost a quadratic function of the zones of the Asian monsoon from south to north, but Ohmori
mean altitude. The denudation rate (DR) is obtained from (1978) points out the absence of serious differences in the denu-
dation rate in northern and southern Japan.
DR = SDR ×1.75 /2.68 ×10−6 ,
DRAINAGE SYSTEMS TO THE TRENCH FLOOR OF
where 1.75 and 2.68 are the mean bulk densities of sediments THE NANKAI TROUGH
and bedrocks, respectively. Then, the relationship between the
denudation rate and the mean altitude is represented by The sediments from the uplifting mountains have not been
directly transported into the trench in many places but are instead
DR = 0.3031×10−9 × H 2.1894 . trapped in the forearc, e.g., on continental shelves, in forearc
basins, and in piggyback slope basins on the accretionary prism.
As reviewed in the previous section, most of the mountains In the eastern part of the Nankai forearc off the Tokai district, the
in Japan are undergoing tectonic uplift. The uplift started in the main submarine canyons—Suruga Trough and Tenryu Canyon—
Pliocene and accelerated at ca. 2 Ma as described. Geologically directly connect the on-land river systems to the trench floor of
long-term uplift is a result of tectonic uplift (U) minus surface the Nankai Trough (Fig. 5). The Sionomisaki Submarine Canyon
erosion. from the Kii Peninsula is the main drainage canyon in the middle
A geological long-term surface uplift rate dH/dt might be part of the Nankai Trough. Other, rather small canyons connect
expressed by the Shikoku and Kyushu islands with the Nankai Trough (Fig. 5).
These canyons are topographic manifestations of faults that
dH /dt = U − DR = U − 0.3031×10−9 × H 2.1894 . cut the accretionary prism (Tokuyama et al., 1999). Detailed topog-
raphy of the mouth of these canyons shows that submarine fans
Integrating the function from a mean altitude h1 to another mean (Tokuyama et al., 1999) composed of turbidites (Ike et al., 2005)
altitude h2, pass through the canyon channels directly from the land (Fig. 6).
137º A 138º 139º 137º B 138º 139º

36º 36º

1.887
2.247

v.
shi Mts

u ri
ry
Kiso riv.

Akai

Ten
3.598 Mt. Fuji
3.267 Fu
ji r
iv.

Abe riv.
35º 35º

Oi riv.
Izu Pen.
Suruga Bay

Kiku riv.
slope
km degree
Pacific Ocean 0 50 0 10 20 30 40 50

137º 138º 139º

Figure 4. (A) Area of mountains higher than 1000 m (gray area). Areas shown in five distinct colors show the drainage systems that supply sediments to the main rivers. Numbers in
solid circles represent denudation rates of each drainage system calculated by Ohmori (1978). (B) Areas with slopes steeper than 20°. Solid line on land shows a contour of 1000 m in
the Akaishi Mountains. Note that areas with slopes steeper than 20°–25° are mostly higher than 1000 m.
398 Kimura et al.

TABLE 1. DENUDATION RATE* AND SURFACE AREA HIGHER


THAN 1000 m ALTITUDE IN THE AKAISHI MOUNTAINS IN
The eastern parts of the Nankai Trough are filled by turbi-
CENTRAL JAPAN dites, which are apparent on seismic profiles (Park et al., 2002a;
River system Denudation rate Surface area Erosion of solids Ike et al., 2005) most of which originate from the canyons men-
–3 6 2 6 3
(10 m/yr) (10 m ) (×10 m /yr) tioned above (see Fig. 7). The trench floor of the Nankai Trough
Fuji 3.598 1,678 6.04 gently deepens from east to west owing to subduction of the
Tenryu 2.247 2,191 4.92
topographic highs in the eastern part (Tokuyama et al., 1999).
Oi 3.267 636 2.08
Abe 3.435 55 0.19
Therefore, the turbidites flow down to the west.
Kiso 1.887 2,125 4.01 The total volume of solid sediments draining directly into the
trench per year, passing through these channels, is ~1.44–1.91 ×
Total erosion of solids 17.24
Note: Denudation rate and each river system are shown in Figure 4.
107 m3/yr, as estimated from the solid volume of denudation in
*(Ohmori, 1978) central Japan, given in the previous section.

244
9711

888
590 3457 253
1257
2903
Japan Sea 5026

Trough
20 183
136 124
69
32

Suruga
Sagami
322 476 370

on
35 Trough
1601 1326 Cany
u

1748 no
Tenry

576
4600 ma
Ku in
s
180
Shi
o ba
130 Can no-mi
yon saki
4992
3867
u gh Pacific Ocean
Tro
1530
652
i
310
n ka
Na

Philippine Sea Plate

Figure 5. Map showing mean sediment delivery rate (SDR) to the reservoirs in central to southwest Japan. Black and gray
solid circles indicate the river systems draining into the Pacific Ocean and Japan Sea, respectively. Numbers in circles
represent the volume (m3/km2/yr), modified from Ohmori (1978).
Mountain building, surface erosion, and growth of an accretionary prism 399

130º 131º 132º 133º 134º 135º 136º 137º 138º 139º 140º
37º

36º

ts.
Akaishi M
Japan Sea

Trough
Honshu
35º

Suruga
Sagami

on
Trough

Cany
Kii Pen.
34º

u
Shikoku

Tenry
no
ma
Ku in
Shi s
o ba

KR9
Can no-m
33º yon isak
i

806
Kyushu 14
gh

-02
1-2 Pacific Ocean
d u
Tro
k ai
32º an
KR

N
98
10
-01

31º

Philippine Sea Plate


30º

m
km
-10000 0 10000 0 100 200 300

Figure 6. Map showing submarine channel systems draining down to the Nankai Trough. Black lines show survey lines
of the seismic profile shown in Figure 7.

ACCRETION RATE OF THE SOLID VOLUME IN THE taper angle varies between 2° and 11°, depending on hydrological
NANKAI ACCRETIONARY PRISM conditions within the wedge and basal décollement (Saffer and
Bekins, 2002; Wang and Hu, 2006; Kimura et al., 2007) (Fig. 7).
The mode of accretion changes at the trench-slope break The frontal accretion rate Vt per unit length along the trench
from an in-sequence and out-of-sequence thrusting mode to an is estimated from
underplating-dominant mode (Kimura et al., 2007). Terrigenous

zd
sediments supplied from the mountainous region on land domi- Vt = v c (1− φ 0e−b0 z )dz
nantly return back to the land owing to frontal accretion in the 0

Nankai Trough (e.g., Moore et al., 2001; Kimura et al., 2007). In


the case of the Nankai Trough the critical taper condition (Davis (Westbrook, 1994; Shibata et al., 2006), where v c is the conver-
et al., 1983) of the accretionary prism is operated trenchward gent rate, 4 cm/yr; φ0 = 0.65, the initial porosity of the trench
from the trench-slope break (Saffer and Bekins, 2002; Wang filling sediment; b0 = 1.54, the constant of the porosity-depth
and Hu, 2006; Kimura et al., 2007), and the accretionary-wedge relationship for the accretionary prism (Bray and Karig, 1985);
400 Kimura et al.

1605 tsunamigenic rupture?

Ashizuri KR9810-01 5.4 7.9 4.6


0 0
Depth (km)

2 Deformation 2
front
4 ST 4
6 OO 6
8
Accretionary prism 8
10 Step down 10
Philippine Sea Plate
12 12
80 70 60 50 40 30 20 10 0
1944 co-seismic rupture
6.3 16.5 7.8
2 Deformation front
Kii 2
Depth (km)

KR9806-02 4 4
)
lt (OOST
Splay fau
6 6
Accretionary prism
8 8
10 Step down 10
Philippine Sea Plate
× 12 12
60 50 40 30 20 10 0
1946 co-seismic rupture
0 0
3.3 8.5 2.1
Depth (km)

2 2
Deformation front
4 OOST 4
Accretionary prism 6
6
8 8
Step down Philippine Sea Plate
10 10
60 50 40 30 20 10 0
Muroto 141-2D
Distance from deformation front (km)
Figure 7. Simplified profiles of the Nankai accretionary prism (Kimura et al., 2007). The age of the prism at a distance of 30 km from the defor-
mation front is ca. 2 Ma (Moore et al., 2001). Out-of-sequence thrusts (OOSTs) are well developed in each accretionary prism.

and zd = 1200 m is the averaged thickness of trench-filling sedi- DISCUSSION


ments involved in frontal accretion in the eastern half of the Nan-
kai Trough (Moore et al., 2001; Park et al., 2002b). We quantitatively compiled data on the present mountain
We obtain Vt = 47.95 m2/yr. This value is the growth rate building, surface erosion, and growth of the accretionary prism
of the area of the profile of the accretionary prism normal to the in southwest Japan and the Nankai Trough. Our treatment is
trench axis. The distance of the accretionary prism along the just a first-order approximation and needs further development,
trench is ~500 km from the mouth of the Suruga Trough to the but these relationships are important toward understanding the
deformation front off Shikoku. The results of ocean drilling off relationship and feedback among these processes, which have
Shikoku show that the trench-filling sediments are dominated been studied separately until now. We discuss in the following
by turbidites directly derived from the land, as mentioned above section how present relationships can be extrapolated to long-
(Taira and Niitsuma, 1986; Taira et al., 1992). However, large term evolution.
amounts of terrigenous sediments are also supplied to the trench
through Shionomisaki Canyon and other canyons to the west, Relationship between Long-Term and Short-Term
diluting the sediment transported from the northeast. Because it Linkages
is difficult to quantify the contributions from various sources, we
consider only turbidites accreted in the trench to the east of the The denudation rate in the mountains reviewed by Ohmori
mouth of Shionomisaki submarine canyon, along ~300 km of the (1978) is averaged over ~10 yr, and the uplift rate in terms of
trench (Fig. 5). Using this value, the total volumetric accretion geodetic observation is averaged over 10–100 yr. It is question-
rate is ~1.68 × 107 m3/yr. able whether these rates are applicable to geologic time scales of
Such a rapid rate of accretion suggests that the accretion- millions of years.
ary prism outboard of the trench-slope break at ~30 km from the Was the denudation rate accelerated or decelerated during
deformation front is composed dominantly of Quaternary turbi- the most recent ice age? It is well known that high mountains in
dites. This is verified by direct drilling into the accretionary prism Japan were capped by glaciers during the last ice age (Minato
(Moore et al., 2001) and by samples using submersible dives et al., 1965). Such climatic conditions appear to have enhanced
(J. Ashi, 2006, personal commun.). surface erosion and thus have produced many more sediments
Mountain building, surface erosion, and growth of an accretionary prism 401

than in the present day, as a wide consensus concludes that aver- plate smooth the surface and make a widely coupled asperity in
age denudation rates have been higher than normal since the start the seismogenic zone. The rupture areas of the 1944 Tonankai
of northern hemispheric glaciation ca. 2.7 Ma. During the last ice and 1944 Nankai earthquakes lie beneath and within the accre-
age the average annual temperature in the Japanese islands was tionary prism (e.g., Park et al., 2002b), which developed from
~7–8 °C lower than that at present. Ohmori (1978) discussed the accretion of sediment transported into the Nankai Trough since
difference in denudation rate between the mountains in northern the late Miocene Epoch. What controls the size of the rupture
Japan from the example of Hokkaido, the northernmost island area of large earthquakes in subduction zones is a complicated
of Japan, and those in southwest Japan. Their locations are sepa- question. One of the hypotheses involves sediment subduction:
rated in latitude by >10°, and the difference in average annual Subducted sediments not only smooth the surface as suggested
temperature is ~10 °C at present. Ohmori’s conclusion is that the by Ruff and Kanamori (1980), but water is released from sedi-
denudation rates are almost the same in northern, central, and ments, especially those that are clayey, dynamically lubricating
southern Japan, as reviewed. Climatic conditions in central and the plate-boundary fault for rupture propagation (e.g., Kimura et
southwest Japan are different from those in northern Japan. A al., 2007). Most of the terrigenous turbidites are accreted at the
heavy rainy season is common in central and southwest Japan toe of the accretionary prism, but old pelagic to hemipelagic sedi-
but not in northern Japan. In contrast, northern Japan has a heavy ments underthrust, and their lithification works for the onset of the
snowy winter, whereas southwest Japan does not have such a seismogenic zone (Moore and Saffer, 2001; Kimura et al., 2007).
severe winter. Such climatic factors may compensate each other, The dynamic weakening resulting from lubrication of the fault
resulting in similar denudation rates. Such present analysis of might be the most likely mechanism to propagate the rupture area,
the denudation rate in the Japanese islands suggests that a much and the size of the fluid-rich area might control the magnitude of
cooler climate during the last ice age might not have resulted in earthquakes. In turn, the shaking of the land and tsunamis caused
higher denudation rates than at present. The East Asian summer by large earthquakes enhance landslides in the mountainous areas
monsoon rains were much weaker in glacial times, and because and promote surface erosion.
rainfall is a major control on erosion rates (Reiners et al., 2003;
Dadson et al., 2003) this means that denudation rates during the Accretion versus Subduction Erosion
last glacial maximum likely were much lower than in the present
interglacial. Thus, the wide consensus is that average denudation One of the unsolved but important questions in the research of
rates higher than normal during an ice age should be reexamined subduction zones concerns the balance between accretion and sub-
on a global scale. duction erosion (e.g., von Huene and Scholl, 1991; Clift and Van-
Another factor controlling denudation rate is sea level, nucchi, 2004). The review in this paper suggests that the amount of
which was much lower during the last ice age (Fairbanks, 1989). underthrust sediments, even in the frontal part of the prism, might
Ohmori (1978) pointed out that the altitude of the mountains is be one or two orders of magnitude smaller than accreted sedi-
the primary factor controlling the erosion rate and the lowest part ments. Only the distal parts of turbidites transported far from the
of land does not seriously contribute to sediment production but trench and deposited as hemipelagic sediments will come back to
just affects transportation and accumulation. the trench and will underthrust, together with biogenic and chemi-
All these observations by Ohmori (1978) suggest that the cally precipitated pelagic sediments overlying oceanic basement
denudation rate during the last ice age was not very different from of the subducting plate. Subducting sediments are much smaller
the present interglacial. The intensification of the Asian monsoon in volume because underplating takes place beneath the frontally
controlled by global-scale atmospheric currents can play back to accreted sediments. Thus the amount of sediment subduction at
the beginning of the Quaternary Period. In addition, the present the toe of the deformation front in the eastern Nankai Trough
tectonic framework of the Japanese islands also dates back to the might be 1–10 × 105 m3/yr along a unit kilometer along the trench,
beginning of the Quaternary (Huzita, 1980; Honza and Fujioka, because the difference between the sediment supply rate of 1.72 ×
2004). Thus the linkage among mountain building, denudation 107 m3/yr and the accretion rate of 1.68 × 107 m3/yr is the amount
rates, and sediment supply into the trench might be applicable at that should subduct along 300 km of the trench axis. This value is
least to the Quaternary Period. This hypothesis is to be tested by one or two orders of magnitude smaller than the value estimated
studying the sedimentation history in the Nankai Trough and the for the erosional margins (von Huene and Scholl, 1991; Clift and
northern Shikoku Basin, where drilling in the framework of the Vannucchi, 2004), although the value is estimated without taking
Integrated Ocean Drilling Program will soon operate. into account the amount of basal erosion or underplating.

Onset of Large Earthquakes, the Seismogenic Zone, and Long-Term Feedback System among Tectonic,
Growth of the Accretionary Prism Sedimentary, and Climatic Processes

The growth of the accretionary prism is important for the A generic relationship among tectonic, sedimentary, and cli-
onset of the rupture of large earthquakes in subduction zones (Ruff matic processes is shown in Figure 8. An important link between
and Kanamori, 1980). Sediments covering the subducting oceanic the tectonic and sedimentary processes is the enhancement of
402 Kimura et al.

Tectonic process Sedimentary process Climatic process


Collision
Glacier
Mountain building Surface erosion
Heavy rainfall
Sedimentary transport Figure 8. Diagram showing linkage
among tectonic, sedimentary, and cli-
Accretionary prism Trench filling matic systems.

Large earthquake Shaking & Tsunami


>Mw=8.0
Smooth the plate interface?
Fluid induced rupture propagation/lubrication?

Change the climate/ocean current system

surface erosion owing to mountain uplift and the resulting growth Huzita, K., 1980, Role of the Median Tectonic Line in the Quaternary tecton-
of the accretionary prism owing to sediment transport into the ics of the Japanese Islands: Memoirs of the Geological Society of Japan,
v. 18, p. 129–153.
trench. The growth of the accretionary prism is a key for large- Ike, T., Park, J., Kaneda, Y., and Moore, G., 2005, Along-strike variations in
earthquake generation in subduction zones, as discussed above, the structure of the Nankai Trough accretionary prism: Correlation with
and then strong shaking and tsunamis further enhance surface variations in subducting basement topography and overlying sediments:
Eos (Transactions, American Geophysical Union), v. 86, Fall Meeting
erosion in the mountains on land. Thus, tectonic events and sedi- Supplement, p. T13B–0461.
ment production are interrelated at long-term geological scales. Kaizuka, S., 1975, Tectonic model for the morphology of arc trench systems,
especially for the echelon ridges and mid-arc fault: Japanese Journal of
Geology and Geography, v. 45, p. 9–28.
ACKNOWLEDGMENTS Kimura, G., and Tamaki, K., 1986, Collision, rotation, and back-arc spreading
in the region of the Okhotsk and Japan Sea: Tectonics, v. 50, p. 389–401.
This study was supported by the Plate Dynamics Program of the Kimura, G., Kitamura, Y., Hashimoto, Y., Yamaguchi, A., Shibata, T., Ujiie, K.,
and Okamoto, S., 2007, Primary switch of plate boundary fault around
Japan Agency for Marine-Earth Science & Technology, and the the up-dip limit of seismogenic subduction zone and its relationship
21st Century Center of Excellence Program of the University of with change in wedge taper: Earth and Planetary Science Letters, v. 255,
Tokyo. Discussions with G.F. Moore, Y. Kaneda, T. Shibata, S. p. 471–484.
Kunimi, T., Takano, Y., Suzuki, M., Saitou, T., Narita, T., and Okamura, S.,
Okamoto, and T. Ike were quite useful. Reviews by A. Draut, C. 2001, Vertical crustal movements in Japan estimated from the leveling
Wobus, P. Clift, and G.F. Moore improved the early draft. We observations data for the past 100 years: Journal of the Geographical Sur-
appreciate all these persons. vey Institute, v. 96, p. 23–37.
Matsuda, T., and Uyeda, S., 1971, On the pacific-type orogeny and its model-
extension of paired belts concept and possible origin of marginal seas:
REFERENCES CITED Tectonophysics, v. 11, p. 5–27, doi: 10.1016/0040-1951(71)90076-X.
Minato, M., Gorai, M., and Hunahashi, M., 1965, The Geological Development
of the Japanese Islands: Tokyo, Tsukiji-Shokan, 442 p.
Ando, M., 1975, Source mechanisms and tectonic significance of historical Moore, G.F., Taira, A., Klaus, A., Becker, L., Boeckel, B., Cragg, B.A., Dean,
earthquakes along Nankai Trough, Japan: Tectonophysics, v. 27, p. 119– A., Fergusson, C.L., Henry, P., Hirano, S., Hisamitsu, T., Hunze, S.,
140, doi: 10.1016/0040-1951(75)90102-X. Kastner, M., Maltman, A.J., Morgan, J.K., Murakami, Y., Saffer, D.M.,
Bray, C.J., and Karig, D.E., 1985, Porosity of sediments in accretionary prisms Sanchez-Gomez, M., Screaton, E.J., Smith, D.C., Spivack, A.J., Steurer,
and some implications for dewatering processes: Journal of Geophysical J., Tobin, H.J., Ujiie, K., Underwood, M.B., and Wilson, M., 2001, New
Research, v. 90, p. 768–778. insights into deformation and fluid flow processes in the Nankai Trough
Clift, P., and Vannucchi, P., 2004, Controls on tectonic accretion versus erosion accretionary prism: Results of Ocean Drilling Program Leg 190: Geo-
in subduction zones: Implications for the origin and recycling of the con- chemistry, Geophysics, Geosystems, v. 2., no. 2001GC000166.
tinental crust: Reviews in Geophysics, v. 42, no. RG2001. Moore, J.C., and Saffer, D., 2001, Updip limit of the seismogenic zone
Dadson, S.J., Hovius, N., Chen, H.G., Dade, W.B., Hsieh, M.L., Willett, S.D., beneath the accretionary prism of southwest Japan: An effect of diage-
Hu, J.C., Horng, M.J., Chen, M.C., Stark, C.P., Lague, D., and Lin, J.C., netic to low-grade metamorphic processes and increasing effective stress:
2003, Links between erosion, runoff variability and seismicity in the Tai- Geology, v. 29, p. 183–186, doi: 10.1130/0091-7613(2001)029<0183:
wan orogen: Nature, v. 426, p. 648–651, doi: 10.1038/nature02150. ULOTSZ>2.0.CO;2.
Davis, J.D., Suppe, F.A., and Dahlen, Q., 1983, Mechanics of fold-and-thrust Nakamura, K., 1983, Possible nascent trench along the eastern Japan Sea as
belts and accretionary wedges: Journal of Geophysical Research, v. 88, the convergent boundary between Eurasian and North American plates:
p. 1153–1172. University of Tokyo, Bulletin of the Earthquake Research Institution, v.
Fairbanks, R.G., 1989, A 17,000-year glacio-eustatic sea level record: Influence 58, p. 711–722 (in Japanese).
of glacial melting rates on Younger Dryas events and deep ocean circula- Nakamura, K., and Uyeda, S., 1980, Stress gradient in arc–back arc regions and
tion: Nature, v. 342, p. 637–642, doi: 10.1038/342637a0. plate subduction: Journal of Geophysical Research, v. 85, p. 6419–6428.
Honza, E., and Fujioka, K., 2004, Formation of arcs and backarc basins inferred Ohmori, H., 1978, Relief structure of the Japanese mountains and their stages
from the tectonic evolution of Southeast Asia since the Late Cretaceous: in geomorphic development: University of Tokyo, Bulletin of the Depart-
Tectonophysics, v. 384, p. 23–53, doi: 10.1016/j.tecto.2004.02.006. ment of Geography, v. 10, p. 31–85.
Mountain building, surface erosion, and growth of an accretionary prism 403

Park, J.O., Tsuru, T., Kodaira, S., Cummins, P.R., and Kaneda, Y., 2002a, Taira, A., Tashiro, M., Okamura, M., and Katto, J., 1980, The geology of the
Splay fault branching along the Nankai subduction zone: Science, v. 297, Shimanto Belt in Kochi Prefecture, Shikoku, Japan, in Taira, A., and
p. 1157–1160, doi: 10.1126/science.1074111. Tashiro, M., eds., Geology and Paleontology of the Shimanto Belt: Kochi,
Park, J.O., Tsuru, T., Takahashi, N., Hori, T., Kodaira, S., Nakanishi, A., Miura, Japan, Rinyakosaikai Press, p. 319–389.
S., and Kaneda, Y., 2002b, A deep strong reflector in the Nankai accretion- Taira, A., Katto, J., Tashiro, M., Okamura, M., and Kodama, K., 1988, The
ary wedge from multichannel seismic data: Implications for underplating Shimanto Belt in Shikoku, Japan—Evolution of Cretaceous to Miocene
and interseismic shear stress release: Journal of Geophysical Research– accretionary prism: Modern Geology, v. 12, p. 5–46.
Solid Earth, v. 107, no. B4, doi:10.1029/2001JB00026. Taira, A., Hill, I., Firth, J., Berner, U., Bruckmann, W., Byrne, T., Chabern-
Reiners, P.W., Ehlers, T.A., Mitchell, S.G., and Montgomery, D.R., 2003, Cou- aud, T., Fisher, A., Foucher, J.-P., Gamo, T., Gieskes, J., Hyndman, R.,
pled spatial variations in precipitation and long-term erosion rates across Karig, D., Kastner, M., Kato, Y., Lallemand, S., Lu, R., Maltaman, A.,
the Washington Cascades: Nature, v. 426, p. 645–647, doi: 10.1038/ Moore, G., Moran, K., Olaffson, G., Owens, W., Pickering, K., Siena,
nature02111. F., Taylor, E., Underwood, M., Wilkinson, C., Yamano, M., and Zhang,
Ruff, L., and Kanamori, H., 1980, Seismicity and the subduction process: J., 1992, Sediment deformation and hydrogeology of the Nankai Trough
Physics of the Earth and Planetary Interiors, v. 23, p. 240–252, doi: accretionary prism: Synthesis of shipboard results of ODP Leg 131: Earth
10.1016/0031-9201(80)90117-X. and Planetary Science Letters, v. 109, p. 431–450, doi: 10.1016/0012-
Saffer, D.M., and Bekins, B.A., 2002, Hydrologic controls on the morphology 821X(92)90104-4.
and mechanics of accretionary wedges: Geology, v. 30, p. 271–274, doi: Takahashi, M., 2006, Tectonic boundary between Northeast and Southwest
10.1130/0091-7613(2002)030<0271:HCOTMA>2.0.CO;2. Japan Arcs during Japan Sea opening: Journal of Geological Society of
Sagiya, T., 2004, A decade of GEONET: 1994–2003—The continuous GPS Japan, v. 112, p. 14–32.
observation in Japan and its impact on earthquake studies (Review): Earth Tokuyama, H., Ashi, J., Soh, W., Kuramoto, S., and Ikeda, Y., 1999, Active Sub-
Planets Space, v. 56, p. xxix–xli. marine Faults off Tokai: Tokyo, University of Tokyo Press, 151 p.
Savostin, L.A., and Karasik, A.M., 1981, Recent plate-tectonics of the Arctic Tsunakawa, H., and Takeuchi, A., 1986, Tertiary paleo-stress field and volcanic
Basin and of Northeastern Asia: Tectonophysics, v. 74, p. 111–145, doi: activity of the Japanese islands: Kagaku, v. 50, p. 624–631.
10.1016/0040-1951(81)90131-1. von Huene, R., and Scholl, D.W., 1991, Observations at convergent margins
Seno, T., 1985, Is northern Honshu a microplate?: Tectonophysics, v. 115, concerning sediment subduction, subduction erosion, and the growth of
p. 177–196, doi: 10.1016/0040-1951(85)90138-6. continental-crust: Reviews of Geophysics, v. 29, p. 279–316.
Seno, T., 1989, Philippine Sea plate kinematics: Modern Geology, v. 14, Wang, K., and Hu, Y., 2006, Accretionary prisms in subduction earthquake
p. 87–97. cycles: The theory of dynamic Coulomb wedge: Journal of Geophysical
Shibata, T., Orihashi, Y., Kimura, G., and Hashimoto, Y., 2006, Intermittent Research, vol. 111, no. B6, doi:10.1029/2005JB004094.
underplating and sediment subduction—Implication of U-Pb dating of Westbrook, G.K., 1994, Growth of accretionary wedges off Vancouver Island
zircons from the Shimanto accretionary complex, SW Japan: EOS Trans. and Oregon, in Westbrook, G.K., Carson, B., and Musgrave R.J., et al.,
AGU, v. 87, Fall Meet. Suppl., Abstract T33C-0524. Proceedings of the Ocean Drilling Program, Initial Reports, Volume 146:
Sugimura, A., 1972, Plate boundaries in the vicinity of Japan: Kagaku, v. 42, College Station, Texas (Ocean Drilling Program), p. 381–388.
p. 192–202 (in Japanese). Yoshikawa, T., 1974, Denudation and tectonic movement in contemporary
Taira, A., and Niitsuma, N., 1986, Turbidite sedimentation in the Nankai Trough Japan: Bulletin of the Department of Geography: University of Tokyo,
as interpreted from magnetic fabric, grain size, and detrital modal analy- v. 6, p. 1–14.
ses: Washington (U.S. Government Printing Office), Initial Reports of the
Deep Sea Drilling Project, v. 87, p. 611–632. MANUSCRIPT ACCEPTED BY THE SOCIETY 24 APRIL 2007

Printed in the USA

You might also like