You are on page 1of 389

PROVENANCE, PALEOGEOGRAPHY, AND MASS-MOVEMENT OF DEEPWATER DEPOSITIONAL SYSTEMS IN ARC-ADJACENT BASINS:

THE CRETACEOUS-PALEOGENE CALIFORNIA FOREARC AND UPPER


MIOCENE MOHAKATINO FORMATION, NEW ZEALAND

A DISSERTATION
SUBMITTED TO THE DEPARTMENT OF
GEOLOGICAL AND ENVIRONMENTAL SCIENCES
AND THE COMMITTEE ON GRADUATE STUDIES
OF STANFORD UNIVERSITY
IN PARTIAL FULFILLMENT OF THE REQUIREMENTS
FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

Glenn Russell Sharman


August 2014

2014 by Glenn Sharman. All Rights Reserved.


Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons AttributionNoncommercial 3.0 United States License.
http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/xy840gy0004


Includes supplemental files:
1. Appendix A-4 Cumulative Detrital Zircon U-Pb Age Distributions (Appendix A-4.pdf)
2. Appendices A-5 through A-8, B-5, and B-6 (Appendix Tables.xlsx)
3. Appendix C-5 Full-size cliff photomosaics - part 1 (Appendix C-5 part 1.jpg)
4. Appendix C-5 Full-size cliff photomosaics - part 2 (Appendix C-5 part 2.jpg)
5. Appendix C-6 Wave-cut platform orthophotos (Appendix C-6.jpg)

ii

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.
Stephan Graham, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.
Martin Grove

I certify that I have read this dissertation and that, in my opinion, it is fully adequate
in scope and quality as a dissertation for the degree of Doctor of Philosophy.
Donald Lowe

Approved for the Stanford University Committee on Graduate Studies.


Patricia J. Gumport, Vice Provost for Graduate Education

This signature page was generated electronically upon submission of this dissertation in
electronic format. An original signed hard copy of the signature page is on file in
University Archives.

iii

ABSTRACT
Plate convergence and associated subduction result in a dynamic setting where the
development and evolution of sedimentary systems are influenced by a variety of
autogenic and allogenic processes. Correspondingly, patterns of sedimentation within
basins that form trench-ward of magmatic arcs (i.e., forearc basins) are often complex as
a result of the interplay between interrelated factors that include: (1) magmatism and
orogenesis in the hinterland, (2) changes in subduction regime (e.g., steep versus shallow
angle of plate descent), and (3) local basin deformation.
This dissertation examines the tectonic and sedimentary processes that acted in
concert to influence the evolution of two arc-adjacent sedimentary systems: the Late
Cretaceous-Eocene California forearc (Chapters 1 and 2) and the late Miocene Taranaki
Basin, New Zealand (Chapter 3). These three chapters include investigation of a variety
of processes that occur at different temporal and spatial scales. Specifically, major themes
of this dissertation include (1) the evolution of sediment dispersal patterns in response to
changes in continental arc magmatism and subduction regime (Chapter 1), (2) the
development and early history of the central San Andreas fault system and the
paleogeographic setting of central California (Chapter 2), and (3) submarine massmovement and patterns of soft-sediment deformation within a basin-floor sedimentary
succession (Chapter 3). The data types and methods used in each chapter are diverse and
include detrital zircon U-Pb geochronology, subsurface datasets (e.g., well logs and
seismic-reflection data), outcrop description, and structural analysis.
Chapter 1 investigates the role of subduction-related tectonism in shaping
paleogeography and sediment dispersal pathways along the California continental margin
during Late Cretaceous-Eocene time. Specifically, our synthesis of new and published
detrital zircon U-Pb ages over a 2,000 km length of the southern Oregon-Californianorthern Baja forearc reveals spatial and temporal changes in sandstone provenance that
reflect evolving sediment dispersal patterns associated with the extinction of continental
margin arc magmatism and transfer of deformation to the continental interior during
latest Cretaceous-early Cenozoic Laramide low-angle subduction.

iv

Measured age distributions from Cenomanian to Campanian forearc strata


indicate the existence of a drainage divide formed by a high-standing mid-Cretaceous
Cordilleran arc that crosscut older, Late Permian-Jurassic arc segments. Progressive
influx of 125-85 Ma detrital zircon in the Great Valley forearc reflects ongoing
denudation of the Sierra Nevada batholith throughout Late Cretaceous-early Paleogene
time. In contrast, age distributions in the Peninsular Ranges forearc indicate early
denudation of the Peninsular Ranges batholith that is hypothesized to have resulted from
the initial collision of an oceanic plateau with the southern California margin; as a result,
these age distributions exhibit little change over time until delivery of extraregional
detritus to the margin in Eocene time. Maastrichtian through middle Eocene strata
preserved south of the Sierra Nevada record a pronounced shift from local to
extraregional provenance caused by the development of drainages that extended across
the breached mid-Cretaceous continental margin batholith to tap the continental interior.
This geomorphic breaching of the mid-Cretaceous arc and associated inland drainage
migration represents the culminating influence of Laramide low-angle subduction on the
continental margin and likely occurred following subduction of the Shatsky conjugate
plateau beneath the western United States.
Chapter 1 is currently in press in the Geological Society of America Bulletin. This
chapter is coauthored by Stephan Graham, Marty Grove, David Kimbrough, and James
Wright. My contributions to this chapter include conception and design of the study;
collection, processing, and analyzing samples; interpretation of the results; and writing
the manuscript. Stephan Graham and Marty Grove also contributed to the conception of
the study, provided abundant guidance, and also assisted with interpretation of the results.
Marty Grove, David Kimbrough, and James Wright contributed to collecting, processing
and analyzing the samples. All coauthors contributed via discussion and by reviewing the
manuscript.
Chapter 2 presents a reappraisal of the early history of the central San Andreas
fault and provides a revised interpretation of Eocene paleogeography in the southern San
Joaquin basin and northern Salinian block. Although the modern San Andreas fault
system is widely considered to have formed in response to the initiation of PacificNorth
v

American plate interaction ca. 27 Ma, its Paleogene history remains controversial. In
particular, 100 km of right-lateral offset between mid-Cretaceous plutonic rocks of the
northern Salinian block and the western edge of Sierra Nevada basement remains
unaccounted for after restoration of Neogene displacement along strike-slip faults of the
San Andreas system. We use detrital zircon U-Pb geochronology to invalidate a key
Paleogene piercing point and demonstrate that displaced portions of the hypothesized
Middle Eocene ButanoPoint of Rocks submarine fan were never contiguous across the
San Andreas fault. Rather, we infer that the Point of Rocks Sandstone was derived from
erosion of the ancestral Sierra Nevada Mountains and is the deep-marine equivalent of
the shelfal Famoso Sandstone and the non-marine Walker Formation. We also speculate
that a submarine canyon was positioned on the Eocene shelf-slope break and delivered
shelfal sands to the adjacent deep-marine basin over much of Middle Eocene time. The
Butano Sandstone and related Paleogene sandstone from the northern Salinian block were
derived from sources within the southeastern Sierra Nevada and/or western Mojave
region. This interpretation is supported by the abundance of Late Permian-Jurassic zircon
in these sands that lack a local source in the Salinian block.
Our revised paleogeographic interpretation implies that the northern Salinian
block was located at least 7550 km farther south in Eocene time than previously
recognized. Our data require (1) pre23 Ma dextral slip along the San Andreas fault in
central California, and/or (2) slip along a predecessor fault that formed prior to Pacific
North American plate interaction. This previously undocumented slip may indicate that
significant PacificNorth American plate interaction propagated from the plate margin
into the continental interior much earlier than conventionally believed. Alternatively, late
Paleogene slip could predate the development of the modern plate boundary and
represent inboard dextral strike-slip displacement along the eastern margin of the Salinian
block, similar to the deformation that occurs today along the strike-slip Sumatra fault
system.
Chapter 2 has been published in GEOLOGY, v. 41, July 2013. This chapter is
coauthored by Stephan Graham, Marty Grove, and Jeremy Hourigan. My contributions to
this chapter include designing the study; collecting, processing, and analyzing the
samples; interpreting the results; and writing the manuscript. Stephan Graham also
vi

contributed by assisting with conception and design of the study, collecting some of the
samples, interpreting results, and providing guidance. Marty Grove assisted with data
reduction and interpreting the results. Jeremy Hourigan assisted with data acquisition and
reduction. All coauthors contributed to the study by discussing the results and reviewing
the manuscript.
A portion of Chapter 2 has also been published in by the American Association of
Petroleum Geologists Pacific Section in the field trip guidebook Westside Turbidite
Field Trip San Joaquin Valley California as part of the 2014 AAPG Pacific Section
Convention held in Bakersfield, California. This article is included as Appendix B-1 and
is coauthored by Stephan Graham and Christopher Sine. Stephan Grahams contributions
are the same as described above. Christopher Sines contributions include assisting with
interpretation, particularly via insights from geophysical interpretation of seismic data in
the southern San Joaquin basin.
Chapter 3 explores the processes and products related to submarine massmovement within the volcaniclastic upper Miocene Mohakatino Formation (Taranaki
Basin, New Zealand). This chapter documents the spectacular exposures of the North
Awakino mass-transport deposit (MTD) that outcrops along the northern Taranaki
coastline for ~11 km in wave-cut platforms and in cliffs up to 100 m high. Spectacular
deformation features associated with submarine mass-movement are developed in
remobilized sediment gravity flow deposits that initially accumulated within a lowgradient, unconfined intra-slope basin. Sedimentary facies within the North Awakino
MTD range from laterally continuous, thick- to thin-bedded volcaniclastic sandstone and
mudstone and correspond with distinct style of deformation: folds developed in thickbedded sandstone are larger (fold heights up to 10s of m) and more laterally continuous
(up to 1 km) than those developed in thinner-bedded facies.
Regional geologic relationships suggest that exposures of the North Awakino
MTD extend nearly across the full width of the MTD and thus provide a rare opportunity
to observe lateral relationships between the marginal and central portions of the MTD.
We conduct a rigorous paleoslope analysis of slump fold, fault, and bedding orientations
using both existing and newly proposed methodologies. Separate analysis of seven
vii

subregions within the North Awakino MTD reveals that the predicted MTD transport
direction varies widely along the outcrop extent. Most notably, slump deformation
structures within the inferred margins have mean orientations that are sub-orthogonal to
those within the central portions of the MTD. This relationship is hypothesized to be a
consequence of edge effects that may be related to lateral compression along the
margins of the MTD. Our analysis demonstrates the importance of accounting for spatial
heterogeneity in deformation structure orientations when determining the paleoslope
orientation through kinematic analysis. Our inference of W-directed translation suggests
that the North Awakino MTD formed in response to a local reversal of the bathymetric
slope that was likely a result of tectonically-induced basin deformation.
Chapter 3 will be submitted to GEOSPHERE. This chapter is coauthored by
Stephan Graham, Larisa Masalimova, Peter King, and Lauren Shumaker. My
contributions to this chapter include conceiving and designing the study, conducting field
work, analyzing and interpreting the results, and writing the manuscript. Stephan Graham
and Peter King assisted with conceiving the study and provided guidance and discussion.
Larisa Masalimova and Lauren Shumaker assisted with the field work and also
contributed via discussion, ideas, and data.

viii

ACKNOLWEDGEMENTS

I have been extremely fortunate to have gotten to know so many great people over
the course of my time at Stanford. These close friends and colleagues have made
graduate school both fun and memorable and also have greatly enriched my PhD
research. First of all, I would like to thank Steve Graham for being a fantastic advisor and
mentor. Steve has been a huge source of support and encouragement, both scientifically
and personally, during my PhD. It has been a pleasure getting to learn from Steves vast
knowledge of California geology, and spending time with Steve in the field has been
particularly memorable (including the numerous trips to Wagon Caves Rock). In
addition, I would like to thank my other committee members, Marty Grove, Don Lowe,
and George Hilley, for their contributions to my research and to my education at
Stanford, in general. Marty was very generous with sharing his knowledge and was
always happy to have extended conversations about California geology. I extend my
thanks to Don for his excellent classes that helped me gain knowledge and confidence in
topics related to sedimentology. I have particularly enjoyed learning from Don in the
field, including in Saudi Arabia, New Zealand, Arkansas, and Death Valley, California. I
thank George for always being willing to discuss my projects, and for challenging me to
become a better scientist. I also thank Tim McHargue for his excellent classes, seminars,
and field trips that he has led over these past five years.
My research would not have been possible without the support and funding
provided by the Stanford Project on Deep-Water Depositional Systems that has included
the following companies: Aera Energy, Anadarko, BHP Billiton, Chevron,
ConocoPhillips, ENI, Hess, Karoon Gas Australia Ltd., Marathon, Neos, Nexen,
Occidental Petroleum, Petrobras, PTT Exploration and Production Public Company Ltd.,
RAG, Reliance, Saudi Aramco, Schlumberger, Shell, Statoil, Talisman Energy, and
Venoco. Additional research funding was provided by the Geological Society of America
and by the Stanford School of Earth Sciences Chevron Fund. I thank the Geological
Society of America, American Association of Petroleum Geologists, International
Association of Sedimentologists, and the Stanford Shell Fund for providing funds to
attend professional meetings and conferences.
ix

There have been numerous people who have contributed to my California


research projects. In particular, Id like to thank Oxy and its California Exploration group
including Kurt Neher, Dave Defelice, Kurt Johnson, Ingvar Alarik, Christopher Sine, Bob
Stupp, Gregg Pyke, Mike Johns, and the rest of the team. Thanks to Bob McLaughlin and
Rick Stanley at the USGS for helpful discussions and for helping me collect samples in
the southern Santa Cruz Mountains. Kathleen Baldwin helped me gain access to surpless
Quad Sandstone at the Stanford University boneyard. I thank Matt Coble for help in the
SHRIMP lab and Jeremy Hourigan for his help at the University of California Santa Cruz
ICPMS lab. Ive benefited from discussions with and feedback from the following
individuals: Al Almgren, Alex Lechler, Andy Barth, Earl Brabb, Liz Cassel, Bill
Dickinson, Trevor Dumitru, Amalia Doebbert, Carl Jacobson, Gail Mahood, Elizabeth
Miller, Jason Saleeby, Kathy Surpless, Joe Wooden, and Nick Van Buer. Many
additional people helped me with collecting, processing, and analyzing samples: Blair
Burgreen, Anne Bernhardt, Elizabeth Cassel, Wells Chang, Julie Fosdick, Nicky Giesler,
Eric Gottlieb, Katie Maier, Matt Malkowski, Sara Maatta, Tim OBrien, Mark Pecha,
Nicole Sarto, and Theresa Schwartz. In particular, I thank Trevor Dumitru for his help in
the mineral separation lab. Thanks to the Twisselman family and the Wildlands
Conservancy for allowing access to private land for sample collecting.
My research in New Zealand was greatly assisted through the help and expertise
of local scientists. I extend a sincere thanks to GNS Science for supporting my field
efforts over the years. Peter King, Malcolm Arnot, and Greg Browne were generous with
their time and knowledge of my field area. I thank Martin Crundwell for taking the time
to visit my field area, helping me collect biostratigraphic samples, and for personally
analyzing the samples. I would like to thank Lorna Strachan, from the University of
Auckland, for helpful advice and for going into the field with me several times. In
addition, Lorna and Tim Debacker were kind enough to take me into the field north of
Auckland to see the spectacular mass-transport deposits that outcrop there.
I owe a considerable debt of gratitude to the many Kiwis who made my time in
New Zealand sweet as. My personal opinion is that New Zealanders are most
consistently friendly and helpful people I have ever come across, and possibly in the
entire world. First of all, my research would not have been possible without permission to
x

access the coast through private land. For such access I thank Karl Reipen at Awakino
Estate, Crain Rain at Paparahia Station, and Alistair and Maree Bryant at Onetai Station.
The Bryant family was particularly welcoming, allowing us to use their quad bike and
guest house. We also remember and say thanks to Tom for his help during my first
field season in 2010.
I also thank Dawn and Neil Colman, who went out of their way to help me during
my first field season in New Zealand; riding up the Mokau River on their historic cream
boat was a major highlight. Alan Jones and his family contributed to my thesis in a major
way by taking me out on a boat to photograph the coastal cliffs. John and Angela Potroz
also made a really big impact on my time spent along the Taranaki coast. Besides
allowing access on their land to look at the lower Mount Messenger Formation, they
allowed us to stay in their shearers quarters during wet weather. John teaching me how
to wrestle a sheep to the ground for shearing will be lasting memory. Id also like to
thank Daimen for storing our equipment in between field seasons, and for teaching me
how to milk a cow. Id like to extend a special thanks to Shane, Jenny, and Graham
Marsden at the Whitebait Inn. This campground, ice cream shop, and restaurant felt like
home after all the weeks spent there, mainly because of the locals who took care of us by
delivering messages, warning us about the weather bomb, and baking some seriously
delicious muffins and whitebait fritters. Id like to specifically thank Shane and his dogs
for being particularly good friends during my time in New Zealand. I enjoyed hanging
out in the evenings, learning about possum and pig hunting, and tasting wild pig (which
was delicious).
The GES office staff has been completely essential for my graduate studies, and I
would like the thank Yvonne Lopez, Leslie Honda, Elaine Anderson, Siegrid Munda,
Robyn Maslin, Arlene Abucay, Alyssa Ferree, Lauren Nelson, Lorraine Sandoval, Javier
Illueca, and Stephanie Matthews. Id also like to thank the staff of the Branner library
including Julie Sweetkind-Singer, Hannah Winkler, Patricia Carbajales, and Jane Ingalls.
My fellow students and other friends at Stanford have made an incredibly positive
impact on my time at Stanford. I thank the past students who made my earlier years at
Stanford really memorable: Zane Jobe, Dominic Armitage, Julie Fosdick, Anne
Bernhardt, Lisa Stright, Katie Maier, Keisha Durant, and Jon Rotzien. I owe particular
xi

thanks to Larisa Masalimova, who has been a great friend and field partner for two
seasons in New Zealand. Larisa endured numerous pre-dawn mornings, countless gate
openings, lots of rain, heavy equipment lugging, and endless strike and dip
measurements. I would also like to thank Larisas mom, for making sure we didnt get
back from the field too late at night and also cooking delicious Russian meals.
Id like to thank my officemates for the past several years: Blair Burgreen, Tess
Menotti, and Theresa Schwartz. First of all, I thank Blair for being a great field partner
during our first field season in 2010. Together we explored a good portion of the North
Islands eastern and western coastlines, and in doing so ended up kayaking, spelunking,
boating, beach walking, and fishing, all in the name of geology. Blair put up with a lot of
occupational hazards in New Zealand, including vicious seals, left-side driving, rogue
geese, inquisitive pigs, big waves, and a tide table-misreading field assistant. Thanks to
Tess who helped me for a few days during my final New Zealand field season. In
particular, thanks for not completely cutting off your finger with my pocket knife, and for
thoroughly documenting the coastal cliffs through numerous selfies. Thank you Theresa
for keeping me on my toes with occasional outbursts of swearing and being willing to
trudge uphill through poison oak bushes to photograph our favorite outcrop. It was really
fun getting to hang out in Argentina this past year and learn what shallow-marine trace
fossils look like.
There have been a host of good friends who have made my time at Stanford what
it has been. Matt Malkowski has been a particularly great friend over the years, and we
have had some really memorable times in Patagonia (for example, hanging out in
Estancia Nibepo Aikes puesto and dining on Oliviers cordero patagonico). Im really
glad that we didnt fall off a cliff, break our ankles on the boulders around Lago Frias, or
get gored by one of those feral bulls. Thanks also to Bethany for bringing out the yogi in
me, and for letting Matt and me win Rook most of the time. I thank Lauren Shumaker
and Nora Nieminski who helped me map my wave-cut platforms for a few days during
my final New Zealand field season. We also had some great adventures exploring the
coastline to the north and bushwhacking through impenetrable New Zealand jungle.
Thanks to Matt Thomas for his dissertation-finishing advice and for the excellent home
brew. I also thank the other students in the Stanfords SedGroup who have been a lot of
xii

fun to spend time with: Sam, Lizzy, Danielle, Moy, Zach, Cody, Nadja, Nilay, Wisam,
and Lauren Schultz. Id also like to thank those in my church small group who have been
so encouraging: the Nelsons, Cromies, Albaneses, Alcorns, Reynolds, McIntoshs, and
Rochas.
Finally, I would also like to acknowledge and thank my family, who has always
been a constant source of support and inspiration: mom, dad, Stephanie, Andrew, Kristel,
Tom, Jonathan, and of course, Sam, Sadie, Kate, and Ian. And most of all, Id like to
thank my lovely wife and best friend, Joni, for her consistent love and support over the
years. Joni has walked beside me every step of the way, despite long field seasons and
internships, and has blessed my life beyond description.

xiii

TABLE OF CONTENTS

ABSTRACT ..................................................................................................................... IV
ACKNOLWEDGEMENTS ........................................................................................... IX
TABLE OF CONTENTS ............................................................................................ XIV
LIST OF TABLES ....................................................................................................... XIX
LIST OF FIGURES ....................................................................................................... XX

CHAPTER 1: DETRITAL ZIRCON PROVENANCE OF THE LATE


CRETACEOUS-EOCENE CALIFORNIA FOREARC: INFLUENCE OF
LARAMIDE LOW-ANGLE SUBDUCTION ON SEDIMENT DISPERSAL AND
PALEOGEOGRAPHY......................................................................................................1
ABSTRACT .........................................................................................................................2
INTRODUCTION ...............................................................................................................3
GEOLOGIC BACKGROUND ............................................................................................5
Cordilleran Magmatism....................................................................................................5
Tectonic Restructuring of the Southern California Margin .............................................6
Upper Cretaceous-Eocene Forearc ...................................................................................7
Palinspastic Reconstruction..............................................................................................7
METHODS AND DATA SOURCES .................................................................................9
Detrital Zircon U-Pb Ages ...............................................................................................9
RESULTS ..........................................................................................................................11
Detrital Zircon Geochronology ......................................................................................11
SEDIMENTARY PROVENANCE ANALYSIS ..............................................................13
Pre-Permian (> ca. 300 Ma) zircon: ...............................................................................13
Permian-Triassic (ca. 285-225 Ma) zircon:....................................................................14
Jurassic to earliest Cretaceous (ca. 180-135 Ma) zircon: ...............................................15
Mid-Cretaceous (ca. 125-85 Ma) zircon: .......................................................................15
xiv

Latest Cretaceous (85-65 Ma) zircon: ............................................................................16


Paleogene (ca. 65-40 Ma) zircon: ..................................................................................16
DISCUSSION ....................................................................................................................17
Temporal and Spatial Evolution of Forearc Provenance and Paleogeography ..............17
Forearc-Subduction Complex-Underplated Schist Comparison ....................................26
Influence of Underlying Geometry of the Jurassic and Mid-Cretaceous Arcs on the
Evolution of Forearc Detrital Zircon Age Distributions ................................................27
Why is Post-100 Ma Zircon Scarce in the Cenomanian-Coniacian Great Valley
Forearc? ..........................................................................................................................28
Influence of Low-Angle Subduction on the Evolution of an Arc-Forearc System ........30
Influence on Future Development of the San Andreas Fault System ............................34
CONCLUSIONS................................................................................................................35
ACKNOWLEDGEMENTS ...............................................................................................36
REFERENCES ..................................................................................................................36
CHAPTER 2: A REAPPRAISAL OF THE EARLY SLIP HISTORY OF THE
CENTRAL SAN ANDREAS FAULT, CENTRAL CALIFORNIA, USA ..................76
ABSTRACT .......................................................................................................................77
INTRODUCTION .............................................................................................................78
PIERCING POINTS ..........................................................................................................79
DETRITAL ZIRCON RESULTS ......................................................................................79
SEDIMENTARY PROVENANCE ...................................................................................80
DISCUSSION ....................................................................................................................81
TECTONIC IMPLICATIONS...........................................................................................82
ACKNOWLEDGEMENTS ...............................................................................................83
REFERENCES ..................................................................................................................84
CHAPTER 3: SPATIAL PATTERNS OF DEFORMATION AND PALEOSLOPE
ESTIMATION WITHIN THE MARGINAL AND CENTRAL PORTIONS OF A
BASIN-FLOOR MASS-TRANSPORT DEPOSIT, TARANAKI BASIN, NEW
ZEALAND ........................................................................................................................96
xv

ABSTRACT .......................................................................................................................97
INTRODUCTION .............................................................................................................98
GEOLOGIC BACKGROUND ........................................................................................100
METHODS ......................................................................................................................101
Measured Stratigraphy .................................................................................................102
Structural Measurements and Kinematic Analysis ......................................................102
Photographic Documentation and 3D Model Generation ............................................103
NORTH AWAKINO MTD .............................................................................................104
Overview ......................................................................................................................104
MTD Thickness and Extent..........................................................................................105
Stratigraphy and Sedimentology ..................................................................................107
Slump Deformation Structures .....................................................................................112
PALEOSLOPE ANALYSIS ............................................................................................115
Methods Based Upon Slump Fold Orientations ...........................................................116
Methods Based Upon Slump Fault Orientations ..........................................................123
Mean Bedding Strike Method ......................................................................................127
Comparison between Paleoslope Methods ...................................................................128
DISCUSSION ..................................................................................................................131
Transport Direction of the NAMTD ............................................................................131
Why Are Folds and Faults within the MTD Margins Oriented Parallel to the Inferred
Downslope Direction? ..................................................................................................133
Spatial Variability in MTD Structural Architecture .....................................................137
Implications for Paleoslope Analysis ...........................................................................137
Paleogeography of the Northeastern Taranaki Basin ...................................................139
Mass-Movement within Basin-Floor Successions .......................................................141
CONCLUSIONS..............................................................................................................142
ACKNOWLEDGEMENTS .............................................................................................144
REFERENCES ................................................................................................................144

xvi

APPENDICES

APPENDIX A: SUPPLEMENTARY MATERIAL FOR CHAPTER 1


APPENDIX A-1: Geologic Map Data Sources..221
APPENDIX A-2: Details Regarding the Palinspastic Reconstruction in Figure 2B.....225
APPENDIX A-3: Supplementary Description of Sample Locations.....232
APPENDIX A-4: Cumulative Detrital Zircon U-Pb Age Distributions (available
electronically)...269
APPENDIX A-5: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of Arizona Laserchron Center (available electronically)270
APPENDIX A-6: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of California Santa Cruz (available electronically).....271
APPENDIX A-7: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of California Los Angeles (available electronically)..272
APPENDIX A-8: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
Stanford-USGS SHRIMP (available electronically)....273
APPENDIX A-9: Detrital Zircon Age Population Proportions..274
APPENDIX B: SUPPLEMENTARY MATERIAL FOR CHAPTER 2
APPENDXI B-1: A New Paleogeographic Model for the Point of Rocks Sandstone, San
Joaquin Basin, California.............281
INTRODUCTION.281
DETRITAL ZIRCON U-Pb GEOCHRONOLOGY.283
SUBSURFACE INVESTIGATION.....286
MIDDLE EOCENE PALEOGEOGRAPHY.288
CONCLUSIONS...290
ACKNOWLEDGEMENTS...291
REFERENCES......291
APPENDIX B-2: Analytical Methods....309
APPENDIX B-3: Kolgomorov-Smirnov Statistics.313
xvii

APPENDIX B-4: Discussion of northern Gabilan Range western San Emigdio


Mountain piercing points...314
APPENDIX B-5: U-Th-Pb isotope composition of detrital zircons analyzed at the
University of Arizona Laserchron Center (available
electronically)....317
APPENDIX B-6: U-Th-Pb Isotopic Compositions of Detrital Zircons Analyzed at the
University of California Santa Cruz (available electronically).318
APPENDIX B-7: Explanation of tectonic and depositional events shown in Figure
4..319
APPENDIX B-8: Explanation of San Andreas fault piercing points used in Figure
4..320
APPENDIX C: SUPPLEMENTARY MATERIAL FOR CHAPTER 3
APPENDIX C-1: Detailed measured sections..324
APPENDIX C-2: Slump fold orientations....342
APPENDIX C-3: Slump fault orientations...346
APPENDIX C-4: Unrestored versus restored slump fault orientations353
APPENDIX C-5: Full-size cliff photomosaics (available electronically)....354
APPENDIX C-6: Wave-cut platform orthophotos (available electronically)..355
APPENDIX C-7: Biostratigraphy.356
APPENDIX C-8: Paleocurrent measurements..364

xviii

LIST OF TABLES
CHAPTER 1
Table 1. Detrital Zircon Samples ..................................................................................... 49
Table 2. Detrital Zircon U-Pb Age Populations ............................................................... 59
CHAPTER 2
Table 1. Sample locations ................................................................................................ 87
Table 2. Kolgomorov-Smirnov (K-S) P-values and D-values of detrital zircon U-Pb age
distributions....................................................................................................................... 88
CHAPTER 3
Table 1. Summary of Paleoslope Analysis Methods...................................................... 214
Table 2. Summary of Paleoslope Analysis Results ........................................................ 217
Table 3. Criteria for Identification of the Lateral/Oblique Portions of MTDs ............... 219

xix

LIST OF FIGURES
CHAPTER 1
Figure 1. Tectonic setting of the western U.S. ........................................................... 60
Figure 2. A: Generalized geologic framework of southern Oregon, California, and
northern Baja California. B: Palinspastic reconstruction of southern California
northern Baja California during Eocene time. ............................................................ 61
Figure 3. Cumulative and normalized distributions of detrital zircon U-Pb ages for
groups of Upper Cretaceous-Eocene forearc strata. ................................................... 63
Figure 4. Zircon U-Pb age distributions (< 300 Ma) for volcanic and plutonic rocks
from the Sierra Nevada batholith, Salinian block, Mojave Desert region, and
Peninsular Ranges batholith (Premo et al., 1998; Silver and Chappell, 1998; Grove et
al., 2008; Chapman et al., 2012; Barth et al., 2013; Kimbrough et al., in press) and
Cenomanian-Eocene forearc strata (see Table 1 for data sources). ............................ 66
Figure 5. Cumulative distributions of detrital zircon U-Pb ages (< 200 Ma) for
regional groups of Late Cretaceous-Eocene forearc strata. ........................................ 67
Figure 6. Cumulative and normalized distributions of detrital zircon U-Pb ages for
arc-derived detrital zircon (i.e., < 300 Ma) from groups of forearc, subduction
complex, underplated schist, intra-arc, and retroarc strata. ........................................ 69
Figure 7. A: Spatial distribution of detrital zircon age populations in CenomanianEocene forearc sandstone. Distances are measured north-to-south parallel to the
reconstructed Eocene margin (Fig. 2B). Age population abundances are linearly
interpolated between sample localities (see Figure 2A for sample location
abbreviations). B: Inferred spatial and temporal distribution of source regions that
supplied detritus to the California forearc (dashed where source contribution is minor
or uncertain). ............................................................................................................... 71
Figure 8. Generalized paleogeographic reconstruction of the Cenomanian-Eocene
continental margin. ..................................................................................................... 73
CHAPTER 2
Figure 1. Paleogeographic reconstruction of central California during Middle Eocene
time ............................................................................................................................. 90
xx

Figure 2. Cumulative and normalized distributions of detrital zircon U-Pb ages...... 91


Figure 3. Proposed tectonic and paleogeographic reconstruction of northern Salinian
block and southern Sierra Nevada during Paleogene time. ........................................ 93
Figure 4. A reappraisal of the tectonic history of the San Andreas fault (SAF) in
central California. ....................................................................................................... 95
CHAPTER 3
Figure 1. Geographic setting of New Zealand and the Taranaki Basin. .................. 151
Figure 2. Map of the northeastern Taranaki Basin showing the location of volcanoes,
major faults, seismic lines, and wells mentioned in the text (modified from Giba et al.,
2013). ........................................................................................................................ 152
Figure 3. Study area map. ........................................................................................ 153
Figure 4. Generalized stratigraphy of the middle-upper Miocene sedimentary
succession of the northeastern Taranaki Basin. ........................................................ 155
Figure 5. Measured stratigraphic sections from within the northern three kilometers
of the NAMTD.......................................................................................................... 156
Figure 6. Measured stratigraphic sections from within the southern eight kilometers
of the NAMTD.......................................................................................................... 157
Figure 7. Measured stratigraphic sections from above, below, and lateral to the
NAMTD. ................................................................................................................... 159
Figure 8. Outcrop photos and measured stratigraphic section at Opito Point.......... 160
Figure 9. Cliff photomosaics and line-drawing interpretations for the northern 5.5 km
of the study area. ....................................................................................................... 162
Figure 10. Example outputs from the 3D model generated from aerial photographs of
the NAMTD south of Waioroko Stream (subregion R4).......................................... 166
Figure 11. Aerial photographs and line-drawing interpretation of the wave-cut
platforms. .................................................................................................................. 168
Figure 12. Selected cross-line (A-B) and in-line (C) amplitude sections from the
Kahu 3D seismic survey. .......................................................................................... 173
Figure 13. Cliff photomosaic and line-drawing interpretation for the southern extent
of the NAMTD, between Paritutu and Pitone Streams. ............................................ 174
xxi

Figure 14. Examples of the thick- to medium-bedded volcaniclastic sandstone


lithofacies. ................................................................................................................. 175
Figure 15. Examples of folds within the NAMTD. ................................................. 177
Figure 16. Examples of the thin- to medium-bedded volcaniclastic sandstone and
mudstone lithofacies. ................................................................................................ 183
Figure 17. Illustration of lateral (north to south) changes in the style of deformation
within the northern 5 km of the NAMTD. ................................................................ 185
Figure 18. Profile view of an upright to recumbent anticline-syncline pair north of
Kopuai Stream. ......................................................................................................... 186
Figure 19. Histograms of fold axis plunge, interlimb angle, and axial plane dip. ... 187
Figure 20. Slump fold orientations. ......................................................................... 188
Figure 21. Fold profile analysis following Ramsay (1967). .................................... 191
Figure 22. Example cross-sections of large folds within the northern 4 kilometers of
the NAMTD. ............................................................................................................. 192
Figure 23. A series of tight to isoclinal, nearly recumbent folds north of Waioroko
Stream. ...................................................................................................................... 193
Figure 24. Examples of slump faults within the NAMTD. ...................................... 195
Figure 25. Histograms of slump fault surface dip and approximate displacement. . 199
Figure 26. Slump fault orientations.......................................................................... 200
Figure 27. Illustration of fold-based paleoslope methods. ....................................... 202
Figure 28. Scatterplots of selected fold orientation attributes for the entire NAMTD
and by subregion. ...................................................................................................... 203
Figure 29. Explanation of the best-fit girdle to fault poles method (GFPM)........... 205
Figure 30. Summary of paleoslope analysis results. ................................................ 206
Figure 31. Schematic and idealized depiction of the NAMTD................................ 208
Figure 32. Plan-view images of MTDs from the literature. ..................................... 209
Figure 33. Schematic paleogeography of the northeastern Taranaki Basin during late
Miocene time. ........................................................................................................... 212

xxii

CHAPTER 1: DETRITAL ZIRCON PROVENANCE OF THE LATE


CRETACEOUS-EOCENE CALIFORNIA FOREARC: INFLUENCE OF
LARAMIDE LOW-ANGLE SUBDUCTION ON SEDIMENT DISPERSAL AND
PALEOGEOGRAPHY

DETRITAL ZIRCON PROVENANCE OF THE LATE CRETACEOUS-EOCENE


CALIFORNIA FOREARC: INFLUENCE OF LARAMIDE LOW-ANGLE
SUBDUCTION ON SEDIMENT DISPERSAL AND PALEOGEOGRAPHY
Glenn R. Sharman1, Stephan A. Graham1, Marty Grove1, David L. Kimbrough2, and
James E. Wright3
1

Department of Geological and Environmental Sciences, Stanford University 94305

Department of Geological Sciences, San Diego State University, San Diego, California

92182
3

Department of Geology, University of Georgia, Athens, Georgia 30602

ABSTRACT
Upper Cretaceous-Eocene forearc strata deposited along the California continental
margin record a complex history of plate convergence that shaped the tectonic
development of the U.S. Cordillera. Synthesis of new and published detrital zircon U-Pb
ages over a 2,000 km length of the southern Oregon-California-northern Baja forearc
clearly demonstrates spatial and temporal changes in sandstone provenance that reflect
evolving sediment dispersal patterns associated with the extinction of continental margin
arc magmatism and transfer of deformation to the continental interior during latest
Cretaceous-early Cenozoic Laramide low-angle subduction.
Measured age distributions from Cenomanian to Campanian forearc strata
indicate the existence of a drainage divide formed by a high-standing mid-Cretaceous
Cordilleran arc that crosscut older, Late Permian-Jurassic arc segments. Progressive
influx of 125-85 Ma detrital zircon in the Great Valley forearc reflects ongoing
denudation of the Sierra Nevada batholith throughout Late Cretaceous-early Paleogene
time. In contrast, age distributions in the Peninsular Ranges forearc indicate early
denudation of the Peninsular Ranges batholith that is hypothesized to have resulted from
the initial collision of an oceanic plateau with the southern California margin; as a result,
these age distributions exhibit little change over time until delivery of extraregional
detritus to the margin in Eocene time. Maastrichtian through middle Eocene strata
2

preserved south of the Sierra Nevada record a pronounced shift from local to
extraregional provenance caused by the development of drainages that extended across
the breached mid-Cretaceous continental margin batholith to tap the continental interior.
This geomorphic breaching of the mid-Cretaceous arc and associated inland drainage
migration represents the culminating influence of Laramide low-angle subduction on the
continental margin and likely occurred following subduction of the Shatsky conjugate
plateau beneath the western United States.
INTRODUCTION
The Cretaceous-Paleogene forearc of California is perhaps one of the best-studied
examples of sedimentation in a convergent setting (e.g., Dickinson, 1995a) and preserves
a detailed record of the complex history of plate convergence that has shaped the tectonic
development of the U.S. Cordillera. Cretaceous to Paleogene subduction of the Farallon
plate beneath North America is recorded in the Sierran-Peninsular Ranges batholith
(remnant magmatic arc), adjacent forearc basins, and Franciscan subduction complex
(Fig. 1). Together these elements comprise a type example of a cross-section through an
ancient convergent margin (Dickinson, 1995a).
The California margin also contains one of the best preserved records of lowangle slab subduction, associated subduction erosion, and tectonic underplating of the
margin batholith (Grove et al., 2003a; Saleeby, 2003; Grove et al., 2008; Ducea et al.,
2009; Jacobson et al., 2011; Chapman et al., 2013). This episode of latest CretaceousPaleogene low-angle subduction has been widely linked with the development of the
Laramide orogeny in the western U.S. Cordillera (Dickinson and Snyder, 1978; Miller et
al., 1992) and may have resulted from subduction of thickened oceanic crust (Livaccari et
al., 1981; Henderson et al., 1984; Saleeby, 2003; Liu et al., 2010). Because low-angle
subduction is a widely-occurring phenomenon along convergent margins (Gutscher et al.,
2000), the California margin provides an important ancient example of the influence of
low-angle subduction on forearc sedimentation and provides valuable insight into arcforearc dynamics in analogous tectonic settings (e.g., Laursen et al., 2002; Fildani et al.,
2008).

Although the Laramide orogeny is well-understood to have been manifested in the


Cordilleran foreland as a series of basement-cored uplifts and partitioned foreland
depocenters (DeCelles, 2004), significant debate exists regarding the influence of this
tectonic event on the continental margin (e.g., Saleeby, 2003; Jacobson et al., 2011).
Margin tectonism attributed to the Laramide orogeny includes 1) ca. 85 Ma cessation of
magmatism in the Sierran-Peninsular Ranges arc accompanied by inland migration of
plutonism in the southwestern U.S. (Chen and Moore, 1982; Lipman, 1992); 2) uplift and
denudation of the mid-Cretaceous arc and an associated pulse of forearc sedimentation
(Grove et al., 2003b, 2008; Saleeby et al., 2010); 3) uplift and shoaling of the forearc and
adjacent subduction complex (Moxon and Graham, 1987; Mitchell et al., 2010); 4)
underplating of the subduction complex beneath the southern California segment of the
Cordilleran arc (Jacobson et al., 1996; Grove et al., 2003a, 2008); and 5) deep
exhumation and structural juxtaposition of the eastern portion of the mid-Cretaceous
batholith against the Franciscan subduction complex across the Nacimiento fault (Hall,
1991; Saleeby, 2003; Dickinson et al., 2005; Ducea et al., 2009).
Our approach is to use sedimentary provenance of forearc sandstone to interpret
the evolution of Late Cretaceous-Eocene sediment dispersal patterns to the southern
Oregon-California-northern Baja forearc (hereafter California forearc) and to consider
the influence of Laramide low-angle subduction on margin paleogeography and
landscape evolution. In particular, we use detrital zircon U-Pb age distributions to refine
previous

interpretations

based

on

sandstone

petrography,

conglomerate

clast

assemblages, and paleocurrent distributions (e.g., Nilsen and Clarke, 1975; Dickinson et
al., 1979; Kies and Abbott, 1982; Ingersoll, 1983; Seiders and Cox, 1992). Because the
age distribution of igneous rocks in California is generally well known (e.g., Irwin and
Wooden, 2001; Fig. 2), detrital zircon U-Pb ages can be directly linked with potential
source regions. Detrital zircon provenance analysis has already been used effectively to
improve understanding of drainage evolution along certain segments of the margin (e.g.,
DeGraaff-Surpless et al., 2002; Cassel et al., 2012). This study expands on previous work
by integrating both new and published U-Pb ages of forearc strata along ~2,000 km of the
continental margin (Fig. 2). In addition, we consider datasets from time-equivalent units
including the Franciscan subduction complex (Dumitru et al., 2010, 2012; Snow et al.,
4

2010), underplated schist (Grove et al., 2003a, 2008; Jacobson et al., 2011; Chapman et
al., 2013), and intra-arc and retroarc foreland basin strata (Barth et al., 2004; Dickinson
and Gehrels, 2008; Lechler and Niemi, 2011; Laskowski et al., 2013). By considering
regional-scale, spatial trends in sedimentary provenance from Late Cretaceous to Eocene
time, we document forearc drainages that progressively migrated inland in response to a
redistribution of topography associated with the subduction of an oceanic plateau beneath
the continental interior. As such, we place the forearc sedimentary record within the
context of an evolving tectonic framework associated with the development of Laramide
low-angle subduction.
GEOLOGIC BACKGROUND
Cordilleran Magmatism
A magmatic arc developed along the western U.S. continental margin during Late
Permian time and continued throughout the Mesozoic with episodes of increased
magmatism occurring during Middle to Late Jurassic (ca. 175-155 Ma) and midCretaceous (ca. 125-85 Ma) time (Figs. 1-2; Ducea, 2001; Walker et al., 2002; Ducea and
Barton, 2007; Barton et al., 2011). The Jurassic arc that extended along the margin from
northwestern Nevada to Sonora, Mexico was likely a low-standing topographic feature
that was dominantly emplaced in an extensional setting (Fig. 2; Walker et al., 2002;
Barton et al., 2011) and was incapable of shielding the continental margin from detritus
transported from retroarc regions (Ingersoll et al., 2013). Triassic and Jurassic
magmatism also occurred in intraoceanic island arcs that were accreted to the margin by
Late Jurassic time in the Klamath Mountains and Foothills belt of the western Sierra
Nevada (Figs. 1-2; Schweickert and Cowan, 1975; Dickinson, 2008).
Voluminous magmatism during mid-Cretaceous time (ca. 125-85 Ma) coincided
with the culmination of the Sevier orogeny and likely resulted in a physiography that
resembled the modern Andes with a high-standing volcano-plutonic arc, inboard elevated
plateau, and a retroarc fold-thrust belt (Figs. 1-2; Ducea, 2001; House et al., 2001;
DeCelles, 2004). A systematic eastward younging of pluton crystallization ages in the
mid-Cretaceous arc reflects progressive migration of magmatism over time associated
with gradual slab flattening (Chen and Moore, 1982; Silver and Chappell, 1988). As a
5

result, the Sierran-Peninsular Ranges arc can be divided into eastern and western zones
that are separated by a 100 Ma isochron (Fig. 2). Particularly high rates of magmatic flux
occurred during Late Cretaceous time that resulted in the Sierra Crest intrusive event (ca.
100-85 Ma;) in the Sierra Nevada and emplacement of the La Posta suite (99-92 Ma) in
the Peninsular Ranges (Walawender et al., 1990; Coleman and Glazner, 1997; Ducea,
2001; Grove et al., 2003b; Todd et al., 2003; Ducea and Barton, 2007; Coleman et al.,
2012). Following extinction of the Sierran-Peninsular Ranges arc, plutonism migrated
inland in the southwestern U.S. and in Sonora, Mexico and continued throughout latest
Cretaceous time (ca. 85-65 Ma) (Fig. 2; Lipman, 1992; McDowell et al., 2001).
Tectonic Restructuring of the Southern California Margin
During late Campanian-early Paleogene time the southern California margin
underwent a fundamental restructuring that resulted in the destruction of the topographic
continuity of the formerly high-standing mid-Cretaceous arc (Saleeby, 2003; Jacobson et
al., 2011; Hall and Saleeby, 2013). The manifestations of this tectonism are most clearly
observed where portions of the eastern mid-Cretaceous batholith (e.g., Salinian block) are
positioned adjacent to the Franciscan subduction complex across the Nacimiento fault in
central California (Figs. 1-2; Dickinson, 1983). By analogy to the Sierra Nevada
batholith-Great Valley forearc-Franciscan Complex triad preserved to the north, this
juxtaposition implies removal of 150-180 km of intervening western batholith, foothills
belt, and forearc basin (Dickinson et al., 2005).
Two competing models have been proposed to explain this tectonic restructuring
of the latest Cretaceous continental margin: 1) large-magnitude (~600-500 km) sinistral
displacement along the Nacimiento fault (Dickinson, 1983; Jacobson et al., 2011) or 2)
west-directed low-angle faulting of the mid-Cretaceous batholith (Hall, 1991; Barth et al.,
2003; Saleeby, 2003; Hall and Saleeby, 2013). Both models are interpreted as
manifestations of the collision of an oceanic ridge or plateau with the continental margin
associated with the development of the Laramide orogeny (Saleeby, 2003; Jacobson et
al., 2011). The Laramide orogeny also coincided with 1) marine transgression that
occurred atop deeply denuded batholithic rocks of the southernmost Sierra-SalinianMojave segment of the margin (Cox, 1982; Grove, 1993; Kidder et al., 2003; Ducea et
6

al., 2009), and 2) a sudden influx of inboard detritus to the trench and forearc strata
preserved atop and adjacent to the Salinian-Mojave segment of the Cretaceous arc
(Jacobson et al., 2011).
Upper Cretaceous-Eocene Forearc
A well-defined forearc basin developed by latest Jurassic-earliest Cretaceous time
along much of the continental margin (Dickinson and Seely, 1979; Ingersol1, 1979;
Bottjer and Link, 1984). The Cenomanian-Eocene forearc strata that are the focus of this
study occur in a NNW-to-SSE-oriented outcrop belt from southern Oregon to Baja
California (Fig. 2A). Sampled stratigraphic units are from several forearc depocenters
that have unique structural and stratigraphic histories (see Appendix A-3 for additional
geologic background on individual forearc depocenters).
The Great Valley and Peninsular Ranges segments of the forearc are wellpreserved and both extend for 100s of kilometers along the margin (Kennedy and
Moore, 1971; Ingersoll, 1979; Bottjer and Link, 1984). Both forearc segments onlap the
western margin of the Sierran-Peninsular Ranges arc and are inferred to have been
confined to the west by an actively accreting subduction complex (Dickinson, 1995a;
Williams and Graham, 2013). Upper Cretaceous-Eocene forearc strata of the Salinian
block and Transverse Ranges are widespread in the California Coast Ranges (Graham,
1976a, b; Grove, 1993; Dickinson, 1995b) where Neogene deformation associated with
the development of the modern strike-slip plate boundary has obscured pre-Neogene
geologic relationships (Atwater, 1989). Forearc strata in these regions were deposited
atop deeply denuded plutonic rocks of the mid-Cretaceous batholithic belt or Franciscan
subduction complex (Dickinson, 1995b; Kidder et al., 2003).
Palinspastic Reconstruction
The original distribution of Upper Cretaceous-Eocene forearc strata has been
greatly modified by deformation associated with the development of the San Andreas
plate boundary (Atwater, 1989), including large-magnitude offset along strike-slip faults
(Hill and Dibblee, 1953; Graham et al., 1989; Dickinson et al., 2005) and microplate
capture and associated transrotation of crustal blocks in the western Transverse Ranges
7

and southern California continental borderland (Luyendyk, 1991; Nicholson et al., 1994;
Dickinson, 1996). Additional disruption occurred by 1) backarc extension within the
Basin and Range province (Wernicke, 1992), 2) local domain rotation within the Mojave
Desert region (Dickinson, 1996), 3) inboard dextral offset (e.g., the Eastern California
Shear Zone, Dokka and Travis, 1990), and 4) local compressional and extensional
tectonism in the California Coast Ranges (Crowell, 2003).
We present a generalized palinspastic reconstruction that integrates previous
interpretations (Grove et al., 2003a; Dickinson et al., 2005; Jacobson et al., 2011) with
updated fault offset estimates (Graymer et al., 2002; Sharman et al., 2013) in an effort to
restore original paleogeologic relationships that existed along the California continental
margin during Eocene time (Fig. 2B). In general, our reconstruction follows that of
Graymer et al. (2002) for the San Franciscan Bay region, Dickinson et al. (2005) for the
Salinian block, and Jacobson et al. (2011) for the Transverse Ranges and Peninsular
Ranges (see Appendix A-2 for additional details). We do not attempt to restore slip on the
Nacimiento fault, given its controversial and uncertain structural history (e.g., Dickinson
et al., 2005; Ducea et al., 2009). Because the Nacimiento fault is thought to have been
active from latest Cretaceous to Paleocene time (ca. 75-60 Ma; Jacobson et al., 2011),
offset along this structure likely influences the relative positions of CenomanianPaleocene forearc strata along the margin as depicted in Figure 2B.
A major difference between our palinspastic reconstruction and those of previous
workers is the addition of ~100 25 km of offset (Sharman et al., 2013) to the
commonly-quoted ~315 km (e.g., Graham, 1978; Graham et al., 1989) along the central
San Andreas fault that restores the Salinian block to an intra-Cordilleran arc position
(Fig. 2B). This reconstruction differs from models that depict the northern Salinian block
as being juxtaposed against the outboard edge of the forearc and subduction complex of
the southern San Joaquin basin during Paleogene time (e.g., Nilsen and Clarke, 1975;
Hall, 1991; Hall and Saleeby, 2013). Our palinspastic reconstruction is consistent with
regional provenance trends that suggest the northern Salinian block was juxtaposed
against the southernmost Sierra Nevada in middle Eocene time (Shaman et al., 2013), and
alignment of the northern extent of the Salinian block (Navarro structural discontinuity)
with the inferred western edge of Sierran basement beneath the fill of the San Joaquin
8

basin (Dickinson et al., 2005). Because the displaced Pinnacles and Neenach volcanic
centers (ca. 23 Ma) are offset by only ~315 km (Matthews, 1976), our restoration implies
that this additional displacement must have occurred between middle Eocene and early
Miocene time (ca. 38-23 Ma; Sharman et al., 2013). Alternatively, portions of the
additional ~100 25 km slip could be accounted for by lengthening (or telescoping) the
Salinian block via internal strike-slip faulting or compression (e.g., routing RelizRinconada fault displacement into the northern Salinian block; Dickinson et al., 2005;
their Fig. 10) or by back-rotating the southernmost Sierra Nevada (Kanter and
McWilliams, 1982), thereby shifting the northern Salinian block southward (Dickinson,
1996).
Our restoration of the Salinian block results in the total displacement on the
central San Andreas fault being greater than estimates for the cumulative displacement
along the San Andreas fault system in southern California (~415 km versus ~325 km),
including offset along the Canton fault, San Gabriel fault, and various strands of the San
Andreas fault (Crowell, 2003). We account for this discrepancy by routing excess slip
through the western Transverse Ranges to a fault located offshore Peninsular California.
A potential candidate for this structure is the San Benito-Tosco-Abreojos fault that is
known to have accommodated right-lateral offset between the Pacific and Northern
American plates between ca. 12-5 Ma and prior to inland migration of the transform
boundary to initiate the opening of the Gulf of California (Spencer and Normark, 1979;
Dickinson, 1996). During this time period the central San Andreas fault must have been
linked with the San Benito-Tosco-Abreojos fault by a fault, or series of faults, that ran
through what is today the western Transverse Ranges and California continental
borderland. Unfortunately, any structures that accommodated pre-middle Miocene slip
have been strongly overprinted by transrotation (Luyendyk, 1991; Dickinson, 1996) and
local transpression in Pliocene-Holocene time (Crowell, 2003).
METHODS AND DATA SOURCES
Detrital Zircon U-Pb Ages
This study presents a compilation of U-Pb crystallization ages of detrital zircons
from Upper Cretaceous-Eocene forearc sandstone from California and areas immediately
9

to the north (southern Oregon) and south (northern Baja California; Fig. 2). In total this
dataset includes ~12,200 detrital zircon U-Pb ages from 211 sandstone samples, of which
4,474 grains from 66 samples are fully published herein for the first time. Table 1
presents a complete list of samples used in this study, and analytical results and a
description of sample locations are provided in Appendices A-3 and A-5 to A-8. We also
present a comparison of forearc strata with time-equivalent units in the Franciscan
subduction complex (681 grains from 10 samples), underplated Upper Cretaceous-lower
Paleogene schist (2027 grains from more than 60 samples), and strata from intra-arc and
retroarc positions (2,550 grains from 31 samples; see Table 1 for data sources).
Because this dataset contains samples collected and analyzed by different authors
for different purposes, the number of grains analyzed per sample varies widely (9 to 178
grains) with the median sample having ~56 grains (Table 1). Most studies analyzed either
~100 or ~60 randomly-selected grains per sample following typical sampling procedures
for detrital zircon provenance studies (e.g., Dickinson and Gehrels, 2009). However,
much of the data from the Salinian block, Transverse Ranges, and Peninsular Ranges was
collected to maximize sample coverage at the expense of the number of grains per sample
analyzed (e.g., Jacobson et al., 2011). In these regions, the median sample has ~30 grains
analyzed.
All zircon grains presented herein were analyzed using either secondary
ionization mass spectrometry (SIMS) or laser ablation-inductively coupled plasma mass
spectrometry (LA-ICPMS). Datasets acquired using LA-ICPMS tend to have more grains
analyzed per sample than those acquired using SIMS due to the relative ease of collecting
large datasets using the former method. Zircon grain ages collected as part of this study
were analyzed using 1) a multi-collector LA-ICPMS at the University of Arizona (see
Gehrels et al., 2008 and Cassel et al., 2012 for a description of methods), 2) a singlecollector LA-ICPMS at the University of California Santa Cruz (Sharman et al., 2013),
and 3) a SIMS Cameca IMS 1270 ion microprobe at the University of California Los
Angeles (Grove et al., 2003a).
We combine individual samples into groups based upon both reconstructed
paleogeography and depositional age to emphasize major provenance trends at the
expense of local variability and to facilitate presentation of data (e.g., LaMaskin, 2012).
10

Sample groups are further combined to form six regional groups that reflect basin
configuration and reconstructed north-south position along the forearc: 1) Oregon forearc
(including the Tyee and Hornbrook basins), 2-3) Great Valley forearc (subdivided into
the Sacramento and San Joaquin basins), 4) northern and central Salinian block (hereafter
Salinian block forearc), 5) southern Salinian block and Transverse Ranges (hereafter
Transverse Ranges forearc), and 6) Peninsular Ranges forearc (Table 1). These
regional groups are defined with the purpose of emphasizing major provenance
transitions along the margin.
In addition to defining geographic groupings, we divide the dataset into four
additional depositional age categories: 1) Cenomanian to Coniacian, 2) Santonian to
Campanian, 3) Maastrichtian to Paleocene, and 4) Early to Middle Eocene (Table 1).
These age divisions were chosen based on periods of major reorganization within the
forearc. For example, the Coniacian-Santonian boundary approximately coincides with
major changes in both sandstone petrology and detrital zircon age populations in the
Great Valley forearc (Ingersoll, 1983; DeGraaff-Surpless et al., 2002). The CampanianMaastrichtian boundary approximately corresponds with a shift to inboard provenance in
the Salinian block and Transverse Ranges (Jacobson et al., 2011). Because displaying
results from individual samples is precluded by the large number of samples in the
dataset, we present cumulative U-Pb age distributions from individual samples within
each location and age group as Appendix A-4.
RESULTS
Detrital Zircon Geochronology
A compilation of normalized and cumulative detrital zircon U-Pb age
distributions is presented in Figure 3 for Upper Cretaceous-Eocene forearc sandstone
samples by location and age. Although the majority of zircon grains are Mesozoic in age,
a wide range of age populations are present that indicate derivation from both local and
extraregional source regions (Table 2; DeGraaff-Surpless et al., 2002; Jacobson et al.,
2011; Sharman et al., 2013). In general, detrital zircon grains can be divided, from most
to least abundant, into three first-order age populations (Table 2): 1) Late PermianCretaceous Cordilleran arc zircon assemblages (ca. 285-65 Ma), 2) pre-arc Paleozoic and
11

Precambrian assemblages (older than ca. 285 Ma), and 3) Paleogene zircon assemblages
(ca. 65-40 Ma). Arc-derived detrital zircons can be divided into subpopulations that
reflect episodic magmatism in the Cordilleran arc (Fig. 4). Major magmatic pulses in the
Sierra Nevada, Salinian block, Mojave Desert region, and Peninsular Ranges occurred
during Permian-Triassic time (ca. 285-225, peak at ca. 252 Ma), Middle-Late Jurassic
time (ca. 180-135 Ma, peaks at ca. 162-148 Ma), and mid-Cretaceous time (ca. 125-80
Ma, peak at ca. 97 Ma) (Fig. 4). Less pronounced but still distinct pulses occur at ca. 77
Ma and ca. 49 Ma (Fig. 4). The peaks of age populations observed in the detrital record
correspond closely with those from exposed volcanic and plutonic rocks (Fig. 4).
Figures 5 and 6 display cumulative and normalized age distributions of regional
(basin-scale) groups, and Figure 7 illustrates the spatial and temporal (CenomanianEocene) evolution of major detrital zircon age populations along ~2,000 km of the
margin from southern Oregon to northern Baja California. Cenomanian-Campanian
forearc strata (3,989 grains from 79 samples) tend to be dominated by Jurassic to midCretaceous (ca. 200-85 Ma) zircon (45-100%; median 93%; Figs. 3AB, 5-7; Appendix A9). In particular, Jurassic-earliest Cretaceous (200-135 Ma) zircon forms a significant
component of Great Valley forearc strata but dramatically decreases in abundance
southward (Figs. 3AB, 7). Late Permian-Triassic zircon is uncommon (0-7%) and pre300 Ma zircon is generally rare (0-48%, median 3%) but constitutes an important
component of some Santonian-Campanian sandstone samples (e.g., Del Puerto Canyon,
Chico, and Cache Creek; Fig. 3B). Strata from the Peninsular Ranges forearc are
characterized by nearly unimodal distributions with peaks at ca. 110-95 Ma that are only
slightly older than the depositional ages of these samples (Fig. 3AB).
Maastrichtian-Paleocene forearc strata (1,652 grains from 41 samples) display a
comparatively wider range of detrital zircon age distributions than their CenomanianCampanian counterparts (Figs. 3C, 7). Although strata deposited within the Great Valley
forearc are still dominated by Jurassic to mid-Cretaceous (200-85 Ma) zircon (87-98%;
median 93%), these samples display higher proportions of 100-85 Ma zircon (8-56%;
median 32%) than their Cenomanian-Campanian counterparts (0-33%; median 2%) (Fig.
7). Maastrichtian-Paleocene forearc strata in the Transverse Ranges and Santa Ana
Mountains are characterized by higher abundances of latest Cretaceous (85-65 Ma)
12

zircon (0-89%; median 21%) and pre-arc detrital zircon (0-92%; median 40%) than timeequivalent strata to the north (Figs. 3C, 7). Samples from San Miguel Island, the San
Diego area, and El Rosario area are dominated by mid-Cretaceous zircon (79-96%; Figs.
3C, 7).
A wide range of detrital zircon age populations are present in early to middle
Eocene forearc strata (6,479 grains from 89 samples; Figs. 3D, 7). Mid-Cretaceous zircon
(ca. 135-85 Ma) is abundant in the Great Valley forearc (19-94%; median 74%) but
decreases in abundance in the northern and central Salinian block (2-87%; median 35%)
and in the Transverse Ranges and northern Santa Ana Mountains (0-25%; median 7%;
Figs. 3D, 7). Jurassic zircon becomes a dominant constituent in the northern and central
Salinian block (12-92%; median 42%) relative to older forearc counterparts (Figs. 3D, 7).
The Transverse Ranges and adjacent northern Peninsular Ranges forearc contain
abundant latest Cretaceous (0-42%; median 17%) and pre-arc zircon (27-75%; median
57%; Figs. 3D, 7).
SEDIMENTARY PROVENANCE ANALYSIS
The following sections provide a discussion of the potential source regions that
we interpret to have been capable of producing the zircon age populations present in the
California forearc (Table 2).
Pre-Permian (> ca. 300 Ma) zircon:
Pre-Permian (i.e., pre-Cordilleran arc) zircon can be divided into a variety of
subpopulations that range from Paleozoic to Archean in age and reflect derivation from a
variety of ultimate sources in Laurentia (Table 2; Dickinson and Gehrels, 2009;
Dickinson et al., 2012). Although some pre-300 Ma age populations could have been
derived from the local Proterozoic framework of the southwestern U.S. (e.g., Late
Paleoproterozoic Yavapai-Mazatzal province), some exotic age populations (e.g.,
Grenville Mesoproterozoic) have no local bedrock source and likely were originally
derived from the Appalachian orogen and transported across the continent via fluvial and
eolian processes to the Cordilleran miogeocline and retroarc foreland (Dickinson and
Gehrels, 2009).
13

Pre-arc detrital zircon in forearc sandstone can be divided in two general


assemblages (Jacobson et al., 2011). The first is characterized by a wide spread of age
peaks that span Paleozoic to Archean ages and is found in Cenomanian-Campanian strata
along the entire California margin and in Maastrichtian-Eocene strata of the Great Valley
forearc (Fig. 3). Although this age assemblage could be derived from a large number of
sources in the western United States (see above), the most likely source for these pre-300
Ma zircon grains is recycling from wall rocks of the Cordilleran Mesozoic arc (Grove et
al., 2008; Jacobson et al., 2011) or from accreted Paleozoic terranes located outboard of
the mid-Cretaceous arc in northern California (Figs. 2, 3; DeGraaff-Surpless et al., 2002;
Surpless and Beverly, 2013). This interpretation is supported by the overwhelming
abundance of arc-derived zircon present in these samples (Figs. 3, 6-7; Jacobson et al.,
2011).
The second pre-arc age assemblage is found in Maastrichtian-Eocene forearc
strata of the Transverse Ranges and Peninsular Ranges and is characterized by a major
peak at ca. 1700 Ma, a secondary peak at ca. 1400 Ma, and a conspicuous absence of
Paleozoic-Neoproterozoic zircon and zircon older than ca. 1850 Ma (Fig. 3). This
assemblage was likely derived from first-cycle erosion of Late PaleoproterozoicMesoproterozoic crystalline rocks of the Mojave Desert region and the Mogollon
highlands of southwestern Arizona (Fig. 2; Jacobson et al., 2011; Dickinson et al., 2012).
Comparison with upper Cretaceous and Paleogene strata deposited on the flanks of the
Arizona Mogollon highlands reveals nearly identical age peaks and proportions
(Dickinson and Gehrels, 2008; Dickinson et al., 2012). This assemblage also frequently
occurs in conjunction with latest Cretaceous (85-65 Ma) zircon (Figs. 3CD, 7), and
igneous rocks of this age are present in the eastern Mojave Desert and Mogollon
highlands regions (Fig. 2).
Permian-Triassic (ca. 285-225 Ma) zircon:
Late Permian and Triassic volcanic and plutonic rocks are relatively uncommon
in California, but are locally present in the eastern Sierra Nevada, Mojave Desert, and
Transverse Ranges (Fig. 4; Dilles and Wright, 1998; Walker et al., 2002; Barth and
Wooden, 2006). Here, the Permian-Triassic volcanic arc lies inboard of the mid14

Cretaceous arc in approximately the same position as the continental Jurassic arc (Fig. 2).
Permian to Triassic rocks are also present in the Mexican segment of the Cordilleran arc
and in the accreted Guerrero superterrane (Dickinson and Lawton, 2001; Centeno-Garca
et al., 2008), although the relative paucity of 300-200 Ma zircon in the Transverse
Ranges and Peninsular Ranges forearc suggest Mexican Permian-Triassic rocks were not
a significant source to the continental margin at this latitude (Kimbrough et al., 2014a).
Jurassic to earliest Cretaceous (ca. 180-135 Ma) zircon:
Arc-related volcanic and plutonic rocks of Jurassic and earliest Cretaceous age are
widespread along the California continental margin and in the southwestern United States
in general (Figs. 2, 4; Irwin and Wooden, 2001; Barton et al., 2011). The present-day
distribution of Jurassic rocks is generally divided into regions outboard and inboard of
the mid-Cretaceous arc (Klamath Mountains and northern Sierra Nevada, and
southeastern Sierra Nevada, Mojave Desert, and Sonoran Desert, respectively; Fig. 2). In
addition, Jurassic to earliest Cretaceous island arcs of the Guerrero superterrane are
present in Baja California (Dickinson and Lawton, 2001; Centeno-Garca et al., 2008),
and Jurassic gneissic granites are present within the central Peninsular Ranges (Shaw et
al., 2003). Together, these rocks are capable of providing a local source of the scarce
180-135 Ma zircon that is found the Peninsular Ranges forearc (Fig. 7; Dickinson and
Lawton, 2001; Shaw et al., 2003). These sources may have also provided Jurassic zircon
to Upper Jurassic forearc strata (Peasquitos Formation) prior to the development of the
mid-Cretaceous Peninsular Ranges arc (Kimbrough et al., 2014a). A magmatic lull (ca.
135-125 Ma) separates Jurassic-earliest Cretaceous from mid-Cretaceous arc volcanism
(Fig. 4; Barton et al., 2011). This gap is also clearly reflected in the detrital record (Fig.
4), although zircon grains in this age range constitute an important age population in
some forearc sandstone samples (e.g., Cenomanian-Campanian Great Valley forearc;
DeGraaff-Surpless et al., 2002; Fig. 3AB).
Mid-Cretaceous (ca. 125-85 Ma) zircon:
Volcanic and plutonic rocks of the mid-Cretaceous Cordilleran arc are widespread
along the margin as a result of voluminous arc magmatism and batholithic emplacement
15

from Idaho to Baja California (Figs. 1, 2; Chen and Moore, 1982; Cowan and Bruhn,
1992; Todd et al., 2003). In particular, mid- to Late Cretaceous arc rocks comprise the
majority of the Sierra Nevada batholith, Salinian block, and Peninsular Ranges batholith
and are common in the Transverse Ranges and Mojave Desert region (Figs. 2, 4; Irwin
and Wooden, 2001; Kistler and Champion, 2001; Walker et al., 2002; Todd et al., 2003).
We further subdivide this age population into pre- and post-100 Ma divisions to reflect an
eastward migration of magmatism over time that is recorded in the Sierran-Peninsular
Ranges arc (Fig. 2; Chen and Moore, 1982).
Latest Cretaceous (85-65 Ma) zircon:
Magmatism in the Sierran-Peninsular Ranges segments of the Cordilleran arc
ended by ca. 85-80 Ma as a result of the onset of Laramide flat-slab subduction (Chen
and Moore, 1982; Irwin and Wooden, 2001). However, plutonic rocks of latest
Cretaceous (ca. 85-65 Ma) age are widespread inboard of the mid-Cretaceous arc in
southeastern California, southwestern Arizona, and Sonora Mexico (Figs. 2, 4; Lipman,
1992; Barth et al., 1995; McDowell et al., 2001; Wells and Hoisch, 2008). These rocks
are likely the source of the latest Cretaceous zircon found in Maastrichtian-Eocene strata
of the Transverse Ranges and Peninsular Ranges forearc (Fig. 3; Jacobson et al., 2011).
This age range also overlaps with the Idaho Batholith (ca 95-53 Ma; Gaschnig et al.,
2011). The association of latest Cretaceous and Paleogene zircon in the Eocene Tyee
basin and Great Valley forearc (Fig. 3D) suggests a distant origin in central Idaho for
these grains (Dumitru et al., 2012).
Paleogene (ca. 65-40 Ma) zircon:
A local source of the Paleogene zircon found in some forearc sandstone samples
(Fig. 3D) is unlikely because there are no known sources of volcanic or plutonic rocks of
this age in California (Fig. 2). However, early Cenozoic igneous rocks are widespread in
the interior of the western U.S. (Lipman, 1992). Dumitru et al. (2012) established that
detritus of this age reached the continental margin via rivers that drained the Idaho
Batholith (98-53 Ma; Gaschnig et al., 2011) and the Challis volcanic field (51-43 Ma;
MGonigle and Dalrymple, 1996; Chetel et al., 2011; Gaschnig et al., 2011) in Eocene
16

time. Eocene-Oligocene zircon (ca. 44-29 Ma) is also present in western flowing
drainages in the northern Sierra Nevada that were likely sourced from volcanic rocks in
northern Nevada (Cassel et al., 2012). However, most of the samples that contain this
young age population are late Eocene or Oligocene in age, based on the youngest zircon
grains and stratigraphic relationships (Cassel et al., 2012). Furthermore, fluvial samples
thought to be as old as middle Eocene in age lack zircon younger than ca. 90-80 Ma (with
a single exception of a 41.5 Ma grain in sample CEC-5; Cecil et al., 2010; Cassel et al.,
2012). This suggests that Sierran fluvial drainages were not a significant source of
Nevadan Paleogene zircon to the Great Valley forearc until late Eocene time. For these
reasons, the central Idaho region is most likely to have been the source of Paleogene
zircon to the early- middle Eocene Great Valley forearc, associated trench-slope basins,
and the subduction complex.
Igneous rocks as young as 50 Ma are also present in southwestern Arizona and
Sonora Mexico (Jacobson et al., 2011). These sources likely account for the few Eocene
grains found in the Poway fluvial system in northern Baja, but otherwise were not a
significant source to other Paleogene forearc strata in southern California.
DISCUSSION
Temporal and Spatial Evolution of Forearc Provenance and Paleogeography
Figure 8 presents a generalized depiction of the paleogeographic evolution of the
Cordilleran margin from Late Cretaceous to middle Eocene time based upon observed
spatial and temporal trends in forearc detrital zircon age signatures (Fig. 7). We also
compare forearc age distributions with those from contemporaneous strata in intra-arc
and retroarc positions (e.g., Barth et al., 2004; Dickinson and Gehrels, 2008; Dickinson et
al., 2012; Fig. 6). The restored north-to-south positions of forearc strata (Figs. 7, 8) are
inferred from our palinspastic reconstruction of the margin (Fig. 2B) and are thus subject
to uncertainties inherent in such reconstructions (see Appendix A-2). In particular, the
structural history of the Nacimiento fault has important implications for pre-Eocene
forearc relationships (e.g., Jacobson et al., 2011). Even so, this compilation reveals
systematic, margin-scale provenance trends in the California forearc that reflect the
tectonic and landscape evolution of the continental margin (Figs. 7, 8).
17

Cenomanian-Coniacian forearc (100-86 Ma)


Although Cenomanian-Coniacian forearc strata are dominated by Jurassic to midCretaceous (200-85 Ma; 83-100%) zircon, the contributions from various arc
subpopulations vary systematically along the forearc, with peak detrital zircon ages
becoming progressively younger towards the south (Figs. 3A, 7). Jurassic-earliest
Cretaceous (200-135 Ma) zircon grains constitute a majority of the forearc sandstone
samples in the Hornbrook and Sacramento basins (41-100%) with samples from Cache
Creek and Mount Diablo displaying nearly unimodal age distributions with peaks at ca.
148 and 140 Ma, respectively (Fig. 3A). The abundance of 200-135 Ma zircon in this
region clearly reflects the presence of accreted Triassic-Jurassic assemblages (Klamath
Mountains and Foothills belt) that occur outboard of the mid-Cretaceous arc at these
latitudes (~43-37N; Figs. 1, 2). These forearc strata also display a conspicuous scarcity
of contemporary (100-86 Ma) zircon despite concurrent emplacement of the voluminous
Sierra Crest intrusive suite in the adjacent Sierra Nevada arc (Chen and Moore, 1982;
Coleman and Glazner, 1997; Ducea, 2001; see discussion below).
Forearc strata south of the Sacramento basin display a systematic shift towards
younger, mid-Cretaceous (125-85 Ma) ages that suggests decreased contribution from
outboard Triassic-Jurassic assemblages and increased contribution from the midCretaceous arc (Fig. 7). In particular, Jurassic-earliest Cretaceous (200-135 Ma) zircon
becomes scarce in the Peninsular Ranges forearc (0-14%; Fig. 7). This observation
suggests that any contribution from Jurassic rocks within the Peninsular Ranges or from
the Guerrero superterrane was overwhelmed by the voluminous mid- to Late-Cretaceous
batholith and its volcanic equivalents. Although 100-85 Ma zircon is nearly absent in the
Great Valley forearc, this population becomes increasingly abundant towards the south
(Fig. 7). The presence of post-100 Ma zircon in the Peninsular Ranges forearc suggests
that the eastern batholith (La Posta suite; Walawender et al., 1990) was a source to the
forearc during this time period (Fig. 7). In particular, Cenomanian to Turonian strata in
the El Rosario area and Vizcaino Peninsula are highly enriched in post-100 Ma zircon
relative to samples to the north (Fig. 3A; Kimbrough et al., 2001).

18

We interpret the dominance of arc-derived (ca. 200-85 Ma) zircon in


Cenomanian-Coniacian forearc strata (Fig. 7) to suggest that the forearc was completely
isolated from inboard source areas and that the mid-Cretaceous arc formed a continuous,
high-standing topographic feature (Fig. 8A). The scarcity of post-100 Ma zircon indicates
little derivation from the eastern portion of the mid-Cretaceous arc, except along the
southernmost portion of the study area (Figs. 2, 7). In addition, age-equivalent retroarc
strata are enriched in post-100 Ma zircon relative to the forearc and also contain abundant
Triassic-Jurassic zircon that suggests derivation from older segments of the Cordilleran
arc that occur inboard of the mid-Cretaceous arc (Figs. 2, 6 Correspondingly, a drainage
divide is inferred to have existed along the axis of the Cretaceous arc that we depict along
the western rim of the elevated Nevadaplano to the north and west of the retroarc McCoy
Mountains basin in the south (Barth et al., 2004; Spencer et al., 2011; Fig. 8A). The
scarcity of arc-derived zircon (3-5%; Dickinson et al., 2012; their Fig. 5) in the foreland
east of the Sevier thrust front suggests that relatively little arc detritus was able to cross
the elevated Nevadaplano to reach the foreland basin to the east.
Santonian-Campanian forearc (86-72 Ma)
Although Santonian-Campanian forearc strata continue to be dominated by
Jurassic to mid-Cretaceous zircon (200-85 Ma; 45-100%), we observe several distinct
changes relative to the older, Cenomanian-Coniacian forearc (Fig. 7). In general, age
distributions indicate a decrease in contribution from outboard Triassic-Jurassic
assemblages relative to the mid-Cretaceous arc (125-85 Ma; Fig. 7). In particular,
widespread influx of 100-85 Ma zircon along the margin signals increased contribution
from the eastern, post-100 Ma portion of the mid-Cretaceous arc relative to the pre-100
Ma, western part of the arc (Figs. 2A, 7).
These changes are most clearly observed in the Hornbrook and Sacramento basins
where a dramatic decrease in the abundance of 200-135 Ma zircon corresponds with an
abrupt increase in the proportion of 135-85 Ma zircon, particularly the 100-85 Ma
subpopulation, at the Coniacian-Santonian boundary (Figs. 3AB, 5, 7). These samples
also contain higher proportions of pre-arc zircon (locally up to 48%, median 15%)
relative to their Cenomanian-Coniacian counterparts (up to 16%, median 2%; Fig. 7).
19

Broadly speaking, the forearc south of the Sacramento basin is characterized by


consistent detrital zircon age distributions dominated by mid-Cretaceous (135-85 Ma)
zircon (Fig. 7). Peak ages shift towards younger populations and correspondingly are
enriched in post-100 Ma zircon relative to their Cenomanian-Coniacian counterparts
(Figs. 3AB, 5, 7). Jurassic-earliest Cretaceous (200-135 Ma) zircon gradually decreases
in abundance southward, and strata from San Miguel Island and the San Diego area are
enriched in 135-100 Ma zircon relative to others along the margin (Figs. 7, 8B).
We infer that increased abundance of 100-85 Ma zircon grains and a widespread
shift towards younger peak ages within the forearc basin as a whole signal an eastward
migration of drainage divides and progressive unroofing of the mid-Cretaceous batholith
immediately following the cessation of arc magmatism (Fig. 8B; see discussion below).
Initial dissection of the extinguished Sierran arc at the Coniacian-Santonian boundary is
also supported by the transition to the Rumsey petrofacies that is characterized by an
increase in quartz and feldspar and a decrease in lithic fragments and plagioclase-to-total
feldspar ratio (Ingersoll, 1983). Increased contribution from the Sierra Nevada batholith
to the Great Valley forearc is also supported by an inferred transition from dominantly
longitudinal to transverse drainages in this time period (Ingersoll, 1979; DeGraaffSurpless et al. 2002), coeval with an increase in sedimentation rate and progradation of
deltas into the basin (Williams and Graham, 2013).
Detrital zircon age populations in the Santonian-Campanian forearc also indicate
continued topographic continuity of the mid-Cretaceous arc and isolation from inboard
source regions along most of the margin (Fig. 8B). This inference is supported by
comparison with retroarc foreland strata that contain ca. 85-72 Ma detrital zircons that
were sourced from Laramide-aged igneous rocks emplaced inboard of the midCretaceous arc (Figs. 2; 8B). Age-equivalent forearc strata generally lack zircon younger
than ca. 85 Ma, with the exception of late Campanian-early Maastrichtian strata in the
Simi Hills and Santa Monica Mountains that contain a few grains as young as ca. 75 Ma
(Figs. 3B, 7, 8B; Jacobson et al., 2011). Also, Jurassic zircon is abundant in the retroarc
but scarce in the forearc (Fig. 6), except in the Hornbrook basin and Great Valley forearc
where Jurassic-earliest Cretaceous zircon was readily supplied from outboard, accreted
terranes (Figs. 1, 2).
20

Maastrichtian-Paleocene forearc (72-56 Ma)


Maastrichtian-Paleocene strata in the Great Valley forearc are characterized by a
consistent detrital zircon signature that is enriched in 100-85 Ma zircon (25-56%) relative
to their Santonian-Campanian counterparts (0-23%), suggesting further eastward
migration of drainages and continued uplift and denudation of the Sierra Nevada
batholith (Figs. 7, 8C). In addition, lower abundances of Jurassic-earliest Cretaceous
(200-135 Ma) zircon suggest diminished input from the Klamath Mountains and accreted
terranes in the Sierra Nevada Foothills belt relative to input from the Sierra Nevada
batholith (Figs. 7, 8C). These findings are consistent with studies of sandstone petrology
that indicate higher proportions of quartz and feldspar and lower proportions of volcanic
lithic grains (Rumsey petrofacies; Ingersoll, 1983). However, although Campanian and
Maastrichtian-Paleocene sandstone are petrologically similar arkose, these units are
clearly distinguished on the basis of their detrital zircon U-Pb age signatures.
Paleocene strata from the northern Salinian block (Gualala bock and La Honda
basin) display provenance signatures that resemble equivalent Great Valley strata in
northern exposures. However, Maastrichtian-Paleocene strata from the central Salinian
block (Santa Lucia Range) are characterized by a wide range of detrital zircon
distributions, including varying proportions of mid-Cretaceous, Jurassic, PermianTriassic, and Proterozoic age populations (Figs. 3C, 7; see also Appendix A-4). When
combined as a group, this region displays detrital zircon age distributions intermediate
between those found in the Great Valley forearc and Transverse Ranges (Figs. 7, 8C).
Thus, Maastrichtian-Paleocene strata of the Santa Lucia Range appear to contain a mixed
provenance signature that suggests derivation from both local exposures of the midCretaceous batholith and from Triassic-Jurassic igneous rocks in the western Mojave
Desert region (Fig. 2).
Forearc strata deposited in the La Panza Range and San Gabriel Mountains
exhibit a pronounced switch to source regions inboard of the mid-Cretaceous arc (Figs. 7,
8C). The abundance of latest Cretaceous (ca. 85-65 Ma) and Proterozoic (ca. 1.8-1.3 Ga)
zircon and lack of mid-Cretaceous and Jurassic zircon are consistent with an inland
drainage that reached into the eastern Mojave Desert and possibly as far as the Mogollon
21

highlands (Jacobson et al., 2011). The observed influx of inboard detritus is less
pronounced in the Santa Ynez Mountains, Simi Hills, and Santa Monica Mountains
where mid-Cretaceous (135-85 Ma) zircon grains still comprise a significant component
of the detrital age populations (Fig. 7). The outboard-to-inboard provenance transition in
the northern Santa Ana Mountains is recorded as an influx of latest Cretaceous (85-65
Ma) zircon to the upper Paleocene Silverado Formation (Fig. 7, 8C). Upper Paleocenelower Eocene sandstone from the San Diego area (Mount Soledad Formation) is
dominated by ca. 135-100 Ma zircon (81-96%; Fig. 7), suggesting that these grains were
derived locally from the western Peninsular Ranges batholith. However, the presence of
Poway porphyritic rhyolite clasts in the Mount Soledad Formation suggests that the
drainage at this latitude must have extended inboard to the Sonoran Desert (Kennedy and
Moore, 1971; Kies and Abbott, 1982; Abbott and Smith, 1989). We infer that the Jurassic
rhyolitic bedrock source of the Poway clasts yielded little sand-sized detritus relative to
local batholithic rocks and is thereby underrepresented in the detrital zircon age
distributions.
The sudden delivery of abundant inboard detritus to the southern California
forearc signals a rapid inland migration of the drainage divide that formerly ran down the
axis of the Salinian-Mojave segment of the mid-Cretaceous arc (Jacobson et al., 2011)
and was accompanied by marine flooding of the arc interior that extended to at least the
San Gabriel block (Figs. 7, 8C; Kooser, 1982). As a consequence, forearc and retroarc
depocenters transitioned from having separate to shared source regions (i.e., Mojave
Desert and Mogollon provinces) during this time period (Jacobson et al., 2011; Dickinson
et al., 2012; Fig. 6). For example, the upper Paleocene-lower Eocene Colton Formation in
the Uinta basin is thought to have been supplied via the hypothetical California paleoriver
from the Mojave Desert segment of the Cordilleran arc (e.g., Dickinson et al., 2012). This
interpretation is supported by comparison of the detrital zircon age distributions from the
Colton Formation with time-equivalent forearc strata from the Salinian block, Transverse
Ranges, and Peninsular Ranges (Fig. 6). Latest Cretaceous zircon (ca. 85-65 Ma) is
present in both the Colton Formation and in the Transverse Ranges forearc, although
Triassic-Jurassic zircon is an important age population present in the Colton Formation
that is nearly absent in Maastrichtian-Paleocene forearc strata (Fig. 6). Extension of the
22

California River headwaters into the western Mojave Desert region is supported by
comparison with the Paleocene intra-arc Goler basin and Eocene Salinian forearc that
contain abundant Jurassic zircon (Figs. 6, 8CD). Triassic zircon is also present in modest
abundances in strata from the Salinian block and most notably comprises up to 22% of
the early Eocene German Rancho Formation (Gualala block) and up to 50% of
Maastrichtian-Paleocene strata in the Santa Lucia Range (Figs. 3D, 7; Barbeau et al.,
2005; Doebbert et al., 2012). Thus the hypothetical California River drainage is inferred
to have extended headward as far as the western Mojave Desert region where TriassicJurassic igneous rocks are known to have been a local source to the forearc (Fig. 8CD).
Lower to middle Eocene forearc (56-36 Ma)
Lower to middle Eocene forearc strata display systematic north-south trends in
sedimentary provenance that reflect the culmination of inboard sediment delivery to the
continental margin and the age distribution of igneous rocks along both the intact and
breached segments of the mid-Cretaceous Cordilleran arc (Figs. 7, 8D). In general,
distinct assemblages of detrital zircon age populations were derived from at least five
adjacent fluvial systems with headwaters that extended into 1) north-central Idaho (Tyee
and Princeton Rivers; Dumitru et al., 2012), 2) the Sierra Nevada batholith (Cecil et al.,
2010; Cassel et al., 2012), 3) the western Mojave region (Lechler and Niemi, 2011), 4)
the eastern Mojave Desert and Mogollon highlands (Jacobson et al., 2011), and 5)
Sonoran Desert and southwestern Arizona (Jacobson et al., 2011).
1. Idaho Batholith and Adjacent Volcanic Fields. Eocene extensional
deformation and associated uplift and erosion in north-central Idaho (ca. 53-40 Ma;
Foster et al., 2007) produced abundant sediment derived from the Idaho batholith and
Challis volcanic field that reached the Oregon and California forearc via the Tyee and
Princeton Rivers (Fig. 8D; Dumitru et al., 2012). The Paleocene-early Eocene Princeton
submarine canyon for a time conveyed Idahoan detritus down the axis of the Sacramento
basin to the Mount Diablo area, which has formed a persistent depocenter in the Great
Valley forearc since Cretaceous time (Moxon, 1990). From this point Idahoan detritus
was longitudinally dispersed to the north and south in trench-slope basins (e.g.,
Underwood, 1983; Snow et al., 2010) and eventually to the trench, where a mixture of
23

Sierran- and Idahoan-derived sediment was subducted and accreted during early Eocene
time to form the Franciscan Coastal Belt (Dumitru et al., 2012). Paleogene zircon is
found as far south as the Temblor Range (Sharman et al., 2013), suggesting that the
Idahoan sediment dispersal system was active over the entire length of the Eocene Great
Valley forearc.
2. Sierra Nevada: Aside from variable contribution from sediment sources in
Idaho, Eocene sandstone in the Great Valley forearc is characterized by a laterally
consistent provenance signature for over ~400 km that is indistinguishable from age
distributions in Maastrichtian-Paleocene Great Valley strata (Figs. 3CD, 5, 7). These
sandstone samples are dominated by mid-Cretaceous (ca. 125-85 Ma) zircon with a
characteristically asymmetrical distribution (peak at ca. 97 Ma; Fig. 3D), have lesser
abundances of Jurassic-earliest Cretaceous zircon (Fig. 7), and are interpreted to have
been derived from both the western and eastern portions of the Sierra Nevada batholith
(Fig. 8D). This signature is present in westward-draining fluvial systems that are
preserved in the northern Sierra Nevada and that indicate significant topography of the
ancestral Sierra Nevada Mountains (Mulch et al., 2006; Cecil et al., 2010; Cassel et al.,
2012). Similar west-flowing drainages are likewise inferred in the southern Sierra Nevada
(Fig. 8D), although incised channels have not been identified and are likely no longer
preserved (Henry et al., 2012). Sierran detritus was distributed to the adjacent shelfal and
deep-water forearc via incised valleys and submarine canyons cut into the Sierran shelf
edge (Graham and Berry, 1979; Sullivan and Sullivan, 2012). Uplift and emergence of
the Franciscan subduction complex along the western margin of the San Joaquin basin
(Schulein, 1993) helped partition thick successions of sand-rich turbidite deposits into
structurally-controlled, ponded depocenters (e.g., Cantua Sandstone and Point of Rocks
Sandstone; Nilsen et al., 1974; Nilsen and Clarke, 1975; Graham and Berry, 1979).
Sierran detritus was subsequently delivered from the forearc to trench-slope basins
preserved atop the San Francisco Bay block through an outlet that is inferred to have
existed in the trench-slope break (Fig. 8D).
3. Western Mojave Region: A major provenance shift occurred in the
southernmost San Joaquin basin at about 35 degrees latitude where an influx of Late
Permian-Jurassic zircon to the margin indicates that forearc drainages extended eastward
24

to the western Mojave Desert and/or southeastern Sierra Nevada regions (Figs. 7, 8D;
Lechler and Niemi, 2011; Doebbert et al., 2012; Sharman et al., 2013). Forearc strata
deposited atop the northern and central Salinian block are dominated by Jurassic zircon
(36-73%) with lesser proportions of Permian-Triassic zircon (4-20%) that indicate
derivation from older arc segments that occurred inboard of the mid-Cretaceous arc (Figs.
2, 7). In particular, the importance of a Jurassic-aged source region along this portion of
the margin is evident in the northern Santa Lucia Range where Jurassic zircon comprises
a large proportion (70-92%) of the middle Eocene The Rocks Sandstone (Figs. 3D, 7).
Maastrichtian-Paleocene forearc strata in these same areas lack abundant Jurassic zircon,
suggesting that this sediment dispersal system did not supply detritus to the forearc until
early Eocene time (Fig. 8CD). The cause of the influx of Jurassic grains to the forearc is
enigmatic but may have resulted from either increased uplift and denudation in the
hinterland or an eastward migration of forearc drainages into the western Mojave Desert
region during Eocene time (Fig. 8D).
An important aspect of our paleogeographic reconstruction is the inference that
strata deposited atop the northern and central Salinian block were derived, in large part,
from drainages that extended inboard of the mid-Cretaceous arc (Fig. 8). This
interpretation departs from previous models (e.g., Nilsen and Clarke, 1975) that have
inferred the central California margin to be characterized by continental borderland
topography with local, uplifted crustal blocks providing sediment to adjacent, bathyal
basins (see also discussion in Grove, 1993; Jacobson et al., 2011). Derivation from local
exposures of the Salinian block would yield detrital zircon age distributions dominated by
mid- to Late Cretaceous zircon (ca. 110-76 Ma; Mattinson, 1990; Figs. 4, 5). However,
detrital zircon age distributions of Cretaceous-Eocene Salinian forearc sandstone are
markedly dissimilar to the distribution of zircon U-Pb ages from Salinian basement (Fig.
5). Instead, the abundance of pre-135 Ma zircon in Eocene successions deposited atop the
Salinian block suggests that these strata were sourced in large part from regions to the
east (Fig. 8D). In addition, widespread marine sedimentation atop the Salinian block
during Paleogene time (Nilsen and Clarke, 1975; Graham, 1976a, b) suggests that this
terrane was largely submerged during this time period (Fig. 8D). For these reasons, the

25

Salinian block was not likely a major sediment source to adjacent Paleogene forearc
depocenters.
4. Eastern Mojave Region/Mogollon highlands: Another abrupt provenance shift
occurred south of the central Salinian block where mid-Cretaceous zircon becomes nearly
absent and is replaced by Proterozoic and Jurassic zircon populations that are found in the
Transverse Ranges forearc (Sierra Madre Mountains, Pine Mountain block, and Orocopia
Mountains; Figs. 7, 8D; Jacobson et al., 2011). Similar distributions are found in forearc
strata from the Santa Ynez Mountains, Simi Hills, and Santa Monica Mountains,
although these regions also include populations of latest Cretaceous (ca. 85-65 Ma)
zircon. These strata were likely sourced from the eastern Mojave Desert and Mogollon
highlands region of southwestern Arizona based on the occurrence of Jurassic and latest
Cretaceous plutons that intrude Proterozoic crustal rocks with age peaks at ca. 1.4 and 1.7
Ga in these regions (Jacobson et al., 2011).
5. Sonoran Desert/Southwestern Arizona: Detrital zircon age distributions in the
Peninsular Ranges forearc are consistent with studies of conglomerate clast compositions
that indicate that drainages extended across the northern Peninsular Ranges batholith into
the Sonoran Desert region and southwestern Arizona (Abbott and Smith, 1978; 1989).
Additional populations of mid-Cretaceous zircon from forearc strata in the San Diego
area and northern Baja California indicate local contribution from the Peninsular Ranges
batholith (Figs. 7; 8D). Detrital zircon age distributions from the late Eocene Ballenas
Gravels are similar to those in the San Diego area, supporting the interpretation of a
through-going fluvial network (Fig. 8D). In general, the amount of extraregional detritus
decreases southward along the Peninsular Ranges forearc, suggesting increasing isolation
from inboard source regions in this direction (Figs. 7, 8D).
Forearc-Subduction Complex-Underplated Schist Comparison
Detritus delivered to the continental margin that was able to bypass the forearc
and eventually reach the trench was either accreted to the margin as part of the
Franciscan subduction complex (Dumitru et al., 2010) or underplated beneath the margin
to form meta-sedimentary schist (e.g., Pelona-Orocopia-Rand schist; Grove et al., 2003a).
In some cases, detrital zircon age distributions in the forearc closely match those in the
26

adjacent subduction complex or underplated schist (e.g., compare Transverse and


Peninsular Ranges forearc with age-equivalent underplated schist; Fig. 6; see also
discussion in Jacobson et al., 2011). However, we also observe that adjacent forearc and
subduction complex segments can display markedly different detrital zircon age
distributions. For example, the Eocene Franciscan subduction complex has a very
different age distribution than that found in the adjacent Sacramento basin (Figs. 2, 6).
Samples from the Franciscan subduction complex inferred to be Cenomanian-Coniacian
in age display the prominent Late Jurassic peak present in the Sacramento basin but also
contain Early Cretaceous ages found in San Joaquin basin (Fig. 6). Similarly, three
Santonian-Campanian(?) Franciscan samples from Pacheco Pass (Ernst et al., 2009) and
Marin County (Prohoroff et al., 2012) broadly resemble those from the Great Valley
forearc, but are enriched in post-100 Ma zircon relative to time-equivalent forearc strata
(Fig. 6).
These observed mismatches in forearc-subduction complex provenance can be
interpreted as manifestations of longitudinal transport of sand along trench-slope basins
and trench floors from variably spaced entry points. This pattern of sediment dispersal is
ubiquitous in modern trenches (Underwood and Moore, 1995), and can introduce a
detrital zircon age mismatch between forearc basins and their associated subduction
complexes. This pattern is well exhibited in the Eocene Franciscan Coastal Belt that has a
distinct provenance character from the adjacent forearc because of the delivery of
abundant Idahoan detritus that reached the trench via an entry point in the vicinity of
Mount Diablo (Dumitru et al., 2012). Alternatively, these differences could be explained
by margin-parallel dextral transport of the Franciscan subduction complex relative to the
adjacent forearc (Ernst et al., 2009).
Influence of Underlying Geometry of the Jurassic and Mid-Cretaceous Arcs on the
Evolution of Forearc Detrital Zircon Age Distributions
One of the most striking aspects of how detrital zircon age distributions evolve
over time in the California forearc is the abundance of Jurassic-earliest Cretaceous (200135 Ma) zircon in the Great Valley forearc during Cenomanian-Coniacian time, the
subsequent decline of this age population relative to mid-Cretaceous zircon, and the
27

following resurgence of Jurassic zircon in the Salinian block and Transverse Ranges in
Eocene time (Figs. 3, 7). This first-order pattern is a clear reflection of the underlying
geometry of the Jurassic-earliest Cretaceous and mid-Cretaceous arcs; specifically, the
latter cross-cuts the former, thereby partitioning preserved Jurassic-earliest Cretaceous
source rocks into regions inboard- and outboard-of the younger mid-Cretaceous arc (Figs.
1, 2). The systematic southward decrease of 200-135 Ma zircon in the ConiacianCampanian forearc (Fig. 7) is a consequence of Jurassic-aged arc rocks primarily lying
outboard of the mid-Cretaceous arc in the north (Klamath Mountains and Sierra Nevada
Foothills belt) and inboard of the mid-Cretaceous arc in the south (southeastern Sierra
Nevada, Mojave Desert, and Sonoran Desert regions; Figs. 1, 2). For this reason, Jurassic
zircon is uncommon in the southern California forearc (< 35N latitude; Fig. 7) until
Eocene time when forearc drainages were able to locally extend far enough inland to tap
the inboard Jurassic arc (Fig. 8D).
Why is Post-100 Ma Zircon Scarce in the Cenomanian-Coniacian Great Valley
Forearc?
An important characteristic of the Cenomanian-Coniacian forearc is the scarcity
of contemporaneous zircon, particular in the Hornbrook and Great Valley forearc
segments (0-15%; median 2%), despite being deposited synchronously with a major
flare-up (ca. 100-85 Ma) in the adjacent mid-Cretaceous arc (Figs. 3A, 7; Ducea, 2001).
The post-100 Ma Sierra Crest intrusive suite accounts for more than 50% of the present
exposures of the Sierra Nevada batholith (Fig. 2; Coleman and Glazner, 1997; Ducea,
2001) and zircon of this age is abundant (up to 44%) in modern rivers that drain the
Sierra Nevada (Kimbrough et al., 2009). The near absence of age-equivalent zircon in the
forearc is quite surprising, particularly because both petrographic and geochemical
studies have suggested that the Sierra Nevada arc was a major source to the
contemporaneous Great Valley forearc (Ingersoll, 1983; Linn et al., 1992).We see three
potential explanations:
1. A drainage divide may have existed west of the active volcanic arc that
prevented detritus from the volcanic carapace from being delivered to the adjacent
forearc basin (DeGraaff-Surpless et al. 2002, their Figure 12B). The presence of a
28

drainage divide wholly west of the mid-Cretaceous arc is precluded by petrographic and
geochemical evidence that indicates that forearc sandstone contains andesitic detritus
derived from the volcanic carapace of the mid-Cretaceous arc (Ingersoll, 1983; Linn et
al., 1992). However, the lack of post-100 Ma detrital zircon could suggest that the
western, pre-100 Ma arc was a major source to the forearc while the volcanic carapace of
the eastern, post-100 Ma volcanic arc was largely partitioned from the forearc (Fig. 2).
Modern silicic calderas that form in continental arcs tend to be preferentially located
behind the arc front (Hughes and Mahood, 2011) and thus are more likely to be a source
to the retroarc region rather than to the forearc. Some support for this interpretation is
provided by the observation that Cenomanian-Coniacian retroarc sandstone is enriched in
post-100 Ma zircon relative to their forearc counterparts (Fig. 6). In addition, local
absence of mid-Cretaceous zircon in the western Sacramento basin (Fig. 3A) can be
explained by predominantly southward-directed, longitudinal sediment transport during
this time period (Ingersoll, 1979). In this case, major sediment contribution to the forearc
was likely from the Klamath Mountains to the north with lesser contribution from the
mid-Cretaceous arc in the Sierra Nevada (Fig. 8A; DeGraaff-Surpless et al., 2002).
2. The post-100 Ma volcanic arc may have been a source to the Great Valley
forearc but yielded little zircon. This interpretation hinges on the inference that felsic to
intermediate volcanic rocks yield fewer and smaller zircon grains than their plutonic
counterparts (Poldervaart, 1956; Watson, 1979). However, this inference is at odds with
the abundance of Early Cretaceous zircon in Upper Cretaceous samples with high
volcanic lithic contents that are inferred to have been derived from the volcanic carapace
of the mid-Cretaceous arc (e.g., Grabast and Los Gatos petrofacies; Ingersoll, 1983;
DeGraaff-Surpless et al., 2002).
3. The paucity of post-100 Ma zircon in the Cenomanian-Coniacian forearc may
reflect an absence of volcanic equivalents to the ca. 98-86 Ma Sierra Crest intrusive suite.
However, although relatively few mid-Cretaceous volcanic sequences have been
identified in the Sierra Nevada, the Minarets and Merced Peak caldera complexes provide
well-preserved examples of ca. 100 to > 93 Ma volcanism likely associated with the
voluminous Sierra Crest intrusive event (Fiske and Tobisch, 1994; Lowe, 1996). In
addition, the presence of volcanic ash deposits of this age in retroarc foreland sequences
29

(e.g., Cadrin et al., 1996) confirms the existence of extrusive, silicic volcanism during
Late Cretaceous time. Regardless of the specific reason why post-100 Ma zircon is so
scarce in the Cenomanian-Coniacian forearc, we interpret the gradual influx of this age
population during Santonian to Paleogene time to signal the progressive erosion into
plutons of the Sierra Crest intrusive suite (Figs. 3B-D, 5, and 7).
Influence of Low-Angle Subduction on the Evolution of an Arc-Forearc System
The Laramide orogeny is widely considered to have resulted from low-angle
subduction that accompanied the collision and subsequent subduction of an oceanic ridge
or plateau beneath North America (Livaccari et al., 1981; Henderson et al., 1984; Saleeby
et al., 2003). Potential candidates include the conjugates to the Hess and Shatsky plateaus
of the present-day northwest Pacific (Tarduno et al., 1985). Both plate motion
reconstructions and inverse convection models predict that the Shatsky conjugate plateau
collided with the North American plate at ca. 90 Ma in the vicinity of southern California
(Liu et al., 2010). We hypothesize that major spatial and temporal trends in forearc
sandstone provenance were greatly influenced by the impingement of the Shatsky
conjugate plateau on the continental margin, and that this event resulted in both
differential uplift and denudation of the mid-Cretaceous volcanic-plutonic arc and
eventual breaching of a ~500 km segment of the southern California mid-Cretaceous
margin batholith (e.g., Saleeby, 2003).
Contrasting Denudation History of the Sierran and Peninsular Ranges Batholiths
Although the Sierran and Peninsular Ranges segments of the mid-Cretaceous
batholith share many commonalities in their magmatic history (e.g., Chen and Moore,
1982; Silver and Chappell, 1988), they differ significantly in the magnitude, distribution,
and timing of post-magmatic uplift and denudation (Ague and Brimhall, 1988; George
and Dokka, 1994; Grove et al., 2003b; Saleeby et al., 2007, 2008). Major differences in
detrital zircon age distributions between the Great Valley and Peninsular Ranges
segments of the forearc (Fig. 5) can be understood to be a consequence of these
contrasting denudation histories, in addition to the underlying control contributed by the
cross-cutting geometry of the mid-Cretaceous arc relative to older arc segments.
30

The Peninsular Ranges underwent rapid denudation following emplacement of the


99-92 Ma La Posta suite that caused an influx of coarse detritus to the CenomanianTuronian forearc and resulted in detrital zircon age distributions that exhibit little time lag
between U-Pb crystallization and depositional ages (Kimbrough et al., 2001; Grove et al.,
2003b; Fig. 3A). The timing of denudation (ca. 91-86 Ma) corresponds closely with the
predicted time (ca. 90 Ma) and location of plateau collision with the margin (Grove et al.,
2003b; Liu et al., 2010), suggesting that these two events were causally linked.
Alternatively, mechanical and thermal effects related to high rates of magmatic flux
during the La Posta intrusive event could have also contributed to denudation of the
Peninsular Ranges arc (Kimbrough et al., 2001; Grove et al., 2003b). Regardless of the
mechanism, depths of the modern level of erosion in the northern Peninsular Ranges
batholith (0-8 km in the west and typically 12-20 km in the east; Ague and Brimhall,
1998; Herzig and Kimbrough, 2014) were nearly attained by latest Cretaceous time
(George and Dokka, 1994; Grove et al., 2003b, 2008). As a result, detrital zircon age
distributions of Cenomanian-Paleocene forearc strata essentially match the distribution of
U-Pb ages from the modern surface of the Peninsular Ranges batholith and exhibit little
change over time until the delivery of inboard detritus to the margin in Eocene time (Figs.
5, 8CD).
The Great Valley forearc exhibits a very different provenance evolution where
Cenomanian-Paleocene forearc strata exhibit increasing abundances of 125-85 Ma zircon
over time that is interpreted to reflect progressive denudation of the Sierra Nevada
batholith (Fig. 5). Cenomanian-Coniacian strata display comparatively larger time lags
between peak detrital zircon U-Pb ages and depositional ages, despite voluminous
concurrent magmatism to the east (Fig. 3A). The influx of 100-85 Ma zircon to the
forearc at the Coniacian-Santonian boundary (Fig. 3B) likely relates to the attainment of
erosional depths great enough to expose plutons of the Sierra Crest intrusive suite in the
eastern Sierra Nevada batholith and coincides with a cooling event in the western Sierra
Nevada (Cecil et al., 2006; Saleeby et al., 2010; Surpless and Beverly, 2013). This
inference is supported by sandstone petrography and geochemistry that also signal
unroofing of Sierra Nevada batholith during this time interval (Ingersoll, 1983; Linn et
al., 1992).

Surprisingly, this pulse of denudation coincides with a decrease in


31

sedimentation rates in the Sacramento basin (Williams, 1997), although evidence for
sediment transport via incised submarine canyons suggest that any pulse of sedimentation
may have bypassed the preserved portion of the Great Valley forearc (Williams et al.,
1998; Williams and Graham, 2013).
Although the ca. 98-86 Ma Sierra Crest intrusions account for a majority of
present day exposure of the Sierra Nevada batholith (Coleman and Glazner, 1997; Ducea,
2001), detrital zircon grains of this age do not become abundant in the forearc until
Maastrichtian-Paleocene time, suggesting a time lag of at least ca. 13 m.y. between final
pluton emplacement and sufficient arc denudation and erosion to expose the full width of
the batholith (Figs. 3C; 5). Modern levels of erosion in the eastern and northern Sierra
Nevada batholith (< 4 km; Ague and Brimhall, 1988) include volcanic levels of arc crust
(e.g., Minarets Caldera, Fiske and Tobisch, 1994), are significantly less than those in the
eastern Peninsular Ranges batholith (typically 12-20 km; Ague and Brimhall, 1988), and
were attained over a much longer period of time. As a result, the eastern batholithic
signature that is present in the ca. 90 Ma Peninsular Ranges forearc is not found in the
Great Valley forearc until ca. 70 Ma (Fig. 5).
We hypothesize that the contrasting denudation history of the Sierra Nevada and
Peninsular Ranges relates to the location of plateau collision with the margin.
Specifically, regions proximal to the collision site (i.e., southernmost Sierra Nevada, and
northern Peninsular Ranges) experienced higher magnitudes and rates of ca. 90 Ma
denudation than regions farther away from the collision site (i.e., central and northern
Sierra Nevada; Grove et al., 2003b; Saleeby et al., 2007, 2010). This interpretation helps
explain why the Great Valley and Peninsular Ranges forearc segments display opposite
trends in detrital zircon age distributions over time despite developing in comparable
tectonic settings (Fig. 5). Great Valley forearc age distributions progressively become
more similar to the age distribution of plutons in the Sierra Nevada batholith as a result of
progressive arc uplift and denudation during Late Cretaceous-early Paleogene time (Fig.
5; Cecil et al., 2006). This contrasts with age distributions in the Peninsular Ranges
forearc that initially were very similar to the age distribution of plutons in the Peninsular
Ranges batholith but become less so during Eocene time as fluvial systems bypassed the
local batholith and delivered extraregional detritus to the margin (Figs. 5, 8). Because
32

rapid denudation favors low time lags between U-Pb crystallization and depositional
ages, youngest zircons can provide a close approximation of depositional age for strata
sourced from a rapidly unroofing arc (e.g., Kimbrough et al., 2001), but can be
considerably older for unroofed or shallowly denuded volcanic arcs (e.g., CenomanianConiacian strata in Cache Creek; DeGraaff-Surpless et al., 2002).
Geomorphic Breaching of the Southern California Margin Batholith
The pronounced and sudden influx of inboard detritus to the southern California
margin at ca. 75 Ma (Fig. 7) is a clear reflection of the development of a geomorphic
breach within the mid-Cretaceous arc and an associated rapid migration of forearc
drainages into the continental interior (Fig. 8). Although these events have been linked
with the collision of an oceanic plateau or aseismic ridge with the southern California
margin (e.g., Saleeby et al., 2003; Jacobson et al., 2011), the outboard-to-inboard
provenance transition observed in the forearc likely postdated the initial oceanic plateau
collision by ca. 15 Ma (Fig. 8). By ca. 75 Ma the Shatsky conjugate plateau is predicted
to have been positioned beneath the Colorado Plateau region (Liu et al., 2010). This
suggests that the geomorphic breaching of the mid-Cretaceous arc and associated inland
drainage migration (Fig. 8) were more likely related to the passage of the oceanic plateau
beneath the continental interior than to its initial collision with the margin.
Forearc detrital zircon age distributions clearly demarcate the north-south extent
of inboard detritus along the margin and demonstrate that the outboard-to-inboard
provenance transition migrated both to the north and south through Maastrichtian-Eocene
time (Fig. 7). This pattern is consistent with a localized collision of the oceanic plateau
along a ~500 km length of the southern California margin (Figs. 7, 8; Saleeby, 2003).
North-to-south variability within the extraregional provenance signature is related to the
distribution of source regions inboard of the mid-Cretaceous arc, rather than from the
mid-Cretaceous arc itself (Figs. 7, 8).
In contrast, provenance trends in areas to the north and south of the oceanic
plateau collision (Sierra Nevada and Peninsular Ranges segments) still reflect derivation
from the adjacent batholith until drainages migrated to the east to allow extraregional
detritus to reach the forearc (Figs. 7, 8; Cassel et al., 2012). Contrasting denudation
33

histories between the Sierra Nevada and Peninsular Ranges batholiths may have
influenced the timing of inboard drainage migration that occurred earlier (late Paleoceneearly Eocene; Kies and Abbott, 1982) in the Peninsular Ranges than in the Sierra Nevada
(late Eocene-Oligocene; Cassel et al., 2012). The preservation of the Great Valley and
Peninsular Ranges forearc segments can be interpreted to be a consequence of slab
segmentation (Saleeby, 2003) that allowed subduction accretion along the Great Valley
forearc while the Salinian-Mojave segment of the margin underwent tectonic
underplating and subduction erosion (sensu Von Huene and Scholl, 1991). Finally, the
influx of Idahoan detritus to the Eocene Tyee and Great Valley forearcs and Franciscan
Coastal Belt demonstrates that far-field tectonism also contributed to shaping forearc
provenance trends (Dumitru et al., 2012).
Influence on Future Development of the San Andreas Fault System
One of the lasting legacies of Laramide flat-subduction on the California margin
was its influence on preconditioning the later development of the modern Pacific-North
American plate boundary. Following tectonic restructuring of the margin batholith in
southern California, the Salinian-Mojave segment of the margin likely formed a western
bulge in the margin (Johnson and Normark, 1974; Fig. 8CD). This vulnerable position
would later be exploited by strike-slip faults of the San Andreas system that translated
and extended the Salinian block and adjacent crustal pieces of the Mojave Desert region
(e.g., the San Gabriel block) northward (Atwater, 1989). Strictly speaking, these
processes could have begun immediately following the tectonic restructuring of the
margin, inasmuch as dextral obliquity of subduction may have promoted trench-parallel
faulting (Fig. 8D; Doubrovine and Tarduno, 2008; Sharman et al., 2013, their Figure 4).
This mechanism may account for the ~100 km of late Paleogene northward translation of
the Salinian block that is required by our reconstruction (Sharman et al., 2013; Fig. 2A).
Additional northward displacement followed Pacific-North American plate contact (ca.
28-26 Ma) and development of the modern San Andreas system that resulted in >550 km
of cumulative northward translation of portions of the Salinian block (Fig. 2; Atwater,
1989; Dickinson et al., 2005).

34

CONCLUSIONS
Spatial and temporal trends in sedimentary provenance within in the California
forearc can be understood as manifestations of Laramide low-angle subduction and
associated oceanic plateau subduction on the landscape evolution of the western U.S.
continental margin. As such, our results yield important insights into the tectonic
evolution of a well-studied arc-forearc system.
In general, pre-Maastrichtian detrital zircon age distributions demonstrate the
existence of a drainage divide formed by a high-standing mid-Cretaceous Cordilleran arc
that effectively shielded the forearc from inboard sediment source regions. North-south
trends in detrital zircon age distributions can be explained in terms of the primary, crosscutting geometry between the mid-Cretaceous and older arc segments and along-strike,
differential uplift and denudation of the mid-Cretaceous batholith that is hypothesized to
be related to the initial impingement of an oceanic plateau on the continental margin.
A pronounced transition from local to extraregional provenance in the
Maastrichtian-Eocene southern California forearc indicates that a geomorphic breach
developed within the formerly high-standing mid-Cretaceous arc that allowed forearc
drainages to migrate into the continental interior. These events likely postdated the initial
collision of the Shatsky conjugate plateau with the margin by ca. 15 Ma, suggesting that
breaching of the southern California margin was more likely related to the passage of the
subducted oceanic plateau beneath the continental interior than to its initial collision with
the margin. Forearc strata north and south of the breached margin continued to be derived
from the adjacent mid-Cretaceous batholith until progressive eastward drainage migration
allowed extraregional detritus to be delivered to the forearc.
The California forearc provides a case study of the utility of detrital zircon
geochronology in resolving landscape evolution in an active margin setting. When
applied to arc-adjacent depocenters, detrital zircon provenance analysis allows highresolution discrimination of potential source regions in a way that is not possible using
traditional petrographic techniques alone. This is particularly true where the source
regions are segments of a volcano-plutonic arc complex with distinct and welldocumented spatial variations in crystallization age. We found the relative abundances of
arc-derived age populations to be a sensitive indicator of provenance, in addition to the
35

presence or absence of distinctive age populations. Thus, detrital zircon geochronology


constitutes a valuable tool for understanding the tectonic and paleogeographic evolution
of arc-forearc systems.
ACKNOWLEDGEMENTS
Funding for detrital zircon analyses was provided by the Stanford Project on
Deep-water Depositional Systems, the Stanford School of Earth Sciences Chevron Fund,
the Department of Geological Sciences at San Diego State, and the National Science
Foundation grant EAR-0409869. 2014 The Geological Society of America. Facilities
at the Arizona Laserchron Center and the University of California are funded by the
National Science Foundations Instrumentation and Facilities Division and Major
Research Instrumentation programs, respectively. We thank the Wildlands Conservancy
and B. Cotton for granting access to private land, R. McLaughlin and R. Stanley for
assistance with sample collecting in the Sierra Azul block of the Santa Cruz Mountains,
K. Baldwin for assistance with collecting the Quad Sandstone at Stanford University,
and P. Abbott for assistance with collection of Upper Cretaceous and Paleocene
Peninsular Ranges and Channel Island samples. This study benefited from discussion
with P. Abbott, I. Alarik, A. Almgren, A. Barth, E. Brabb, E. Cassel, W. Dickinson, A.
Doebbert, T. Dumitru, C. Jacobson, M. Johns, G. Mahood, M. Malkowski, R.
McLaughlin, E. Miller, J. Saleeby, C. Sine, R. Stanley, J. Wooden, and N. Van Buer. We
thank the many individuals that provided assistance with processing, analyzing, and
reducing data: B. Burgreen, A. Bernhardt, E. Cassel, W. Chang, M. Coble, T. Dumitru, J.
Fosdick, N. Giesler, E. Gottlieb, J. Hourigan, J. Kimbrough, B. Mahoney, K. Maier, M.
Malkowski, S. Maatta, T. OBrien, M. Pecha, N. Sarto, and T. Schwartz. We thank C.
Jacobson, K. Surpless, and W. Dickinson for providing constructive feedback that greatly
improved the readability and content of this chapter.
REFERENCES
Abbott, P.L., and Smith, T.E., 1978, Trace element comparison of clasts in Eocene
conglomerates, southwestern California and northwestern Mexico: Journal of
Geology, v. 86, p. 753762.

36

Abbott, P.L., and Smith, T.E., 1989, Sonora, Mexico, source for the Eocene Poway
Conglomerate of southern California: Geology, v. 17, p. 329332.
Ague, J.J., and Brimhall, G.H., 1988, Magmatic arc asymmetry and distribution of
anomalous plutonic belts in the batholiths of California: Effects of assimilation,
crustal thickness, and depth of crystallization: Geological Society of America
Bulletin, v. 100, p. 912927.
Atwater, T.M., 1989, Plate tectonic history of the northeast Pacific and western North
America, in Winterer, E.L., Hussong, D.M., and Decker, R.W., eds., The Eastern
Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America,
Geology of North America, v. N, p. 2172.
Barbeau, D.L., Jr., Ducea, M.N., Gehrels, G.E., Kidder, S., Wetmore, P.H., and Saleeby,
J.B., 2005, U-Pb detrital-zircon geochronology of northern Salinian basement and
cover rocks: Geological Society of America Bulletin, v. 117, p. 466-481, doi:
10.1130/B25496.1.
Barth, A.P., and Wooden, J.L., 2006, Timing of magmatism following initial
convergence at a passive margin, southwestern U.S. Cordillera, and ages of lower
crustal magma sources: The Journal of Geology, v. 114, p. 231-245.
Barth, A. P., Wooden, J. L., Tosdal, R. M., and Morrison, J., 1995, Crustal contamination
in the petrogenesis of a calc-alkalic rock series: Josephine Mountain intrusion,
California: Geological Society of America Bulletin, v. 107, p. 201211.
Barth, A.P., Wooden, J.L., Grove, M., Jacobson, C.E., and Pedrick, J.N., 2003, U-Pb
zircon geochronology of rocks in the Salinas Valley region of California: A
reevaluation of the crustal structure and origin of the Salinian block: Geology, v.
31, p. 517520.
Barth, A.P., Wooden, J.L., Jacobson, C.E., and Probst, K., 2004, U-Pb geochronology
and geochemistry of the McCoy Mountains Formation, southeastern California: A
Cretaceous retro-arc foreland basin: Geological Society of America Bulletin, v.
116, p. 142153, doi: 10.1130/B25288.1.
Barth, A.P., Wooden, J.L., Jacobson, C.E., and Economos, R.C., 2013, Detrital zircon as
a proxy for tracking the magmatic arc system: The California arc example:
Geology, v. 41, p. 223-226, doi: 10.1130/G33619.1.
Barton, M.D., Girardi, J.D., Kreiner, D.C., Seedorff, E., Zurcher, L., Dilles, J.H., Haxel,
G.B., and Johnson, D.A., 2011, Jurassic igneous-related metallogeny of
southwestern North America, in Steinger, R.L., and Pennell, B., eds., Great Basin
Evolution and Metallogeny: Geological Society of Nevada, Reno, Nevada, p. 373396.
Bottjer, D.J., and Link, M.H., 1984, A synthesis of Late Cretaceous southern California
and northern Baja California paleogeography, in Crouch, J.K., and Bachman,
S.B., eds., Tectonics and Sedimentation Along the California Margin: Pacific
Section, SEPM (Society for Sedimentary Geology), v. 38, p. 171188.
Cadrin, A.A.J., Kyser, T.K., Caldwell, W.G.E., and Longstaffe, F.J., 1996, Isotopic and
chemical compositions of bentonites as paleoenvironmental indicators of the
Cretaceous Western Interior Seaway: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 119, p. 301-320.

37

Cassel, E.J., Grove, M., and Graham, S.A., 2012, Eocene drainage evolution and erosion
of the Sierra Nevada batholith across northern California and Nevada: American
Journal of Science, v. 312, p. 117-144, doi: 10.2475/02.2012.03.
Cecil, M.R., Ducea, M.N., Reiners, P.W., and Chase, C.G., 2006, Cenozoic exhumation
of the northern Sierra Nevada, California, from (U-Th)/He thermochronology:
Geological Society of America Bulletin, v. 118, p. 1481-1488,
doi:10.1130/B25876.1.
Cecil, M.R., Ducea, M.N., Reiners, P., Gehrels, G., Mulch, A., Allen, C., and Campbell,
I., 2010, Provenance of Eocene river sediments from the central northern Sierra
Nevada and implications for paleotopography: Tectonics, v. 29, TC6010,
doi:10.1029/2010TC002717, 13 p.
Centeno-Garca, E., Guerrero-Suastegui, M., and Talavera-Mendoza, O., 2008, The
Guerrero Composite Terrane of western Mexico: Collision and subsequent rifting
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.
Chapman, A.D., Saleeby, J.B., Wood, D.J., Piasecki, A., Kidder, S., Ducea, M.N., and
Farley, K.A., 2012, Late Cretaceous gravitational collapse of the southern Sierra
Nevada batholith, California: Geosphere, v. 8, p. 314-341.
Chapman, A.D., Saleeby, J.B., and Eiler, J., 2013, Slab flattening trigger for isotopic
disturbance and magmatic flare-up in the southernmost Sierra Nevada batholith,
California: Geology, v. 41, p. 1007-1010.
Chen, J.H., and Moore, J.G., 1982, Uranium-lead isotopic ages from the Sierra Nevada
batholith: Journal of Geophysical Research, v. 87, p. 47614784, doi:
10.1029/JB087iB06p04761.
Chetel, L.M., Janecke, S.U., Carroll, A.R., Beard, B.L., Johnson, C.M., and Singer, B.S.,
2011, Paleogeographic reconstruction of the Eocene Idaho River, North American
Cordillera: Geological Society of America Bulletin, v. 123, p. 7188,
doi:10.1130/B30213.1.
Coleman, D.S., and Glazner, A.F., 1997, The Sierra Crest magmatic event: rapid
formation of juvenile crust during the Late Cretaceous in California: International
Geology Review, v. 39, p. 768-787.
Coleman, D.S., Bartley, J.M., Glazner, A.F., Pardue, M.J., 2012, Is chemical zonation in
plutonic rocks driven by changes in source magma composition or shallow-crustal
differentiation?: Geosphere, v. 8, p. 1568-1587.
Cowan, D.S., and Bruhn, R.L., 1992, Late Jurassic to early Late Cretaceous geology of
the U.S. Cordillera, in Burchfiel, B.C., Lipman, P.W., and Zoback, M.L., eds.,
The Cordilleran Orogen: Boulder, Colorado, Geological Society of America, The
Geology of North America, v. G-3, p. 169189.
Cox, B.F., 1982, Stratigraphy, sedimentology, and structure of the Goler Formation
(Paleocene), El Paso Mountains, California: Implications for Paleogene tectonism
on the Garlock Fault Zone [Ph.D. thesis]: Riverside, University of California, 248
p.
Crowell, J.C., 2003, Tectonics of Ridge Basin region, southern California, in Crowell,
J.C., ed., Evolution of Ridge Basin, southern California: An interplay of

38

sedimentation and tectonics: Boulder, Colorado, Geological Society of America


Special Paper 367, p. 157203.
DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and
foreland basin system, western U.S.A.: American Journal of Science, v. 304, p.
105168, doi: 10.2475/ajs.304.2.105.
DeGraaff-Surpless, K., Graham, S.A., Wooden, J.L., and McWilliams, M.O., 2002,
Detrital zircon provenance analysis of the Great Valley Group, California:
Evolution of an arc-forearc system: Geological Society of America Bulletin, v.
114,
p.
15641580,
doi:
10.1130/00167606(2002)114<1564:DZPAOT>2.0.CO;2.
Dickinson, W.R., 1983, Cretaceous sinistral strike slip along Nacimiento fault in coastal
California: The American Association of Petroleum Geologists Bulletin, v. 67, p.
624645.
Dickinson, W.R., 1995a, Forearc basins, in Busby, C.J., and Ingersoll, R.V., eds.,
Tectonics of sedimentary basins: Cambridge, Massachusetts, Blackwell Science,
p. 221261.
Dickinson, W.R., 1995b, Paleogene depositional systems of the western Transverse
Ranges and adjacent southernmost Coast Ranges, California, in Fritsche, A.E.,
ed., Cenozoic Paleogeography of the Western United StatesII: Los Angeles,
Pacific Section, SEPM (Society for Sedimentary Geology), Book 75, p. 5883.
Dickinson, W.R., 1996, Kinematics of transrotational tectonism in the California
Transverse Ranges and its contribution to cumulative slip along the San Andreas
transform fault system: Geological Society of America Special Paper 305, 46 p.
Dickinson, W.R., 2008, Accretionary Mesozoic-Cenozoic expansion of the Cordilleran
continental margin in California and adjacent Oregon: Geosphere, v. 4, p. 329353, doi: 10.1130/GES00105.1.
Dickinson, W.R., and Gehrels, G.E., 2008, Sediment delivery to the Cordilleran foreland
basin: Insights from U-Pb ages of detrital zircons in upper Jurassic and
Cretaceous strata of the Colorado Plateau: American Journal of Science, v. 308, p.
1041-1082, doi: 10.2475/10.2008.01.
Dickinson, W.R., and Gehrels, G.E., 2009, U-Pb ages of detrital zircons in Jurassic eolian
and associated sandstones of the Colorado Plateau: Evidence for transcontinental
dispersal and intraregional recycling of sediment: Geological Society of America
Bulletin, v. 121, p. 408433.
Dickinson, W.R., and Lawton, T.F., 2001, Carboniferous to Cretaceous assembly and
fragmentation of Mexico: Geological Society of America Bulletin, v. 113, p.
11421160.
Dickinson, W.R., and Seely, D.R., 1979, Structure and stratigraphy of forearc regions:
The American Association of Petroleum Geologists Bulletin, v. 63, p. 2-31.
Dickinson, W.R., and Snyder, W.S., 1978, Plate tectonics of the Laramide orogeny, in
Matthews, V., III, ed., Laramide Folding Associated with Basement Block
Faulting in the Western United States: Geological Society of America Memoir
151, p. 355-366.
Dickinson, W.R., Ingersoll, R.V., and Graham, S.A., 1979, Paleogene sediment dispersal
and paleotectonics in northern California: Geological Society of America
Bulletin, v. 90, pt. I, p. 897898; pt. II, p. 14581528.
39

Dickinson, W.R., Ducea, M., Rosenberg, L.I., Greene, H.G., Graham, S.A., Clark, J.C.,
Weber, G.E., Kidder, S., Ernst, W.G., and Brabb, E.E., 2005, Net dextral slip,
Neogene San GregorioHosgri Fault Zone, Coastal California: Geologic Evidence
and Tectonic Implications: Geological Society of America Special Paper 391, 43
p.
Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A.,
2012, Provenance of the Paleogene Colton Formation (Uinta Basin) and
Cretaceous-Paleogene provenance evolution in the Utah foreland: Evidence from
U-Pb ages of detrital zircons, paleocurrent trends, and sandstone petrofacies:
Geosphere, v. 8, p. 854-880, doi:10.1130/GES00763.1.
Dilles, J.H., and Wright, J.E., 1988, The chronology of early Mesozoic arc magmatism in
the Yerington district of western Nevada and its regional implications: Geological
Society of America Bulletin, v. 100, p. 644-652.
Doebbert, A.C., Carroll, A.R., and Johnson, C., 2012, The sandstone-derived provenance
record of the Gualala basin, northern California, U.S.A.: Journal of Sedimentary
Research, v. 82, p. 841-858, doi: 10.2110/jsr.2012.72.
Dokka, R.K., and Travis, C.J., 1990, Late Cenozoic strike-slip faulting in the Mojave
Desert, California: Tectonics, v. 9, p. 311340.
Doubrovine, P.V., and Tarduno, J.A., 2008, A revised kinematic model for the relative
motion between Pacific oceanic plates and North America since the Late
Cretaceous: Journal of Geophysical Research, v. 113, B12101,
doi:10.1029/2008JB005585.
Ducea, M., 2001, The California arc: Thick granitic batholiths, eclogitic residues,
lithospheric-scale thrusting and magmatic flare-ups: GSA Today, v. 11, p. 410.
Ducea, M.N., and Barton, M.D., 2007, Igniting flare-up events in Cordilleran arcs:
Geology, v. 35, p. 1047-1050, doi: 10.1130/G23898A.1.
Ducea, M.N., Kidder, S., Chesley, J.T., and Saleeby, J.B., 2009, Tectonic underplating of
trench sediments beneath magmatic arcs, the central California example:
International Geology Review, v. 51, p. 126, doi:10.1080/00206810802602767.
Dumitru, T.A., Wakabayashi, J., Wright, J.E., and Wooden, J.L., 2010, Early Cretaceous
transition from nonaccretionary behavior to strongly accretionary behavior within
the Franciscan subduction complex: Tectonics, v. 29, TC5001,
doi:10.1029/2009TC002542.
Dumitru, T.A., Ernst, W.G., Wright, J.E., Wooden, J.L., Wells, R.E., Farmer, L.P., Kent,
A.J.R., and Graham, S.A., 2012, Eocene extension in Idaho generated massive
sediment floods into the Franciscan trench and into the Tyee, Great Valley, and
Green River basins: Geology, v. 41, p. 187-190, doi: 10.1130/G33746.1.
Ernst, W.G., Martens, U., and Valencia, V., 2009, U-Pb ages of detrital zircons in
Pacheco Pass metagraywackes: Sierran-Klamath source of mid- and Late
Cretaceous Franciscan deposition and underplating: Tectonics, v. 28, TC6011,
doi:10.1029/2008TC002352.
Fildani, A., Hessler, A.M., and Graham, S.A., 2008, Trench-forearc interactions reflected
in the sedimentary fill of Talara basin, northwest Peru: Basin Research, v. 20, p.
305-331.

40

Fiske, R.S., and Tobisch, O.T., 1994, Middle Cretaceous ash-flow tuff and calderacollapse deposit in the Minarets caldera, east-central Sierra Nevada, California:
Geological Society of America Bulletin, v. 106, p. 582-593.
Foster, D.A., Doughty, P.T., Kalakay, T.J., Fanning, C.M., Coyner, S., Grice, W.C., and
Vogl, J., 2007, Kinematics and timing of exhumation of metamorphic core
complexes along the Lewis and Clark fault zone, northern Rocky Mountains,
USA, in Till, A.B., Roeske, S.M., Sample, J.C., and Foster, D.A., eds.,
Exhumation Associated with Continental Strike-Slip Fault Systems: Geological
Society of America Special Paper 434, p. 207-232.
Gaschnig, R.M., Vervoort, J.D., Lewis, R.S., and Tikoff, B., 2011, Isotopic evolution of
the Idaho batholith and Challis intrusive province, northern US Cordillera:
Journal of Petrology, v. 52, p. 23972429, doi:10.1093/petrology/egr050.
Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy,
efficiency, and spatial resolution of U-Th-Pb ages by LA-MC-ICPMS:
Geochemistry,
Geophysics,
Geosystems,
v.
9,
Q03017,
doi:
10.1029/2007GC001805.
George, P.G., and Dokka, R.K., 1994, Major Late Cretaceous cooling events in the
eastern Peninsular Ranges, California, and their implications for Cordilleran
tectonics: Geological Society of America Bulletin, v. 106, p. 903914.
Graham, S.A., 1976a, Tertiary sedimentary tectonics of the central Salinian block of
California [Ph.D. dissertation]: Stanford, California, Stanford University, 510 p.
Graham, S.A., 1976b, Tertiary stratigraphy and depositional environments near Indians
Ranch, Monterey County, California, in Fritsche, A.E., Ter Best, H., Jr., and
Wornardt, W.W., eds., The Neogene symposium: Los Angeles, Pacific Section,
SEPM (Society for Sedimentary Geology), p. 125136.
Graham, S.A., 1978, Role of Salinian block in evolution of San Andreas fault system,
California: The American Association of Petroleum Geologists Bulletin, v. 62, p.
22142231.
Graham, S.A., and Berry, K.D., 1979, Early Eocene paleogeography of the central San
Joaquin Valley: Origin of the Cantua Sandstone, in Armentrout, J.M., Cole, M.R.,
and Ter Best, H., Jr., eds., Cenozoic paleogeography of the western United States:
Pacific Section, SEPM (Society for Sedimentary Geology), Pacific Coast
Paleogeography Symposium 3, p. 119127.
Graham, S.A., Stanley, R.G., Bent, J.V., and Carter, J.B., 1989, Oligocene and Miocene
paleogeography of central California and displacement along the San Andreas
fault: Geological Society of America Bulletin, v. 101, p. 711730.
Graymer, R.W., Sarna-Wojcicki, A.M., Walker, J.P., McLaughlin, R.J., and Fleck, R.J.,
2002, Controls on timing and amount of right-lateral offset on the East Bay fault
system, San Francisco Bay region, California: Geological Society of America
Bulletin, v. 114, p. 1471-1479.
Grove, K., 1993, Latest Cretaceous basin formation within the Salinian terrane of westcentral California: Geological Society of America Bulletin, v. 105, p. 447463.
Grove, M., Jacobson, C.E., Barth, A.P., and Vucic, A., 2003a, Temporal and spatial
trends of Late CretaceousEarly Tertiary underplating of Pelona and related
schists beneath southern California and southwestern Arizona, in Johnson, S.E., et

41

al., eds., Tectonic Evolution of Northwestern Mexico and the Southwestern USA:
Geological Society of America Special Paper 374, p. 381406.
Grove, M., Lovera, O.M., and Harrison, T.M., 2003b, Late Cretaceous cooling of the
east-central Peninsular Ranges Batholith (33 degrees N): Relationship to La Posta
Pluton emplacement, Laramide shallow subduction, and forearc, in Johnson, S.E.,
et al., eds., Tectonic Evolution of Northwestern Mexico and the Southwestern
USA: Geological Society of America Special Paper 374, p. 355379.
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.G., 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. 335362.
Gutscher, M.A., Spakman, W., Bijward, H., and Engdahl, E.R., 2000, Geodynamics of
flat subduction: Seismicity and tomographic constraints from the Andean margin:
Tectonics, v. 19, p. 814-833, doi: 10.1029/1999TC001152.
Hall, C.A., Jr., 1991, Geology of the Point SurLopez Point Region, Coast Ranges,
California: A Part of the Southern California Allochthon: Geological Society of
America Special Paper 266, 40 p.
Hall, C.A., Jr., and Saleeby, J.B., 2013, Salinia revisited: a crystalline nappe sequence
lying above the Nacimiento fault and dispersed along the San Andreas fault
system, central California: International Geology Review, v. 55, p. 1575-1615.
Henderson, L.J., Gordon, R.G., and Engebretson, D.C., 1984, Mesozoic aseismic ridges
on the Farallon plate and southward migration of shallow subduction during the
Laramide orogeny: Tectonics, v. 3, p. 121-132.
Henry, C.D., Hinz, N.H., Faulds, J.E., Colgan, J.P., John, D.A., Brooks, E.R., Cassel,
E.J., Garside, L.J., Davis, D.A., and Castor, S.B., 2012, EoceneEarly Miocene
paleotopography of the Sierra NevadaGreat BasinNevadaplano based on
widespread ash-flow tuffs and paleovalleys: Geosphere, v. 8, p. 127, doi:10.1130
/GES00727.1.
Herzig, C.T., and Kimbrough, D.L., 2014, Santiago Peak volcanics: Early Cretaceous arc
volcanism of the western Peninsular Ranges batholith, Baja California and
Southern California, in Morton, D.M., and Miller, F.K., eds., Peninsular Ranges
Batholith, Baja California and Southern California Geological Society of America
Memoir 211, p. 345-364.
Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults,
California: Geological Society of America Bulletin, v. 64, p. 443458.
House, M.A., Wernicke, B.P., and Farley, K.A., 2001, Paleo-geomorphology of the
Sierra Nevada, California, from (U-Th)/He ages in apatite: American Journal of
Science, v. 301, p. 77102.
Hughes, G.R., and Mahood, G.A., 2011, Silicic calderas in arc settings: Characterisitcs,
distribution, and tectonic controls: Geological Society of America Bulletin, v.
123, p. 1577-1595.
Ingersoll, R.V., 1979, Evolution of the Late Cretaceous forearc basin, northern and
central California: Geological Society of America Bulletin, v. 90, p. 18131826.

42

Ingersoll, R.V., 1983, Petrofacies and provenance of late Mesozoic forearc basin,
northern and central California: The American Association of Petroleum
Geologists Bulletin, v. 67, p. 11251142.
Ingersoll, R.V., Grove, M., Jacobson, C.E., Kimbrough, D.L., and Hoyt, J.F., 2013,
Detrital zircons indicate no drainage link between southern California rivers and
the Colorado Plateau from mid-Cretaceous through Pliocene: Geology, v. 41, p.
311-314, doi: 10.1130/G33807.1.
Irwin, W.P., and Wooden, J.L., 2001, Map showing plutons and accreted terranes of the
Sierra Nevada, California with a tabulation of U/Pb isotopic ages: U.S. Geological
Survey Open-File Report 2001-229, scale 1:1,000,000.
Jacobson, C.E., Oyarzabal, F.R., and Haxel, G.B., 1996, Subduction and exhumation of
the Pelona-Orocopia-Rand schists, southern California: Geology, v. 24, p. 547
550.
Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsaglia, K.M., Gehrels, G.E.,
and Nourse, J.A., 2011, Late Cretaceous-early Cenozoic tectonic evolution of the
southern California margin inferred from provenance of trench and forearc
sediments: Geological Society of America Bulletin, v. 123, p. 485506,
doi:10.1130/B30238.1.
Johnson, J.D., and Normark, W.R., 1974, Neogene tectonic evolution of the Salinian
block, west-central California: Geology, v. 2, p. 1114.
Kanter, L.R., and McWilliams, M.O., 1982, Rotation of the southernmost Sierra Nevada,
California: Journal of Geophysical Research, v. 87, p. 3819-3830.
Kennedy, M.P., and Moore, G.W., 1971, Stratigraphic relations of Upper Cretaceous and
Eocene formations, San Diego coastal area, California: The American Association
of Petroleum Geologists Bulletin, v. 55, p. 709722.
Kidder, S., Ducea, M., Gehrels, G., Patchett, P.J., and Vervoort, J., 2003, Tectonic and
magmatic development of the Salinian Coast Ridge belt, California: Tectonics, v.
22, no. 5, 1058, doi: 10.1029/2002TC0001409, p. 121 to 1220.
Kies, R.P., and Abbott, P.L., 1982, Sedimentology and paleogeography of lower
Paleogene conglomerates, southern California Continental Borderland, in Fife,
D.L., and Minch, J.A., eds., Geology and mineral wealth of the California
Transverse Ranges: South Coast Geological Society, p. 337-349.
Kimbrough, D.L., Smith, D.P., Mahoney, B.J., Moore, T.E., Grove, M., Gastil, R.G.,
Ortega-Rivera, A., and Fanning, M.C., 2001, Forearc-basin sedimentary response
to rapid Late Cretaceous batholith emplacement in the Peninsular Ranges of
southern and Baja California: Geology, v. 29, p. 491494.
Kimbrough, D.L., Grove, M., Mahoney, J., Barnard, P., and Keller, B., 2009, Detrital
zircon U-Pb geochronology of the modern Sacramento-San Joaquin River delta,
California and implications for interpretation of Late Cretaceous Great Valley
Group sandstone provenance: Geological Society of America Abstracts with
Programs, v. 41, no. 7, p. 657.
Kimbrough, D.L., Abbott, P.L., Balch, D.C., Bartling, S.H., Grove, M., Mahoney, J.B.,
and Donohue, R.F., 2014a, Upper Jurassic Peasquitos FormationForearc basin
western wall rock of the Peninsular Ranges batholith, in Morton, D.M., and
Miller, F.K., eds., Peninsular Ranges Batholith, Baja California and Southern
California: Geological Society of America Memoir 211, p. 625644.
43

Kimbrough, D.L., Grove, M., and Morton, D.M., 2014b, Timing and significance of
gabbro emplacement within two distinct plutonic domains of the Peninsular
Ranges batholith, Southern and Baja California, Geological Society of America
Bulletin, (in press).
Kistler, R.W., and Champion, D.E., 2001, Rb-Sr whole-rock and mineral ages, K-Ar,
40
Ar/39Ar, and U-Pb mineral ages, and strontium, lead, neodymium, and oxygen
isotopic compositions for granitic rocks from the Salinian composite terrane,
California: U.S. Geological Survey Open-File Report 01-453, 84 p.
Kooser, M., 1982, Stratigraphy and sedimentology of the type San Francisquito
Formation, southern California, in Crowell, J.C., and Link, M.H., eds., Geologic
History of Ridge Basin, Southern California: Los Angeles, Pacific Section, SEPM
(Society for Sedimentary Geology), p. 5361.
LaMaskin, T.A., 2012, Detrital zircon facies of Cordilleran terranes in western North
America: GSA Today, v. 22, p. 4-11, doi: 10.1130/GSATG142A.1.
Laskowski, A.K., DeCelles, P.G., and Gehrels, G.E., 2013, Detrital zircon geochronology
of Cordilleran retroarc foreland basin strata, western North America: Tectonics, v.
32, p. 1027-1048.
Laursen, J., Scholl, D.W., and von Huene, R., 2002, Neotectonic deformation of the
central Chile margin: Deepwater forearc basin formation in response to hot spot
ridge
and
seamount
subduction:
Tectonics,
v.
21,
1038,
doi:10.1029/2001TC901023, p. 2-1 to 2-27.
Lechler, A.R., and Niemi, N.A., 2011, Sedimentologic and isotopic constraints on the
Paleogene paleogeography and paleotopography of the southern Sierra Nevada,
California: Geology, v. 39, p. 379382, doi:10.1130/G31535.1.
Linn, A.M., DePaolo, D.J., and Ingersoll, R.V., 1992, Nd-Sr isotopic, geochemical, and
petrographic stratigraphy and paleotectonic analysis: Mesozoic Great Valley
forearc sedimentary rocks of California: Geological Society of America Bulletin,
v. 104, p. 12641279.
Lipman, P.W., 1992, Magmatism in the Cordilleran United States; Progress and
problems, in Burchfield, B.L., Zoback, M.L., and Lipman, P., eds., The
Cordilleran Orogen: Conterminous U.S.: Boulder, Colorado, Geological Society
of America, The Geology of North America, v. G-3, p. 107168.
Liu, L., Gurnis, M., Seton, M., Saleeby, J., Muller, R.D., and Jackson, J.M., 2010, The
role of oceanic plateau subduction in the Laramide orogeny: Nature Geoscience,
v. 3, p. 353357, doi:10.1038/ngeo829.
Liu, L., Spasjevic, S., and Gurnis, M., 2008, Reconstructing Farallon plate subduction
beneath North America back to the Late Cretaceous: Science, v. 322, p. 934-938.
Livaccari, R.F., Burke, K., and Sengor, A.M.C., 1981, Was the Laramide orogeny related
to subduction of an oceanic plateau?: Nature, v. 289, p. 276278.
Lowe, T.K., 1996, Petrogenesis of the Minarets and Merced Peak volcanic-plutonic
complexes, Sierra Nevada, California [Ph.D. thesis]: Stanford University, 157 p.
Luyendyk, B.P., 1991, A model for Neogene crustal rotations, transtension, and
transpression in southern California: Geological Society of America Bulletin, v.
103, p. 15281536.

44

Matthews, V., III, 1976, Correlation of Pinnacles and Neenach volcanic fields and their
bearing on San Andreas fault problem: The American Association of Petroleum
Geologists Bulletin, v. 60, p. 21282141.
Mattinson, J.M., 1990, Petrogenesis and evolution of the Salinian magmatic arc, in
Anderson, J.L., ed., The nature and origin of Cordilleran magmatism: Geological
Society of America Memoir 174, p. 237250.
McDowell, F.W., Roldn-Quintana, J., and Connelly, J.N., 2001, Duration of Late
Cretaceousearly Tertiary magmatism in east-central Sonora, Mexico: Geological
Society of America Bulletin, v. 113, p. 521531.
MGonigle, J.W., and Dalrymple, G.B., 1996, 40Ar/39Ar Ages of Challis volcanic rocks
and the initiation of Tertiary sedimentary basins in southwestern Montana: U.S.
Geological Survey Bulletin 2132, 17 p.
Miller, E.L., Miller, M.M., Wright, J.E., and Madrid, R., 1992, Late Paleozoic
paleogeographic and tectonic evolution of the Western U.S. Cordillera, in
Burchfiel, C., Lipman, W., and Zoback, S., eds., The Cordilleran Orogen;
Conterminous U.S.: Geological Society of America, Geology of North America,
v. G-3, p. 57106.
Mitchell, C., Graham, S.A., and Suek, D.H., 2010, Subduction complex uplift and
exhumation and its influence on Maastrichtian forearc stratigraphy in the Great
Valley Basin, northern San Joaquin Valley, California: Geological Society of
America Bulletin, v. 122, p. 2063-2078, doi: 10.1130/B30180.1.
Moxon, I.A.,1990, Stratigraphic and structural architecture of the San Joaquin and
Sacramento valleys [Ph.D. thesis]: Stanford University, 371 p.
Moxon, I.W., and Graham, S.A., 1987, History and controls of subsidence in the Late
CretaceousTertiary Great Valley forearc basin, California: Geology, v. 15, p.
626629.
Mulch, A., Graham, S.A., and Chamberlain, C.P., 2006, Hydrogen isotopes in Eocene
river gravels and paleoelevation of the Sierra Nevada: Science, v. 313, p. 87-89.
Nicholson, C., Sorlien, C.C., Atwater, T., Crowell, J.C., and Luyendyk, B.P., 1994,
Microplate capture, rotation of the western Transverse Ranges, and initiation of
the San Andreas transform as a low-angle fault system: Geology, v. 22, p. 491
459.
Nilsen, T.H., and Abbott, P.L., 1981, Paleogeography and sedimentology of Upper
Cretaceous turbidites, San Diego, California: The American Association of
Petroleum Geologists Bulletin, v. 65, p. 12561284.
Nilsen, T.H., and Clarke, S.H., Jr., 1975, Sedimentation and tectonics in the early
Tertiary continental borderland of central California: U.S. Geological Survey
Professional Paper 925, 64 p.
Nilsen, T.H., Dibblee, T.W., Jr., and Simoni, T.R., Jr., 1974, Stratigraphy and
sedimentology of the Cantua Sandstone Member of the Lodo Formation,
Vallecitos area, California, in Payne, M., ed., The Paleogene of the Panoche
Creek-Cantua Creek area central California, Pacific Section, SEPM (Society for
Sedimentary Geology), Fall Field Trip Guidebook, p. 38-68.
Poldervaart, A., 1956, Zircon in rocks: 2. Igneous rocks: American Journal of Science, v.
254, p. 521554.

45

Premo, W., Morton, D.M., Snee, L., Naeser, N.D., and Fanning, C.M., 1998, Isotopic
ages, cooling histories, and magmatic origins for Mesozoic tonalitic plutons from
the N. Peninsular Ranges batholith, S. California: Geological Society of America
Abstracts with Programs, v. 30, no. 5, p. 5960.
Prohoroff, R., Wakabayashi, J., and Dumitru, T.A., 2012, Sandstone matrix olistostrome
deposited on intra-subduction complex serpentinite, Franciscan Complex, western
Marin County, California: Tectonophysics, v. 568-569, p. 296-305.
Saleeby, J., 2003, Segmentation of the Laramide slabEvidence from the southern
Sierra Nevada region: Geological Society of America Bulletin, v. 115, p. 655
668.
Saleeby, J., Farley, K.A., Kistler, R.W., and Fleck, R., 2007, Thermal evolution and
exhumation of deep-level batholithic exposures, southernmost Sierra Nevada,
California, in Cloos, M., Carlson, W.D., Gilbert, M.C., Liou, J.G., and Sorensen,
S.S., eds., Convergent Margin Terranes and Associated Regions: A Tribute to
W.G. Ernst: Geological Society of America Special Paper 419, p. 3966, doi:
10.1130/2006.2419(01).
Saleeby, J.B., Ducea, M.N., Busby, C., Nadin, E., and Wetmore, P.H., 2008, Chronology
of pluton emplacement and regional deformation in the southern Sierra Nevada
batholith, California, in Wright, J.E., and Shervais, J.W., eds., Ophiolites, arcs,
and batholiths: A tribute to Cliff Hopson: Geological Society of America Special
Paper 438, p. 397427.
Saleeby, J.B., Saleeby, Z., Liu, L., and Maheo, G., 2010, Mid-Cretaceous regional
exhumation of the Sierra NevadaGreat Valley Batholith and a possible tectonic
driving mechanism: Abstracts with Programs Geological Society of America, v.
42, no. 4, p. 67.
Schulein, B.J., 1993, Sedimentation and tectonics of the upper lower to lower middle
Eocene Domengine Formation Vallecitos syncline, California [M.S. thesis]:
Stanford, California, Stanford University, 343 p.
Schweickert, R.A., and Cowan, D.S., 1975, Early Mesozoic tectonic evolution of the
western Sierra Nevada, California: Geological Society of America Bulletin, v. 86,
p. 13291336.
Seiders, V.M., and Cox, B.T., 1992, Place of origin of the Salinian block, California, as
based on clast compositions of Upper Cretaceous and lower Tertiary
conglomerates: U.S. Geological Survey Professional Paper 1526, 80 p.
Shaw, S.E., Todd, V.R., and Grove, M., 2003, Jurassic peraluminous gneissic granites in
the axial zone of the Peninsular Ranges, southern California, in Johnson, S.E., et
al., eds., Tectonic Evolution of Northwestern Mexico and the Southwestern USA:
Geological Society of America Special Paper 374, p. 157-183.
Sharman, G.R., Graham, S.A., Grove, M., and Hourigan, J.K., 2013, A reappraisal of the
early slip history of the San Andreas fault, central California, USA: Geology, v.
41, p. 727-730, doi: 10.1130/G34214.1.
Silver, L.T., and Chappell, B., 1988, The Peninsular Ranges batholith: An insight into the
Cordilleran batholiths of southwestern North America: Transactions of the Royal
Society of Edinburgh, Earth Sciences, v. 79, p. 105121.
Snow, C.A., Wakabayashi, J., Ernst, W.G., and Wooden, J.L., 2010, Detrital zircon
evidence for progressive underthrusting in Franciscan metagraywackes, west46

central California: Geological Society of America Bulletin, v. 122, p. 282291,


doi:10.1130/B26399.1.
Spencer, J.E., and Normark,W.R., 1979, Tosco-Abreojos fault zone: A Neogene
transform plate boundary within the Pacific margin of southern Baja California:
Geology, v. 7, p. 554557.
Spencer, J.E., Richard, S.M., Gehrels, G.E., Gleason, J.D., and Dickinson, W.R., 2011,
Age and tectonic setting of the Mesozoic McCoy Mountains Formation in western
Arizona, USA: Geological Society of America Bulletin, v. 123, p. 12581274,
doi:10.1130/B30206.1.
Sullivan, R., and Sullivan, M.D., 2012, Sequence stratigraphy and incised valley
architecture of the Domengine Formation, Black Diamond Mines Regional
Preserve and the southern Sacramento basin, California, U.S.A.: Journal of
Sedimentary Research, v. 82, p. 781-800, doi: 10.2110/jsr.2012.66.
Sundberg, F.A., and Cooper, J.D., 1982, Late Cretaceous (Campanian) paleogeography
of southern California, in Bottjer, D.J., Coldburn, I.P., and Cooper, J.D., eds.,
Late Cretaceous Depositional Environments and Paleogeography, Santa Ana
Mountains, Southern California, Pacific Section, SEPM (Society for Sedimentary
Geology), Field Trip Volume and Guidebook, p. 115-121.
Surpless, K.D., and Beverly, E.M., 2013, Understanding a critical basinal link in
Cretaceous Cordilleran paleogeography: Detailed provenance of the Hornbrook
Formation, Oregon and California: Geological Society of America Bulletin, v.
125, p. 709-727, doi: 10.1130/B30690.1.
Tarduno, J.A., McWilliams, M., Debiche, M.G., Sliter, M.V., and Blake, M.C., Jr., 1985,
Franciscan Complex Calera limestones: accreted remnants of Farallon Plate
oceanic plateaus: Nature, v. 317, p. 345-347.
Todd, V.R., Shaw, S.E., and Hammarstrom, J.M., 2003, Cretaceous plutons of the
Peninsular Ranges Batholith, San Diego and westernmost Imperial Counties,
California: Intrusion across a Late Jurassic continental margin, in Johnson, S.E.,
et al., eds., Tectonic Evolution of Northwestern Mexico and the Southwestern
USA: Geological Society of America Special Paper 374, p. 185235.
Underwood, M.B., 1983, Depositional setting of the Paleogene Yager Formation,
Northern Coast Ranges of California, in Larue, D.K., and Steel, R.J., eds.,
Cenozoic marine sedimentation, Pacific margin, U.S.A.: Los Angeles, Pacific
Section, Society of Economic Paleontologists and Mineralogists, p. 81101.
Underwood, M.B., and Moore, G.F., 1995, Trenches and trench-slope basins, in Busby,
C.J., and Ingersoll, R.V., eds., Tectonics of sedimentary basins: Cambridge,
Massachusetts, Blackwell Science, p. 179-219.
Von Huene, R., and Scholl, D.W., 1991, Observations at convergent margin concerning
sediment subduction erosion, and the growth of continental crust: Reviews of
Geophysics, v. 29, p. 279-316.
Walawender, M.J., Gastil, R.G., Clinkenbeard, J.P., McCormick, W.V., Eastman, B.G.,
Wernicke, R.S., Wardlaw, M.S., Gunn, S.H., and Smith, B.M., 1990, Origin and
evolution of the zoned La Posta-type plutons, eastern Peninsular Ranges batholith,
southern and Baja California, in Anderson, J.L., ed., The nature and origin of
Cordilleran magmatism: Geological Society of America Memoir 174, p. 118.

47

Walker, J.D., Martin, M.W., and Glazner, A.F., 2002, Late Paleozoic to Mesozoic
development of the Mojave Desert and environs, California, in Glazner, A.F.,
Walker, J.D., and Bartley, J.M., eds., Geologic Evolution of the Mojave Desert
and Southwestern Basin and Range: Geological Society of America Memoir 195,
p. 118.
Watson, E.B., 1979, Zircon saturation in felsic liquids: Experimental results and
applications to trace element geochemistry: Contributions to Mineralogy and
Petrology, v. 70, p. 407419.
Wells, M.L., and Hoisch, T.D., 2008, The role of mantle delamination in widespread Late
Cretaceous extension and magmatism in the Cordilleran orogen, western United
States: Geological Society of America Bulletin, v. 120, p. 515530, doi:
10.1130/B26006.1.
Wernicke, B.P., 1992, Cenozoic extensional tectonics of the U.S. Cordillera, in Burchfiel,
B.C., Lipman, P.W., and Zoback, M.L., eds., The Cordilleran Orogen:
Conterminous U.S.: Boulder, Colorado, Geological Society of America, The
Geology of North America, v. G-3, p. 553581.
Williams, T.A., 1997, Basin-fill architecture and forearc tectonics: Cretaceous Great
Valley Group, Sacramento basin, northern California [Ph.D. thesis]: Stanford,
California, Stanford University, 412 p.
Williams, T.A., and Graham, S.A., 2013, Controls on forearc basin architecture from
seismic and sequence stratigraphy of the Upper Cretaceous Great Valley Group,
central Sacramento Basin, California: International Geology Review, v. 55, p.
2030-2059.
Williams, T.A., Graham, S.A., and Constenius, K.N., 1998, Recognition of a Santonian
submarine canyon, Great Valley Group, Sacramento Basin, California:
Implications for petroleum exploration and sequence stratigraphy of deep-marine
strata: The American Association of Petroleum Geologists Bulletin, v. 82, p.
1575-1595.

48

Sample
Oregon forearc
Tyee
LF020
LF024
LF023
LF021
JV395
JV437
JW290
07-HB-05
07-HB-26
07-HB-21
06-KS-11
07-HB-11
07-HB-13
07-HB-17
06-KS-06
06-KS-04

Formation

Age

Age
Group

Location (abbreviation)

Dumitru et al., 2012


Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013
Surpless and Beverly, 2013

Reference

Table 1. Detrital Zircon Samples


n

Tyee basin (TB)


Tyee basin (TB)
Tyee basin (TB)
Tyee basin (TB)
Tyee basin (TB)
Tyee basin (TB)
Tyee basin (TB)
Tyee basin (TB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)
Hornbrook basin (HB)

DeGraaff-Surpless et al., 2002


DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012
Dumitru et al., 2012

Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
S-C
S-C
S-C
S-C
C-C
C-C
C-C
C-C
C-C

Redding (R)
Redding (R)
Redding (R)
Chico (CH)
Chico (CH)
Chico (CH)
Cache Creek (CC)
Cache Creek (CC)
Cache Creek (CC)
Cache Creek (CC)
Cache Creek (CC)
Franciscan Coastal Belt# (FCB)
Franciscan Coastal Belt# (FCB)
Franciscan Coastal Belt# (FCB)
Franciscan Coastal Belt# (FCB)

early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
late Campanian-Maastrichtian
late Campanian-Maastrichtian
late Campanian-Maastrichtian
Coniacian-Campanian
late Albian-early Coniacian
late Albian-early Coniacian
late Albian-early Coniacian
Albian-Turonian
Albian-Turonian
S-C
C-C
C-C
S-C
S-C
C-C
S-C
C-C
C-C
C-C
C-C
Eoc
Eoc
Eoc
Eoc

Tyee Fm
Tyee Fm
Tyee Fm
Tyee Fm
Tyee Fm
Tyee Fm
Tyee Fm
Tyee Fm
Hornbrook Fm - Blue Gulch Ms Mbr
Hornbrook Fm - Blue Gulch Ms Mbr
Hornbrook Fm - Blue Gulch Ms Mbr
Hornbrook Fm - Rocky Gulch Ss Mbr
Hornbrook Fm - Osburger Gulch Ss Mbr
Hornbrook Fm - Osburger Gulch Ss Mbr
Hornbrook Fm - Osburger Gulch Ss Mbr
Hornbrook Fm - Klamath River Cg Mbr
Hornbrook Fm - Klamath River Cg Mbr
Santonian
Coniacian
Turonian
Campanian
Santonian
Coniacian
Santonian
Coniacian
Coniacian
Turonian
Turonian
Eocene
Eocene
Eocene
Eocene

48
49
24
26
9
53
49
38
82
98
98
97
99
98
61
74
93

Great Valley - Sacramento forearc


KDS11
56
Bear Creek Ss
KDS13
55
Frazier Siltstone
KDS10
56
Bellavista Ss
KDS23
55
Ten Mile
KDS21
60
Musty Buck
KDS22
57
Ponderosa Way
KDS3
56
Guinda Ss
GV44
56
Sites Ss
GV45
56
Yolo Shale
GV42
56
Venado Ss
GV40
56
Fiske Creek Shale
CR-2
93
Franciscan Coastal terrane
CR-25
89
Franciscan Yager terrane
CR-85
69
Franciscan Yager terrane
CR-90
96
Franciscan Yager terrane

49

CR-55
SN07-060NC
SN09-001DF
SN09-006YB
SN09-007RD
SN09-070HP
CEC-3a
CEC-4
CEC-5
CEC-1
CEC-2
GVG09-8
GVG09-7
GVG09-9
TD1F-105
GVG09-18
GVG09-0.5
GVG09-22
GVG09-23
GVG09-24
GVG09-17
GVG09-16
GVG09-15
GVG09-14
GVG09-13
GVG09-12

92
97
92
106
97
95
102
99
100
90
96
95
92
83
90
87
89
85
94
96
94
94
100
89
93
93

Franciscan Coastal terrane


Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Eocene "auriferous gravels"
Ione Fm
Ione Fm
Ione Fm
Markley Ss
Markley Ss
Domengine Ss
Domengine Ss
Domengine Ss
Domengine Ss
Meganos Ss
Meganos Ss
Deer Valley Fm
Unnamed
Unnamed
Unnamed

Great Valley - San Joaquin forearc


GVG09-1
97
Telsa Fm
99-65
36
Panoche Fm
GVG09-2
86
Panoche Fm
SG-20
99
Domengine Ss
SG-21
97
Cantua Ss
SG-22
99
San Carlos Ss
GV13
56
Joaquin Ridge
GV8
56
Joaquin Ridge
GV21
54
Upper Los Gatos Creek
GV33
56
Lower Los Gatos Creek

Paleocene-early Eocene
Campanian
Campanian
early Eocene
early Eocene
Paleocene
Campanian
Campanian
Santonian
Turonian

Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
Eocene
middle Eocene
middle Eocene
middle Eocene
middle Eocene
middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early Eocene
late Paleocene-early Eocene
Maastrictian
Campanian-Maastrichtian
late Campanian
Turonian
M-P**
S-C
S-C
Eoc
Eoc
M-P
S-C
S-C
S-C
C-C

Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
M-P**
M-P**
M-P
S-C
S-C
C-C

Del Puerto Canyon (DPC)


Del Puerto Canyon (DPC)
Del Puerto Canyon (DPC)
Vallecitos Syncline (VS)
Vallecitos Syncline (VS)
Vallecitos Syncline (VS)
Coalinga (C)
Coalinga (C)
Coalinga (C)
Coalinga (C)

Franciscan Coastal Belt# (FCB)


Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Northern Sierra Nevada (NSN)
Central Sierra Nevada (CSN)
Central Sierra Nevada (CSN)
Central Sierra Nevada (CSN)
Central Sierra Nevada (CSN)
Central Sierra Nevada (CSN)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)
Mount Diablo (MD)

this study
Jacobson et al., 2011
this study
this study
this study
this study
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002
DeGraaff-Surpless et al., 2002

Dumitru et al., 2012


Cassel et al., 2012
Cassel et al., 2012
Cassel et al., 2012
Cassel et al., 2012
Cassel et al., 2012
Cecil et al., 2010
Cecil et al., 2010
Cecil et al., 2010
Cecil et al., 2010
Cecil et al., 2010
this study
this study
this study
Dumitru et al., 2012
Dumitru et al., 2012
this study
this study
this study
this study
this study
this study
this study
this study
this study
this study

50

GV24
SG-4
SG-23
SG-47
SBM
SG-46
SG-6
POR-1
POR-2
POR-3
SG-40
SEF-1
TEJ-2
TEJ-1
SEM-1

56
98
99
100
55
95
97
100
97
101
97
99
97
100
104

Salinian block forearc


ST08
103
GCV2
161
GC1B
178
GP01
125
SBSS
136
German Rancho Fm
German Rancho Fm
German Rancho Fm
German Rancho Fm
Anchor Bay

Grabast
Whiskey Hill Fm
Greystone Fm
Mount Chual Ss
San Bruno Mt. Ss
Pilarcitos Ss
Pilarcitos Ss
Point of Rocks Ss
Point of Rocks Ss
Point of Rocks Ss
San Emigdio Fm
San Emigdio Fm
Tejon Fm - Metralla Ss Mbr
Tejon Fm - Uvas Cg Mbr
Tejon Fm - Uvas Cg Mbr

Cenomanian
early-middle Eocene
early-middle Eocene
early Eocene
early Eocene (?)
early-middle Eocene (?)
early-middle Eocene (?)
middle Eocene
middle Eocene
middle Eocene
late Eocene
late Eocene
middle Eocene
early-middle Eocene
early-middle Eocene

Eoc
Eoc
Eoc
M-P
S-C

C-C
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc

Gualala block (GB)

Gualala block (GB)


Gualala block (GB)
Gualala block (GB)
Gualala block (GB)
Gualala block (GB)

Coalinga (C)
San Francisco Bay block (SFB)
San Francisco Bay block (SFB)
San Francisco Bay block (SFB)
San Francisco Bay block (SFB)
Pilarcitos block* (PB)
Pilarcitos block* (PB)
Temblor Range (TR)
Temblor Range (TR)
Temblor Range (TR)
San Emigdio Mtns (SEM)
San Emigdio Mtns (SEM)
San Emigdio Mtns (SEM)
San Emigdio Mtns (SEM)
San Emigdio Mtns (SEM)

Doebbert et al., 2012

Doebbert et al., 2012


Doebbert et al., 2012
Doebbert et al., 2012
Doebbert et al., 2012
Doebbert et al., 2012

DeGraaff-Surpless et al., 2002


this study
this study
this study
Snow et al., 2010
this study
this study
Sharman et al., 2013
Sharman et al., 2013
Sharman et al., 2013
this study
Sharman et al., 2013
Sharman et al., 2013
Sharman et al., 2013
Lechler and Niemi, 2011

Anchor Bay

S-C

111

Doebbert et al., 2012

WB03

Gualala block (GB)

Stewarts Point

S-C

105

Doebbert et al., 2012

SM19

Gualala block (GB)

Stewarts Point

S-C

86

Sharman et al., 2013


Sharman et al., 2013
Sharman et al., 2013
Sharman et al., 2013
Sharman et al., 2013
Jacobson et al., 2011
Sharman et al., 2013
Jacobson et al., 2011
Jacobson et al., 2011

SMSS

La Honda Basin (LHB)


La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)
La Honda Basin (LHB)

Butano Ss
Butano Ss
Butano Ss
Butano Ss
Butano Ss
Butano Ss
San Juan Bautista Fm
Point San Pedro Fm
Point San Pedro Fm

Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
M-P
M-P

100
99
100
99
96
28
96
12
14

early-middle Eocene
early-middle Eocene
early-middle Eocene
Paleocene
late Campanian-early
Maastrichtian
late Campanian-early
Maastrichtian
late Campanian-early
Maastrichtian
late Campanian-early
Maastrichtian
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
early-middle Eocene
Paleocene
Paleocene

BUT-1
BUT-2
BUT-3
BUT-4
BUT-5
BB1
SJB-1
MM1
MM1A

51

MM2
MM3
SG-31
SG-28
SG-29
IR3
IR4
IR5
IR1
03-04
MO1
ES1
PS1
PT1
ROJO
ESALEN
ROCKY
POINT

16
20
100
99
98
22
28
21
19
16
13
13
9
13
79
104
89

Transverse Ranges forearc


PP1
12
PP2
14
PP3
13
03-01
14
03-03
18
99-66A
32
99-68
14
YM1
14
YM3
13
YM5
14
YM4
26
LP1
15
LP9
15
TP1
25
LP10
16
LP11
15
BS1
13
BS1A
16
Pigeon Point Fm
Pigeon Point Fm
Pigeon Point Fm
Cambria slab
Cambria slab
Cambria slab
Atascadero Fm
Atascadero Fm
Atascadero Fm
Atascadero Fm
Atascadero Fm
Unnamed
Unnamed
Buckhorn Ss
Unnamed
Unnamed
Pfeiffer slab
Pfeiffer slab

Point San Pedro Fm


Point San Pedro Fm
The Rocks Ss
The Rocks Ss
The Rocks Ss
Church Creek Fm
Church Creek Fm
Merle Fm
Unnamed
Carmelo Fm
Carmelo Fm
Unnamed
Unnamed
Unnamed
Unnamed
Unnamed
Unnamed

Campanian
Campanian
Campanian
Campanian
Campanian
Campanian
Campanian
Campanian
Campanian
Campanian
Cenomanian-Turonian
Campanian
Campanian(?)
Campanian-early Maastrichtian
Cenomanian-Turonian
Cenomanian-Turonian
Coniacian(?)-Campanian
Coniacian(?)-Campanian

Paleocene
Paleocene
middle Eocene
middle Eocene
middle Eocene
late Eocene
late Eocene
Paleocene
Maastrichtian
early Eocene
early Eocene
Maastrichtian
Maastrichtian
Maastrichtian
Campanian(?)-Maastrichtian
Maastrichtian
Campanian(?)-Maastrichtian

S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
C-C
S-C
S-C
S-C
C-C
C-C
S-C
S-C

M-P
M-P
Eoc
Eoc
Eoc
Eoc
Eoc
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P

Pigeon Point block (PP)


Pigeon Point block (PP)
Pigeon Point block (PP)
Cambria slab (CS)
Cambria slab (CS)
Cambria slab (CS)
Atascadero (A)
Atascadero (A)
Atascadero (A)
Atascadero (A)
Atascadero (A)
San Rafael Mtns (SRM)
San Rafael Mtns (SRM)
San Rafael Mtns (SRM)
San Rafael Mtns (SRM)
San Rafael Mtns (SRM)
Point Sur block (PSB)
Point Sur block (PSB)

La Honda Basin (LHB)


La Honda Basin (LHB)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)
Santa Lucia Range (SLR)

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011

Jacobson et al., 2011


Jacobson et al., 2011
this study
this study
this study
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Barbeau et al., 2005
Barbeau et al., 2005
Barbeau et al., 2005

52

SM8

SM6
SM6A
SM2
SM4
SM3

PS2
SF3
PZ1
PZ3
PZ4
SML1
SF1
SF2
SF4
SF5
VA1
OR432
OR433
MP1
MP2
MP3
LV2
LV3
LP3
LP7
LP13
LP14
LP15
LP12
LP21
SH6
SH2
SH8
SH1
SH7

67

15
15
17
15
22

14
13
9
14
12
15
20
13
13
15
12
70
59
15
12
12
12
14
15
16
14
15
14
29
72
48
15
72
15
74

Tuna Canyon Fm

Llajas Fm
Llajas Fm
Santa Susana Fm
Las Virgenes Ss
Tuna Canyon Fm

Pfeiffer slab
"San Francisquito Fm"
Unnamed
Unnamed
Unnamed
Unnamed
San Francisquito Fm
San Francisquito Fm
San Francisquito Fm
San Francisquito Fm
San Francisquito Fm
Maniobra Fm
Maniobra Fm
Unnamed
Unnamed
Unnamed
Juncal Fm
Juncal Fm
Juncal Fm
Cozy Dell Sh
Coldwater Ss
Cozy Dell Sh
Matilija Ss
Jalama Fm?
Jalama Fm
Llajas Fm
Simi Cg
Santa Susana Fm
Chatsworth Fm
Chatsworth Fm

Coniacian(?)-Campanian
Maastrichtian
Maastrichtian
Maastrichtian
Paleocene-Eocene(?)
Maastrichtian
Maastrichtian
Maastrichtian or Paleocene
early-middle Paleocene
early-middle Paleocene
middle or late Paleocene
early Eocene
early Eocene
early Eocene
early Eocene
Paleocene-early Eocene
early Eocene
early Eocene
early Eocene(?)
middle Eocene
middle Eocene
middle Eocene
middle Eocene
late Campanian-Maastrichtian
Maastrichtian
early or middle Eocene
middle Paleocene
late Paleocene
late Campanian
late Campanian or early
Maastrichtian
early or middle Eocene
early or middle Eocene
late Paleocene
late Paleocene
late Campanian or early
Maastrichtian
late Campanian or early
S-C

Eoc
Eoc
M-P
M-P
S-C

S-C
-M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
M-P
M-P
Eoc
M-P
M-P
S-C
S-C

Santa Monica Mountains (SM)

Santa Monica Mountains (SM)


Santa Monica Mountains (SM)
Santa Monica Mountains (SM)
Santa Monica Mountains (SM)
Santa Monica Mountains (SM)

Point Sur block (PSB)


-La Panza Range (LP)
La Panza Range (LP)
La Panza Range (LP)
La Panza Range (LP)
San Gabriel bock (SG)
San Gabriel bock (SG)
San Gabriel bock (SG)
San Gabriel bock (SG)
San Gabriel bock (SG)
Orocopia Mountains (O)
Orocopia Mountains (O)
Sierra Madre Mountains (M)
Sierra Madre Mountains (M)
Sierra Madre Mountains (M)
Pine Mountain block (P)
Pine Mountain block (P)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Santa Ynez Mountains (SYM)
Simi Hills (SH)
Simi Hills (SH)
Simi Hills (SH)
Simi Hills (SH)
Simi Hills (SH)

Jacobson et al., 2011

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011

53

ROS-2

JOLLA-5
JOLLA-1
JOLLA-2
JOLLA-3
JOLLA-4
ROSA-1
ROSA-2
ROSA-3
ROSA-4
ROSA-5
PALM-1
PALM-2
PALM-3
ROS-1

60

65

64
45
50
45
48
30
55
30
58
54
49
52
45
33

Rosario Fm

Rosario Fm

Peninsular Ranges forearc


SNTGU
88
Santiago Fm
SNTGL
98
Santiago Fm
SLVU
51
Silverado Fm
SLVL
31
Silverado Fm
WLM
30
Williams Fm - Shultz Ranch Mbr
MUST
26 Ladd Fm - Mustang Spring Conglomerate
Mbr
BAK
29 Ladd Fm - Baker Canyon Conglomerate
Mbr
TRB
32
Trabuco Fm
SMI0406
53
unnamed
SMI0404
30
unnamed
SMI0403
34
unnamed
SMI0402
34
unnamed
SMI0401
35
unnamed
POW-1
74
Pomerado Conglomerate of the Poway
Group
POW-2
53
Stadium Conglomerate of the Poway
Group
Scripps Fm of the La Jolla Group
Scripps Fm of the La Jolla Group
Delmar Fm of the La Jolla Group
Mount Soledad Fm of the La Jolla Group
Mount Soledad Fm of the La Jolla Group
Cabrillo Fm of the Rosario Group
Point Loma Fm of the Rosario Group
Point Loma Fm of the Rosario Group
Point Loma Fm of the Rosario Group
Point Loma Fm of the Rosario Group (?)
Las Palmas Gravels
Las Palmas Gravels
Las Palmas Gravels
Rosario Fm

ROS-3

late early-early middle Eocene


late early-early middle Eocene
late Paleocene
late Paleocene
late Campanian
middle Campanian
C-C

Eoc
Eoc
M-P
M-P
S-C
S-C

Santa Ana Mtns (SAM)


San Miguel Island (SMI)
San Miguel Island (SMI)
San Miguel Island (SMI)
San Miguel Island (SMI)
San Miguel Island (SMI)
San Diego (SD)

Santa Ana Mtns (SAM)

Santa Ana Mtns (SAM)


Santa Ana Mtns (SAM)
Santa Ana Mtns (SAM)
Santa Ana Mtns (SAM)
Santa Ana Mtns (SAM)
Santa Ana Mtns (SAM)

this study; Jacobson et al., 2011


this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011

this study; Jacobson et al., 2011

this study; Jacobson et al., 2011


this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011

Maastrichtian

late Turonian

C-C
Eoc
M-P
S-C
C-C
C-C
Eoc

this study; Jacobson et al., 2011


this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011
this study; Jacobson et al., 2011

Cenomanian-Turonian
early Eocene
Maastrichtian
Campanian
Coniacian
Turonian
late Eocene

San Diego (SD)


San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
San Diego (SD)
Northern Baja (NB)
Northern Baja (NB)
Northern Baja (NB)
Northern Baja (NB)

this study; Jacobson et al., 2011

this study; Jacobson et al., 2011


Eoc
Eoc
Eoc
M-P
M-P
S-C
S-C
S-C
S-C
S-C
Eoc
Eoc
Eoc
S-C

Northern Baja (NB)

this study; Jacobson et al., 2011

San Diego (SD)

S-C

Northern Baja (NB)

Eoc

S-C

middle to late Eocene


late early Eocene
Eocene
late early Eocene
late Paleocene-early Eocene
late Paleocene-early Eocene
late Campanian
late Campanian
middle Campanian
Campanian
Campanian(?)
middle-late Eocene
middle-late Eocene
middle-late Eocene
late Campanian-early
Maastrichtian
late Campanian-early
Maastrichtian
late Campanian-early

54

RF-2
RF-1
EG-2
EG-1
PB
LBR
9JBM-98

Guad01
Guad02
57
54
60
58
60
59
55

30
60
Rosario Fm
Rosario Fm
El Gallo Fm
El Gallo Fm
Punta Baja Fm
La Bocana Roja Fm
Unnamed

Redonda Fm
Redonda or Rosario Fm(?)

Peninsular Ranges interior


MINE
57
Ballenas gravels (equivalent)

97

Goler Fm

Southern Sierra Nevada intra-arc basins


TM-1
96
Witnet Fm
EPM-1
94
Goler Fm
EPM-2

MC4
30
28

30

Upper McCoy Mountains Fm


Upper McCoy Mountains Fm

Upper McCoy Mountains Fm

McCoy Mountains retroarc basin


3-26-04-2,3
99
Upper McCoy Mountains Fm
MC5
32
Upper McCoy Mountains Fm

MC3
01-327

Cordilleran retroarc foreland


COL-3
98
Horse Bench Mbr of Green River Fm
COL-4
90
Green River Fm
COL-2
86
DeBeque Fm
WADC401
38
DeBeque Fm
COL-6
89
Colton Fm
COL-7
89
Colton Fm
COL-8
94
Colton Fm
CO22
61
Colton Fm

Oligocene-Cretaceous (?)
Paleocene

late Eocene

Maastrichtian
Cenomanian(?)-Turonian
Cenomanian-early
Maastrichtian(?)
early Maastrichtian
late Campanian
late Campanian
late Campanian
Campanian
Turonian
Turonian

M-P

-M-P

Eoc

M-P
S-C
S-C
S-C
S-C
C-C
--

C-C
C-C

-Southernmost Sierra Nevada


(EPM)
Southernmost Sierra Nevada
(EPM)

Mine Wash (MW)

El Rosario (ER)
El Rosario (ER)
El Rosario (ER)
El Rosario (ER)
El Rosario (ER)
El Rosario (ER)
Vizcaino Peninsula (VP)

Northern Baja (NB)


Northern Baja (NB)

this study

this study
this study
this study
this study
this study
this study
Kimbrough et al., 2001

this study; Jacobson et al., 2011


this study; Jacobson et al., 2011

Campanian (?) (> ca. 80 Ma)


Santonian-Campanian (?) (> ca.
84 Ma)
Coniacian-Santonian (?) (> ca. 87
Ma)
Turonian (?) (> ca. 91 Ma)
Turonian (?) (> ca. 91 Ma)

Eoc
Eoc
Eoc
Eoc
M-P
M-P
M-P
M-P

C-C
C-C

S-C

S-C
S-C

Retroarc foreland (RF)


Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)

McCoy Mountains backarc (MM)


McCoy Mountains backarc (MM)

McCoy Mountains backarc (MM)

McCoy Mountains backarc (DR)


McCoy Mountains backarc (MM)

Dickinson et al., 2012


Dickinson et al., 2012
Dickinson et al., 2012
Davis et al., 2009
Dickinson et al., 2012
Dickinson et al., 2012
Dickinson et al., 2012
Davis et al., 2009

Barth et al., 2004


Barth et al., 2004

Barth et al., 2004

Spencer et al., 2011


Barth et al., 2004

Lechler and Niemi, 2011

Lechler and Niemi, 2011


Lechler and Niemi, 2011
Paleocene

early Eocene
early Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
late Paleocene
Paleocene-Eocene

55

M-P
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
S-C
C-C
C-C
C-C
C-C

Franciscan Complex (FC)


Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)
Franciscan Complex (FC)

Retroarc foreland (RF)


Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)
Retroarc foreland (RF)

Prohoroff et al., 2012

Ernst et al., 2009


Ernst et al., 2009
Ernst et al., 2009
Ernst et al., 2009
Snow et al., 2010
Snow et al., 2010
Snow et al., 2010
Snow et al., 2010
Prohoroff et al., 2012

Davis et al., 2009


Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Lawton and Bradford, 2011
Dickinson and Gehrels, 2008
Dickinson et al., 2012
Dickinson and Gehrels, 2008
Dickinson and Gehrels, 2008
Dickinson et al., 2012

late Paleocene
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
late Campanian
Campanian
Turonian
Turonian
Turonian
Cenomanian
S-C
C-C
C-C
C-C
S-C
C-C
C-C
C-C
S-C

Franciscan Complex (FC)

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011

Flagstaff Limestone
Kaiparowits Fm (upper)
Kaiparowits Fm (upper)
Kaiparowits Fm (middle)
Kaiparowits Fm (middle)
Kaiparowits Fm (lower)
Tucher Fm (middle)
Tucher Fm (lower)
Farrer Fm (middle)
Farrer Fm (lower)
Bluecastle Mbr of Castlegate Ss
Upper Mbr of Wahweap Fm
Ferron Ss
Gallup Ss Mbr of Mancos Sh
Ferron Ss Mbr of Mancos Sh
Dakota Fm

71
96
99
98
96
97
89
97
92
100
95
92
92
87
99
95

C-C

Underplated schist (SCH)


Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)

GRF
KK0407
KK0507
KK0607
KK0207
KK0107
MT0907
MT0107
MF0407
MF0607
MB1007
CP39
COL-11
CP23
CP33
COL-10

Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc

Pacheco Pass metagraywackes


Pacheco Pass metagraywackes
Pacheco Pass metagraywackes
Pacheco Pass metagraywackes
Novato Quarry terrane
Hunters Point shear zone
upper El Cerrito
lower El Cerrito
Marin County metagraywackes
Marin County metagraywackes

Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene

Franciscan Complex
Q-24
96
A-1
95
X206
88
Q58
23
NQT
55
HPSZ
55
UEC
45
LEC
41
JW-1Z-006
84
99

Peter Kane Mountain-Gavilan Hills schist


Peter Kane Mountain-Gavilan Hills schist
Peter Kane Mountain-Gavilan Hills schist
Ocoropia schist
Ocoropia schist
Ocoropia schist
Ocoropia schist

Santonian(?) (<~85 Ma)


Coniacian(?) (< ~89 Ma)
Cenomanian(?) (< ~101 Ma)
Cenomanian(?) (< ~101 Ma)
Campanian(?) (< ~83 Ma)
Cenomanian(?) (< ~97 Ma)
Cenomanian(?) (< ~101 Ma)
Cenomanian(?) (< ~101 Ma)
Santonian-Campanian (?) (< ca.
86 Ma)
Cenomanian(?) (<~97 Ma)

OHR-10
Underplated schist
PK114B
27
UG1417A
43
UG1500
9
KE22431
26
KE24033
23
KE12702
24
OR15A
24

56

OR77B
OR113
OR307
OR312A
OR314
OR337
MW10
TR23
TR24A
YN17
KE4
KE6
NR1
NR2
PM1
PM3
PM4
PM9
PM11
98-241
SP10
SP25D
SP60
SP70
SP71
SP81
98-240
BR218
98-237
SG69
SG532
SG533
CD7
CD9
02-342
02-345
02-358
DR-1656

26
33
29
27
24
25
31
16
16
29
12
18
31
35
25
25
24
27
24
50
24
24
25
22
26
28
43
27
30
18
30
32
33
14
35
23
35
27

Ocoropia schist
Ocoropia schist
Ocoropia schist
Ocoropia schist
Ocoropia schist
Ocoropia schist
Picacho-Trigo Mts. schist
Picacho-Trigo Mts. schist
Picacho-Trigo Mts. schist
Picacho-Trigo Mts. schist
SW & SE Castle Dome Mts. schist
SW & SE Castle Dome Mts. schist
Neversweat Ridge schist
Neversweat Ridge schist
Mount Pinos schist
Mount Pinos schist
Mount Pinos schist
Mount Pinos schist
Mount Pinos schist
Sierra Pelona schist
Sierra Pelona schist
Sierra Pelona schist
Sierra Pelona schist
Sierra Pelona schist
Sierra Pelona schist
Sierra Pelona schist
Blue Ridge schist
Blue Ridge schist
East Fork schist
East Fork schist
East Fork schist
East Fork schist
SW & SE Castle Dome Mts. schist
SW & SE Castle Dome Mts. schist
Sierra de Salinas schist
Sierra de Salinas schist
Sierra de Salinas schist
Sierra de Salinas schist

Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Paleocene-Eocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Late Campanian-early Paleocene
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian

Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
Eoc
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
M-P
S-C
S-C
S-C
S-C

Underplated schist (SCH)


Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Barth et al., 2003
Barth et al., 2003
Barth et al., 2003
Jacobson et al., 2011

57

JM-80-102
RA58
RA89
RA169
RA170
RA175
PR150
99-57
04-SE1B
SE03
06SE23
07SE34
08SE467
Cat-lws-bs
Cat-alb-act
Cat-lws-alb
Cat-ep-bs

18
14
14
46
37
55
37
35
24
23
24
96
70
164
59
82
30

Sierra de Salinas schist


Rand schist
Rand schist
Rand schist
Rand schist
Rand schist
Portal Ridge schist
San Emigdio schist
San Emigdio schist
San Emigdio schist
San Emigdio schist
San Emigdio schist
San Emigdio schist
Catalina schist
Catalina schist
Catalina schist
Catalina schist

Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Santonian-Campanian
Cenomanian-Turonian
Cenomanian-Turonian
Cenomanian-Turonian
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)
Cenomanian-Turonian(?)

S-C
S-C
S-C
S-C
S-C
S-C
S-C
C-C
C-C
C-C
C-C
C-C
C-C
C-C
C-C
C-C
C-C

Underplated schist (SCH)


Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)
Underplated schist (SCH)

Jacobson et al., 2011


Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Jacobson et al., 2011
Chapman et al., 2013
Chapman et al., 2013
Chapman et al., 2013
Grove et al., 2008
Grove et al., 2008
Grove et al., 2008
Grove et al., 2008

Notes
* The Pilarcitos block (PB) is included with the La Honda basin (LHB) group in Figures 5-8 based on similarities with the Butano Sandstone (see discussion in Appendix A-3)
The San Emigdio Mountain (SEM) samples are included with the Salinian block forearc based on provenance affinities with other Eocene sandstones in this group
Depositional ages of Franciscan Complex samples based mainly on youngest detrital zircon ages
# The Franciscan Coastal Belt samples are included with the "Franciscan Complex" rather than with the "San Joaquin basin" group in Figure 6.
** Included with the Maastrichtian-Paleocene group, although may be as young as early Eocene in
age
Includes samples originally published in Jacobson et al. (2000) and Grove et al. (2003)
Samples grouped by metamorphic facies (individual samples not differentiated)

Abbreviations: C-C-Cenomanian-Coniacian; Cg-Conglomerate; Eoc-Eocene; M-P-Maastrichtian-Paleocene; Mbr-Member; Mtn-Mountain; Ms-Mudstone; S-C-SantonianCampanian; Ss-Sandstone; Sh-Shale

58

Magmatic gap

IIc. Jurassic to earliest Cretaceous

Magmatic gap

II. Cretaceous-Permian grains


IIa. Late Cretaceous
IIb. Mid- to Late-Cretaceous

I. Paleogene grains
Ia. Early to middle Eocene
Ib. Paleocene to early Eocene

300-700

225-285

180-225

135-180

125-135

65-85
80-125

43-52
53-65

Approx. age
range (Ma)

n.a.

n.a.

ca. 252

n.a.

ca. 148-162

n.a.

n.a.
ca. 97

ca. 50
n.a.

Peak age(s) (Ma)

Anorogenic granitic plutons (southwest Laurentia)

Grenville province (eastern-southern Laurentia)

Appalachian orogen, accreted Paleozoic terranes, and rift plutons

Idaho Batholith or "Laramide-aged" plutons in southwest U.S.


Idaho Batholith (<98 Ma) or mid-Cretaceous Cordilleran arc

Challis volcanic center


Idaho Batholith

Inferred source region(s)

Table 2. Detrital Zircon U-Pb Age Populations

IId. Late Permian-Triassic

912-1310

n.a.

Yavapai-Mazatzal province (southwest Laurentia)


Multiple northern Laurentian age provinces
Superior province of northeast Laurentia

Age population

IIIb. Grenville Mesoproterozoic

1311-1579

n.a.
n.a.
n.a.

Late Permian-Triassic Cordilleran arc

Jurassic Cordilleran arc or accreted Jurassic oceanic island arcs

IIIc. Pre-Grenville Mesoproterozoic

1581-1855
1855-2430
2470-3058

III. Pre-Permian grains*


IIIa. Paleozoic to Neoproterozoic grains

IIId. Late Paleoproterozoic


IIIe. Older Paleoproterozoic
IIIf. Archean

Notes:
* Pre-Permian grain age populations modified from Dickinson et al. (2012)

59

Figure 1. Tectonic setting of the western U.S. (modified from Dickinson, 1996,
DeCelles, 2004; Dickinson, 2008; Grove et al., 2008; Dickinson and Gehrels, 2008, and
Surpless and Beverly, 2013). BM-Blue Mountains; FB-Foothills belt; GC-Gulf of
California; HB-Hornbrook basin; IB-Idaho batholith; KM-Klamath Mountains; MDMojave Desert; NCB-Nacimiento block; PR-Peninsular Ranges; SAF-San Andreas fault;
TB-Tyee basin; TR-Transverse Ranges.
60

Figure 2. A: Generalized geologic framework of southern Oregon, California, and


northern Baja California. B: Palinspastic reconstruction of southern California northern
Baja California during Eocene time. See Appendix A-1 for an explanation of data sources
in Figure 2A and see text and Appendix A-2 for details regarding the palinspastic
reconstruction shown in Figure 2B.

61

62

Figure 3. Cumulative and normalized distributions of detrital zircon U-Pb ages for
groups of Upper Cretaceous-Eocene forearc strata. Cumulative distributions are colored
according to regional group (green-Oregon forearc; black-Great Valley forearc; redSalinian forearc; yellow-Transverse Ranges forearc; blue-Peninsular Ranges forearc).
Number of samples/grains shown in parentheses. The vertical scale of normalized
distributions greater than 300 Ma are displayed at 1/10th scale. See Table 1 for additional
information on sample groups and data sources. Ng-Neogene, Pg-Paleogene, KCretaceous.

63

64

65

Figure 4. Zircon U-Pb age distributions (< 300 Ma) for volcanic and plutonic rocks from
the Sierra Nevada batholith, Salinian block, Mojave Desert region, and Peninsular
Ranges batholith (Premo et al., 1998; Silver and Chappell, 1998; Grove et al., 2008;
Chapman et al., 2012; Barth et al., 2013; Kimbrough et al., in press) and CenomanianEocene forearc strata (see Table 1 for data sources).

66

Figure 5. Cumulative distributions of detrital zircon U-Pb ages (< 200 Ma) for regional
groups of Late Cretaceous-Eocene forearc strata. See Table 1 for additional information
on sample groups and data sources. Thick, red lines are zircon U-Pb age distributions
from plutonic and igneous rocks from the Sierra Nevada batholith, Salinian block,
Mojave Desert region, and Peninsular Ranges batholith (see data sources in Figure 4).
Abbreviations: Cen-Cenomanian; Cmp-Campanian; Con-Coniacian; Eoc-Eocene; MaMaastrichtian; MD-Mojave Desert; Pc-Paleocene; PRB-Peninsular Ranges batholith;
San-Santonian; SB-Salinian block; SNB-Sierra Nevada batholith.

67

68

Figure 6. Cumulative and normalized distributions of detrital zircon U-Pb ages for arcderived detrital zircon (i.e., < 300 Ma) from groups of forearc, subduction complex,
underplated schist, intra-arc, and retroarc strata. See Table 1 for samples and data
sources. Numbers in parentheses indicate the number of samples/grains younger than 300
Ma and the percentage of this subpopulation relative to the total population.
Abbreviations: Cum-Cumulative; FC-Franciscan Complex; MRA-McCoy Retroarc, OROregon, PR-Peninsular Ranges, Prob-Probability; RAF-Retroarc foreland, SB-Salinian
block, SC-Sacramento basin, SCH-underplated schist; SJ-San Joaquin basin, TRTransverse Ranges.

69

70

Figure 7. A: Spatial distribution of detrital zircon age populations in CenomanianEocene forearc sandstone. Distances are measured north-to-south parallel to the
reconstructed Eocene margin (Fig. 2B). Age population abundances are linearly
interpolated between sample localities (see Figure 2A for sample location abbreviations).
B: Inferred spatial and temporal distribution of source regions that supplied detritus to the
California forearc (dashed where source contribution is minor or uncertain).
Abbreviations: Camp-Campanian; Cen-Cenomanian; Con-Coniacian; HB-Hornbrook
basin; Ma-Maastrichtian; Pc-Paleocene; PR-Peninsular Ranges; Sant-Santonian; SBSalinian block; SCB-Sacramento basin; SJB-San Joaquin basin; TB-Tyee basin; TRTransverse Ranges.

71

72

Figure 8. Generalized paleogeographic reconstruction of the Cenomanian-Eocene


continental margin. The age of potential source rocks are generalized from Figure 2.
Detrital zircon U-Pb age populations are shown as pie diagrams (see Figure 7 for key).
Hypothetical fluvial-submarine sediment dispersal pathways are shown as blue lines. The
continental drainage divide is shown as a dashed black line. The approximate location of
an inferred oceanic plateau is from Liu et al. (2010). The relative plate motion between
the Farallon and Pacific plates is shown as a black arrow with convergence velocity
indicated (Cenomanian-Coniacian: Liu et al., 2008; Santonian-Eocene: Doubrovine and
Tarduno, 2008). Aspects of this reconstruction are modified from the following data
sources: Dickinson et al., 1979; Ingersoll, 1979; Nilsen and Abbott, 1981; Kies and
Abbott, 1982; Sundberg and Cooper, 1982; DeGraaff-Surpless et al., 2002; Dickinson et
al., 2012; Sharman et al., 2013; Surpless and Beverly, 2013. See Figure 2 for abbreviated
forearc sample locations. Abbreviations: NCB-Nacimiento block; DR-Dome Rock; MMMcCoy Mountains retroarc; RF-Retroarc foreland; SB-Salinian block; WTR-Western
Transverse Ranges.

73

74

75

CHAPTER 2: A REAPPRAISAL OF THE EARLY SLIP HISTORY OF THE


CENTRAL SAN ANDREAS FAULT, CENTRAL CALIFORNIA, USA

76

A REAPPRAISAL OF THE EARLY SLIP HISTORY OF THE CENTRAL SAN


ANDREAS FAULT, CENTRAL CALIFORNIA, USA
Glenn R. Sharman1, Stephan A. Graham1, Marty Grove1, and Jeremy K. Hourigan2
1. Geological and Environmental Sciences, Stanford University, Stanford, California
94305
2. Earth and Planetary Sciences, University of California Santa Cruz, Santa Cruz,
California 95064

ABSTRACT
The modern San Andreas fault system (California, United States) is widely
considered to have formed in response to the initiation of PacificNorth American
plate interaction ca. 27 Ma. Although there is general consensus on the magnitude and
timing of Neogene displacement along the San Andreas system, its Paleogene history
remains unresolved. In particular, 100 km of right-lateral offset between midCretaceous plutonic rocks of the northern Salinian block and the western edge of
Sierra Nevada basement remains unaccounted for after restoration of Neogene
displacement along strike-slip faults of the San Andreas system. Our detrital zircon
data invalidate a key Paleogene piercing point by demonstrating that displaced
portions of the hypothesized Middle Eocene ButanoPoint of Rocks submarine fan
were never contiguous across the San Andreas fault. We instead show that the Eocene
provenance characteristics exhibited by northern Salinian strata closely match those of
the southern Sierra Nevada and northwestern Mojave Desert. This implies that the
northern Salinian block was located at least 7550 km farther south in Eocene time
than previously recognized. Our data require (1) pre23 Ma dextral slip along the San
Andreas fault in central California, and/or (2) slip along a predecessor fault that
formed prior to PacificNorth American plate interaction. This previously
undocumented slip may indicate that significant PacificNorth American plate
interaction propagated from the plate margin into the continental interior much earlier

77

than conventionally believed. Alternatively, late Paleogene slip could predate the
development of the modern plate boundary and represent inboard dextral strike-slip
displacement along the eastern margin of the Salinian block, similar to the
deformation that occurs today along the strike-slip Sumatra fault system.
INTRODUCTION
The latest Cretaceous and Cenozoic tectonic evolution of the central California
(United States) margin has been profoundly shaped by (1) the structural emplacement
of the Salinian block in latest Cretaceous to Paleocene time (Suppe, 1970; Page, 1981;
Hall, 1991; Saleeby, 2003; Chapman et al., 2012) and (2) the development of the
modern San Andreas fault (SAF) system in Neogene time (Atwater, 1989). The
Salinian block is a displaced fragment of the mid-Cretaceous continental margin
batholith that is out of place relative to the classic arc-forearc-subduction complex
sequence that characterizes most of the Mesozoic Cordilleran margin (Graham, 1978;
Page, 1981). The incongruous juxtaposition of the Salinian block with Great Valley
forearc strata and the underlying Franciscan subduction complex led early workers to
recognize and quantify the significant right-lateral slip that has occurred along the
SAF system (e.g., Hill and Dibblee, 1953). Recognition that such large lateral
displacements of continental crust occurred permitted reconciliation of oceanic and
continental geologic relationships in the burgeoning theory of plate tectonics (Atwater,
1970).
Although there is general consensus regarding the amount (315 km) of
Neogene displacement along the SAF in central California (Atwater, 1989; Graham et
al., 1989), an additional minimum displacement of at least 100 km is required to
restore the northern Salinian block to an intra-arc position (Dickinson et al., 2005). To
account for this discrepancy, past researchers have postulated that a dextral protoSan
Andreas fault coincident with the modern SAF was active between latest Cretaceous
and Paleocene time (Suppe, 1970; Graham, 1978; Page, 1981; Dickinson et al., 2005).
Others have proposed that westward-directed thrusting and/or extensional collapse

78

occurred during a similar time frame (Hall, 1991; Saleeby, 2003; Chapman et al.,
2012).
PIERCING POINTS
The magnitude and timing of offset along strike-slip faults are determined by
correlated paleogeologic features, or piercing points, that originally extended across
the trace of the fault. A number of Neogene piercing points document progressively
greater displacement with time along the SAF (e.g., Graham et al., 1989). Perhaps the
most widely accepted of these is the correlation of the ca. 23 Ma Pinnacles and
Neenach volcanic centers on either side of the SAF in central California that is based
on similarities in age, stratigraphy, petrography, and geochemistry (Matthews, 1976;
Dickinson, 1997; Fig. 1).
The most accepted Paleogene tie across the central SAF is based upon
correlation of the Early to Middle Eocene Butano Sandstone with the Middle Eocene
Point of Rocks Sandstone Member of the Kreyenhagen Formation (hereafter Point of
Rocks Sandstone) (Clarke, 1973; Fig. 1). The Butano Sandstone and Point of Rocks
Sandstone were interpreted to be parts of the same northeast- to northwest-flowing
submarine fan system with a common provenance derived from nearby exposures of
Salinian block plutonic rocks based on petrographic similarity, thickness trends, facies
relationships, and paleocurrent indicators (Clarke, 1973; Critelli and Nilsen, 1996; Fig.
1). Because the ca. 23 Ma Pinnacles and Neenach volcanic centers (Matthews, 1976;
Dickinson, 1997) and the ca. 4838 Ma ButanoPoint of Rocks submarine fan
(Clarke, 1973) have approximately the same offset (315 km), previous workers
concluded that no slip occurred on the SAF in central California during late Paleogene
time (e.g., Graham et al., 1989) (ca. 3823 Ma; Fig. 1).
DETRITAL ZIRCON RESULTS
We use detrital zircon U-Pb geochronology to (1) test the correlation of the
Butano and Point of Rocks Sandstones, and (2) establish regional Eocene provenance
relationships in central California. Cumulative and normalized U-Pb age distributions

79

are presented in Figure 2 for 1184 grains measured in 12 samples (see Appendices A-1
and A5-8). Variable proportions of mid-Late Cretaceous zircon (ca. 12580 Ma) and
earliest Cretaceous to Jurassic zircon (ca. 175140 Ma) are present in all samples (Fig.
2). PermianTriassic zircons (ca. 280230 Ma) occur in some samples. Minor (1%
9%) Eocene grains (5543 Ma) are present in the Point of Rocks Sandstone. All
samples have subordinate (4%13%) pre300 Ma zircon (Fig. 2).
SEDIMENTARY PROVENANCE
The abundance of Mesozoic zircon clearly indicates that the Paleogene forearc
strata were derived from the Mesozoic Cordilleran arc (Jacobson et al., 2011; Lechler
and Niemi, 2011; Fig. 3). Two distinct batholithic source regions are apparent. The
first is represented by the Point of Rocks Sandstone and has 67%80% late Early to
Late Cretaceous zircon with minor (10%13%) Jurassic grains (Fig. 2). This
provenance signature is consistent with the distribution of zircon U-Pb crystallization
ages from the Cretaceous batholith and local Jurassic intrusive rocks of the Sierra
Nevada (Irwin and Wooden, 2001; Fig. 3). The interpretation that the Point of Rocks
sands were derived from this region is supported by comparison with Eocene fluvial
strata in the northern and central Sierra Nevada that have very similar detrital zircon
age distributions (Cecil et al., 2010; Cassel et al., 2012; Fig. 2). An additional
indication of a northern source is the presence of Eocene zircons that appear to have
been supplied from the Idaho batholith region (e.g., the 5143 Ma Challis volcanic
field; Dumitru et al., 2012).
The second provenance signature typifies Eocene forearc strata from the
northern Salinian block (e.g., Butano Sandstone) and the San Emigdio Mountains
(e.g., Tejon Formation; Fig. 1). The subequal proportions of Jurassic and Cretaceous
zircon, small but distinctive population of PermianTriassic zircon (5%17%), and
lack of Eocene grains in these sandstones define an age distribution that is
quantitatively distinguished from the Point of Rocks Sandstone and other Eocene
Great Valley sandstones (see Table 2 and Appendix B-3). One sample (SEM-1)
collected from the basal Tejon Formation by Lechler and Niemi (2011) yields a

80

provenance signature different from our nearby, equivalent sample (TEJ-1),


suggesting local provenance variability in the basal Tejon Formation (Fig. 2).
Because Jurassic plutons are very uncommon in the northern Salinian block
(Chapman et al., 2012), the abundance of Jurassic zircon in the Butano Sandstone
(38%62%) rules out a local Salinian source (Critelli and Nilsen, 1996; Fig. 3).
Although the provenance of the Butano Sandstone is also dissimilar to zircon U-Pb
crystallization ages in the main Sierra Nevada batholith, it strongly resembles those of
the southeastern Sierra Nevada and northwestern Mojave Desert region (Fig. 3).
JurassicPermian plutonic and volcanic rocks are present in the southeastern Sierra
Nevada and northwestern Mojave (Walker et al., 2002; Chapman et al., 2012; Fig. 3).
Significant populations of Jurassic and Triassic detrital zircon grains characterize the
Paleogene Witnet and Goler Formations from the southernmost Sierra Nevada
(Lechler and Niemi, 2011; Fig. 2). Based upon these relationships, we conclude that
the Eocene forearc strata of the northern Salinian block were most likely derived from
the southeastern Sierra Nevada and northwestern Mojave region (Fig. 3). The
significant contributions from the inboard JurassicPermian arcs indicate that the
topographic continuity of the mid-Cretaceous arc was locally breached during Eocene
time at the latitude of the southern Sierra Nevada by an extraregional river system that
delivered abundant Jurassic and older arc detritus from the continental arc interior to
the Paleogene forearc deposited atop the northern Salinian block (Fig. 3).
DISCUSSION
Although the ButanoPoint of Rocks correlation has been widely accepted and
used to constrain paleogeographic and tectonic reconstructions (Clarke, 1973; Graham
et al., 1989; Dickinson et al., 2005), our results clearly demonstrate that the Butano
and Point of Rocks Sandstones do not share a common provenance and could not have
been part of a contiguous submarine fan that extended across the future SAF.
Rejection of the ButanoPoint of Rocks piercing point removes a critical constraint on
the pre23 Ma displacement history of the SAF system (Fig. 4).

81

Although gabbroic rocks (Ross, 1984; Chapman et al. 2012) and overlying
Early to Middle Eocene strata exposed in the San Emigdio Mountains (Tejon
Formation) and near the northern Gabilan Range (San Juan Bautista Formation;
Nilsen, 1984) have also been suggested as piercing points (Fig. 1), several lines of
evidence, including poor outcrop quality and limited exposures, suggest that these
correlation are also suspect (Appendix B-4). There is thus no compelling evidence that
requires inactivity of the SAF between 38 and 23 Ma in central California (Fig. 4).
Linking the Butano Sandstone with a southeastern Sierra Nevadanorthwestern
Mojave source region requires that the northern Salinian block was located farther
south in Middle Eocene time than currently recognized in tectonic reconstructions
(Fig. 3). Our preferred reconstruction juxtaposes the Butano Sandstone against strata
with similar detrital zircon provenance signatures in the San Emigdio Mountains and
provides access to PermianJurassic arc rocks in the southeastern Sierra Nevada and
Mojave region (Fig. 3). Although 5075 km of pre23 Ma dextral displacement is
sufficient to honor provenance ties, a total of 100 km would more satisfactorily restore
northern Salinian basement with the inferred western edge of Sierra Nevada basement
beneath the Great Valley (Dickinson et al., 2005; Fig. 1). Our restoration thus implies
50100 km of ca. 3823 Ma dextral slip on the SAF in central California (Fig. 4).
TECTONIC IMPLICATIONS
Because current tectonic models assume zero late Paleogene displacement on
the SAF in central California, our results require a careful reappraisal of the early
history of the SAF system. We propose that the 50100 km of pre-Miocene
displacement resulted from (1) an earlier than recognized response to PacificNorth
American plate interaction, and/or (2) slip on an older dextral strike-slip fault that
predated the development of the Neogene SAF system (Fig. 4).
The first scenario requires displacement to occur prior to the eruption of the ca.
23 Ma Pinnacles and Neenach volcanic centers (Matthews, 1976; Dickinson, 1997),
but after the contact of the Pacific and North American plates following the
subduction of the intervening Farallon plate ca. 27 Ma (Atwater, 1989; Fig. 4). This

82

scenario represents a significant revision of how the PacificNorth American plate


boundary evolved to establish the SAF system by (1) invoking an episode of early
(i.e., pre23 Ma) inboard dextral displacement, and (2) differing markedly from
previous interpretations that indicated a progressive inland migration of the plate
boundary from the continental margin over time (i.e., after ca. 18 Ma; Atwater, 1989).
In the second scenario, late Paleogene displacement along the trend of the
modern SAF was unrelated to PacificNorth American plate contact and instead
resulted from strain partitioning during subduction of the Farallon plate beneath North
America (Fig. 4). In this case, dextral slip would have been accommodated inboard of
the trench along the eastern margin of the Salinian block, analogous to the modern
strike-slip Sumatra fault system, which accompanies oblique subduction of the Indian
plate beneath the Sunda plate (Fitch, 1972). This explanation is consistent with revised
plate reconstructions (Doubrovine and Tarduno, 2008) that indicate significant dextral
obliquity of subduction along the central California margin during Paleogene time
(Fig. 4). Regardless of the mechanism by which pre23 Ma slip occurred, it is clear
that important aspects of the Late CretaceousPaleogene history of the central
California margin remain poorly understood and likely involved more protracted
tectonism than was previously recognized in models proposed for the initiation of
PacificNorth American plate interaction and the subsequent development of the SAF
system.
ACKNOWLEDGEMENTS
Funding for this study was provided by the Stanford Project on Deep-water
Depositional Systems and the Stanford School of Earth Sciences Chevron Fund. The
Arizona LaserChron Center is funded in part by the National Science Foundation
(NSF) Instrumentation and Facilities Division. Instrumentation used at the University
of California at Santa Cruz was funded by the NSF Major Research Instrumentation
Program. 2013 The Geological Society of America. We thank the Twisselman
family and the Wildlands Conservancy for allowing access to privately owned land.
This work benefited from discussion with A. Almgren, E. Brabb, B. Dickinson, A.

83

Doebbert, T. Dumitru, C. Jacobson, R. McLaughlin, E. Miller, and J. Saleeby. Many


individuals helped with preparation and analysis of samples, including B. Burgreen, A.
Bernhardt, L. Cassel, W. Chang, M. Coble, T. Dumitru, J. Fosdick, E. Gottlieb, K.
Maier, M. Malkowski, N. Sarto, and T. Schwartz. We thank the staff of the Arizona
LaserChron Center. This manuscript benefited from comments by A. Lechler and two
other anonymous reviewers.
REFERENCES
Atwater T., 1970, Implications of plate tectonics for the Cenozoic tectonic evolution
of western North America: Geological Society of America Bulletin, v. 81, p.
35133536, doi:10.1130/0016-7606(1970)81[3513:IOPTFT]2.0.CO;2.
Atwater T.M., 1989, Plate tectonic history of the northeast Pacific and western North
America, in Winterer E.L., et al., eds., The eastern Pacific Ocean and Hawaii:
Boulder, Colorado, Geological Society of America, Geology of North
America, v. N, p. 2172.
Cassel E.J., Grove M., Graham S.A., 2012, Eocene drainage evolution and erosion of
the Sierra Nevada batholith across northern California and Nevada: American
Journal of Science, v. 312, p. 117144, doi:10.2475/02.2012.03.
Cecil M.R., Ducea M.N., Reiners P., Gehrels G., Mulch A., Allen C., Campbell I.,
2010, Provenance of Eocene river sediments from the central northern Sierra
Nevada and implications for paleogeography: Tectonics, v. 29, TC6010,
doi:10.1029/2010TC002717.
Chapman A.D., Saleeby J.B., Wood D.J., Piasecki A., Kidder S., Ducea M.N., Farley
K.A., 2012, Late Cretaceous gravitational collapse of the southern Sierra
Nevada batholith, California: Geosphere, v. 8, p. 314341,
doi:10.1130/GES00740.1.
Clarke S.H. Jr., 1973, The Eocene Point of Rocks Sandstone: Provenance, mode of
deposition and implications for the history of offset along the San Andreas
fault in central California [Ph.D. thesis]: Berkeley, University of California,
302 p.
Critelli S., Nilsen T.H., 1996, Petrology and diagenesis of the Eocene Butano
Sandstone, La Honda basin, California: Journal of Geology, v. 104, p. 295
315, doi:10.1086/629826.
Dickinson W.R., 1997, Tectonic implications of Cenozoic volcanism in coastal
California: Geological Society of America Bulletin, v. 109, p. 936954,
doi:10.1130/0016-7606(1997)109<0936:OTIOCV>2.3.CO;2.
Dickinson W.R., Ducea M., Rosenberg L.I., Greene H.G., Graham S.A., Clark J.C.,
Weber G.E., Kidder S., Ernst W.G., Brabb E.E., 2005, Net dextral slip,
Neogene San GregorioHosgri fault zone, Coastal California: Geologic
evidence and tectonic implications: Geological Society of America Special
Paper 391, 43 p., doi:10.1130/0-8137-2391-4.1.
84

Doubrovine P.V., Tarduno J.A., 2008, A revised kinematic model for the relative
motion between Pacific oceanic plates and North America since the Late
Cretaceous: Journal of Geophysical Research, v. 113, B12101,
doi:10.1029/2008JB005585.
Dumitru T.A., Ernst W.G., Wright J.E., Wooden J.L., Wells R.E., Farmer L.P., Kent
A.J.R., Graham S.A., 2012, Eocene extension in Idaho generated massive
sediment floods into Franciscan trench and into Tyee, Great Valley, and Green
River basins: Geology, doi:10.1130/G33746.1.
Fitch T.J., 1972, Plate convergence, transcurrent faults, and internal deformation
adjacent to Southeast Asia and the western Pacific: Journal of Geophysical
Research, v. 77, p. 44324460, doi:10.1029/JB077i023p04432.
Graham S.A., 1978, Role of Salinian block in evolution of San Andreas fault system,
California: American Association of Petroleum Geologists Bulletin, v. 62, p.
22142231.
Graham S.A., Stanley R.G., Bent J.V., Carter J.B., 1989, Oligocene and Miocene
paleogeography of central California and displacement along the San Andreas
fault: Geological Society of America Bulletin, v. 101, p. 711730,
doi:10.1130/0016-7606(1989)101<0711:OAMPOC>2.3.CO;2.
Hall C.A. Jr., 1991, Geology of the Point SurLopez Point region, Coast Ranges,
California: A part of the Southern California allochthon: Geological Society of
America Special Paper 266, 40 p.
Hill M.L., Dibblee T.W. Jr., 1953, San Andreas, Garlock, and Big Pine faults,
California: Geological Society of America Bulletin, v. 64, p. 443458,
doi:10.1130/0016-7606(1953)64[443:SAGABP]2.0.CO;2.
Irwin W.P., Wooden J.L., 2001, Map showing plutons and accreted terranes of the
Sierra Nevada, California with a tabulation of U/Pb isotopic ages: U.S.
Geological Survey Open-File Report 2001229, scale 1:100,000.
Jacobson C.E., Grove M., Pedrick J.N., Barth A.P., Marsaglia K.M., Gehrels G.E.,
Nourse J.A., 2011, Late Cretaceousearly Cenozoic tectonic evolution of the
southern California margin inferred from provenance of trench and forearc
sediments: Geological Society of America Bulletin, v. 123, p. 485506,
doi:10.1130/B30238.1.
Lechler A.R., Niemi N.A., 2011, Sedimentologic and isotopic constraints on the
Paleogene paleogeographic and paleotopography of the southern Sierra
Nevada, California: Geology, v. 39, p. 379382, doi:10.1130/G31535.1.
Matthews V. III., 1976, Correlation of Pinnacles and Neenach volcanic fields and their
bearing on San Andreas fault problem: American Association of Petroleum
Geologists Bulletin, v. 60, p. 21282141.
Nilsen T.H., 1984, Offset along the San Andreas fault of Eocene strata from the San
Juan Bautista area and western San Emigdio Mountains, California: Geological
Society of America Bulletin, v. 95, p. 599609, doi:10.1130/00167606(1984)95<599:OATSAF>2.0.CO;2.
Nilsen T.H., Simoni T.R., 1973, Deep-sea fan paleocurrent patterns of the Eocene
Butano Sandstone, Santa Cruz Mountains, California: U.S. Geological Survey
Journal of Research, v. 1, p. 439452.
85

Page B.M., 1981, The southern Coast Ranges, in Ernst W.G., ed., The geotectonic
development of California: Rubey Volume I: Englewood Cliffs, New Jersey,
Prentice-Hall, p. 329417.
Ross D.C., 1984, Possible correlations of basement rocks across the San Andreas, San
GregorioHosgri, and RinconadaRelizKing City faults, California: U.S.
Geological Survey Professional Paper 1317, 37 p.
Saleeby J., 2003, Segmentation of the Laramide slabEvidence from the southern
Sierra Nevada region: Geological Society of America Bulletin, v. 115, p. 655
668, doi:10.1130/0016-7606(2003)115<0655:SOTLSF>2.0.CO;2.
Suppe J., 1970, Offset of late Mesozoic basement terrains by the San Andreas fault
system: Geological Society of America Bulletin, v. 81, p. 32533258,
doi:10.1130/0016-7606(1970)81[3253:OOLMBT]2.0.CO;2.
Walker J.D., Martin M.W., Glazner A.F., 2002, Late Paleozoic to Mesozoic
development of the Mojave Desert and environs, California, in Glazner A.F., et
al., eds., Geologic evolution of the Mojave Desert and southwestern Basin and
Range: Geological Society of America Memoir 195, p. 118, doi:10.1130/08137-1195-9.1.

86

Table 1. Sample locations


Sample
POR-3
POR-2
POR-1
SJB-1
BUT-5
BUT-4
BUT-3
BUT-2
BUT-1
SEF-1
TEJ-2
TEJ-1

Stratal Unit
Point of Rocks Sandstone
Point of Rocks Sandstone
Point of Rocks Sandstone
San Juan Bautista Formation
Butano Sandstone
Butano Sandstone
Butano Sandstone
Butano Sandstone
Butano Sandstone
San Emigdio Formation
Tejon Formation - Metralla Sandstone Member
Tejon Formation - Uvas Conglomerate Member

87

Coordinates (WGS84)
Latitude (N)
Longitude (W)
35.38175
-119.77914
35.43772
-119.84809
35.72727
-120.01812
36.86272
-121.60069
37.37850
-122.26370
37.19889
-122.19639
37.23231
-122.23231
37.15930
-122.23500
37.14267
-122.19014
34.91706
-119.16808
34.90597
-119.15653
34.89947
-119.15006

N
101
97
100
99
97
100
96
96
99
100
99
100

N
101
97
100
99
97
100
96
96
99
100
99
100

POR-1
0.035
0.076
-0.316
0.303
0.368
0.334
0.322
0.515
0.316
0.305
0.454

POR-3 POR-2 POR-1


-0.803 1.000
0.803
-0.936
1.000
0.936
-0.000
0.000 0.000
0.000
0.001 0.000
0.000
0.000 0.000
0.000
0.001 0.000
0.000
0.000 0.000
0.000
0.000 0.000
0.000
0.000 0.000
0.000
0.002 0.000
0.000
0.000 0.000

POR-3 POR-2
-0.091
0.091
-0.035
0.076
0.324
0.293
0.301
0.280
0.377
0.345
0.369
0.287
0.343
0.299
0.514
0.492
0.340
0.292
0.340
0.262
0.456
0.431

SEF-1
0.000
0.000
0.000
-0.999
0.518
0.795
0.511
0.040
0.991
0.612
0.050

SEF-1
0.324
0.293
0.316
-0.054
0.116
0.093
0.118
0.199
0.062
0.108
0.192

TEJ-2
0.000
0.001
0.000
0.999
-0.214
0.862
0.561
0.023
0.989
0.680
0.066

TEJ-2
0.301
0.280
0.303
0.054
-0.151
0.087
0.114
0.213
0.063
0.103
0.186

D-values
TEJ-1 SJB-1
0.377 0.369
0.345 0.287
0.368 0.334
0.116 0.093
0.151 0.087
-0.089
0.089
-0.080 0.031
0.147 0.204
0.150 0.060
0.128 0.047
0.100 0.144

P-values
TEJ-1 SJB-1
0.000 0.000
0.000 0.001
0.000 0.000
0.518 0.795
0.214 0.862
-0.829
0.829
-0.914 1.000
0.233 0.034
0.212 0.994
0.393 1.000
0.695 0.265

BUT-5
0.343
0.299
0.322
0.118
0.114
0.080
0.031
-0.193
0.076
0.058
0.132

BUT-5
0.000
0.000
0.000
0.511
0.561
0.914
1.000
-0.053
0.942
0.997
0.360

BUT-4
0.514
0.492
0.515
0.199
0.213
0.147
0.204
0.193
-0.200
0.230
0.100

BUT-4
0.000
0.000
0.000
0.040
0.023
0.233
0.034
0.053
-0.038
0.011
0.697

BUT-3
0.340
0.292
0.316
0.062
0.063
0.150
0.060
0.076
0.200
-0.046
0.158

BUT-3
0.000
0.000
0.000
0.991
0.989
0.212
0.994
0.942
0.038
-1.000
0.164

Table 2. Kolgomorov-Smirnov (K-S) P-values and D-values of detrital zircon U-Pb age distributions
Sample
Point of Rocks Sandstone (POR-3)
Point of Rocks Sandstone (POR-2)
Point of Rocks Sandstone (POR-1)
San Emigdio Formation (SEF-1)
Tejon Formation - Metralla ss. mbr. (TEJ-2)
Tejon Formation - Uvas cngl. mbr. (TEJ-1)
San Juan Bautista Formation (SJB-1)
Butano Sandstone (BUT-5)
Butano Sandstone (BUT-4)
Butano Sandstone (BUT-3)
Butano Sandstone (BUT-2)
Butano Sandstone (BUT-1)

Sample
Point of Rocks Sandstone (POR-3)
Point of Rocks Sandstone (POR-2)
Point of Rocks Sandstone (POR-1)
San Emigdio Formation (SEF-1)
Tejon Formation - Metralla ss. mbr. (TEJ-2)
Tejon Formation - Uvas cngl. mbr. (TEJ-1)
San Juan Bautista Formation (SJB-1)
Butano Sandstone (BUT-5)
Butano Sandstone (BUT-4)
Butano Sandstone (BUT-3)
Butano Sandstone (BUT-2)
Butano Sandstone (BUT-1)

BUT-2
0.000
0.002
0.000
0.612
0.680
0.393
1.000
0.997
0.011
1.000
-0.117

BUT-2
0.340
0.262
0.305
0.108
0.103
0.128
0.047
0.058
0.230
0.046
-0.169

BUT-1
0.000
0.000
0.000
0.050
0.066
0.695
0.265
0.360
0.697
0.164
0.117
--

BUT-1
0.456
0.431
0.454
0.192
0.186
0.100
0.144
0.132
0.100
0.158
0.169
--

88

Notes:
P-values use error in the cumulative distribution function
Samples that are statistically distinct at a 95% confidence level (P-value < 0.05) are shaded
Abbreviations: ss: sandstone; cngl: conglomerate; mbr: member; N: number of grain analyses
See also Appendix B-3

89

Figure 1. Paleogeographic reconstruction of central California during Middle Eocene


time. Removal of 315 km of slip on San Andreas fault (SAF) juxtaposes the northern
Salinian block against the southern Great Valley forearc (Clarke, 1973; Graham et al.,
1989; Dickinson et al., 2005). Outcrop extent and inferred depositional limit (shaded) of
Eocene forearc strata are shown in yellow (Clarke, 1973; Dickinson et al., 2005).
Paleocurrent directions (Clarke, 1973; Nilsen and Simoni, 1973) are shown as rose
diagrams (scale bar is 10 measurements). Mean paleoflow vector is shown as black
arrow. Detrital zircon sample locations are labeled (SEF, TEJ, SJB). BUTButano
Sandstone, GRGabilan Range; NCNeenach volcanic center; PCPinnacles volcanic
center; PORPoint of Rocks Sandstone; SCMSanta Cruz Mountains; SEMSan
Emigdio Mountains; SSNBsouthern Sierra Nevada batholith.

90

Figure 2. Cumulative and normalized distributions of detrital zircon U-Pb ages. A:


Combined cumulative density functions (CDFs) for Eocene forearc strata in Great Valley
forearc, San Emigdio Mountains, and the northern Salinian block (California). Shaded
regions encompass combined CDFs of each individual sample with these groups. These
distributions are compared with Paleocene strata of southernmost Sierra Nevada (Goler
and Witnet Formations; Lechler and Niemi, 2011), Eocene Tejon Formation (SEM-1;
Lechler and Niemi, 2011), and Eocene fluvial sands of northern and central Sierra
Nevada (Cecil et al., 2010; Cassel et al., 2012). B: Probability density functions with
histograms for each sample. Colors correspond to age of potential source rocks (Fig. 3).
Histogram tick marks are in increments of 1 (thin) and 5 (thick) age measurements. Note
that horizontal scale changes at 300 Ma. NgNeogene; PgPaleogene; KCretaceous;
FmFormation; MbrMember.

91

92

Figure 3. Proposed tectonic and paleogeographic reconstruction of northern Salinian


block and southern Sierra Nevada during Paleogene time. Tectonic reconstructions
include restoration of 390 km slip on San Andreas fault (SAF) and 50 km slip on
Garlock fault (Graham et al., 1989; Dickinson et al., 2005; Chapman et al., 2012; this
study). Northern Salinian block is arbitrarily rotated 15 counterclockwise to minimize
amount of space (gray shading) created by the bend in SAF. Generalized age and
lithologic zones of Mesozoic arc rocks are modified from Chapman et al. (2012).
Additional 75 km of right-lateral slip juxtaposes Butano Sandstone against southern
Sierra Nevada and provides access to sediment dispersal system with source in
southeastern Sierra Nevada and northwestern Mojave region (Lechler and Niemi, 2011).
BABakersfield arch; BUTButano Sandstone; GFGarlock fault; PORPoint of
Rocks Sandstone; SSNBsouthern Sierra Nevada batholith.

93

94

Figure 4. A reappraisal of the tectonic history of the San Andreas fault (SAF) in central
California. A: Key tectonic and depositional events; dashed where uncertain or
approximate (see Appendix B-7). Sssandstone. B: Margin-parallel (dextral) component
of plate motion relative to fixed North America assuming margin azimuth of 330 and
location (33N, 117W) near San Diego, California (Doubrovine and Tarduno, 2008). C:
Displacement-time plot of SAF; key cross-fault ties (boxes labeled with lowercase letters;
see Appendix B-8) are used to constrain displacement scenarios (lines) within regions of
possible displacement-time space (gray shading). Two-stage slip history (blue line;
Suppe, 1970) includes both protoSan Andreas and modern San Andreas fault separated
by period of no slip ca. 3823 Ma. Two alternative scenarios (red dashed lines; this
study) represent possible end-member slip histories that restore the northern Salinian
block with southern Sierra Nevada (piercing point j) and require 50100 km of late
Paleogene displacement.

95

CHAPTER 3: SPATIAL PATTERNS OF DEFORMATION AND PALEOSLOPE


ESTIMATION WITHIN THE MARGINAL AND CENTRAL PORTIONS OF A
BASIN-FLOOR MASS-TRANSPORT DEPOSIT, TARANAKI BASIN, NEW
ZEALAND

96

SPATIAL PATTERNS OF DEFORMATION AND PALEOSLOPE


ESTIMATION WITHIN THE MARGINAL AND CENTRAL PORTIONS OF A
BASIN-FLOOR MASS-TRANSPORT DEPOSIT, TARANAKI BASIN, NEW
ZEALAND
Glenn R. Sharman1, Stephan A. Graham1, Larisa Masalimova2, Lauren Shumaker1,
and Peter King3
1

Department of Geological and Environmental Sciences, Stanford University 94305,

USA
2

Shell Exploration & Production Company, Houston, TX, 77079, USA

GNS Science, Lower Hutt, New Zealand

ABSTRACT
Submarine mass-movement within the upper Miocene Mohakatino Formation
(Taranaki Basin, New Zealand) provides an exceptional opportunity to study patterns
of soft-sediment deformation within a well-exposed mass-transport deposit (MTD).
The North Awakino MTD is at least 55 meters thick and outcrops along the northern
Taranaki coastline for ~11 km in wave-cut platforms and in cliffs up to 100 m high.
Spectacular deformation features are developed in remobilized sediment gravity flow
deposits that initially accumulated within a low-gradient, intra-slope basin.
Sedimentary facies within the North Awakino MTD range from laterally extensive,
thick- to thin-bedded volcaniclastic sandstone and mudstone and correspond with
distinct style of deformation: folds developed in thick-bedded sandstone are larger
(fold heights up to 10s of m) and more laterally continuous (up to 1 km) than those
developed in thinner-bedded facies.
Regional geologic relationships suggest that exposures of the North Awakino
MTD extend nearly across the full width of the MTD and thus provide a rare
opportunity to observe lateral relationships between the marginal and central portions
of the MTD. We conduct a rigorous paleoslope analysis of slump fold, fault, and
97

bedding orientations using both existing and newly proposed methodologies. Separate
analysis of seven subregions within the North Awakino MTD reveals that the
predicted MTD transport direction varies widely along the outcrop extent. Most
notably, slump deformation structures within the inferred margins have mean
orientations that are sub-orthogonal to those within the central portions of the MTD.
This relationship is hypothesized to be a consequence of edge effects that may be
related to lateral compression along the margins of the MTD. Our analysis
demonstrates the importance of accounting for spatial heterogeneity in deformation
structure orientations when determining the paleoslope orientation through kinematic
analysis. Our inference of W-directed translation suggests that the North Awakino
MTD formed in response to a local reversal of the bathymetric slope that was likely a
result of tectonically-induced basin deformation.
INTRODUCTION
Submarine landslides, or mass-transport deposits (MTDs), have become welldocumented in bathymetric and seismic-reflection datasets and are recognized to be
important agents of sediment delivery and reorganization along the worlds
continental slopes (Nardin et al., 1979; Haflidason et al., 2004; Bull et al., 2009;
Posamentier and Martinsen, 2011; and references within). However, exposures of
MTDs of similar scale as observed on the seafloor or within the subsurface are
relatively uncommon in the ancient record (Woodcock, 1979a; MacDonald, 1993;
Lucente and Pini, 2008; Odonne et al., 2011). Most ancient outcrops only preserve a
fraction of the original MTD extent, and thus published outcrop studies lack detailed
information of the large-scale geometries that are commonly observed in remote
sensing datasets (e.g., Posamentier and Martinsen, 2011). Although many outcrop
studies of MTDs have documented longitudinal changes in deformation from headwall
to toe (e.g., Farrell, 1984; Gawthorpe and Clemmey, 1985; Farrell and Eaton, 1987),
few have documented the lateral relationships between the marginal and central
portions of MTDs (Debacker et al., 2009). In general, knowledge of how MTD
deformation structures (e.g., folds and faults) change in character or orientation

98

laterally is often hampered by a lack of regional context and/or subsequent tectonic


deformation.
This paper aims to document the internal architecture, kinematic history, and
spatial patterns of deformation within a seismic-scale MTD that is spectacularly
exposed for ~11 km along the northern Taranaki coastline of New Zealand (hereafter
referred to as the North Awakino MTD; King et al., 2011; Fig. 1). The North
Awakino MTD (NAMTD) consists of remobilized sediment gravity flow deposits that
were deposited within a deep-marine (bathyal) depocenter within the larger Taranaki
Basin during late Miocene time (Utley, 1987; King et al., 1993). Subsequent
deformation during submarine mass-movement has resulted in a wide range of ductile
and brittle deformation structures that include spectacular slump folds and faults
(sensu Martinsen, 1994). Several factors contribute to the importance and uniqueness
of the NAMTD: (1) preservation of a nearly complete lateral cross-section through the
compressional domain that preserves both marginal and central portions of the MTD;
(2) excellent exposure in cliffs (up to ~100 m high) and wave-cut platforms that
extend semi-continuously along the study area; (3) limited tectonic deformation; and
(4) regional context provided by 2D and 3D seismic-reflection surveys that are located
3 km offshore of the study area (e.g., King et al., 2011).
Preservation of the margin-to-center relationships within the NAMTD provides
critical insight into lateral geometries that are typically lacking in outcrop studies of
ancient MTDs. For example, we observe how deformation structure orientations vary
with regard to their location and consider implications for estimation of the paleoslope
orientation via analysis of slump folds and faults (i.e., paleoslope analysis;
Woodcock, 1979b). Correspondingly, a major goal of this study is to critically
examine proposed methods of paleoslope analysis as they have been applied to the
lateral and central portions of MTDs (e.g., Debacker et al., 2009). In doing so, we
highlight both the utility and drawbacks of particular paleoslope methods and
demonstrate the importance of incorporating spatial context into kinematic analysis of
slump deformation features.

99

An important secondary goal of this study is to consider the implications of


mass-movement on the paleogeographic setting and tectonic history of the
northeastern Taranaki Basin. We aim to resolve an ongoing debate about the transport
direction of the NAMTD that has been variously interpreted as E-, S-, or W-directed
(Utley 1987; Nagel, 2010; King et al., 2011) by conducting a comprehensive
paleoslope analysis using both existing and new methodologies (e.g., Woodcock,
1979b; Strachan and Alsop, 2006). We also incorporate sedimentological observations
from both within the NAMTD and from surrounding units to constrain the
depositional setting prior to subsequent remobilization. In doing so we document the
internal architecture associated with soft-sediment deformation of sand-rich,
deepwater lobes (sensu Prlat et al., 2009) and provide a type example of tectonicallyinduced basin deformation that resulted in mass-movement of a basin-floor
sedimentary succession.
GEOLOGIC BACKGROUND
The NAMTD is located within the northeastern Taranaki Basin (Fig. 1), a
Cretaceous to Holocene sedimentary succession that has undergone a complex and
multi-phase basin history. The Taranaki Basin initiated during Late Cretaceous time
when the New Zealand sub-continent rifted from Gondwanaland (eastern Australia)
and thereafter developed into a passive margin setting (King and Thrasher, 1992,
1996; King, 2000). The fledgling modern Australian-Pacific plate boundary
subsequently propagated through the New Zealand region from middle Eocene time
(King, 2000). Convergence in the north started during late Eocene time (Bache et al.,
2012) and was manifested in the Taranaki Basin as emplacement of the W-verging
Taranaki thrust fault that displaced basement more than 7 km vertically (Figs. 1 and 2;
King and Thrasher, 1996; Stagpoole and Nicol, 2008). The eastern Taranaki Basin
may have evolved into an under-filled foreland basin during this time as a response to
crustal loading to the east (King and Thrasher, 1992; Holt and Stern, 1994). Major
activity on the Taranaki fault in the vicinity of the study area is thought to have ended
by early Miocene time (Stagpoole and Nicol, 2008), and a sedimentary succession

100

subsequently developed within a bathyal depocenter defined by uplifted basement to


the east and southeast (Herangi-Patea-Tongaporutu submarine highs) and by a
submarine, andesitic volcanic arc to the west and north (Figs. 2 and 3; Nodder et al.,
1990b; King and Thrasher, 1992; King et al., 1993, 2007).
The NAMTD was emplaced within the basal portion (Mohakatino Formation)
of a ~2 km thick, upwards-shoaling sedimentary succession of late Miocene
(Tortonian) age that is well-exposed along the Taranaki coast for over 65 km south of
Tirua Point (Figs. 3 and 4). The Mohakatino Formation comprises volcaniclastic
gravel, sandstone, and mudstone that are thought to have been derived from the
submerged chain of stratovolcanoes (ca. 15-6.5 Ma) to the north and west (King and
Thrasher, 1996; Giba et al., 2013; Figs. 1 and 2). Foraminiferal assemblages suggest
that the Mohakatino Formation was deposited in lower bathyal water depths (>1,500
m), and past workers have suggested that this unit is comprised of both sediment
gravity flow deposits and suspension sedimentation of volcanic debris through the
water column within a basin-floor fan or fan-apron setting (Utley, 1987; Nodder
et al., 1990b; King et al., 1993, 2011). The Mohakatino Formation is overlain by the
Mount Messenger, Urenui, and Kiore Formations that record a pronounced
progradation of the continental slope to the north and west from ca. 10.5-9 Ma (King
et al., 1993; Browne et al., 2000; Browne and Slatt, 2002; Maier, 2012; Masalimova,
2013; Rotzien et al., 2014; Fig. 4). Subsequent tectonic deformation along the northern
Taranaki coastline is restricted to slight regional tilting and the development of
occasional normal faults with generally low (< 20 m) amounts of displacement
(Nodder, 1987; King and Thrasher, 1996; Childs et al., 2007, 2009; Giba et al., 2010).
METHODS
Our study of the NAMTD includes a variety of approaches that include both
outcrop- and subsurface-based investigations. The primary datasets used in this study
include (1) structural and sedimentological observations and measurements collected
in the field; (2) photo documentation of the cliffs and wave-cut platforms and 3D
model generation of the study area using photogrammetric techniques; and (3) open-

101

source, industry-acquired 2D and 3D seismic-reflection surveys and wells located


offshore the study area (Fig. 2).
Measured Stratigraphy
In total, we measured ~245 m of stratigraphy from 18 sections at a cm- to mscale from adjacent to and within the NAMTD (Figs. 5-7; detailed sections are
provided in Appendix C-1). Sections were strategically collected to document the
stratigraphic context of the MTD, correlate key marker beds along the outcrop extent,
and assist with interpreting the sedimentary facies and depositional setting of MTD
stratigraphy. Sections 1, 15, 17, and 18 document the sedimentary successions that
overlie, underlie, or are laterally adjacent to the NAMTD (Figs. 7 and 8). Sections 2-7
were measured from the same stratigraphic interval of thick- to medium-bedded
sandstone that is repeatedly deformed along the northern 3 km of the study area (Fig.
5). Sections 8-14 and 16 are representative of stratigraphy along the southern 8 km of
the MTD outcrop extent (Fig. 6). The stratigraphic relationship of these sections to
each other is uncertain as a result of intense MTD-related deformation that precludes
correlation between sections.
Structural Measurements and Kinematic Analysis
A structural dataset of 134 slump fold and 242 slump fault orientations was
collected for kinematic analysis of the MTD (Appendices C-2 and C-3). Although
some fold orientations were measured directly in the field, most were calculated
stereographically following the procedures outlined in Davis et al. (2012). When
possible, one or more bedding measurements were collected from each limb and hinge
of the fold. The fold axis orientation is taken to be either the - or -axis (Davis et al.,
2012), and the axial plane orientation is taken to be the plane that bisects the fold
limbs. Asymmetric folds are assumed to verge in the up-dip direction of the fold axial
plane, and the asymmetry sense (clockwise or counterclockwise) is measured looking
down-plunge of the fold axis. Folds that displayed significant non-cylindrical behavior
were excluded from the kinematic analysis. Because fold and fault orientations vary

102

spatially along the outcrop extent, we subdivide the NAMTD into 7 subregions (R1-7;
Fig. 3). Each subregion is considered independently to test for systematic spatial
variability in the estimated transport direction of the MTD.
The orientation, displacement sense (normal or reverse), and approximate
magnitude of displacement were collected for each slump fault. In some cases, slump
faults are cross-cut (tilted) by subsequent folding. In most other cases, the relative
timing between normal faults, reverse faults, and folds is unclear due to a lack of
convincing cross-cutting relationships. Thus, it is often ambiguous whether fault
surfaces need to be structurally restored to bedding horizontal in order for them to
attain their original orientation at the time of formation. About a third of the fault
measurements were collected from bedding surfaces that were dipping more than 10
degrees (Appendix C-3). These measurements correspondingly possess an added
degree of uncertainty regarding their attitude at time of formation. We take the
approach of computing both un-restored and restored fault orientations, and comparing
the resulting distributions. In general, restored fault orientations exhibit fewer outliers
in both orientation and dip magnitude (Appendix C-4). This finding is consistent with
the observation that some faults are locally cross-cut by subsequent folding, and
suggests that many faults formed prior to the latest phase of folding. For this reason,
we use restored fault orientations in our kinematic analysis (see discussion below).
Photographic Documentation and 3D Model Generation
A high-resolution photographic database was collected that includes complete
coverage of the coastal cliffs (taken from a boat ~300 to 500 m offshore) and wave-cut
platforms (taken vertically downward from an airplane). Individual photos of the
coastal cliffs were stitched together using Adobe Photoshop to form a nearly
seamless photomosaic of the entire study area (Fig. 9; Appendix C-5). Bedding
contacts, fault traces, and the upper surface of the MTD were traced on the cliff
photomosaics (Fig. 9).
A complete 3D model of the study area was created using the photogrammetry
software Agisoft PhotoScan Professional (Fig. 10; Appendix C-6). The model was

103

georeferenced using ground control points collected in the field using a differential
GPS unit and laser-rangefinder. In total, 2,711 individual point coordinates were
collected from the coastal cliffs and over 100 points were collected from the wave-cut
platforms. The median ground control point has 15 and 27 cm horizontal and vertical
precision, respectively. The aerial photos were draped on a textured 3D model (mesh)
of the coastline (Fig. 10), and georeferenced orthophotos and digital elevation models
were produced for the entire study area (Appendix C-6). The wave-cut platform
orthophotos provide a basemap onto which we traced bedding and the traces of slump
fold axial surfaces (Fig. 11). In total, we outlined ~48 km of bedding contacts on the
wave-cut platforms.
NORTH AWAKINO MTD
Overview
The NAMTD collectively refers to an exposed interval of extensive softsediment deformation related to submarine mass-movement that extends for ~11 km
along the coastline between Pitone and Paparahia Streams (Henderson and Ongley,
1923; Utley, 1987; King et al., 2011; Fig. 3). The majority of the NAMTD can be
classified as a submarine slump on the basis of bedding being largely preserved, even
within zones that have been intensely folded (Martinsen, 1994). However, portions of
the NAMTD are only slightly deformed, and thus this MTD also displays
characteristics of a translational slide (sensu Posamentier and Martinsen, 2011).
Exposures of the NAMTD are nearly continuous along the entire study area, but the
coastal cliffs are considerably higher in the northern 6 km of the study area (up to
~100 m; Fig. 9) than compared to the southern 5 km (typically < 10 m). A welldeveloped wave-cut platform extends along nearly the entire coastline and is typically
150-200 m wide (Figs. 3 and 11).
The upper surface of the MTD is well-exposed over most of its outcrop extent
and can be traced over long distances (Fig. 9). Deformed strata within the MTD are
often abruptly truncated at this contact (Fig. 9). Although this surface is quite flat,
some vertical relief (typically up to several meters) is present over broad distances

104

(Fig. 9). Such relief can be locally observed to have been healed by deposition of
overlying turbidite sand (e.g., north of Waihi Stream; Fig. 9E). For example, a broad,
shallow channel-form (~3 m deep and ~80 m wide) incises into the upper MTD
surface north of Waiakapua Stream and is filled with current-structured sandstone
(Fig. 7).
The flatness of the NAMTD upper surface is somewhat puzzling because MTD
upper surfaces are often marked by significant rugosity following emplacement
(Lucente and Pini, 2008; Bull et al., 2009; Armitage et al., 2009). However, similarly
flat and apparently erosional upper surfaces have been described in other MTDs
observed in both outcrop and the subsurface (Woodcock, 1976; Farrell, 1984; Farrell
and Eaton, 1988; Alsop and Marco, 2011; King et al., 2011). The upper surface of the
NAMTD may have been modified by erosion following emplacement by either
turbidity currents or possibly strong contour currents (King et al., 2011). Alternatively,
the upper surface may have been levelled in a process described as relaxation by
Alsop and Marco (2011). In this scenario, seafloor topography generated by submarine
mass-movement was subject to secondary translation of material from highs into
adjacent troughs (Alsop and Marco, 2011). It is possible that one or both of these
mechanisms may have played a role in reducing relief atop the NAMTD.
MTD Thickness and Extent
The thickness of the MTD can only be constrained to a minimum estimate
(locally 55 m) along much of its outcrop extent because the basal detachment
surface is generally not exposed above beach level (Fig. 9). However, we have
identified and mapped a local exposure of the basal detachment surface for ~400
meters along the coastline at Paritutu Stream where both the upper and lower
bounding surfaces of the MTD are clearly visible in the wave-cut platform (Figs. 3 and
11). The calculated thickness of the MTD in this region is ~40 meters based on a
measured bedding dip of ~11 degrees to the southwest (Fig. 11).
The areal extent of the NAMTD is similarly uncertain. The eastern margin of
the MTD is poorly constrained due to dense vegetation that precludes detailed

105

mapping of the MTD inland from the coast, although isolated exposures of the MTD
are locally observed inland (Utley, 1987). The offshore region west of the MTD
outcrop extent is covered by a 3D seismic-reflection survey that extends to within ~3
km of the study area (Fig. 2). However, poor data quality within the eastern portion of
the 3D volume precludes confident identification of mass-transport related seismic
facies (Fig. 12). Although data quality improves ~12 km west of the coastline, the
Kahu 3D volume lacks strong evidence for the presence of a MTD within the
Mohakatino-Mount Messenger interval in this region (Fig. 12). Thus, we infer that the
NAMTD does not extend past ~12 km west of the coastline or has a thickness that is
below seismic resolution. The theoretical vertical resolution of the Mohakatino
Formation within the Kahu 3D survey is ~15 m (1/4 of the dominant frequency) based
on an estimated seismic velocity of 2,650 m/s (King et al., 2011).
Semi-continuous exposures of the Mohakatino Formation along the coastline
allow the N-S extent of the MTD to be better constrained than its E-W dimensions
(Fig. 3). Field relationships suggest that the southern margin of the MTD is likely
located just north of Pitone Stream (Fig. 13). This interpretation is based on the
observation that the MTD is absent in the vicinity of Pitone Stream where the basal
contact of the Mohakatino Formation is exposed (Figs. 4 and 13) and implies that the
MTD thickness decreases from ~40 to 0 meters in less than 1 kilometer (Fig. 13). This
interpretation is at odds with the hypothesis that the NAMTD is cogenetic with the
Otukehu Composite MTD that is exposed ~15 km to the south between the
Mohakatino and Tongaporutu Rivers (King et al., 2011). Although these MTDs are
similar in scale and deformational characteristics (King et al., 2011; Masalimova,
2013), we infer that these MTDs are not correlative on the basis of different
emplacement timing. Specifically, distinctive foraminifera in strata immediately
overlying the NAMTD yield an estimated depositional age of 10.9-10.5 Ma (Appendix
C-7). This age range is distinctly older than a volcanic ash bed sampled from within
the Otukehu Composite MTD (9.66 0.28 Ma; Maier, 2012). In addition, exposures
of the Mohakatino Formation south of the Awakino River lack evidence for being
remobilized, suggesting that the MTD does not extend into this region (Fig. 3).

106

The northern margin of the NAMTD is not exposed, but several lines of
evidence suggest that the MTD does not extend much farther north than Paparahia
Stream (Fig. 3). Coastal and inland exposures of the Mohakatino Formation are
widespread north of Paparahia Stream and are only rarely deformed by submarine
mass-movement (Nodder, 1987). In particular, a thick (>80 m), fault-bounded
sequence of undeformed Mohakatino Formation is present at Opito Point, just ~2.5 km
north of exposures of the NAMTD (Fig. 8). Because the estimated thickness of the
Mohakatino Formation is only ~120 m in this area (Nodder, 1990a), it is likely that the
NAMTD does not extend as far north as Opito Point. Similarly, the NAMTD is not
present at Waikawau beach where the Mohakatino-Mount Messenger contact is wellexposed and is undeformed. These observations suggest that the NAMTD likely only
extends up to 2.5 km, and not likely past 6 km, from its northernmost outcrop location
(Fig. 3). Taken together, these observations suggest that the outcrop exposures of the
NAMTD likely include a minimum of 60-85% of its original N-S extent (Fig. 3).
Stratigraphy and Sedimentology
Lithofacies
We designate two general end-member lithofacies that comprise the majority
of the NAMTD and of surrounding exposures of undeformed Mohakatino Formation:
1) thick- to medium-bedded volcaniclastic sandstone (LF1) and 2) thin-bedded
volcaniclastic sandstone and mudstone (LF2). These lithofacies are defined based on
associations

of

lithologies

characterized

by

distinctive

bedding

thickness,

mineralogical composition, grain size, sedimentary structures, and inferred processes


of deposition (Ghosh and Lowe, 1993). There is significant spatial variation in the
distribution of these two lithofacies within the NAMTD: LF1 predominates in the
northern ~5 km and gradually transitions to LF2 that becomes dominant south of
Poporotaupo Stream (Figs. 5 and 6). This same general facies transition is also
observed within surrounding, undeformed exposures of the Mohakatino Formation.
LF1 is abundant in northern outcrops at Waikawau beach and at Opito Point whereas
LF2 is more abundant to the south where exposed along the banks of the Awakino and

107

Mokau Rivers (Fig. 3). The commonality between lithofacies types and proportions
between the MTD and surrounding undeformed intervals strongly suggests that the
NAMTD originated locally and likely underwent limited translation.
LF1: Thick- to medium-bedded volcaniclastic sandstone
Description: Thick- (>50 cm) to medium-bedded (~20-50 cm) volcaniclastic
sandstone (LF1) is the most abundant lithofacies in the northern 5 km of the NAMTD,
between Ounutae and Paparahia Streams (Figs. 5, 6, and 14). Although LF1 is
dominated by medium- to very coarse-grained sand and gravel, interbedded thin beds
of fine volcaniclastic sandstone and mudstone are also present. Coarse-grained,
volcaniclastic sandstone beds occur as amalgamated packages up to 10 meters in
thickness (Fig. 14). Normal grading (including coarse-tail grading) and massive
texture are typical, and common sedimentary structures include water escape
structures (sheets, pillars, and dish structures); sole marks (flutes); and planar, ripple,
and convoluted laminations. Intraformational clasts of sand and mud (up to
approximately a meter in diameter) are abundant in the thick-bedded sandstone and
often form intraclast and lithic fragment breccio-conglomerates (Fig. 5).
LF1 is characterized by a high degree of lateral bedding continuity. Individual
beds can be correlated between outcrops within the northern 3 kilometers of the MTD
and demonstrate little bedding thickness changes over this distance despite significant
structural shortening (Fig. 5). This same lithofacies can be observed undeformed
several kilometers north of the MTD at Opito Point where individual beds can be
traced for over two kilometers (Fig. 8).
LF1 also contains occasional beds of non- or mixed-volcaniclastic composition
that appear to have a Mount Messenger Formation affinity (Utley, 1987). These
siltstone and sandstone beds are conspicuous because they weather to a grey color
instead of the brown color typical of the surrounding volcaniclastic units. Because
these beds are relatively uncommon, they constitute useful marker beds that aid
correlation along the coastline (e.g., green-highlighted marker bed; Figs. 5 and 9).
Interestingly, these isolated beds are only found within LF1 north of Poporotaupo

108

Stream, but are also present within undeformed Mohakatino strata at Waikawau beach,
~ 6 km north of the NAMTD.
A ~22 m thick package of intensely-deformed, broken and chaotic medium- to
thick-bedded sandstone is present in the upper portions of the NAMTD north of
Kopuai Stream (Figs. 9 and 15F). A detachment surface separates this unit from
underlying beds that are only slightly deformed by long-wavelength folding (Fig. 9B).
Although precise timing between emplacement of this unit and the NAMTD is
ambiguous, the intensely-deformed package is cross-cut by folds within the NAMTD.
We correspondingly interpret this package to have been a precursor MTD within the
original stratigraphy that was subsequently remobilized within the NAMTD.
Interpretation: Thick-bedded strata of LF1 are interpreted to have been
deposited primarily by high-density turbidity currents forming S2 and S3 units of Lowe
(1982) and Tb units of Bouma (1962). Medium-bedded strata are interpreted to have
been deposited both by high- and low-density turbidity currents, characterized by Ta,
Tb, and Tc divisions (Bouma, 1962). In addition, banded slurry-flow deposits (sensu
Lowe and Guy, 2000) have also been identified locally within LF1.
LF2: Thinly-bedded volcaniclastic sandstone and mudstone
Description: Thin-bedded ( 20 cm) volcaniclastic sandstone and mudstone
(LF2) becomes abundant south of Waioroko Stream and is the dominant lithofacies
south of Poporotaupo Stream (Fig. 6). This lithofacies is very well-bedded and has a
notably bimodal grain size distribution; medium sand to granule bases are capped by
fine sand to mud tops (Fig. 16). Sedimentary structures include normal grading;
massive texture; planar and convoluted laminations; ripple laminations; and sole
marks (flame and loading structures; Fig. 16). LF2 is also characterized by a high
degree of bedding continuity; packages of thin-bedded sandstone can be traced for
over 2 km at Opito Point (Fig. 8). Bioturbation is frequently destroys primary
sedimentary structures and often mixes medium to very coarse volcaniclastic sand
within the fine-grained matrix. This lithofacies is only rarely observed to have
undergone mass-movement prior to the NAMTD event, and lacks the interbedded
non-volcaniclastic beds found within LF1.

109

Interpretation: The thin-bedded sandstone and mudstone of LF2 are


interpreted to have been primarily deposited by one of two mechanisms. Beds that
have current structures and display Bouma (1962) Tb and Tc divisions are indicative of
deposition from waning, low-density turbidity currents. Other beds, characterized by
normal grading and otherwise massive texture, are indicative of suspension
sedimentation of volcanic ash through the water column.
Interpreted Depositional Environment
Several lines of evidence suggest that both LF1 and LF2 were deposited as
sheets or lobes within a low-gradient, laterally unconfined deep-marine environment:
(1) the abundance of massive sandstone beds with dish structures (S3 divisions; Lowe,
1982) suggesting collapsing flows and rapid sedimentation rates due to a loss of
confinement and/or decrease in slope; (2) the high degree of lateral bedding continuity
(Johnson et al., 2001; Fig. 8); (3) the presence of slurry-flow or hybrid-event beds that
have elsewhere been largely associated with unconfined depositional settings
(Hodgson, 2009; Haughton et al., 2009); and (4) the absence of incisional or largescale erosional features that suggest channelized confinement (e.g., Fig. 8).
Although the channels that fed the Mohakatino lobes were not observed in
outcrop, we suspect that these conduits were located nearby to LF1. Abundant
intraformational rip-up clasts within LF1 suggest that considerable erosion occurred
upstream prior to deposition. Also, the coarse-grained, thick-bedded, and amalgamated
nature of LF1 is consistent with deposition within the proximal (inner) or axial
portions of a submarine lobe (Fig. 14; Prlat et al., 2009). The sedimentary character
and fabric of LF2 suggest that this lithofacies was likely deposited in a lobe off-axis to
fringe setting (Prlat et al., 2009) based on several observations relative to LF1: (1) a
decrease in grain size and bedding thickness, (2) an increase in bioturbation, and (3)
more abundant preservation of volcanic ash horizons that may have been deposited by
suspension sedimentation through the water column. Thus LF2 likely was deposited in
a more quiescent environment relative to LF1, although LF2 sedimentation was

110

periodically interrupted by deposition of medium-bedded, current-structured sandstone


(Fig. 6).
Evaluation of vertical stacking patterns between LF1 and LF2 within the
NAMTD is hindered by intense soft-sediment deformation (e.g., Fig. 9). However, the
in situ vertical relationship between LF1 and LF2 can be observed at Opito Point
where LF1 constitutes ~40% of a 46 m section (Fig. 8). Approximately six poorlydefined upward-thickening and coarsening cycles are present that range from 6 to 11
m thick (Fig. 8). Each interpreted cycle has a fine-grained lower portion of
interbedded thin-bedded sandstone and/or mudstone (LF2) and is capped by thickbedded sandstone (LF1; Fig. 8). This cyclic stacking of LF1 and LF2 could represent
vertical alternation of lobe and interlobe settings that we speculatively attribute to
compensational switching of lobe depocenters over time (Fig. 8).
Paleocurrents
Measurements of paleocurrent indicators from undeformed strata within the
Mohakatino Formation demonstrate a wide variety of flow directions (Figs. 5-7;
Appendix C-8). Paleocurrent measurements collected at Opito Point suggest flow
predominantly toward the S or SE (ripples and flute casts) and toward the N (ripples;
Fig. 8). W-directed flow is indicated by measurement of ripples and numerous dunes
within the basal Mohakatino Formation along the north bank of Pitone Stream (Fig.
7). Paleocurrent measurements from immediately above the MTD include NE-directed
flow in a broad, shallow (~3 m) channel-form at Waiakapua Stream and S-directed
flow at Paparahia Stream (Fig. 7).
Paleocurrent measurements from within the NAMTD should be viewed with
caution because of the potential for vertical-axis rotation during submarine massmovement. For this reason, we collected relatively few paleocurrent measurements
within the MTD. However, we did measure flute casts from the bases of two
distinctive, thick sand beds where they are repeatedly exposed at beach level several
times over a distance of two kilometers within subregions R1-3 (measurements from
sections 2, 4, and 5; Fig. 5). In all instances, the flute casts indicate flow toward the

111

eastern quadrant. However, there appears to be a systematic change from NE- to SEdirected flow from north to south along the outcrop extent (Fig. 5). The mean
paleocurrent vector changes from 54 7 (section 2) to 121 10 (section 5),
suggesting an apparent flow direction that varies up to ~67 (Fig. 5). This pattern
could result from spatial variability in the current flow direction while the sand bed
was being deposited. Alternatively, the flute casts can be assumed to have been
originally parallel and that section 2 was subsequently rotated in a counter-clockwise
direction relative to section 5.
Slump Deformation Structures
Slump Folds
Slump folds form some of the most conspicuous and exquisite deformational
features present within the NAMTD (e.g., Figs. 15 and 17). These features exhibit a
considerable variety of shapes, sizes, and orientations that vary systematically along
the outcrop transect. The largest folds have heights up to several 10s of m and are
most commonly found in the north (subregions R1-4) where LF1 predominates (e.g.,
Figs. 15A-B, 15G-H, 17, and 18). These folds tend to occur within semi-continuous
fold trains (Figs. 11A-B), and individual folds locally extend ~1 km laterally (e.g.,
Figs. 9C and 11A-B). Folds in the south (subregions R5-7) are typically disharmonic,
tend to occur as rootless individuals (Figs. 15D-F), and have an average fold height
that is less than found in subregions R1-4. Continuous fold trains are uncommon
except where present locally in subregion R7 (Figs. 11D and 15H).
Fold axes within the NAMTD are typically shallowly plunging (<25) and
display a wide range of axis orientations (Figs. 19 and 20). The best-fit girdle to fold
axes correspondingly dips shallowly to the west (192/12W; Fig. 20). Fold axes are
typically assumed to be dispersed within the plane of the MTD and thus can be used to
infer the overall orientation of the deposit (Woodcock, 1979b). However, we suspect
that the W-dipping girdle is likely an artifact of the slight (typically 10) westward,
regional tilting of the Taranaki coastal exposures. A low amount (generally <20) of
fold axis curvature has been observed in some large folds (e.g., Figs. 15A,G-H).

112

Sheath folds (fold axis curvature > 90; Ramsay and Huber, 1987) have not been
identified within the NAMTD.
Fold attitudes range from upright to nearly recumbent and interlimb angles are
typically less than 60 or greater than 100 (Fig. 19). Although many fold axial
surfaces are approximately planar at the scale of outcrop, some fold axial surfaces
display considerable amounts of curvature (e.g., Figs. 15G-H and 18). For example,
the upwards-convex geometry observed in some folds (e.g., Fig. 18) may reflect
modification during layer-parallel simple shear (e.g., Farrell and Eaton, 1987; their
Fig. 3A). Similar structures to the large folds shown in Figure 18 have been
experimentally produced by deformation of layered sand and clay during simple shear
(Dasgupta, 2008).
Folds demonstrate variable thickening or thinning of limbs relative to the
hinge, and fold layer shapes vary from class 1A to Class 2 (Ramsey, 1967).
Calculation of dip isogons for several folds of varying scales demonstrates that most
folds exhibit a shape (class 1C) that is intermediate between parallel and similar ideals
(Fig. 21; Ramsey, 1967). Correspondingly, both flexural-slip and flexural-flow folding
(Donath and Parker, 1964) occurred within many folds; interbedded fine-grained
lithologies tend to preferentially flow relative to the coarser-grained lithologies (Figs.
18 and 22). However, even thick sandstone beds exhibit thickening in the hinge region
(class 1C folds), and thus have been subjected to a certain amount of flexural-flow
during folding (Figs. 21 and 22). Fold detachment is repeatedly observed where
parallel folding produces a space problem within the hinge zones (e.g., Fig. 22). Flow
of finer-grained lithologies into the fold hinge creates a dcollement zone that often
contains sandstone boudins within a pervasively sheared, fine-grained matrix (Figs. 18
and 23; Davis et al., 2012). Within some isoclinal folds found in subregions R1-3, the
dcollement zone can extend for more than 50 meters in profile view, thereby
separating hinges of formerly adjacent sandstone beds (Fig. 22A). One well-exposed
series of recumbent, nearly isoclinal folds in a cliff face (subregion R3) demonstrates
that fold detachment can result in the same sandstone beds being structurally repeated
up to three times (Figs. 15B and 23).

113

Deformation intensity varies significantly within the NAMTD (e.g., Fig. 17).
In general, the most intensely deformed regions are found within the central portion of
the outcrop extent (subregions R4-6). However, zones of high strain often transition
laterally or vertically into low-strain domains that are characterized by flat- to
shallowly-dipping bedding, occasional low-angle discontinuities, and relatively few
open to tight folds. These regions can be clearly identified on the wave-cut platforms
as having continuous bedding that is tilted to the west and only disrupted by upright,
long-wavelength folding (Fig. 11). Low-strain zones are particularly well developed in
three locations: north of Kopuai Stream, between Poporotaupo and Waiongaro
Streams, and between Paritutu and Ngatupaku Streams (Figs. 3 and 11). Upwardsincreasing strain is also observed in several locations where folds become tighter,
more disharmonic, and increasingly dismembered towards the upper surface of the
MTD (e.g., subregions R4 and R7). Other regions have highly deformed intervals that
are intercalated with low-strain domains (e.g., Fig. 15D).
Slump Faults
Two general groups of faults can be identified within the NAMTD: those that
accommodated brittle deformation during submarine mass-movement (both normal
and reverse slump faults) and those that postdate the MTD and are of a tectonic origin
(exclusively normal faults; Childs et al., 2007; Giba et al., 2010, 2013). The younger
group of tectonic normal faults generally strike NE-SW or E-W (Utley, 1987; Nodder
et al., 1990a; Giba et al., 2010), and can be identified in outcrop because they extend
through the MTD and cross-cut the overlying succession (e.g., Fig. 9). Slump normal
faults can be distinguished from the tectonic faults based on several defining
characteristics: (1) they are confined to bedding within the MTD; (2) they often have
welded surfaces that lack mineralization or fractures (Fig. 24A); (3) they have
surfaces that are often irregular and curved (Fig. 24B); and (4) they usually
preferentially displace coarser-grained units and tip out within fine-grained intervals
(Figs. 24C-D; Debacker et al., 2009). The following discussion will focus exclusively
on slump faults.

114

Slump faults are abundant in the NAMTD and demonstrate a wide range of
orientations and styles of deformation (Figs. 24, 25, and 26). Planar discontinuities
commonly separate rootless folds in regions that have undergone significant amounts
of deformation (Fig. 15F; Dasgupta, 2008). Similar discontinuities bound coherent
blocks of stratigraphy and commonly have low dip angles (Fig. 24G). These
features may represent low-angle detachment or thrust faults that have likely been
modified by progressive deformation during translation. Because the displacement
magnitude is typically greater than the scale of the outcrop, the relative sense of
displacement cannot usually be determined. Other faults are preferentially located in
fold hinge zones, suggesting that these faults formed during folding (e.g., Fig. 24I).
We measured the orientation and relative displacement sense of 242 slump
faults that displace bedding and can be identified as having a relative normal or
reverse sense of motion (Figs. 25 and 26). We avoided measuring faults that appeared
to have formed locally as a result of folding (e.g., Fig. 24I). Normal faults typically
dip between 30 and 70 degrees, and reverse faults dip between 5 and 35 degrees (Fig.
25). Although fault displacements can be as great as several meters or more, the large
majority of measured slump faults within the NAMTD have apparent displacement
magnitudes that are less than about 20 cm (Fig. 25). Thus, the scale of brittle
deformation is typically up to several orders of magnitude less than the scale of ductile
strain accommodated in folds. Indicators of fault slip direction (e.g., slickenlines) were
not observed, and thus fault displacement magnitude and sense can only be
approximately constrained.
PALEOSLOPE ANALYSIS
The orientations of slump deformation features have long been considered to
be related to the orientation of the underlying bathymetric slope (e.g., Hahn, 1913;
Jones, 1939). For example, early workers noted that some slump fold axes tend to
align in an alongslope direction in a pattern analogous to a corrugated carpet
(Woodcock, 1979b). Although slump folds have been most commonly used in
paleoslope analysis, slump faults have also been demonstrated to provide useful

115

kinematic information (e.g., Farrell, 1984; Martinsen and Bakken, 1990; Debacker et
al., 2009). A challenging aspect to paleoslope analysis is that slump deformation
features can be variably orientated with respect to the underlying paleoslope (Hansen,
1971; Woodcock, 1979b; Farrell and Eaton, 1987). For example, slump folds axes that
initially formed parallel to the strike of slope (hereafter alongslope folds) can
develop oblique or downslope orientations during progressive simple shear (e.g.,
Farrell and Eaton, 1987). Oblique or downslope folds can also be generated in situ as a
result of layer-normal shear or (Alsop and Marco, 2013) or as a result of shortening
across the width of the MTD.
Our approach is to conduct a rigorous paleoslope analysis by using as many
fold- and fault-based methods as applicable and then comparing results of each
method. We also test for spatial variability in the NAMTD by estimating the MTD
transport direction separately within seven divided subregions and for the MTD as a
whole (Fig. 3). Each of the seven subregions was defined according to north to south
position along the outcrop extent and in some cases noted changes in the orientations
of slump folds and faults. The following sections provide an explanation of each foldand fault-based paleoslope method used in this study (Table 1).
Methods Based Upon Slump Fold Orientations
Mean Axis, Separation Arc, and Downslope Average Axis Methods
1. Overview
The mean axis method (MAM) postulates that the mean fold axis orientation
approximates the paleoslope strike (Jones, 1939; Fig. 27; Table 1). The MAM assumes
that fold hinges have undergone little downslope rotation and that the MTD underwent
little strike-parallel length changes (Woodcock, 1979b). This method yields two
possible MTD transport directions that are 180 apart (Fig. 27). The paleoslope dip
direction is usually chosen based on regional paleogeographic constraints, fold
asymmetry (folds are assumed to verge in the downslope direction), and/or fold facing
directions (beds become younger in a downslope direction within fold hinges;
Woodcock, 1976; Bell, 1981). The MAM also has the advantage of being statistically

116

robust as a result of being based on average properties of fold measurements


(Woodcock, 1979b).
A major disadvantage to the MAM is that it does not allow for fold axes that
are oriented parallel or oblique to the downslope direction, even though such
distributions are thought to occur in MTDs (Hansen, 1971; Lajoie, 1972). Downslope
or oblique fold axes can result from (1) progressive layer-parallel simple shear (Farrell
and Eaton, 1987; Bradley and Hanson, 1998), (2) vertical-axis rotation of folds, (3)
lateral compression, and (4) marginal effects within the lateral-to-oblique portions of a
MTD (Woodcock, 1979b; Debacker et al., 2009). In these cases, the MAM can yield
an estimate of the paleoslope orientation that is up to 90 different from the true value
(Woodcock, 1979b).
The separation arc method (SAM; Hansen, 1965) is based on both fold axis
orientation and sense of asymmetry and has the advantage of allowing for downslopeoriented fold axes (Fig. 27; Table 1). This method was originally developed based on
observations from a small tundra landslide and is predicated on the assumption that
folds with opposing vergence will plot in two distinct groups on an equal-area
stereonet (Hansen, 1971; Fig. 27). An angle, or arc, is drawn between groups of
counterclockwise- and clockwise-verging folds on an equal area stereonet, and the
MTD transport direction is assumed to lie within this angle (or separation arc; Hansen,
1965, 1971; Fig. 27). Major disadvantages of the SAM include its reliance on the
extreme rather than average properties of the dataset and its incompatibility with upslope verging folds (Woodcock, 1979b). Addition or subtraction of a single outlier can
dramatically change the resulting separation arc (Woodcock, 1979b; his Fig. 2). In
addition, this method can only be confidently applied to datasets with non-overlapping
fold axis asymmetry senses (Fig. 27).
The downslope average axis method (DAM; Woodcock, 1979b) is a variant of
the MAM that allows for downslope-oriented fold axes (Fig. 27; Table 1). The DAM
assumes that the mean fold axis is orientated at a high angle to the paleoslope strike.
Woodcock (1979b) suggested that an alongslope or downslope fold axis mean can be
distinguished from each other by examining the distribution of fold asymmetry about

117

the mean fold axis (Fig. 27). Folds with an alongslope mean axis will ideally plot in
two clusters of either Z- or S-folds, and folds with a downslope mean axis will plot in
two clusters that each has populations of S- and Z-folds that are positioned on either
side of the mean (Woodcock, 1979b; Fig. 27).
2. Interpretation
Fold axes within the NAMTD are markedly dispersed and show little preferred
orientation (Fig. 20). This observation suggests that folds are variably oriented with
respect to the underlying paleoslope. Furthermore, it is ambiguous whether the mean
fold axis (279/38) is oriented parallel, perpendicular, or possibly obliquely to the
bathymetric slope. Although the MAM predicts a N- or S-directed transport direction
(9 or 189), the wide distribution of fold axes casts doubt on whether a reliable fold
axis mean can be calculated (Woodcock, 1979b). Similarly, the complete overlap
between Z- and S-folds precludes the use of the SAM or DAM in determining the
paleoslope orientation (Figs. 20 and 27).
Although interpreting fold axes within the entire NAMTD is ambiguous, six of
the seven subregions (R1-5 and R7) have interpretable fold axis distributions that yield
markedly different mean values from each other (Fig. 20). Application of the MAM
suggests either N- or S-directed MTD transport for subregions R1 and R7 (14-194
and 179-359, respectively; Fig. 20). These results differ from the E-W or SE-NW
transport directions inferred for subregions R2-5 (from 72-252 to 148-328; Fig.
20). Subregions R1-2 display a dominant vergence direction that suggests S-and SWdirected transport, respectively (Fig. 20). However, the other subregions (R3-7) are
characterized by bimodal fold vergence distributions that only allow the transport
direction to be constrained to one of two possibilities (Fig. 20). The SAM and DAM
cannot be applied because either only S-folds were measured (subregions R1-2) or Zand S-folds do not occupy distinct fields (subregions R3-6; Fig. 20). Subregion R7
displays only partial overlap of opposing vergence senses; most S-folds occur north of
the 89-269 mean vector (7 of 10) and most Z-folds occur south of the mean vector (5
of 8). Although a separation angle cannot be defined, the DAM indicates an E-W
transport direction for subregion R7 (Fig. 20).

118

Axial-Planar and Axial-Planar Intersection Methods


1. Overview
Fold axial surfaces (approximated as axial planes) yield important kinematic
information in a similar way as fold axes (Woodcock, 1976, 1979b; Farrell and Eaton,
1987). Axial plane orientations have the added benefit of including the sense of fold
vergence that is inferred to be in the up-dip direction. Poles to axial planes tend to fan
in a great circle about the mean fold axis, and thus the strike of the best-fit girdle to
these poles gives a bi-directional estimate of the MTD transport direction (axial-planar
method, APM; Woodcock, 1979b; Figs. 20 and 27; Table 1). Furthermore, imbrication
of the mean axial plane with respect to the overall MTD orientation can be used to
infer the paleoslope dip direction (Woodcock, 1979b). This assumes that the majority
of folds verge in the downslope direction and thus the trend of the mean axial plane
pole will be located in the direction of MTD transport (Woodcock, 1979b; his Fig. 4).
A similar method assumes that the MTD transport direction is related to the
intersection of the mean S- and Z-fold axial planes (axial-planar intersection method,
APIM; Alsop and Holdsworth, 2002; Strachan and Alsop, 2006; Table 1). This
method was initially developed on folds from mid-crustal shear zones (Alsop and
Holdsworth, 2002) but has also been successfully applied to paleoslope analysis of
ancient MTDs (Strachan and Alsop, 2006). The MTD transport direction is inferred to
be the trend of the S- and Z-fold mean plane intersection (designated APIM-N) for
regions characterized by layer-normal shear (e.g., wrench strains at the margin of a
submarine landslide; Strachan and Alsop, 2006). However, the transport direction is
orthogonal to the intersection trend for regions characterized by layer-parallel shear
(designated APIM-P; Strachan and Alsop, 2006). Thus, the APIM yields three
possible transport directions. We have additionally considered the fourth direction for
axial-planar intersections with low plunges (< 10) because the point of intersection is
within error of shifting 180 on an equal-area stereonet (e.g., Fig. 20).
It is worth noting that the APIM is subject to several major limitations that do
not apply to the APM. (1) A slight change in the orientation of a shallowly-plunging

119

fold axis can switch the sense of asymmetry from clockwise to counter-clockwise, or
vice versa (e.g., doubly-plunging folds; Fig. 15G). The asymmetry senses of such
folds are therefore also sensitive to any structural restoration applied in tectonically
deformed regions. (2) The APIM can only be applied where both S- and Z-folds are
present. (3) The APIM excludes folds that lack a sense of asymmetry in the paleoslope
analysis; such folds can constitute a large proportion of slump fold datasets (e.g., 43%
of the NAMTD fold dataset; Appendix C-2). Thus, the APIM is most applicable to
datasets with sufficient numbers of both S- and Z-folds that are not susceptible to
reversals in plunge direction.
2. Interpretation
As with fold axes, axial planes within the NAMTD are widely dispersed and
show little preferred orientation (Fig. 20). The best-fit girdle of the poles is oriented
349/64E, and thus the APM predicts a transport direction parallel to 169-349 (Fig.
20; Table 2). Application of the APIM to all axial planes in the NAMTD yields an
estimated transport direction of 10-190 (APIM-P) or towards 280 (APIM-N; Table
2). However, the best-fit girdle is poorly constrained in such scattered distributions as
a result of being sensitive to outlying points (Fig. 20). Thus, the results from these
methods should be viewed with caution.
The APM can be applied to each subregion within the MTD and yields similar
results as the MAM (Fig. 20). Subregions R1 and R6-7 yield north or south transport
directions (Fig. 20). Subregions R2-R5 display a variety of predicted transport
directions that range from NE-SW (53-233) to SE-NW (158-338; Fig. 20, Table
2). The mean axial plane variably dips to the west or east in these regions. The APIM
can be applied to 4 of 7 subregions (R3-4 and R6-7) and yields similar results (within
25) to the APM, with the exception of subregion R6 which is 75 different from the
APM (Fig. 20; Table 2).

120

Methods Based on Modification of Fold Geometry during Progressive Simple Shear


1. Overview
Farrell and Eaton (1987) proposed a model in which the orientation and
deformational style of slump folds can be related to progressive deformation during
submarine mass-movement. Contractional strain (pure shear) initially forms folds that
have upright axial surfaces, alongslope fold axes, and moderate (open to tight)
interlimb angles (Farrell and Eaton, 1987). Subsequent progressive layer-parallel
simple shear modifies the original fold geometry in several ways: (1) axial surfaces are
rotated to become parallel with the bounding surfaces of the MTD, (2) fold axes are
rotated in a downslope direction, and (3) folds progressively tighten (Farrell and
Eaton, 1987). This process culminates in the development of dismembered isoclinal
folds and sheath folds that are elongate downslope (Bradley and Hanson, 1998;
Dasgupta, 2008).
Because folds can progressively tighten and become recumbent during
progressive deformation, the relationship between fold hinge azimuth and fold
interlimb angle (HIM; Strachan and Alsop, 2006) or axial plane dip (HAM; Farrell
and Eaton, 1987) can be used to identify downslope hinge rotation and determine the
direction of MTD transport (Fig. 27, Table 1). In both cases, the downslope direction
can be identified on a plot of interlimb angle or axial plane dip versus fold axis
azimuth as a V-shape with the apex coinciding with the direction of MTD transport
(Figs. 27 and 28; Farrell and Eaton, 1987; Strachan and Alsop, 2006).
Correspondingly, folds that have undergone little hinge rotation will not display a
systematic relationship between fold axis azimuth and interlimb angle or axial plane
dip (Fig. 27). Debacker et al. (2009) proposed a similar method that uses the same
relationship between axial plane orientation and fold interlimb angle to infer
downslope fold rotation with progressive fold tightening (axial surface strike and
interlimb angle method; ASIM).
2. Interpretation
Neither the HIM, HAM, nor ASIM can be confidently applied to the entire
NAMTD or any subregion within because there is no clear relationship between fold

121

interlimb angle and the attitude of fold axes or axial planes (Fig. 28). These
observations suggest that folds underwent limited systematic downslope hinge rotation
(Strachan and Alsop, 2006). A single possible exception may be subregion R5 whose
fold axes and axial planes display a poorly-developed V-pattern when plotted against
interlimb angle with the apex oriented towards ~148 (Fig. 28). Subregions R1-4 and
R7 have fold orientations that plot in one of two groups that each has a range of
interlimb angles (Fig. 28). Application of the HIM, HAM, and ASIM in these regions
yields results that are functionally equivalent to the MAM and APM. The wide scatter
of fold attitudes within subregion R6 precludes the application of HIM, HAM, or
ASIM in this area (Fig. 28).
The Axial Surface Dip and Dip Direction Method
1. Overview
Debacker et al. (2009) proposed that the relationship between the dip
magnitude and dip direction of fold axial surfaces is related to the transport direction
in the lateral to oblique portions of MTDs (axial surface dip and dip direction method;
DDM; Table 1). This method assumes several conditions are present in the
lateral/oblique portions of MTDs: (1) layer-normal shear predominates, (2) fold axial
surfaces are steeper than in the central/frontal portions of the MTD, and (3) fold axes
and axial planes fan about the paleoslope dip direction (Debacker et al., 2009; Alsop
and Marco, 2013). Axial surfaces are presumed to steepen with progressive layernormal shear, and this process is manifested as a negative relationship between
interlimb angle and axial surface dip (Debacker et al., 2009; their Fig. 15C).
2. Interpretation
A major challenge to the application of the DDM is the requirement of
prerequisite knowledge about relative position within the MTD (Table 1). As such,
this method can only be confidently applied when independent evidence indicates a
position within the lateral/oblique portion of the MTD. The margins of the NAMTD
are not exposed, but geologic relationships suggest that subregions R1 and R7 likely
lie within the lateral portions the MTD (Figs. 3 and 13). Application of the DDM to

122

subregions R1 and R7 yields E- or W- directed transport (Fig. 20). However, several


considerations suggest that subregions R1 and R7 do not strictly meet the assumptions
of the DDM (Table 1). (1) Plots of interlimb angle versus axial plane dip show a
positive correlation that suggests slump folds become more reclined during
progressive fold tightening. This pattern is typical of slump folds that have been
deformed by layer-parallel simple shear (e.g., Farrell and Eaton, 1987), but differs
from the negative relationship predicted to occur at MTD margins (Debacker et al.,
2009). (2) Although fold axial surfaces are predicted to be steeper along the MTD
margins than in the center (Debacker et al., 2009; Table 1), this relationship is not well
developed along the inferred margins of the NAMTD. The mean axial plane dip is
slightly greater for subregion R7 than other subregions, but this is not so for subregion
R1 (Fig. 28). (3) There is no clear, systematic relationship between the orientation and
dip of fold axial planes (Fig. 28). Subregions R1 and R7 display little (10)
difference between fold axial surface attitudes above and below the mean dip (Fig.
28). Taken together, these observations suggest that the DDM is not likely an
applicable method for determining the transport direction of the NAMTD.
Methods Based Upon Slump Fault Orientations
Mean Fault Orientation Method
1. Overview
The mean fault orientation method (MFOM) simply presumes that the MTD
transport direction is sub-parallel to the mean fault dip direction (up-dip for reverse
faults and down-dip for normal faults; Farrell, 1984; Martinsen and Bakken, 1990;
Debacker et al., 2009). This method assumes that there is not a significant component
of oblique- or strike-slip fault motion (Debacker et al., 2009). Unfortunately, the
degree of oblique- or strike-slip fault movement is difficult to assess because slip
direction indicators (e.g., slicken-lines) were not observed on slump fault surfaces.
2. Interpretation
Reverse faults within the NAMTD display little preferred orientation when
combined together (Fig. 26). The mean reverse fault is nearly flat-lying (320/1 NE),

123

and thus the MFOM cannot be reliably applied to the NAMTD as a whole. Normal
faults similarly display a wide range of orientations but tend to be clustered into two
main groups that are dipping to the east and north (Fig. 26). As a result, the mean fault
surface dips towards the northeast (332/19 NE; Fig. 26). Smaller populations of S- and
W-dipping faults are also present (29% of total, Fig. 26). Application of the MFOM to
all normal faults within the NAMTD suggests a transport direction towards the
northeast (43; Fig. 26; Table 2).
As with slump folds, reverse faults orientations vary within different
subregions of the NAMTD (Fig. 26). In general, poles to reverse fault planes form a
steeply-dipping girdle with poles clustering in opposing quadrants (Fig. 26). This
pattern reflects the presence of conjugate sets of reverse faults that have sub-parallel
strike orientations but opposing dip directions (Fig. 26). As a result, the MFOM can
only constrain the direction of MTD translation to one of two possiblities in regions
where there is not a dominant vergence orientation (Figs. 26). This is the case for all
subregions except R4 where most (9 of 11) measurements are dipping to the east (Fig.
26). Estimated transport directions range from NE-SW to SE-NW for R1-5 (from 36216 to 116-296) and N-S for R6-7 (169-349 and 117-357, respectively; Fig. 26;
Table 2).
Normal faults display comparatively less systematic variation in orientation
from north to south than reverse faults (Fig. 26). Most subregions have major
populations of E- and N-dipping faults with subordinate populations of S- and/or Wdipping faults (Fig. 26). As a result, the best-fit girdles dip to the W, SW, or S (Fig.
26). Application of the MFOM to normal faults is somewhat less reliable due to the
complex nature of fault orientation distributions, and low numbers of measurements in
the north preclude robust analysis of normal faults in these areas (i.e., subregions R12). Even so, the MFOM yields consistent NE- and E-directed transport directions (36102) for subregions R3-7, assuming that the most abundant populations of normal
faults are synthetic with respect to the paleoslope (Fig. 26, Table 2). These results are
inconsistent with the results from the fold- and reverse fault-based methods with the
possible exception of subregion R4 (Figs. 20 and 26).

124

Best-fit Girdle to Fault Poles Method


1. Overview
A similar approach to the MFOM is to approximate the MTD transport
direction as the strike of the best-fit girdle to fault plane poles (GFPM; Fig. 29; Table
1). This method has previously been applied to slump faults that were interpreted to
have been oriented obliquely to the paleoslope strike (Debacker et al., 2009).
However, this method is equally applicable to faults that strike at low angles to the
paleoslope (alongslope faults). We present a modified version of the GFPM that is
applicable to both normal and reverse faults, allows for both alongslope faults and
faults with orientations that fan obliquely about the paleoslope strike, and can be
applied when both synthetic and antithetic faults are present (Fig. 29).
The modified GFPM is analogous to the fold-based APM. The estimated MTD
transport direction is taken to be in the strike direction of the best-fit girdle for faults
with inferred alongslope orientations (GFPM-P; Figs. 29A-B). Such faults can be
recognized by poles that are ideally dispersed along a steeply-dipping girdle (Figs.
29A-B). Synthetic and antithetic faults will be positioned on opposite sides of the
stereonet center point. The MTD transport direction is taken to be in the down- or updip direction for the synthetic population of normal and reverse faults, respectively
(Fig. 29A). The transport direction can only be constrained to one of two possibilities
if there is ambiguity about which fault population is synthetic versus antithetic. Note
that the GFPM should be applied with caution for tightly clustered (unimodal) datasets
(e.g., Fig. 29A) because the best-fit girdle is imprecise for such distributions; the
MFOM yields a more reliable result in this case.
For faults with orientations that fan about and are oblique to the paleoslope
strike, the MTD transport direction is taken to be parallel to the dip direction of the
best-fit girdle (GFPM-N; Figs. 29C-D). Faults with obliquely fanning axes can be
recognized by poles that are ideally aligned along a moderately-dipping girdle (Figs.
29C-D). The sense of displacement is taken to be the dip direction for normal faults
and the opposite of the dip direction for reverse faults (Debacker et al., 2009).

125

Application of the GFPM-N becomes more complicated if both synthetic and


antithetic faults are present (Fig. 29D). Ideally, separate best-fit girdles can be drawn
if groups of synthetic and antithetic fault poles can be identified, and these girdles will
have similar strike directions but opposite dips (Fig. 29D). The transport direction is
constrained only to one of two possibilities if uncertainty exists regarding which set is
synthetic versus antithetic (Fig. 29D).
2. Interpretation
Use of the GFPM requires evaluation of normal and reverse fault pole
distributions to determine (1) whether faults have alongslope or obliquely fanning
orientations and (2) if both synthetic and antithetic populations are present (Fig. 29).
Normal fault distributions within the NAMTD and most subregions (excluding R4)
have moderately-dipping (38-78) girdles and a dominant set of E- and N-dipping
faults (Fig. 26). Most subregions also contain a second set of S- and W-dipping faults
(Fig. 26). These two populations can be interpreted to represent two conjugate sets of
fanning normal faults (e.g., Fig. 29D). If correct, the GFPM-N is applicable, and this
method yields consistent SW or NE (40-54 or 220-234) transport directions for the
entire NAMTD and for subregions R3 and R5-7 (Fig. 26). If the most abundant
conjugate set (E- and N-dipping) is taken to be synthetic to the paleoslope, the inferred
transport direction is toward the northeast. Normal fault poles in subregion R4 are
aligned along a steeply-dipping girdle, and we thus infer that these normal faults have
primarily alongslope orientations (e.g., Fig. 29B). For this reason, we use the GFPM-P
that yields an inferred transport direction towards the east (110; Fig. 26), assuming
that the dominant E-dipping fault population is synthetic to the paleoslope.
Reverse faults within most subregions of the NAMTD have poles aligned
along steeply-dipping girdles (82-89; Fig. 26), and thus appear to have alongslope
orientations (Figs. 29A-B). Application of the GFPM-P yields very similar results as
the MFOM (Fig. 26), and uncertainty about which population of faults is synthetic
versus antithetic in subregions R1-3 and R6-7 limits the interpreted transport direction
to one of two possibilities (Figs. 26 and 29B). Fault poles in subregion R5 display a

126

comparatively wide spread and are not clearly aligned along a girdle (Fig. 26); thus the
GFPM-P yields an ambiguous result in this case.
Mean Bedding Strike Method
1. Overview
In a similar way to folds, bedding attitudes within a MTD are expected to be
preferentially oriented with respect to the paleoslope. Although individual bedding
orientations will vary widely within any given MTD, we propose that the transport
direction can be related to the mean strike of bedding (mean bedding strike method;
MBSM). This method is analogous to the MAM because the transport direction is
taken to be parallel to the mean strike for MTDs with alongslope fold axes. Ideally, the
mean bedding strike will also be sub-parallel with the mean fold axis and axial plane
strike.
The MBSM can be applied by measuring a sufficient number of randomlyselected bedding attitudes within a MTD and calculating the mean bedding
orientation. We use a modification of this approach that utilizes the well-exposed
wave-cut platforms along the study area (Fig. 11). Because the wave-cut platform
surface is nearly horizontal, bedding traces are close approximates for bedding strike
(Fig. 11). Bedding traces are outlined on georeferenced aerial photos of the wave-cut
platforms (Fig. 11), and strike is plotted on a bi-directional, length-weighted rose
diagram from which the mean vector and associated uncertainty are calculated (Fig.
20). In addition, we measured the orientation of the axial surface traces for large folds
visible on the wave-cut platforms (Fig. 11). In general, the orientations of fold axial
surfaces show a close correspondence with the mean bedding strike, with the
exception of subregion R6 (Table 2). This approach has the advantage of including a
large number of measurements without having to collect numerous individual
bedding attitudes in the field.
A disadvantage to our approach is that any subsequent tilting of the MTD will
change the orientation of bedding planes and cause flat-lying beds to strike in one
direction on a wave-cut platform. For example, the NAMTD has been gently tilted to

127

the west ( 10), and originally flat-lying beds now strike in a N-S direction (Fig. 11).
Bedding orientations cannot be easily corrected for post-emplacement tilting or
folding because bedding dip magnitudes are generally not known with our approach.
Because the NAMTD has only been tilted slightly (< 10), we avoid this potential bias
by excluding regions of the wave-cut platform that are characterized by shallowlydipping strata (< ~20).
Interpretation
Bedding strike orientations from the entire NAMTD display a wide range of
attitudes but have a dominant N-S trend when plotted on a length-weighted rose
diagram (Fig. 20). Correspondingly, the MBSM yields an interpreted E-W transport
direction for the entire NAMTD (81-261; Fig. 20; Table 2). However, as with fold
and reverse fault orientations, bedding strike distributions change systematically
between subregions (Fig. 20). Bedding in the far north (R1) and south (R7) is E-W
striking (transport towards 4-14or 184-194), and bedding in the central portion of
the study area (R2-R6) ranges from NE-SW to SE-NW striking (transport towards 62112 or 242-292; Fig. 20). The MBSM shows close agreement with the results from
fold-based methods in most subregions, and has a comparatively smaller margin of
error (Table 2). In addition, the mean bedding strike closely agrees with the mean fold
axial surface orientation for subregions excepting R6 (Table 2).
Comparison between Paleoslope Methods
Rather than relying on one or two methods to infer the MTD transport
direction, we conduct a rigorous comparison between the results of all applicable
paleoslope methods for the entire NAMTD and for each of the seven subregions (Fig.
30). This approach has the advantage of allowing a more realistic assessment of
uncertainty about the inferred paleoslope orientation based on the overlap between
multiple, independent paleoslope estimations (e.g., Debacker et al., 2009). For
example, the overlap of fold- and reverse fault-based results in subregions R1-4 and
R7 provides increased confidence to the paleoslope interpretation in these regions
(Fig. 30). Regions with non-overlapping results (e.g., subregion R6) should be

128

interpreted with caution (Fig. 30). Several important conclusions can be drawn from
comparing the results of each paleoslope method applied to the NAMTD (Fig. 30):
1. Our paleoslope analysis is inconclusive for the NAMTD as a whole due to
non-overlapping results of the few methods that are applicable (Fig. 30). Normal faultbased methods suggest NE-directed transport, but this direction is inconsistent with
fold-based methods and the MBSM (Fig. 30).
2. Although analysis of the entire MTD yields an ambiguous result, subregions
R1-4 and R7 have overlapping results that allow a defensible paleoslope interpretation
to be made in each of these regions (Fig. 30). In particular, methods based on folds,
reverse faults, and bedding all show good agreement (Fig. 30). However, the
interpreted transport direction varies by up to 90 between these subregions (e.g., R1
versus R3; Fig. 30). This change can be readily observed in the wave-cut platforms in
the northern portion of the NAMTD where E-W trending folds transition rapidly into
NE-SW and finally NW-SE trending folds (Figs. 11A-B).
3. The transport direction can only be constrained to one of two possible
directions that are 180 apart in subregions R3-5 and R7 (Fig. 30). This ambiguity is
primarily a consequence of both folds and reverse faults having opposing vergence
senses in nearly equal abundances except for subregions R1 and R2 that only have Sand W- or SW-verging folds, respectively (Figs. 15F-G, 20, and 30).
4. Subregions R5 and R6 show little overlap between paleoslope methods
suggesting that the transport direction is poorly constrained in these regions. Although
the cause of ambiguity in subregions R5 and R6 is unclear, we suspect that the wide
spread in fold and fault orientations relates to the characteristically intense
deformation that LF2 has undergone within the central portions of the NAMTD. For
example, past workers have noted that fold hinge orientations become increasingly
scattered with increasing deformation as a result of several factors that include the
development of multiple generations of cross-cutting folds, folding and faulting in
response to locally-generated stresses rather than to the paleoslope gradient, and
variable fold rotation towards the transport direction (Alsop and Marco, 2013). The
difficulty in interpreting folds and faults in subregions R5-6 relative to other areas

129

within the NAMTD suggests that some portions of the MTD are better suited for
paleoslope analysis than others. These differences likely relate to deformation
intensity that in turn is controlled by the stratigraphic character of the deformed strata
or possibly the relative position within the MTD.
5. Our paleoslope analysis suggests that some methods are more applicable and
yielded more consistent results than others (Fig. 30). For example, the APIM-N does
not overlap with other fold-based methods and probably does not yield an accurate
estimate of the downslope direction (Fig. 30). In addition, methods that assume the
mean fold axis is parallel to the downslope direction (e.g., SAM, DAM) are generally
not applicable or yield ambiguous results. Fold axes are correspondingly interpreted to
have an alongslope mean in at least subregions R1-4 and R7. This conclusion is
supported by the observation that thrust faults verge sub-orthogonally to the trend of
the mean fold axes (Fig. 30). In addition, there is no compelling evidence for
significant downslope rotation of fold axes (e.g., development of sheath folds or a
systematic relationship between fold hinge azimuth and interlimb angle; Fig. 28). For
these reasons, methods that rely on progressive layer-parallel simple shear during
deformation (i.e., HIM, HAM, and ASIM) were found to be ineffective. It is worth
noting that the MBSM yielded the most constrained results for all subregions and was
in good agreement with most other methods (Figs. 20 and 30). Whereas fold- and
fault-based methods were ambiguous in subregions R5-6, the MBSM yields a
generally E- or W-directed transport that is consistent with adjacent subregions R2-4
(Figs. 20 and 30).
6. Results from normal fault-based methods rarely overlap with those based on
folds, reverse faults, and bedding (Fig. 30). Although normal faults are typically
assumed to strike on average in a direction parallel to the paleoslope (Farrell, 1984;
Martinsen and Bakken, 1990), normal faults within the NAMTD have orientations that
are markedly oblique to the inferred transport direction based on the other methods
(Fig. 26). It is possible that some normal faults formed early and were subsequently
overprinted by subsequent deformation that altered their original orientations.
Alternatively, some normal faults could have formed in response to local stresses

130

within the deforming MTD. For example, Alsop and Marco (2011) documented
conjugate sets of normal faults that were oriented obliquely to the presumed transport
direction and suggested that these faults formed in response to lateral spreading along
the width of the slump sheet. In such cases, interpretation of normal fault orientations
alone is likely to yield an ambiguous or erroneous estimate of the paleoslope
orientation. We instead rely on the consistency between folds, reverse faults, and
bedding orientations to aid in interpreting the overall transport direction of the
NAMTD (Fig. 30).
DISCUSSION
Transport Direction of the NAMTD
Kinematic analysis of slump folds and faults within the NAMTD reveals that
the predicted transport direction varies significantly from north to south within the
MTD outcrop extent (Fig. 30). Moreover, the variability in transport direction appears
to be systematic: north or south transport directions are predicted along the inferred
lateral margins of the MTD (subregions R1 and R7) and generally east or west
transport directions are predicted for the central portions of the MTD (subregions R2-4
and possibly R5; Fig. 30). In general, these observations can be explained by two endmember scenarios: (1) the NAMTD is composed of several separate MTDs that have
different kinematic histories, or (2) the NAMTD is one large MTD that is
characterized by spatially heterogeneous deformation structure orientations.
The first scenario may be viable in the south due to the complex nature of
deformation and occasional breaks in outcrop continuity that make it difficult to
conclude whether the NAMTD represents a single, large mass-movement event or if it
is actually a composite of two or more juxtaposed events. However, field observations
suggest that the first scenario is not tenable in the northern portions of the MTD (Figs.
6 and 9AD). S-verging folds north of Kopuai Stream (subregion R1) can be directly
traced into W-verging folds to the south (subregion R2) with no break in outcrop
continuity (Figs. 9A-B). These same beds can then be correlated across a short break
(~200 m) in outcrop to NW- and SE-verging folds in subregion R3 and finally to W-

131

and E-verging folds in subregion R4 (Figs. 6, 9A-D, and 11A-B). The continuity and
repeated deformation of the same stratigraphic interval throughout subregions R1-4
strongly suggests that the northern 4 km of the NAMTD was emplaced during a single
mass-movement event. Thus the dramatic north to south differences in fold and fault
orientations must be related to spatial heterogeneity within the deforming MTD rather
than separate, juxtaposed MTDs. Moreover, only subregions within the inferred lateral
portions of the MTD (subregions R1 and R7) display fold, fault, and bedding
orientations that are systematically oriented in an E-W direction (Figs. 20, 26, and 30).
Thus, we contend that the orientations of slump deformation features in subregions R1
and R7 most likely diverge from the rest of the NAMTD as a result of edge effects
near the lateral portions of a single, large MTD (Fig. 31).
The recognition of distinct central and lateral portions of the MTD greatly
clarifies interpretation of the overall transport direction. Because portions of the MTD
inferred to be in a central position generally yield E- or W-directed transport (67-123
or 247-303), we infer that the MTD was likely emplaced along a N-S striking
paleoslope (Figs. 30 and 31). Determining whether the MTD traveled towards the west
or east is somewhat ambiguous because most paleoslope methods are only able to
constrain MTD transport to one of two possible directions (Fig. 30). However, several
lines of evidence suggest that the NAMTD was most likely emplaced in a westerly
direction.
1. Both bedding strike and fold orientations can be observed to fan ~35 degrees
about a westerly direction in the wave-cut platforms on either side of Waioroko
Stream (subregions R3-4; Figs. 11A-B and 30). This fanning pattern is highly
reminiscent of the typically arcuate pattern of pressure ridges found nearly
ubiquitously on MTD surfaces on the modern seafloor and in the subsurface (Fig. 32).
Because this fanning pattern occurs over a distance of only 1.5-2 km, this region likely
represents a subordinate slump lobe within the larger MTD. We interpret the
westward convexity of this slump lobe to be a strong kinematic indicator of Wdirected translation in subregions R3-4 (Figs. 11A-B).

132

2. Only W- or SW-verging structures are present in subregion R2, although


relatively few fold measurements are available (Fig. 20). Two large, nearly recumbent
folds in this region have fold axes that curve towards the west, suggesting propagation
in this direction (Fig. 15A).
3. Thrust faults in subregion R1 verge slightly obliquely (~25 clockwise) to
the large slump folds in this area (Fig. 30). This westward obliquity is consistent with
an overall W-directed transport direction.
4. The nearby offshore presence of the NAMTD is equivocal due to
problematic data quality in the Kahu 3D seismic survey, but it does not appear to
project westward from ~12 km west of the coastline at least, where the Mohakatino
interval is well imaged (Figs. 3 and 12). Had the MTD originated in the west, we
might expect to see evidence for either the headwall scarp or deposits of this MTD in
the area covered by the Kahu 3D survey.
5. Angular relationships between the Mohakatino Formation and the
unconformably underlying Manganui-Mokau Group (undifferentiated) suggests that
westward tilting occurred between middle and late Miocene time (Utley, 1987; his
Fig. 6.3). West-directed translation of the NAMTD is consistent with the continuation
of W-directed tilting into the early part of late Miocene time (see discussion below).
6. Other MTDs present in the northeastern Taranaki Basin within overlying
stratigraphic intervals (e.g., the Otukehu MTD, lower Mount Messenger Formation)
have inferred W- to N-directed transport directions based upon both deformation
structure orientations and kinematic indicators observed in offshore seismic surveys
(Strachan, 2002; King et al., 2011; Fig. 4).
Why Are Folds and Faults within the MTD Margins Oriented Parallel to the
Inferred Downslope Direction?
Slump deformation structures within the lateral margins of the MTD
(subregions R1 and R7) have an average orientation that is sub-parallel to the inferred
downslope dip direction, based on an assumed overall W-directed MTD transport
direction (Figs. 30 and 31). Thus, the NAMTD provides important verification that

133

downslope-parallel deformation structures occur in large MTDs inasmuch as


published models of margin-to-center structural relationships within MTDs have been
largely based on terrestrial analogues, including a small tundra landslide (Hansen,
1971) and a slumped snow bed (Lajoie, 1972). We propose that the downslopeparallel deformation structures within the margins of the NAMTD could have formed
as a result of two end-member mechanisms that each can be accommodated via
several interrelated processes: (1) Rotation of originally alongslope structures via (a)
downslope rotation during progressive layer-parallel shear or (b) bulk vertical-axis
rotation, and/or (2) in situ formation via (a) layer-normal shear, (b) shortening parallel
to the paleoslope strike, or (c) variable slope bathymetry.
Rotation of Deformation Structures into a Downslope-Parallel Orientation
Downslope rotation of originally alongslope deformation structures can occur
via progressive layer-parallel simple shear (e.g., Farrell and Eaton, 1987) and/or bulk
vertical-axis rotation. The former process seems unlikely to explain patterns of
deformation within the NAMTD because folds within the marginal portions
(subregions R1 and R7) show no evidence for progressive rotation with increasing
deformation (e.g., lack of sheath folds and no systematic relationship between hinge
orientation and interlimb angle; Fig. 28). However, bulk vertical-axis rotation at the
NAMTD margins could conceivably account for oblique- or downslope-oriented
deformation structures within subregions R1 and R7. This process likely contributes to
the typically arcuate pattern of downslope-fanning pressure ridges that are commonly
observed within MTDs that undergo greater amounts of material translation in their
central portions than in their margins (e.g., Fig. 31). Approximately 90 of clockwise
and counter-clockwise rotation would have to have occurred along the margins of the
NAMTD within subregions R1 and R7, respectively, to account for slump fold and
fault orientations (Figs. 20 and 26). Although the extent of vertical-axis rotation is
poorly constrained along the southern extent of the NAMTD, several lines of evidence
suggest that such large amounts of vertical-axis rotation did not likely occur along the
northern margin of the NAMTD.

134

1. Flute casts on the base of a correlated sand bed in subregions R1 and R3 all
have flow towards the eastern quadrant, suggesting limited vertical-axis rotation
between these subregions (Fig. 5). The observed ~60 divergence within flute casts
measured at sections 2 and 5 could be interpreted as counter-clockwise rotation of
subregion R1 relative to R3. However, this interpretation is inconsistent with a
preferred W-directed overall transport direction that would suggest that subregion R1
was rotated clockwise relative to R3. Even if the MTD was emplaced towards the east,
the 60 flute divergence is not great enough to account for the 110 of counterclockwise rotation required by fold orientations (Fig. 20).
2. Folds within subregion R1 verge toward the south and not towards the north,
as expected if originally west-verging folds had been rotated clockwise to become
downslope-parallel at the northern margin of the NAMTD.
3. The transition from downslope- to alongslope-parallel fold axes occurs
abruptly south of Kopuai Stream (Fig. 11). The absence of structures with transitional
orientations seems to be at odds with gradual rotation to a downslope orientation at the
northern margin of the NAMTD (Fig. 31).
4. Large scale rotation of subregion R1 relative to R2-3 would likely create a
space problem that is not observed in outcrop. As described above, the same
stratigraphic interval can be traced from S- to W-verging folds without any apparent
suggestion of rotation.
In Situ Formation of Downslope-Parallel Deformation Structures
Past workers have suggested that slump folds can form with orientations at low
angles to the downslope direction (e.g., Debacker et al,. 2009). In particular, folds
formed within the margins of MTDs can undergo layer-normal shear that is predicted
to result in steeply-dipping axial surfaces and fold hinges that fan about the downslope
direction (Alsop and Holdsworth, 2002; Debacker et al., 2009; Alsop and Marco,
2011). However, it is questionable whether the slump deformation structures within
the margins of the NAMTD were formed by layer-normal shear (Table 3). For
example, fold distributions within subregions R1 and R7 display a positive

135

relationship between interlimb angle and axial surface dip (Fig. 28), opposite to the
relationship predicted to occur as a result of layer-normal shear (Debacker et al., 2009;
Table 3). Large slump folds within subregion R1 show evidence for layer-parallel
simple shear (e.g., upwards-convex axial surfaces; Fig. 23) and display similar
geometries as folds within subregions R3-4 (Fig. 22). Thus it seems unlikely that
downslope-parallel folds within subregion R1 and possibly R7 resulted from layernormal shear.
Alternatively, downslope-parallel deformation structures could have formed as
a response to shortening across the width of the MTD (i.e., lateral compression) and/or
variably sloping seafloor bathymetry. These processes could have acted in concert as
changes in the seafloor gradient to the north and south of the NAMTD may have
resulted in local confinement and inward-directed compression along the margins of
the MTD (Fig. 31). We favor this interpretation for several reasons. (1) This scenario
is consistent with the abrupt transition in fold orientations and bedding strike between
subregions R1-3 (e.g., Figs. 11A-B) and (2) may also help explain why the large folds
present in subregion R1 are all verging south towards the interior of the MTD and not
towards the north as one would expect based on clockwise vertical-axis rotation of
originally W-verging structures (Fig. 9). (3) Confinement of the northern margin of
the NAMTD by a S-dipping paleoslope is supported by paleocurrent measurements
collected at Opito Point and Ngarupupu Point (Fig. 2; L. Shumaker, personal
communication). (4) Comparable examples of lateral compression at the margins of
MTDs have been identified in seismic-reflection datasets (e.g., Posamentier and
Walker, 2006; their Figs. 152 and 160).
In our preferred model, the NAMTD was emplaced towards the west (~245300) onto a low-relief, ponded basin floor (Figs. 30 and 31). Lateral compression
along the MTD margins resulted from interaction of the expanding sediment mass
with basin confinement to the north and possibly to the south (Fig. 31). Downslopeparallel slump deformation structures were thereby formed in response to lateral
shortening that was focused within the marginal portions of the MTD (Fig. 31).

136

Spatial Variability in MTD Structural Architecture


As with deformation structure orientations, the internal structural architecture
of the NAMTD varies widely along its outcrop extent (e.g., Figs. 9, 15, and 17).
However, there is a less clearly-defined relationship between spatial position within
the margin-to-center transect and the style of MTD deformation. Rather, the lithologic
character of the host stratigraphy likely played a major role in influencing the type and
scale of slump deformation features. In general, regions dominated by LF1
(subregions R1-4) tend to form the largest and most continuous folds found in the
NAMTD (Figs. 15 and 17). Deformation structures formed in LF2 are marked by an
increase in the intensity and complexity of deformation that includes generation of
typically smaller-scale rootless, disharmonic folds that are intercalated with low-strain
zones. Thus, we hypothesize that the dramatic change in MTD structural architecture
that occurs south of subregions R3 is primarily a result of the decrease in LF1 relative
to LF2 within subregions R4-5 (Fig. 17).
The relative abundances of normal and reverse slump faults also changes
markedly from north to south. Normal faults are comparatively uncommon in
subregions R1-3 relative to reverse faults (15 versus 60 normal and reverse fault
measurements, respectively; Fig. 26). The opposite is true for subregions R4-7 (129
normal vs 48 reverse measurements; Fig. 26). The scarcity of normal faults in
subregions R1-3, except where developed around fold hinges, suggests that this
portion of the MTD did not undergo significant extension prior to evolving to a
compressional regime (Farrell, 1984).
Implications for Paleoslope Analysis
The NAMTD demonstrates that significant spatial heterogeneity in
deformation structure orientations and internal architecture can greatly complicate
determination of the paleoslope orientation via kinematic analysis. This point is
illustrated by the ambiguity that results from conducting a paleoslope analysis of all
folds and faults without regard for their relative location within the NAMTD (Fig. 30).
We found that a robust paleoslope interpretation was only made possible by

137

accounting for spatial changes in fold and fault orientations along the MTD exposure
(Fig. 30). This finding may account for the conflicting results yielded by previous
estimates of the MTD transport direction that did not account for spatial variability
along the outcrop extent (E- and S-directed transport; Utley, 1987; Nagel, 2010). Thus
the NAMTD highlights the importance of interpreting slump deformation structures
within their spatial context and demonstrates that failure to do so can yield an
erroneous paleoslope interpretation.
Our analysis also highlights the challenges of interpreting slump deformation
structures characterized by widely dispersed fold and fault orientations and senses of
asymmetry. In some cases, scattered fold and fault orientations may be a consequence
of complex and intense patterns of deformation (e.g., subregion R6; Alsop and Marco,
2013). Alternatively, such distributions may result from failure to account for
systematic spatial variations in slump structure orientations. In such cases, we strongly
recommend that the dataset be examined for systematic spatial variability, and fold
and fault measurements be interpreted within their spatial context.
In the case of the NAMTD, discrimination between the lateral and central
portions of the MTD is critical for interpreting the overall transport direction and is
largely made possible by regional geologic constraints. Unfortunately, recognition of
MTD lateral margins in ancient outcrops is typically hampered by a lack of exposure
and by subsequent tectonic deformation. Although criteria for recognizing lateral
margins of MTDs based on slump deformation structures have been proposed (e.g.,
Debacker et al., 2009), many of these criteria are not fulfilled in the regions that we
infer to be within the margins of the NAMTD (Table 3). We suspect that this
incongruity is a result of lateral compression along the margins of the MTD, rather
than deformation via layer-normal shear. Thus the NAMTD demonstrates an
important methodological limitation of some paleoslope methods that generally
assume little shortening across the width of the MTD (Woodcock, 1979b) and
suggests that it can be difficult to confidently discriminate between lateral and central
portions of MTDs based upon the characteristics of slump deformation features alone,
particularly when MTDs undergo lateral compression due to basin confinement.

138

Paleogeography of the Northeastern Taranaki Basin


Our study of the NAMTD documents evidence for a spectacular episode of
mass-movement in the northeastern Taranaki Basin during late Miocene time and in
equal part corroborates, clarifies, and provides new insights into the previously
inferred paleogeographic setting and tectonic evolution of the region (e.g., Nodder et
al., 1990b; King et al., 1993). The pre-deformation sheet-like geometry and lateral
continuity of deep-water Mohakatino Formation beds exposed along the coast between
Opito Point and the Awakino River suggest that deposition by sediment gravity flows
was primary within an unconfined basin-floor setting. However at a slightly broader
scale, this depocenter was likely surrounded on all sides by bathymetric highs (Fig.
33). The submarine volcanic arc that was the source of Mohakatino detritus formed
seafloor topography to the west and north (Figs. 1 and 2). The Herangi submarine high
likely provided an eastern flank to the sub-basin and a backstop for Mohakatino fan
deposition (King et al., 1993). Containment to the south is suggested by the pinch-out
of coarse-grained, thick-bedded lobes (LF1) in this direction and the inferred location
of the late Miocene slope that was actively prograding northward during and following
Mohakatino Formation deposition (Masalimova, 2013; Rotzien et al., 2014). A
depositional position near the sub-basin center is supported by variable paleocurrent
directions within the Mohakatino Formation (primarily W-, S-, and E-flowing; Figs. 57) and local interfingering of the Mount Messenger and Mohakatino Formations that
suggest flows entered the basin from a variety of directions and with varying
provenance (Figs. 5-7 and 33; Utley, 1987; Nodder et al., 1990a, b). Although
depositional facies of the Mohakatino Formation suggest a basin-floor setting, the
ultimate base of slope lay far to the west within the New Caledonian Basin (Fig. 1;
Uruski and Wood, 1991). Thus the NAMTD was likely positioned within a local basin
that was perched, or possibly ponded, within the western continental slope of New
Zealand (sensu Prather 2003; Prather et al., 2012; Fig. 33).
A surprising finding of our study is that the NAMTD was likely translated
back towards the submarine volcanic arc that was the original source of sediment

139

(Figs. 1 and 2). Although the transport distance of the NAMTD is uncertain, the
distribution of lithofacies within the NAMTD closely matches that of the surrounding,
undeformed intervals. This strongly suggests that the NAMTD was generated locally
and probably did not travel very far (Fig. 33). Our interpreted W-directed MTD
transport direction suggests that a local reversal of the seafloor slope gradient occurred
following deposition of the E- or SE-flowing LF1 lithofacies of the Mohakatino
Formation (Fig. 5). In addition, SE-directed flutes measured just to the north of the
NAMTD at Opito Point confirm that the original paleoslope was east-facing at the
time of Mohakatino Formation deposition (Fig. 8). It is likely that westward tilting of
the basin occurred even prior to deposition of the Mohakatino Formation based on
local observations of a subtle angular relationship between middle Miocene bathyal
mudstone and the overlying basal Mohakatino Formation (Utley, 1987). Within the
past few million years strata within the entire northern Taranaki coastal region have
been progressively uplifted and tilted to the west or southwest within a regional
monocline (King and Thrasher, 1996). Taken together, these observations suggest that
the Taranaki Miocene sedimentary succession underwent a long history of regional
westward tilting, both during and after deposition of the Mohakatino Formation.
Although the precise cause of local slope reversal within the northeastern
Taranaki Basin is unclear, it is likely that tectonically-induced basin deformation
played a major role. We speculate that the eastern basin margin (the Herangi
submarine high) was uplifted during late Miocene time prior to ca. 10.5-10.9 Ma
(Appendix C-7) and that this tectonism resulted in W-directed tilting and an associated
slight reconfiguration of the sub-basin depocenter (Fig. 33). A possible driver for
uplift of the eastern sub-basin margin is activity on the buried Taranaki fault that
nominally forms the eastern boundary of the Taranaki Basin (Figs. 1 and 2). This
interpretation is supported by the location of the NAMTD atop the leading edge of the
Taranaki fault hanging wall (Fig. 2). However, major displacement on the Taranaki
fault in this vicinity is thought to predate ~18-19 Ma (Stagpoole and Nicol, 2008), and
thus any continued activity during late Miocene time was probably relatively minor.

140

Alternatively, disruption of the paleoslope may have been triggered by the coeval
active volcanism and associated normal faulting that was occurring to the west
(Figs. 2 and 33; Giba et al., 2013).
Mass-Movement within Basin-Floor Successions
The NAMTD provides an important example of mass-movement within a
stratal succession originally deposited within a low-gradient basin floor, possibly in a
perched or ponded basin. Mass failure in this setting is surprising because MTDs most
commonly originate in outer shelf, upper slope, and mid-slope settings where seafloor
gradients are typically higher (Posamentier and Martinsen, 2011). Correspondingly,
documented examples of mass-movement within basin-floor facies are somewhat
uncommon and have generally been ascribed to one of two processes: (1) basin
deformation that alters local bathymetric gradients and produces slope instabilities in
basin-floor facies (Trincardi and Argnani, 1990; Haughton, 2000; Lucente and Pini,
2008) and (2) entrainment and/or deformation of basin-floor facies within a MTD that
originated higher in the slope profile (van der Merwe et al., 2009; Posamentier and
Martinsen, 2011). We disfavor the latter option because the NAMTD is comprised
exclusively of basin-floor facies and lacks evidence for strata derived from slope or
shelf settings (e.g., van der Merwe et al, 2009).
We contend that the NAMTD provides a type example of mass-movement of a
basin-floor succession related to tectonically-induced basin deformation. Basin tilting
created a slightly steeper and de-stabilized slope gradient which caused sand-rich,
basin-floor lobes of the Mohakatino Formation to be remobilized downslope to form
the NAMTD (Fig. 33). The Oligocene-Miocene Apennine foredeep provides a
comparable example where basin margin uplift created a temporary slope from
which basin-floor facies were remobilized orthogonally to the regional paleoflow
direction (Lucente and Pini, 2008). Because of the inherently low slope gradients in
basin-floor settings any tectonic movements can readily change the seafloor slope
orientation during basin-floor deformation. Correspondingly, it is common for MTDs
developed within basin-floor successions to have transport directions that are quite

141

different from the regional paleoflow direction (e.g., Miyata, 1990; Haughton, 2000;
Lucente and Pini, 2008). This scenario helps account for the coarse-grained sand-rich
nature of the NAMTD that reflects the lithologic character of the staging area that
included proximal portions of coarse-grained lobe deposits. This example suggests
that MTDs that comprise basin-floor lobes can be sand-rich relative to those that
originate on the continental slope where facies are more prone to be mud-dominated
(Posamentier and Martinsen, 2011).
CONCLUSIONS
The NAMTD is spectacularly exposed in coastal cliffs and wave-cut platforms,
provides a world-class example of mass-movement in a deepwater basin-floor setting,
and exemplifies the scale, style, and heterogeneity that is possible within this type of
deposit. Moreover, the fortuitous coincidence of W-directed translation with N-S
oriented coastal exposures has resulted in a lateral cross-section that extends across
nearly the full width of the compressional domain of the NAMTD. This provides an
exceptional, and possibly unique, opportunity to study the spatial variability in
deformation structure orientations and internal architecture from both the marginal and
central portions of a MTD greater than 10 km in width.
The NAMTD developed in strata deposited as sheets or lobes within a perched
or ponded intra-slope basin along the western paleo-continental margin of New
Zealand. Sedimentary lithofacies within the NAMTD can be categorized into two endmembers that are interpreted to have been deposited within both inner lobe and outer
lobe/lobe fringe environments: thick- to medium-bedded volcaniclastic sandstone
(LF1) and thin-bedded volcaniclastic sandstone and mudstone (LF2), respectively.
LF1 predominates in the northern 5 km of the study area but transitions gradually to
LF2 towards the south. These lithofacies correspond with distinctly different styles of
soft-sediment deformation. Slump folds developed in LF1 are typically larger
(amplitudes up to 10s of m), often form semi-continuous fold trains, and contain
dcollement zones within the fold hinges. Slump folds developed in LF2 are typically

142

smaller, disharmonic, and occur as rootless individuals. Intensely deformed zones are
often intercalated vertically or horizontally with low-strain domains.
Kinematic analysis of slump folds and faults compiled for the entire NAMTD
yields an ambiguous interpretation of the transport direction. However, a clearer
picture has emerged through separate analysis of seven subregions within the NAMTD
that allows a robust interpretation of the paleoslope orientation to be made for five of
the subregions that display overlap of multiple paleoslope estimations. Populations of
both upslope- and downslope-verging folds and reverse faults are present throughout
much of the NAMTD, and the paleoslope dip direction can only be constrained to two
possibilities in these regions. By discriminating and accounting for these spatial
changes in slump structure orientations and considering regional geologic constraints,
we demonstrate that the NAMTD was likely emplaced within an overall W-facing
paleoslope. Moreover, slump structures within the inferred margins of the MTD have
orientations that are sub-perpendicular to those in the central portions of the MTD.
This relationship is best expressed in the northern extent of the MTD where the same
stratigraphic interval is repeatedly deformed in a series of spectacular folds with
orientations that systematically change by ~110 over 1.5 km. Thus the NAMTD
verifies that downslope-parallel fold axes can occur within the lateral portions of
MTDs, possibly as a result of lateral compression along the margins. Our analysis
demonstrates the importance of studying slump structures in their spatial context and
using multiple, independent methods of estimating the paleoslope orientation.
The submarine Mohakatino volcanic arc is the acknowledged source of
volcaniclastic sediment to the sub-basin; the inferred presence of an E- or SE-facing
paleoslope formed on the volcanic flanks is supported by paleocurrent measurements
within and adjacent to the MTD. Our interpretation of W-directed transport of the
NAMTD, back towards the volcanic arc, suggests that a local reversal in the
orientation of the bathymetric slope must have occurred prior to emplacement of the
MTD. We speculate that this occurred via local basin tilting that may have resulted
from uplift of the Herangi submarine high to the east. Possible tectonic drivers of
basin deformation include continued activity on the buried Taranaki fault and/or

143

differential subsidence associated with active volcanism and normal faulting to the
west. Such tectonism helps account for the development of a large MTD in sand-rich
strata that were originally deposited in a likely ponded sub-basin with very low
seafloor gradients. The NAMTD may provide an analog for other unconfined,
deepwater deposits in intra-slope basins that have undergone mass-movement
associated with basin deformation via tectonic or gravitational processes (e.g., salt or
mobile shale movement).
ACKNOWLEDGEMENTS
Financial support for this study was provided by a Geological Society of
America graduate research grant and by the Stanford Project on Deep-water
Depositional Systems. We also thank the individuals from GNS Science who
contributed to this study, including Greg Browne, Rob Funnell, Malcolm Arnot,
Martin Crundwell, and Andy Nicol. This study benefited from discussion with Tim
Debacker and Lorna Strachan. We thank Blair Burgreen, Tess Menotti, and Nora
Nieminski for assistance in the field. We also extend a sincere thanks to the many
New Zealanders who made this project possible including Craig Rain at Paparahia
Station; Alistair and Maree Bryant at Onetai Station; Karl Reipen at Awakino Estate;
Dawn and Neil Colman; John and Angela Potroz; Shane, Jenny, and Graham
Marsden; and Alan Jones and his family.
REFERENCES
Alsop, G.I., and Holdsworth, R.E., 2002, The geometry and kinematics of flow
perturbation folds: Tectonophysics, v. 350, p. 99-125.
Alsop, G.I., and Marco, S., 2011, Soft-sediment deformation within seismogenic
slumps of the Dead Sea Basin: Journal of Structural Geology, v. 33, p. 433457.
Alsop, G.I., and Marco, S., 2013, Seismogenic slump folds formed by gravity-driven
tectonics down a negligible subaqueous slope: Tectonophysics, v. 605, p. 4869.
Armitage, D.C., Romans, B.W., Covault, J.A., and Graham, S.A., 2009, The influence
of mass-transport-deposit surface topography on the evolution of turbidite
architecture: The Sierra Contreras, Tres Pasos Formation (Cretaceous),
Southern Chile: Journal of Sedimentary Research, v. 79, p. 287-301.
144

Bache, F., Sutherland, R., Stagpoole, V.M., Herzer, R.H., Collot, J., and Rouillard, P.,
2012, Stratigraphy of the southern Norfolk Ridge and the Reinga Basin: a
record of initiation of Tonga-Kermadec-Northland subduction in the southwest
Pacific: Earth and Planetary Science Letters, v. 321-322, p. 41-53.
Bell, A.M., 1981, Vergence: an evaluation: Journal of Structural Geology, v. 3, p. 197202.
Bouma, A.H., 1962, Sedimentology of some flysch deposits: Elsevier, Amsterdam,
168 p.
Bradley, D., and Hanson, L., 1998, Paleoslope analysis of slump folds in the Devonian
flysch of Maine: The Journal of Geology, v. 106, p. 305-318.
Browne, G.H., and Slatt, R.M., 2002, Outcrop and behind-outcrop characterization of
a Late Miocene Slope Fan (Channel-Levee Complex), Mt Messenger
Formation, New Zealand: American Association of Petroleum Geologists
Bulletin, v. 86, p. 841-862.
Browne, G.H., Slatt, R.M., and King, P.R., 2000, Contrasting styles of basin floor fan
and slope fan deposition: Mount Messenger Formation, New Zealand, in
Bouma, A. H., and Stone, C.G., eds., Fine-Grained Turbidite Systems, AAPG
Memoir v. 72/SEPM Special Publication v. 68, p.143-152.
Bull, S., Cartwright, J., and Huuse, M., 2009, A review of kinematic indicators from
mass-transport complexes using 3D seismic data: Marine and Petroleum
Geology, v. 26, p. 1132-1151.
Childs, C., Manzocchi, T., Walsh, J.J., Bonson, C.G., Nicol, A., and Schoepfer,
M.P.J., 2009, A geometric model of fault zone and fault rock thickness
variations: Journal of Structural Geology, v. 31, p. 117-127.
Childs, C., Walsh, J.J., Manzocchi, T., Strand, J., Nicol, A., Tomasso, M., Schoepfer,
M.P.J., Aplin, A.C., 2007, Definition of a fault permeability predictor from
outcrop studies of a faulted turbidite sequence, Taranaki, New Zealand, in
Jolley, S.J., Barr, D., Walsh, J.J., and Knipe, R.J., eds., Structurally Complex
Reservoirs, Geological Society, London, Special Publication 292, p. 235-258.
Coward, M.P., and Kim, J.H., 1981, Strain within thrust sheets, in, McClay, K.R., and
Price, N.J., eds., Thrust and Nappe Tectonics, Geological Society, London,
Special Publications, v. 9, p. 275-292.
Coward, M.P., and Potts, G.J., 1983, Complex strain patterns developed at the frontal
and lateral tips to shear zones and thrust zones: Journal of Structural Geology,
v. 5, p. 383-399.
Dasgupta, P., 2008, Experimental decipherment of the soft-sediment deformation
observed in the upper part of the Talchir Formation (Lower Permian), Jharia
Basin, India: Sedimentary Geology, v. 205, p. 100-110.
Davis, G.H., Reynolds, S.J., and Kluth, C.F., 2012, Structural Geology of Rocks and
Regions: John Wiley & Sons, Inc., Hoboken, New Jersey, Third Edition, 839
p.
Debacker, T.N., Dumon, M., and Matthys, A., 2009, Interpreting fold and fault
geometries from within the lateral to oblique parts of slumps: Journal of
Structural Geology, v. 31, p. 1525-1539.

145

Donath, F.A., and Parker, R.B., 1964, Folds and folding: Geological Society of
America Bulletin, v. 58, p. 54-72.
Edbrooke, S.W., 2005, Geology of the Waikato area: Institute of Geological and
Nuclear Sciences, Lower Hutt, New Zealand, 1:250,000 geological map + 68
p.
Farrell, S.G., 1984, A dislocation model applied to slump structures, Ainsa Basin,
South Central Pyrenees: Journal of Structural Geology, v. 6, p. 727-736.
Farrell, S.G., and Eaton, S., 1987, Slump strain in the Tertiary of Cyprus and the
Spanish Pyrenees. Definition of palaeoslopes and models of soft-sediment
deformation, in, Jones, M.E., and Preston, R.M.F., eds., Deformation of
Sediments and Sedimentary Rocks. Geological Society, London, Special
Publications, v. 29, p. 181-196.
Farrell, S.G., and Eaton, S., 1988, Foliations developed during slump deformation of
Miocene marine sediments, Cyprus: Journal of Structural Geology, v. 10, p.
567-576.
Martinsen, J., Cartwright, J., and James, D., 2006, Frontally confined versus frontally
emergent submarine landslides: A 3D seismic characterization: Marine and
Petroleum Geology, v. 23, p. 585-604.
Gawthorpe, R.L., and Clemmey, H., 1985, Geometry of submarine slides in the
Bowland Basin (Dinantian) and their relation to debris flows: Journal of the
Geological Society, London, v. 142, p. 555-565.
Ghosh, B., and Lowe, D.R., 1993, The architecture of deep-water channel complexes,
Cretaceous Venado Sandstone Member, Sacramento Valley, California, in
Graham, S.A., and Lowe, D.R., eds., Advances in the Sedimentary Geology of
the Great Valley Group, Sacramento Valley, California: Pacific Section SEPM,
Guidebook 73, p. 51-65.
Giba, M., Nicol, A., and Walsh, J.J., 2010, Evolution of faulting and volcanism in a
back-arc basin and its implications for subduction processes: Tectonics, v. 29,
TC4020.
Giba, M., Walsh, J.J., Nicol, V., Mouslopoulou, V., and Seebeck, H., 2013,
Investigation of the spatio-temporal relationship between normal faulting and
arc volcanism on million-year time scales: Journal of the Geological Society,
London, v. 170, p. 951-962.
Haflidason, H., Sejrup, H.P., Nygard, A., Minert, J., Bryn, P., Lien, R., Forsberg, C.F.,
Berg, K., and Masson, D., 2004, The Storegga Slide: architecture, geometry
and slide development: Marine Geology, v. 213, p. 201-234.
Hahn, F., 1913, Untermeerische gleitung bei Trenton falls (Nord-America) und ihr
Verhaltnis zu ahnlinchen storungsblidern: Neues Jahrbuch Mineralogie,
Beilage Bd, v. 36, p. 1-41.
Hansen, E., 1965, Methods of deducing slip-line orientations from the geometry of
folds: Year Book of the Carnegy Institute Washington, v. 65, p. 387-405.
Hansen, E., 1971, Strain Facies: Springer-Verlag, New York, 207 p.
Haughton, P.D.W., 2000, Evolving turbidite systems on a deforming basin floor,
Tabernas, SE Spain: Sedimentology, v. 47, p. 497-518.

146

Haughton, P., Davis, C., McCaffrey, W., Barker, S., 2009, Hybrid sediment gravity
flow deposits classification, origin, and significance: Marine and Petroleum
Geology, v. 26, p. 1900-1918.
Henderson, J., and Ongley, M., 1923, The geology of Mokau Subdivision with an
account of adjoining areas and of the Te Kuiti district, Pirongia and Taranaki
Divisions: New Zealand Geological Survey Bulletin, v. 24, 83 p.
Hodgson, D.M., 2009, Distribution and origin of hybrid beds in sand-rich submarine
fans of the Tanqua depocenter, Karoo Basin, South Africa: Marine and
Petroleum Geology, v. 26, p. 1940-1956.
Holt, W.E., and Stern, T.A., 1994, Subduction, platform subsidence, and foreland
thrust loading: the late Tertiary development of Taranaki Basin, New Zealand:
Tectonics, v. 13, p. 1068-1092.
Johnson, S.D., Flint, S., Hinds, D., and Wickens, H.D.V., 2001, Anatomy, geometry
and sequence stratigraphy of basin floor to slope turbidite systems, Tanqua
Karoo, South Africa: Sedimentology, v. 48, p. 987-1023.
Jones, O.T., 1939, The geology of the Colwyn Bay district: a study of submarine
slumping during the Salopian period: Quarterly Journal of the Geological
Society of London, v. 380, p. 335-382.
King, P.R., 2000, Tectonic reconstructions of New Zealand: 40 Ma to present: New
Zealand Journal of Geology and Geophysics, v. 43, p. 611-638.
King, P.R., Browne, G.H., Arnot, M.J., Slatt, R.M., Helle, K., and Stromsoyen, I.,
2007, A 2-D oblique-dip outcrop transect through an entire third-order,
progradational, deep-water clastic succession (Late Miocene Mount
Messenger-Urenui Formations), Taranaki Basin, New Zealand, in Nilsen, T.H.,
Shew, R.D., Steffens, G.S., and Sudlick, J.R.J., eds., Atlas of Deep-Water
Outcrops: American Association of Petroleum Geologists, Studies in Geology
56, p. 238-240.
King, P.R., Ilg, B.R., Arnot, M., Browne, G.H., Strachan, L.J., Crundwell, M., and
Helle, K., 2011, Outcrop and seismic examples of mass-transport deposits
from a Late Miocene deep-water succession, Taranaki Basin, New Zealand, in
Shipp, R.C., Weimer, P., and Posamentier, H.W., eds., Mass-transport deposits
in deepwater settings, Society for Sedimentary Geology (SEPM), Special
Publication 96, p. 311-347.
King, P.R., Scott, G.H., and Robinson, P.H., 1993, Description, correlation and
depositional history of Miocene sediments outcropping along the North
Taranaki coast: Institute of Geological and Nuclear Sciences Monograph 5,
145 p.
King, P.R., and Thrasher, G.P., 1992, Post-Eocene development of the Taranaki Basin,
New Zealand: Convergent overprint of a passive margin, in Watkins, J.S.,
Zhiqiang, F., and McMillen, K.J., eds., Geology and geophysics of continental
margins: American Association of Petroleum Geologists Memoir, v. 53, p. 93118.
King, P.R., and Thrasher, G.P., 1996, Cretaceous-Cenozoic geology and petroleum
systems of the Taranaki Basin, New Zealand: Institute of Geological and
Nuclear Sciences Monograph 13, 244 p.
147

Lajoie, J., 1972, Slump fold axis orientations: an indication of paleoslope?: Journal of
Sedimentary Petrology, v. 42, p. 584-586.
Lowe, D.R., 1982, Sediment gravity flows: II. Depositional models with special
reference to the deposits of high-density turbidity currents: Journal of
Sedimentary Petrology, v. 52, p. 279-297.
Lowe, D.R., and Guy, M., 2000, Slurry-flow deposits in the Britannia Formation
(Lower Cretaceous), North Sea: A new perspective on the turbidity current and
debris flow problem: Sedimentology, v. 47, p. 31-70.
Lucente, C.C., and Pini, G.A., 2008, Basin-wide mass-wasting complexes as markers
of the Oligo-Miocene foredeep-accretionary wedge evolution in the Northern
Apennines, Italy: Basin Research, v. 20, p. 49-71.
MacDonald, D.I.M., Moncrieff, A.C.M., and Butterworth, P.J., 1993, Giant slide
deposits from a Mesozoic fore-arc basin, Alexander Island, Antarctica:
Geology, v. 21, p. 1047-1050.
Maier, K.L., 2012, Depositional architecture of deep-water slope systems: Examples
from the Quaternary Lucia Chica channel system, offshore central California
and the Upper Miocene Urenui Formation, New Zealand [Ph.D. thesis]:
Stanford University, Stanford, California, 412 p.
Martinsen, O.J., 1994, Mass movements, in Maltman, A., ed., The Geological
Deformation of Sediments: London, Chapman & Hall, p. 127-165.
Martinsen, O.J., and Bakken, B., 1990, Extensional and compressional zones in
slumps and slides in the Namurian of County Clare, Ireland: Journal of the
Geological Society, London, v. 147, p. 153-164.
Masalimova, L., 2013, Stratigraphic architecture and flow dynamics of deep-water
turbidite deposits: The Miocene Lower Mount Messenger Formation in the
Taranaki Basin in New Zealand, and the Oligocene Puchkirchen Formation in
the Molasse Basin in Austria [Ph.D. thesis]: Stanford University, Stanford,
California, 193 p.
Miyata, T., 1990, Slump strain indicative of paleoslope in Cretaceous Izumi
sedimentary basin along Median tectonic line, southwest Japan: Geology, v.
18, p. 392-394.
Nagel, S., 2010, An integrated outcrop, modeling and seismic investigation of mass
transport complexes from the Taranaki Basin, New Zealand [M.S. thesis]:
Auckland, New Zealand, University of Auckland, 92 p.
Nardin, T.R., Hein, F.J., Gorsline, D.S., and Edwards, B.D., 1979, A review of mass
movement processes, sediment and acoustic characteristics and contrasts in
slope and base-of-slope systems versus canyon-fan-basin-floor basins, in
Doyle, L.J. and Pilkey, O.H., eds., Geology of Continental Slopes: Society of
Economic Paleontologists and Mineralogists (SEPM) Special Publication 27,
p. 6173.
Nodder, S.D., 1987, The mid-Miocene geology of the Waikawau region, north
Taranaki, New Zealand: Catastrophic sedimentation in a restricted slope basin
[Ph.D. thesis]: Waikato, New Zealand, University of Waikato, 302 p.
Nodder, S.D., Nelson, C.S., and Kamp, P.J.J., 1990a, Middle Miocene formational
stratigraphy (Mokau-Mohakatino Groups) at Waikawau, northeastern Taranaki
148

Basin margin, New Zealand: New Zealand Journal of Geology and


Geophysics, v. 33, p. 585-598.
Nodder, S.D., Nelson, C.S., and Kamp, P.J.J., 1990b, Mass-emplaced siliciclasticvolcaniclastic-carbonate sediments in Middle Miocene shelf-to-slope
environments at Waikawau, northern Taranaki, and some implications for
Taranaki Basin development: New Zealand Journal of Geology and
Geophysics, v.33, p. 599-615.
Odonne, F., Callot, P., Debroas, E., Sempere, T., Hoareau, G., and Maillard, A., 2011,
Soft-sediment deformation from submarine sliding: Favourable conditions and
triggering mechanisms in examples from the Eocene Sobrarbe delta (Ainsa,
Spanish Pyrenees) and the mid-Cretaceous Ayabacas Formation (Andes of
Peru): Sedimentary Geology, v. 235, p. 234-248.
Prather, B.E., 2003, Controls on reservoir distribution, architecture, and stratigraphic
trapping in slope settings: Marine and Petroleum Geology, v. 20, p. 529-545.
Prather, B.E., Booth, J.R., Steffens, G.S., and Craig, P.A., 1998, Classification,
lithologic calibration, and stratigraphic succession of seismic facies of
intraslope basins, deep-water Gulf of Mexico: American Association of
Petroleum Geologists Bulletin, v. 82, p. 701-728.
Prather, B.E., Pirmez, C., Sylvester, Z., and Prather, D.S., 2012, Stratigraphic response
to evolving geomorphology in a submarine apron perched on the upper Niger
Delta slope, in Application of the Principles of Seismic Geomorphology to
Continental-Slope and Base-of-Slope Systems: Case Studies from Seafloor and
Near-Seafloor Analogues, Prather, B.E., Deptuck, M.E., Mohrig, M.E., Van
Hoorn, B., and Wynn, R.B., eds., Society for Sedimentary Geology (SEPM)
Special Publication 99, p. 145-161.
Prlat, A., Hodgson, D.M., Flint, S., 2009, Evolution, architecture and hierarchy of
distributary deep-water deposits: a high-resolution outcrop investigation from
the Permian Karoo Basin, South Africa: Sedimentology, v. 56, p. 2132-2154.
Prior, D.B., Bornhold, B.D., and John, M.W., 1984, Depositional characteristics of a
submarine debris flow: The Journal of Geology, v. 92, p. 707-727.
Posamentier, H.W., and Martinsen, O.J., 2011, The character and genesis of
submarine mass-transport deposits: Insights from outcrop and 3D seismic data,
in, Shipp, R.C., Weimer, P., and Posamentier, H.W., eds., Mass-transport
deposits in deepwater settings, Society for Sedimentary Geology (SEPM),
Special Publication 96, p. 7-38.
Posamentier, H.W., and Walker, R.G., 2006, Deep-water turbidites and submarine
fans, in, Posamentier, H.W., and Walker, R.G., eds., Facies Models Revisited,
SEPM Special Publication, v. 84, p. 399-520.
Ramsay, J.G., 1967, Folding and fracturing of rocks: McGraw-Hill Book Company,
New York, 560 p.
Ramsay, J.G., and Huber, M.I., 1987, The Techniques of Modern Structural Geology,
Volume 2: Folds and Fractures, Academic Press, Waltham, Massachusetts, 391
p.
Rotzien, J.R., Lowe, D.R., King, P.R., and Browne, G.H., 2014, Stratigraphic
architecture and evolution of a deep-water slope channel-levee and overbank
149

apron: The Upper Miocene Upper Mount Messenger Formation, Taranaki


Basin: Marine and Petroleum Geology, v. 52, p. 22-41.
Smith, J.V., 2000, Flow pattern within a Permian submarine slump recorded by
oblique folds and deformed fossils, Ulladulla, south-eastern Australia:
Sedimentology, v. 47, 357-366.
Stagpoole, V., and Nicol, A., 2008, Regional structure and kinematic history of a large
subduction back thrust: Taranaki Fault, New Zealand: Journal of Geophysical
Research, v. 113, B01403.
Strachan, L.J., 2002, Geometry to genesis: a comparative field study of slump deposits
and their modes of formation [PhD dissertation]: Cardiff, United Kingdom,
University of Cardiff, 412 p.
Strachan, L.J., and Alsop, G.I, 2006, Slump folds as estimators of palaeoslope: a case
study from the Fisherstreet Slump of County Clare, Ireland: Basin Research, v.
18, p. 451-470.
Strogen, D.P., 2011, Updated paleogeographic maps for the Taranaki Basin and
surrounds: GNS Science, Lower Hutt, New Zealand, GNS Science Report
2010/53, 83 p.
Trincardi, F., and Argnani, A., 1990, Gela submarine slide: A major basin-wide event
in the Plio-Quaternary foredeep of Sicily: Geo-Marine Letters, v. 10, p. 13-21.
Uruski, C., and Wood, R., 1991, A new look at the New Caledonia Basin, an extension
of the Taranaki Basin, offshore North Island, New Zealand: Marine and
Petroleum Geology, v. 8, p. 379-391.
Utley, J.P., 1987, The middle to late Miocene Mohakatino Group in the coastal
Herangi region, northern Taranaki, New Zealand: punctuated sedimentation in
continental slope-slope basin environments [M.S. thesis]: Waikato, New
Zealand, University of Waikato, 221 p.
Van der Merwe, W.C., Hodgson, D.M., and Flint, S.S., 2009, Widespread synsedimentary deformation on a muddy deep-water basin-floor: the Vischkuil
Formation (Permian), Karoo Basin, South Africa: Basin Research, v. 21, p.
389-406.
Woodcock, N.H., 1976, Structural style in slump sheets: Ludlow Series, Powys,
Wales: Journal of the Geological Society of London, v. 132, p. 399-415.
Woodcock, N.H., 1979a, Sizes of submarine slides and their significance: Journal of
Structural Geology, v. 1, p. 137-142.
Woodcock, N.H., 1979b, The use of slump structures as palaeoslope orientation
estimators: Sedimentology, v. 26, p. 83-99.

150

Figure 1. Geographic setting of New Zealand and the Taranaki Basin. Modified from
Edbrooke et al. (2005) and Giba et al. (2010, 2013). The study area lies on the western
coast of the North Island (red circle). The 2000 m isobath is demarcated by a thin, dashed
line. Onshore faults are shown as thin, black lines. Arrows indicate the relative motion
between the Pacific and Australian plates.

151

Figure 2. Map of the northeastern Taranaki Basin showing the location of volcanoes,
major faults, seismic lines, and wells mentioned in the text (modified from Giba et al.,
2013). Volcanoes older and younger than ca. 8 Ma are colored dark and light red,
respectively (age of volcanoes from Giba et al., 2013).
152

Figure 3. Study area map. A) Geologic map of the study area (modified from Edbrooke
et al., 2005). Faults are shown as black lines. Location of Otukehu MTD from
Masalimova (2013). Formation colors shown in (B). B) Chronostratigraphic diagram of
the Taranaki and King Country Basins (modified from Edbrooke et al., 2005).
Approximate section location shown in (A) as a dashed, gray line. C) Detailed map of the
study area. R1-7 labels demarcate the location of subregions discussed in the text.
Mapped exposures of the NAMTD are shown as hachured. Measured sections are shown
as gray circles (see Figs. 5-7). Wave-cut platform outlines are shown as black polygons.
Abbreviations: KS: Kopuai Stream; NS: Ngatupaku Stream; OS: Ounutae Stream; PRS:
Paparahia Stream; PTS: Paritutu Stream; PS: Pitone Stream; PPS: Poporotaupo Stream;
WKS: Waiakapua Stream; WS: Waihi Stream; WGS: Waiongaro Stream; WRS:
Waioroko Stream.

153

154

Figure 4. Generalized stratigraphy of the middle-upper Miocene sedimentary succession


of the northeastern Taranaki Basin. Modified from Maier (2012), Masalimova (2013),
and Rotzien et al. (2014). Abbreviations: Fm: Formation; Ss: Sandstone.
155

Figure 5. Measured stratigraphic sections from within the northern three kilometers of
the NAMTD. Correlated beds are shown as horizontal lines (see Fig. 9). Note subtle bed
thinning toward the south. See Appendix C-1 for full-size sections.

156

Figure 6. Measured stratigraphic sections from within the southern eight kilometers of
the NAMTD. Sections cannot be confidently correlated due to intense soft-sediment
deformation in this region. See Figure 5 for measured section key and Appendix C-1 for
full-size sections.

157

158

Figure 7. Measured stratigraphic sections from above, below, and lateral to the NAMTD.
See Figure 5 for measured section key and Appendix C-1 for full-size sections.

159

Figure 8. Outcrop photos and measured stratigraphic section at Opito Point. A)


Panorama of Opito Point. Photographs were collected offshore from a boat. Note the
lateral continuity of individual sandstone beds, which can be traced for over two
kilometers in this outcrop. Arbitrary bedding contacts are shown on the line-drawing
interpretation below. Tectonic normal faults are shown in brown. B) Measured section of
major lithofacies present at Opito Point. Note several apparent upward-coarsening cycles.
Paleocurrent measurements are shown as gray rose diagrams. C) Inset photograph
showing the character of the Mohakatino Formation where it is not deformed by postdepositional soft-sediment deformation. This interval is dominated by thickly-bedded
volcaniclastic sandstone, similar to what is found to several kilometers to the south in the
northern portion of the NAMTD.

160

161

Figure 9. Cliff photomosaics and line-drawing interpretations for the northern 5.5 km of
the study area. The upper surface of the NAMTD is shown as a red line. Other colored
lines correspond to correlated marker beds (see Figure 5). Tectonic normal faults are
shown as brown lines. Strata that immediately overly the MTD and in-fill local
topography are shaded orange. Contacts are dashed where covered or inferred. Shaded
gray regions indicate vegetated or talus cover. Full size photomosaics are available as
Appendix C-5. Photomosaic locations are shown in Figure 11.

162

163

164

165

Figure 10. Example outputs from the 3D model generated from aerial photographs of the
NAMTD south of Waioroko Stream (subregion R4). A) Sparse point cloud. B) Dense
point cloud. C) Mesh model. D) Photographs draped atop the mesh model. E)
Orthophoto. F) Digital elevation model.

166

167

Figure 11. Aerial photographs and line-drawing interpretation of the wave-cut platforms.
Full-size, georeferenced orthophotos are available in Appendix C-6.

168

169

170

171

172

Figure 12. Selected cross-line (A-B) and in-line (C) amplitude sections from the Kahu
3D seismic survey. See locations in Figure 2. Faults are shown as black lines. The black
arrows denote the base of the Mohakatino Formation as correlated from the nearby
Mokau-1 well (King et al., 2011).

173

Figure 13. Cliff photomosaic and line-drawing interpretation for the southern extent of the NAMTD, between Paritutu and Pitone
Streams.

174

Figure 14. Examples of the thick- to medium-bedded volcaniclastic sandstone


lithofacies. A) Measured section from south of Paparahia Stream (a portion of section 2).
See key in Figure 5. B) Photo of the same amalgamated sandstone bed at 23.5 m in
section 2, but here south of Waioroko Stream. C) Large zones of intraclasts and volcanic
lithic fragments within a thick-bedded volcaniclastic sandstone bed. D) Downwards
injection of sand and intraclasts north of Waioroko Stream. Jacobs staff marked in 10 cm
increments. Abbreviations: md-mudstone; vf-very fine; f-fine; m-medium; c-coarse; vcvery coarse; g-gravel.

175

176

Figure 15. Examples of folds within the NAMTD. A) A series of two large, recumbent
folds south of Kopuai stream. The fold axes run approximately parallel to the cliff (15195) and are slightly curved toward the west. Note that the thick sandstone at beach
level is also structurally repeated in the top of the cliff within the hinge of the upper fold.
The white arrows indicate stratigraphic up, and the white outlines are geologists for scale.
The uppermost folded beds are approximately 50 meters above beach level. B) A series
of two tight, gently inclined folds along the south bank of Waioroko Stream. These folds
are likely a continuation of those present just to the north of Waioroko Stream (Fig. 23).
The lower limb is observed to again be overturned just to the west in the wave-cut
platform (not pictured). C) Rootless, disharmonic folds within the thin-bedded lithofacies
south of Poporotaupo Stream. D) Complex deformation south of Waiongaro Stream
where intensely deformed units are intercalated with less deformed zones. Sheep for scale
(red ovals). E) Complex folding north of Waihi Stream within interbedded thick- and
thin-bedded lithofacies. F) Dismembered, rootless folds within a MTD north of Kopuai
Stream. Geologic relationships suggest that this MTD was emplaced prior to being
remobilized as part of the NAMTD (the thickness of this MTD is ~22 meters). G) A
series of NNW-SSE trending folds north of Ounutae Stream. Aerial photograph (above)
and line-drawing interpretation (below). Bedding strike and dip measurements are shown
in black. Fold axial traces are shown as dashed red lines and the dip of the axial surface is
labeled in red. Arrows indicate direction of fold plunge (see also key in Fig. 11). These
folds are doubly plunging and have variably oriented axial surfaces that are observed to
change dip direction within the same fold (see also cross-section f-f in Fig. 22). Thus,
different portions of the same fold verge in opposing directions. H) A series of longwavelength, E-W trending folds south of Paritutu Stream. Aerial photograph (above),
line-drawing interpretation (middle), and photograph taken from the adjacent coastal cliff
(below). Note geologist for scale (white outline with red oval). The yellow circle and x
demarcate the same location. Folds have straight to curved axial surfaces. Both S- and Nverging folds are present. The basal detachment surface can be locally observed at this
location (solid red line).

177

178

179

180

181

182

Figure 16. Examples of the thin- to medium-bedded volcaniclastic sandstone and


mudstone lithofacies. A) Measured section from north of Waiakapua Stream (from
section 13). See Figure 5 for the key. B) Photo taken south of Waiakapua Stream that
shows the thin-bedded character of this lithofacies. C) Photo taken from the south bank of
Paritutu Stream (below the MTD) that shows how individual beds often have thin,
coarse-grained bases overlain by a finer-grained matrix (from section 15). Arrows
indicate coarse-grained bed bases. D) Convoluted laminations within fine-grained,
laminated sandstone. Photo taken north of Pitone Stream above the NAMTD (from
section 17). E) Current structured sandstone with planar lamination and climbing ripple
cross-lamination (marked by black arrows). Jacobs staff marked in 10 cm increments.
Abbreviations: md-mudstone; vf-very fine; f-fine; m-medium; c-coarse; vc-very coarse;
g-gravel.

183

184

Figure 17. Illustration of lateral (north to south) changes in the style of deformation within the northern 5 km of the NAMTD. A)

Line-drawing interpretation of the NAMTD (vertical exaggeration 5X; see also Fig. 9). B) Selected examples; locations shown in
(A).

185

Figure 18. Profile view of an upright to recumbent anticline-syncline pair north of


Kopuai Stream. A) Photo of the folds taken from a boat offshore (see Fig. 9A). B) Linedrawing interpretation. The dashed black lines are fold axial surfaces that show
significant downward-concavity. Black arrows show dcollement zones below more
competent sandstone beds.

186

Figure 19. Histograms of fold axis plunge, interlimb angle, and axial plane dip. See
Appendix C-2 for a complete datatable of slump fold orientations.

187

Figure 20. Slump fold orientations. See Figure 3 for subregion locations and Table 2 for
paleoslope method abbreviations. Results from the HIM, HAM, and ASIM are not shown
if less than 5 different from the results from the MAM or APM. Predicted transport
directions queried where uncertain or disfavored (see Table 2).

188

189

190

Figure 21. Fold profile analysis following Ramsay (1967). Six measured fold profiles all
plot as class 1C folds that are intermediate between ideal parallel (class 1B) and similar
(class 2) folds.

191

Figure 22. Example cross-sections of large folds within the northern 4 kilometers of the NAMTD. See Figure 11 for cross-section

locations. Folding is characterized by tight to isoclinal interlimb angles, large scale (up to 10s of meters) wavelengths and fold

heights, and axial planes that are inclined to recumbent (and often curved). Thick-bedded sandstone beds primarily deform via

flexural-slip folding whereas fine-grained intervals accommodate detachment through flexural-flow. Dcollement zones are observed

within fold hinges. Sections a-a, b-b, c-c, and d-d, display the same sequence of stratigraphy repeatedly deformed along the outcrop

extent (see Figure 5 for measured stratigraphic sections of this interval). Figures 18, 15B, and 23 show several field photographs of
these folds. Contacts are dashed where inferred or approximate.

192

Figure 23. A series of tight to isoclinal, nearly recumbent folds north of Waioroko
Stream. A) Cliff photograph taken offshore from a boat. B) Line-drawing interpretation
of the cliff photograph. See key in Figure 9. Black arrows are stratigraphic up.
Dcollement zones are shaded grey. C) Oblique block diagram showing observed fold
geometries (dashed where inferred or uncertain).

193

194

Figure 24. Examples of slump faults within the NAMTD. A) Normal faults with
welded contacts. Pen for scale. B) Reverse fault with a concave-upward surface. White
outline is a crouching geologist. Jake staff is 1.5 meters. C) Reverse fault at the base of a
thick sandstone bed. D) A thick sand bed is reverse faulted and structurally repeated. This
fault is exposed within a cave formed by an anticline within a gravel-based thick sand
bed (navy marker bed; 23.5 m on section 2; Fig. 5). E) Reverse faulting (white arrows)
within thick sand beds south of Paparahia Stream. Note a brown-and-white weathering
sand bed at the top of the cliff is structurally repeated at least four times (see also Fig. 9).
Geologist (white outline) for scale. F) A boudinaged sand bed within a finer-grained
volcaniclastic matrix. Jacobs staff is 1.5 meters. G) Example of a low-angle
detachment surface near Waiakapua Stream. Geologist (white outline) for scale. H)
Low-angle reverse(?) fault and associated drag fold. I) Reverse faults within the hinge of
an isoclinal fold. Photo taken north of Waioroko Stream.

195

196

197

198

Figure 25. Histograms of slump fault surface dip and approximate displacement. Black
histograms are faults that have been restored to bedding horizontal. The dashed gray line
shows fault that have not been restored (see discussion in text).

199

Figure 26. Slump fault orientations. See Figure 3 for subregion locations and Table 2 for
paleoslope method abbreviations. Predicted transport directions queried where uncertain
or disfavored (see Table 2).

200

201

Figure 27. Illustration of fold-based paleoslope methods. Arrows denote MTD transport
direction as inferred from each paleoslope method. A-B) Fold axes plotted on an equalarea stereonet illustrating the mean axis, separation arc, and downslope-average axis
methods (modified from Woodcock, 1979b). C) Poles to fold axial planes that illustrate
the axial-planar and axial-planar intersection methods (modified from Strachan and
Alsop, 2006). D-E) Illustration of methods that utilize progressive fold deformation
during layer-parallel simple shear for cases of progressive fold rotation (D) and no fold
rotation (E). Modified from Strachan and Alsop (2006). F) Illustration of the axial surface
dip and dip direction method. See text and Table 1 for an explanation of each method.

202

Figure 28. Scatterplots of selected fold orientation attributes for the entire NAMTD and
by subregion. A) Fold interlimb angle plotted versus fold axis trend (green boxes) and
fold axial surface strike (blue circles). B) Fold axial plane dip magnitude versus dip
direction. Mean fold dip shown as a dashed black line. C) Fold axial plane dip magnitude
versus fold axis trend. D) Fold interlimb angle versus axial plane dip magnitude. Linear
interpolations (dashed line) all have positive slopes.

203

204

Figure 29. Explanation of the best-fit girdle to fault poles method (GFPM). A) Normal
and reverse faults have alongslope orientations and are all synthetic with respect to the
paleoslope. B) Normal and reverse faults have alongslope orientations. Both synthetic
and antithetic fault populations are present. C) Normal and reverse faults have
orientations that fan obliquely about the paleoslope strike. All faults are synthetic with
respect to the paleoslope. D) Normal and reverse faults have orientations that fan
obliquely about the paleoslope strike. Both synthetic and antithetic fault populations are
present. Black arrow indicates MTD transport direction.

205

Figure 30. Summary of paleoslope analysis results. A) The interpreted transport direction
and associated uncertainty is depicted as an arc for each paleoslope method for the entire
NAMTD and for each subregion (Tables 1 and 2). Arcs are dashed where the resulting
transport direction is disfavored or where the paleoslope analysis result is inconclusive
(see text). The gray shaded region encompasses the estimated transport direction of the
MTD, and the thick, black line and arrow indicates the preferred transport direction based
on maximum arc overlap. Paleoslope methods are identified by letter (see Table 2): a)
MBSM, b) GFPM-N (N), c) GFPM-N (R), d) GFPM-P (N), e) GFPM-P (R), f) MFOM
(N), g) MFOM (R), h) DDM, i) ASIM, j) HIM and HAM, k) APIM-P (black) and APIMN (gray), l) APM, m) MAM (black) and DAM (gray). B) Map of the study area, showing
the location of subregions (see Figure 3 for key and abbreviations).

206

207

Figure 31. Schematic and idealized depiction of the NAMTD. The MTD was emplaced within an overall W-

facing paleoslope, perhaps not very far from its original site of deposition. E-W trending fold axes (shown in

red) are interpreted to likely have resulted from lateral compression along the margins of the MTD. Fold axes

within the central portions of the MTD tend to be sub-parallel with the paleoslope, although downslope-

parallel axes are also present locally. Bathymetric slopes are exaggerated for illustration. Block diagram
modified from Bradley and Hanson (1998).

208

Figure 32. Plan-view images of MTDs from the literature. A) Time structure map of a
MTD upper surface from a 3D seismic survey offshore Norway (Bull et al., 2009).
Elongate, arcuate pressure ridges fan about the toe region of this MTD. These ridges are
interpreted to be the topographic expression of multiple curved thrust faults (Bull et al.,
2009). B) Flattened coherence slice from a large MTD imaged in a 3D seismic survey
offshore Israel (Frey-Martnez et al., 2006). Arcuate structures (marked C) are
interpreted to reflect curved thrust faults within the compressional domain of this MTD
(Frey-Martnez et al., 2006). C) Seismic impedance along a time slice from a 3D seismic
survey in the Gulf of Mexico, United States (Prather et al.,1998). Several well-imaged
MTDs display highly arcuate structures that are likely pressure ridges. D) Side-scan sonar
image of pressure ridges within a MTD that developed in response to delta-front collapse
within a fjord in Norway (Prior et al., 1984).

209

210

211

Figure 33. Schematic paleogeography of the northeastern Taranaki Basin during late
Miocene time. A) Plan-view map. Location of stratovolcanoes from Giba et al. (2013).
Position of Herangi bathymetric high from King et al. (1993). Paleobathymetry modified
from Strogen (2011). Depiction of Mount Messenger Formation based on Masalimova
(2013) and Rotzien et al. (2014). Transport direction of the Otukehu MTD from King et
al. (2011). B) Schematic cross-sections (not to scale). Location of cross-sections shown
as a dashed black line in (A).

212

213

Separation Arc Method (MAM)

Mean Axis Method (MAM)

1. The downslope direction lies within the arc that separates folds with
opposing symmetry
2. Does not allow for folds that verge upslope
3. Folds are formed by heterogeneous simple shear about the paleoslope
dip direction
4. Assumes that fold hinges have not been rotated
5. Excludes symmetric folds from analysis

1. The mean fold axis is oriented parallel to the paleoslope strike


2. MTD undergoes little length change parallel to slope
3. Fold axes are parallel to slope and undergo little rotation
4. Downslope direction indicated by fold vergence and/or facing

Assumptions/Limitations

Woodcock (1979b)

Hansen (1971)

Hahn (1913),
Jones (1939),
Woodcock (1979b)

Source(s)

Table 1. Summary of Paleoslope Analysis Methods

Downslope Average Axis


Method (DAM)

1. The mean fold axis is oriented parallel to the downslope dip direction
2. Fold asymmetry senses oppose each other about a downslope average
axis
3. Maximum elongation direction is aligned downslope

Woodcock (1979b)

Method

Axial-planar Method (APM)

1. Fold axial planes fan about the paleoslope strike


2. Poles to axial planes plot in a great circle about the mean fold axis
3. The mean axial plane dips upslope
4. Assumes that axial planes have not been rotated

Alsop and Holdsworth (2002),


Strachan and Alsop (2006)

Fold-based methods

Axial-planar Intersection Method


(APIM)

1. The mean axial-planar intersection of folds with opposing vergence is


parallel or perpendicular to the downslope direction for folds formed
by layer-normal or layer-parallel shear, respectively
2. Fold hinges will verge and face in a statistical arc about the MTD
transport direction (layer-normal shear)
3. Requires both S- and Z-folds and excludes symmetric folds

214

Axial Surface Dip and Dip


Direction Method (DDM)

Axial Surface Strike and


Interlimb Angle Method (ASIM)

Fold Hinge Azimuth and


Interlimb Angle Method (HIM)

Fold Hinge Azimuth and Axial


Surface Dip Method (HAM)

1. Fold axial surfaces are steeper at the lateral margins of MTDs than in
the central portions
2. Fold axial surfaces in the lateral/oblique portions of a MTD will strike
sub-parallel to the downslope direction
3. Assumes dominantly layer-normal shear within the margins of MTDs
4. Requires prerequisite knowledge about relative position within the
MTD

1. Fold axial surfaces initially strike parallel to the paleoslope


2. With applied shear stress folds tighten and axial surfaces rotate into a
transport parallel direction

1. Initial fold hinge orientations are parallel to the paleoslope


2. With applied shear stress, folds tighten and hinges rotate into a
transport parallel direction
3. Hinges rotate at either end to form curvilinear geometries

1. Folds initially form with steeply-dipping axial surfaces and fold hinge
orientations that are parallel to the paleoslope
2. With applied shear stress, axial surfaces rotate parallel to the bouding
surfaces of the MTD and hinges rotate into a transport parallel
direction

Debacker et al. (2009)

Debacker et al. (2009)

Farrell and Eaton (1987),


Strachan and Alsop (2006)

Farrell and Eaton (1987)

Mean Fault Orientation Method


(MFOM)

1. The mean fault dip direction is parallel to the MTD transport direction
2. Reverse faults dip upslope and normal faults dip downslope
3. Assumes that there is no significant component of oblique- or strikeslip fault motion

Debacker et al. (2009),


This study

Farrell (1984),
Martinsen and Bakken (1990),

Fault-based methods

Best-fit Girdle to Fault Poles


Method (GFPM)

1. The downslope direction is parallel to the strike of the best-fit girdle


for faults with alongslope orientations
2. The downslope direction is perpendicular to the strike of the best-fit
girdle for faults with obliquely fanning orientations
3. Allows for both synthetic and antithetic fault populations

215

Bedding-based method
Mean Bedding Strike Method
(MBSM)

1. The mean bedding strike is parallel to the paleoslope strike


2. Assumes that folds have alongslope orientations
3. Downslope direction must be constrained via other methods

Note: Table modified from Strachan and Alsop (2006)

This study

216

D1
9
14
84
123
72
148
-179

D2
189
194
264
303
252
328
-359

DAM

APIM-P

APIM-N

HIM

-8
30
10
18
33
-32

Table 2. Summary of Paleoslope Analysis Results

APM

D2
-194
258
300
252
--352

GFPM-P (R) GFPM-N (R) GFPM-P (N) GFPM-N (N)


D1 D2 D1 D2
D1 D2 D1 D2
--- -- --- -- --- -- 46 226 20
-- -- --- -- --- -47 227 20 --- -- --- -- --- -41 221 20 --- -- --- -- 50 230 20
116 296 20 -99 279 20 --- -- 110 290 20 --- ---- -- --- -- --- -- 40 220 20
-- -- --- -- 54 234 20
1 181 20 --- -- --- -- 44 224 20
177 357 20 --

* D1 D2 * D1 D2 D1 D2 D1 D2 D1
20 99 279 -- 169 349 20 10 190 20 100 280 20 -9.4 ---7 187 20 --- -- 14
-- -- -28 ---- 53 233 20 --- -- 78
-- -- -14 ---- 119 299 20 126 306 20 36 216 20 120
6.8 ---- 72 252 20 94 274 20 4 184 20 72
18 ---- 136 316 20 --- -- --- -- 149
----- 158 338 20 83 263 20 173 -- 20 -9.1 89 269 9.1 180 360 20 172 352 20 82 262 20 172

Methods based on slump folds


MAM
Location
All data
R1
R2
R3
R4
R5
R6
R7
Methods based on slump faults
MFOM (R)
MFOM (N)
D1 D2 * D1 D2 *
---- 43 223 8.5
--36 216 15 ---47 227 21 -116 296 15 68 248 32
104 284 19 102 282 16
47 227 21 47 227 17
169 349 21 46 226 20
177 357 19 36 216 20
Location
All data
R1
R2
R3
R4
R5
R6
R7

D1
-14
78
120
72
147
-164

HAM

ASIM

MBSM
#
D2 *
261 2.5 -4
194 6
0
242 3.2 -292 4.6 -3
259 1.8 3
271 5.7 -9
292 8.4 -51
184 5.7 21

DDM

D2
-277
-----261

*
-29
-----12

Preferred
Transport
D1 D2
--- -30 210 24
67 247 34
123 303 17
84 264 30
94 274 63
--- -176 356 16

D2 D1 D2 D1
------194 8.2 10 190 9.2 97
258 30 75 255 26 -300 10 119 299 6.3 -252 18 73 253 16 --- 33 148 -- 37 -------344 32 173 353 27 81

D1
81
14
62
112
79
91
112
4

Notes:
* Error is the 95% confidence limit
Error is the standard deviation of the mean
Error is arbitrarily assigned
# Difference in degrees between the mean bedding strike and the mean fold axial surface orientation
-- Not applicable
D1 and D2 are possible MTD transport directions
Bolded values represent the preferred transport direction(s)
Preferred transport direction(s) based on maximum overlap of paleoslope methods (see Fig. 18 and discussion in text)

217

Abbreviations: APIM-N: Axial-planar intersection method (layer-normal shear); APIM-P; Axial-planar intersection method (layer-parallel shear); APM:
Axial-planar method; ASIM: Fold axial surface strike and interlimb angle method; DAM: Downslope average axis method; DDM: Axial surface dip and dip
direction method; HAM: Fold hinge azimuth and axial surface dip method; HIM: Fold hinge azimuth and interlimb angle method; MAM: Mean axis
method; MBSM: Mean bedding strike method; MFOM: Mean fault orientation method (fault type shown in parentheses); GFPM-P: Best-fit girdle to fault
poles method (alongslope faults; fault type shown in parentheses); GFPM-N: Best-fit girdle to fault poles method (obliquely fanning faults; fault type shown
in parentheses)

218

Table 3. Criteria for Identification of the Lateral/Oblique Portions of MTDs


Criterion

Subregion
R1
R7

Fold axes are oriented sub-parallel or obliquely to the inferred paleoslope dip direction
(Gawthorpe and Clemmey, 1985; Smith, 2000; Debacker et al, 2009, citing
Woodcock, 1979b; Lajoie, 1972)1

Yes

Yes

Fold axes plunge steeply (Farrell and Eaton, 1987) or downwards towards the MTD
margins (Smith, 2000)

No

No

Fold axes are not tightly clustered (Debacker et al., 2009)

No

No

Axial plane strike fans about the inferred paleoslope dip direction (Debacker et al.,
2009; Alsop and Marco, 2011)1

Yes

Yes

Layer-normal shear dominates (Debacker et al., 2009; Alsop and Marco, 2011)

No

No

Yes?

No

No

No

Yes3

?4

?4

No5

No

Yes

Yes

No

No

NA2

No6

Yes5

No

Axial planes are steeper than in central portions of MTDs (Debacker et al., 2009, citing
Coward and Kim, 1981; Coward and Potts, 1983)
Folds tighten as they become more steep (Debacker et al., 2009, their figure 15, citing
Strachan and Alsop, 2006)
Layer-normal shear results in bimodal fold vergence distributions (Debacker et al.,
2009; Alsop and Marco, 2011)
Presence of strike- or oblique-slip faults (Farrell, 1984; Farrell and Eaton, 1987;
Maltman, 1994; Debacker et al., 2009)
Reverse faults are inconsistent with mean fold axis orientation (Debacker et al., 2009)
Reverse faults strike at low angles to the inferred paleoslope dip direction (Debacker et
al., 2009)1
Falling axis (-axis) of reverse faults has a significant plunge and may deviate from the
mean fold axis (Debacker et al., 2009)
Intersection of reverse and normal faults is at high angles to the inferred MTD transport
direction (plunge azimuth relative to slump sense is counter-clockwise for dextral,
clockwise for sinistral; Debacker et al., 2009) 1
MTD transport direction inferred from faults is clockwise (dextral) or clockwise
(sinistral) relative to the folds (Debacker et al., 2009)

Notes:
1
Paleoslope dip direction must be independently known.
2
There are too few normal faults in subregion R1 to evaluate this criterion.
3
Bimodal fold vergence distributions also observed in areas inferred to be within the central/frontal
portions of the MTD (subregions R2-6).
4
Slump fault slip direction indicators were not observed. Thus the amount of oblique- or strike-slip fault
motion is poorly constrained.
5
Predicted MTD transport based on reverse faults is ~20 to 40 degrees west of the predicted MTD transport
direction based on folds (Table 2).
6
Intersection of the mean normal and reverse fault for subregion R7 is 113/3 (trend and plunge) this is
oriented at a low angle to the inferred transport direction.

219

APPENDIX A: SUPPLEMENTARY MATERIAL FOR CHAPTER 1

Summary of Contents
Appendix A-1: Geologic Map Data Sources
Appendix A-2: Details Regarding the Palinspastic Reconstruction in Figure 2B
Appendix A-3: Supplementary Descriptions of Sample Locations
Appendix A-4: Cumulative Detrital Zircon U-Pb Age Distributions
Appendix A-5: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of Arizona Laserchron Center
Appendix A-6: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of California Santa Cruz
Appendix A-7: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
University of California Los Angeles
Appendix A-8: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the
Stanford-USGS SHRIMP
Appendix A-9: Detrital Zircon Age Population Proportions

Appendices A-4 through A-8 are available electronically as supplemental files on


Stanford Universitys library website

220

APPENDIX A-1: Geologic Map Data Sources


Figure 2 was compiled from a variety of data sources (see below). Ludington et
al. (2007) was used as a base geologic map for California (1:750,000), Oregon
(1:500,000), Nevada (1:500,000), and Arizona (1:1,000,000). Additional sources of
information are presented by region below:

Oregon (Walker and MacLeod, 1991; DeGraaff-Surpless et al., 2002; Miller et al., 2003)
Nevada (Stewart and Carlson, 1976; John, 1983; Crafford, 2007; Van Buer and Miller,
2010; Van Buer, pers. comm.)
Arizona (Richard et al., 2000; Jacobson et al., 2011)
Baja and Sonora Mexico (Grove et al., 2008; Jacobson et al., 2011)
California (Jennings et al., 1977; Ludington et al., 2007, and below)
Klamath Mountains (DeGraaff-Surpless et al., 2002; Dickinson, 2008)
Sierra Nevada Mountains (Saleeby and Sams, 1987; Kistler and Ross, 1990;
Irwin and Wooden, 2001; Lackey et al., 2008; Saleeby et al., 2008; Lechler and Niemi,
2011; Chapman et al., 2012; Van Buer, pers. comm.)
Salinian block (Mattinson, 1978, 1990; James, 1992; Kistler and Champion,
2001; Barth et al., 2003; Jacobson et al., 2011; Chapman et al., 2012)
Mojave and southeastern Sierra Nevada (Allen et al., 1995; Miller and
Glazner, 1995; Miller et al., 1995; Walker et al., 2002a,b; Barth et al., 2004; Jacobson et
al., 2011; Chapman et al., 2012; Barth, pers. comm.)
Eastern Transverse Ranges (Barth et al., 1995; Barth et al., 2004; Barth et al.,
2008, Barth and Wooden, 2010; Jacobson et al., 2011)
Peninsula Ranges Batholith (Barth et al., 2004; Grove et al., 2008; Jacobson et
al., 2011)
221

In addition, Barth et al. (1990) and Diles and Wright (1998) were used for
interpretation of the general distribution of Jurassic and Triassic arc rocks.
References
Allen, C.M., Wooden, J.L., Howard, K.A., Foster, D.A., and Tosdal, R.M., 1995, Sources
of the Early Cretaceous plutons in the Turtle and west Riverside Mountains,
California: Anomalous Cordilleran interior intrusions: Journal of Petrology, v. 36,
p. 1675-1700.
Barth, A.P., and Wooden, J.L., 2010, Coupled elemental and isotopic analysis of
polygenetic zircons from granitic rocks by ion microprobe, with implications for
melt evolution and the sources of granitic magmas: Chemical Geology, v. 277, p.
149-159.
Barth, A. P., Tosdal, R. M., and Wooden, J. L., 1990, A petrologic comparison of
Triassic plutonism in the San Gabriel and Mule Mountains, southern California:
Journal of Geophysical Research, v. 95, p. 2007520096.
Barth, A.P., Wooden, J.L., Grove, M., Jacobson, C.E., and Pedrick, J.N., 2003, U-Pb
zircon geochronology of rocks in the Salinas Valley region of California: A
reevaluation of the crustal structure and origin of the Salinian block: Geology, v.
31, p. 517-520.
Barth, A.P., Wooden, J.L., Jacobson, C.E., and Probst, K., 2004, U-Pb geochronology
and geochemistry of the McCoy Mountains Formation, southeastern California: A
Cretaceous retro-arc foreland basin: Geological Society of America Bulletin, v.
116, p. 142153.
Barth, A. P., Wooden, J. L., Tosdal, R. M., and Morrison, J., 1995, Crustal contamination
in the petrogenesis of a calc-alkalic rock series: Josephine Mountain intrusion,
California: Geological Society of America Bulletin, v. 107, p. 201211.
Barth, A.P., Wooden, J.L., Howard, K.A., and Richards, J.L., 2008, Late Jurassic
plutonism in the southwest U.S. Cordillera: Geological Society of America
Special Paper 438, p. 379-396.
Chapman, A.D., Saleeby, J.B., Wood, D.J., Piasecki, A., Kidder, S., Ducea, M.N., and
Farley, K.A., 2012, Late Cretaceous gravitational collapse of the southern Sierra
Nevada batholith, California: Geosphere, v. 8, p. 314-341.
Crafford, A.E.J., 2007, Geologic map of Nevada: U.S. Geological Survey Data Series
249, 1 CD-ROM, 46 p., 1 plate.
DeGraaff-Surpless, K., Graham, S.A., Wooden, J.L., and McWilliams, M.O., 2002,
Detrital zircon provenance analysis of the Great Valley Group, California:
Evolution of an arc-forearc system: Geological Society of America Bulletin, v.
114, p. 1564-1580.
Dickinson, W.R., 2008, Accretionary Mesozoic-Cenozoic expansion of the Cordilleran
continental margin in California and adjacent Oregon: Geosphere, v. 4, p. 329353.
Diles, J.H., and Wright, J.E., 1998, The chronology of early Mesozoic arc magmatism in
the Yerington district of western Nevada and its regional implications: Geological
Society of America Bulletin, v. 100, p. 644-652.

222

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.G., 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. 335362.
Irwin, W.P., and Wooden, J.L., 2001, Plutons and accreted terranes of the Sierra Nevada,
California: U.S. Geological Survey Open-File Report 99-0374, 1 sheet.
Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsaglia, K.M., Gehrels, G.E.,
and Nourse, J.A., 2011, Late Cretaceous-early Cenozoic tectonic evolution of the
southern California margin inferred from provenance of trench and forearc
sediments: Geological Society of America Bulletin, v. 123, p. 485-506.
James, E.W., 1992, Cretaceous metamorphism and plutonism in the Santa Cruz
Mountains, Salinian block, California, and correlation with the southernmost
Sierra Nevada: Geological Society of America Bulletin, v. 104, p. 1326-1339.
Jennings, C.W., Strand, R.G., and Rogers, T.H., 1977, Geologic map of California:
California Division of Mines and Geology, scale 1:750,000.
John, D.A., 1983, Distribution, ages, and petrographic characteristics of Mesozoic
plutonic rocks, Walker Lake 1 by 2 quad., Calif. and Nev.: U.S. Geological
Survey Miscellaneous Field Studies, Map MF-1382-B.
Kistler, R.W., and Champion, D.E., 2001, Rb-Sr whole-rock and mineral ages, K-Ar,
40Ar/39Ar, and U-Pb mineral ages, and strontium, lead, neodymium, and oxygen
isotopic compositions for granitic rocks from the Salinian Composite Terrane,
California: U.S. Geological Survey Open-File Report 01-453, 80 p.
Kistler, R.W., and Ross, D.C., 1990, A strontium isotopic study of plutons and associated
rocks of the southern Sierra Nevada and vicinity, California: U.S. Geological
Survey Bulletin, v. 1920, 20 p.
Lackey, J.S., Valley, J.W., Chen, J.H., and Stockli, D.F., 2008, Dynamic magma systems,
crustal recycling, and alternation in the central Sierra Nevada batholith: the
oxygen isotope record: Journal of Petrology, v. 49, p. 1397-1426.
Lechler, A.R., and Niemi, N.A., 2011, Sedimentologic and isotopic constraints on the
Paleogene paleogeographic and paleotopography of the southern Sierra Nevada,
California: Geology, v. 39, p. 379-382.
Ludington, S., Moring, B.C., Miller, R.J., Stone, P.A., Bookstrom, A.A., Bedford, D.R.,
Evans, J.G., Haxel, G.A., Nutt, C.J., Flyn, K.S., and Hopkins, M.J., 2007,
Preliminary integrated geologic map databases for the United States Western
States: California, Nevada, Arizona, Washington, Oregon, Idaho, and Utah: U.S.
Geological Survey Open-File Report, v. 2005-1305, version 1.3.
Mattinson, J.M., 1978, Age, origin, and thermal histories of some plutonic rocks from the
Salinian block of California: Contributions to Mineralogy Petrology, v. 67, p.
233-245.
Mattinson, J.M., 1990, Petrogenesis and evolution of the Salinian magmatic arc, in
Anderson, J.L., ed., The nature and origin of Cordilleran magmatism: Geological
Society of America Memoir 174, p. 237-250.

223

Miller, J.S., and Glazner, A.F., 1995, Jurassic plutonism and crustal evolution in the
central Mojave Desert, California: Contributions to Mineralogy and Petrology, v.
118, p. 379-395.
Miller, J.G., Miller, R.B., Wooden, J.L., and Harper, G.D., 2003, Geochronologic links
between the Ingalls ophiolite, North Cascades, Washington, and the Josephine
ophiolite, Klamath Mountains, Oregon and California: Geological Society of
America Abstracts with Pro- grams, v. 35, no. 6, p. 113.
Miller, J.S., Glazner, A.F., Walker, J.D., and Martin, M.W., 1995, Geochronologic and
isotopic evidence for Triassic-Jurassic emplacement of the eugeoclinal allochthon
in the Mojave Desert region, California: Geological Society of America Bulletin,
v. 107, p. 1141-1457.
Richard, S.M., Reynolds, S.J., Spencer, J.E., and Pearthree, P.A., 2000, compilers,
Geologic Map of Arizona: Arizona Geological Survey Map 35, scale 1:1,000,000,
1 sheet.
Saleeby, J.B., and Sams, D.B., 1987, U/Pb zircon, strontium, and oxygen isotopic and
geochronological study of the southernmost Sierra Nevada batholith, California:
Journal of Geophysical Research, v. 92, p. 10,443-10,466.
Saleeby, J.B., Ducea, M.N., Busby, C., Nadin, E., and Wetmore, P.H., 2008, Chronology
of pluton emplacement and regional deformation in the southern Sierra Nevada
batholith, California, in Wright, J.E., and Shervais, J.W., eds., Ophiolites, arcs,
and batholiths: A tribute to Cliff Hopson: Geological Society of America Special
Paper 438, p. 397427.
Stewart, J.H., and Carlson, J.E., 1978, Geologic map of Nevada: U.S. Geological Survey
in collaboration with Nevada Bureau of Mines and Geology, scale 1:500,000.
Walker, G.W., and MacLeod, N.S., 1991, Geologic map of Oregon: U.S. Geological
Survey, scale 1:500,000.
Walker, J.D., Berry, A.K., Davis, P.J., Andrew, J.E., and Mitsdarfer, J.M., 2002a,
Geologic map of northern Mojave Desert and southwestern Basin and Range,
California: Geological Society of America Memoir 195, 1 sheet.
Walker, J.D., Martin, M.W., and Glazner, A.F., 2002b, Late Paleozoic to Mesozoic
development of the Mojave Desert and environs, California: Geological Society
of America Memoir 195, 18 p.
Van Buer, N.J., and Miller, E.L., 2010, Sahwave Batholith, NW Nevada: Cretaceous arc
flare-up in a basinal terrane: Lithosphere, v. 2, p.423-446.

224

APPENDIX A-2: Details Regarding the Palinspastic Reconstruction in Figure 2B

OVERVIEW
The California margin has undergone pervasive and complex deformation since
Eocene time as a result of 1) the development of the modern San Andreas transform
system (Atwater, 1970, 1989; Powell, 1993) and 2) extension within the Basin and Range
province (e.g., Wernicke, 1992). Our goal is to reconstruct the margin to its Paleogene to
Cretaceous configuration using known geological constraints. However, any attempt at
such a reconstruction involves considerable simplification and inherit uncertainty given
the complex nature of deformation that has afflicted the continental margin. Additional
ambiguity results from conflicting estimates of the offset along margin-parallel faults of
the San Andreas system (e.g., Dickinson et al., 2005; Burnham, 2009; Langenheim et al.,
2012).
While compressional, extensional, translational, and rotational deformation is
pervasive in California, we prioritize the reconstruction of large-magnitude, marginparallel transport of crustal blocks that contain Cretaceous-Eocene forearc strata. This
style of deformation has the most influence on the along-strike variations in sedimentary
provenance that is the focus of this study (e.g., Figure 7). In particular, we follow Grove
et al. (2003) and Jacobson et al. (2011) in only restoring 1) margin-parallel offset on
strike-slip faults of the San Andreas fault system, and 2) transrotational domains of the
western Transverse Ranges. As such, our restoration does not account for 1) tectonic
shortening related to folding and thrust faulting, 2) extension within the Basin and Range
province or basins within the California Coast Ranges, 3) rotation of domains within the
Mojave Desert region (e.g., Eastern Transverse Ranges, Tehatchapi Mountains, and
northeastern Mojave block), 4) left-lateral or right-lateral offset of faults in the Mojave
Desert region (e.g., Garlock fault), or 5) displacement along the Eastern California Shear
Zone (Dickinson, 1996) . Consideration of these additional complexities is beyond the
scope of this study.
Also, we make no effort to restore slip on the Nacimiento fault given its
controversial and inherently uncertain structural history (e.g., Ducea et al., 2009;
Jacobson et al., 2011). Because the Nacimiento fault was likely active from latest
225

Cretaceous to Paleocene time (ca. 75-60 Ma; Jacobson et al., 2011), offset along this
structure likely influences the relative positions of Cenomanian-Paleocene forearc strata
along the margin as depicted in Figures 7 and 8.
RECONSTRUCTION STEPS
The palinspastic reconstruction presented in Figure 2B was created using the
following series of steps using Ludington et al. (2007) as a base map. We first restore slip
on faults of central and northern California using Graymer et al. (2002), Dickinson et al.
(2005), and Sharman et al. (2013) as a guide (Steps 1-8). We then integrate this
reconstruction with the Transverse Ranges and the Peninsular Ranges following Grove et
al. (2003) and Jacobson et al. (2011) (Steps 9-12).
Step 1: Shift the Gualala, Point Reyes, Pigeon Point, Santa Lucia, Point Sur, and
Nacimiento blocks 40 km southward to restore dextral offset along the Reliz-Rinconada
fault (Graham, 1978). This reconstruction assumes that Reliz-Rinconada slip is routed
completely to the San Gregorio-Hosgri fault in the vicinity of the Monterey Bay fault
zone (e.g., Dickinson, 1996), rather than being completely or partially absorbed by intraSalinian block deformation in the vicinity of the Zayante-Vergeles fault zone (e.g.,
Dickinson et al., 2005; Langenheim et al., 2012).
Step 2: Shift the Gualala, Point Reyes, Pigeon Point, and Point Sur blocks 115 km
southward by removing dextral offset along the San Gregorio-Hosgri fault (Dickinson et
al., 2005), exclusive of slip contributed by the Reliz-Rinconada fault. The Gualala block
and Point Reyes block are rotated 15 degrees clockwise to eliminate overlap of the Point
Reyes and La Honda blocks created by the bend in the trace of the San Gregorio-Hosgri
fault at ~39.7 degrees latitude. The Pigeon Point block is positioned adjacent to the
southern Santa Lucia Range as shown in Figure 10 of Dickinson et al. (2005), and the
Point Sur block is positioned northwest of Point Buchon as shown in Figure 12C of
Langenheim et al. (2012).
Note that the location of the Gualala block during Cretaceous and Paleogene time
is controversial, and some authors have argued that the Gualala block should not be
226

restored adjacent to Montara Mountain as is shown in Figure 2B based on the


incompatibility of basement types and differences between the overlying sedimentary
successions (e.g., Dickinson et al., 2005). Alternatively, the Gualala block could be
bounded on its western margin by a fault (the Gualala fault) that decouples this terrane
from inferred Salinian block basement to the west (Dickinson et al., 2005). The existence
of such a fault would allow the Gualala block to accumulate less slip than the ~155 km
known to occur on the San Gregorio-Hosgri fault, thereby restoring this terrane farther
north than shown in our Figure 2B (see also Figure 13 of Dickinson et al., 2005).
However, given the similarity of the detrital zircon provenance signatures between the
Eocene German Rancho Formation and the Eocene Butano Sandstone (Doebbert et al.,
2012; Sharman et al., 2013), and the lack of conclusive evidence showing displacement
on the Gualala fault, we restore the Gualala block outboard of the northern La Honda
basin (e.g., Graham and Berry, 1979; Burnham, 2009). We disfavor any reconstruction
that juxtaposes the Gualala block against the Temblor Range across the future San
Andreas fault, since the detrital zircon provenance character of the Eocene German
Rancho Formation is incompatible with that of the Eocene Point of Rocks Formation and
other Great Valley forearc sandstones to the north (Doebbert et al., 2012; Sharman et al.,
2013).
Step 3: Shift the Salinian block (including the Gualala, Point Reyes, La Honda, Santa
Lucia, Gabilan, Nacimiento, Atascadero, Pigeon Point, Point Sur, La Panza, Sierra
Madre, and Pine Mountains blocks) southward by removing 28 km of slip along the San
Juan-Chimineas-Russell fault (Dickinson, 1996; Langenheim et al., 2012).
Step 4: Shift the Salinian block southward by removing 415 km of dextral offset along
the trend of the central San Andreas-Pilarcitos-northern San Andreas faults (Sharman et
al., 2013). This includes ~315 km of Neogene (Graham et al., 1989) and ~100 km of
Paleogene (Sharman et al., 2013) right-lateral offset. The Salinian block is flexed around
the big bend in the San Andreas fault to eliminate the gap that otherwise occurs
between the Salinian block and Mojave Desert region and to avoid overlap with the San
Gabriel block.
227

Step 5: Shift the Pilarcitos block southward by removing 26 km of dextral offset along
the San Francisco Peninsula segment of the San Andreas fault (Dickinson et al., 2005;
and references within). Apply a 5 degree counterclockwise rotation to eliminate overlap
with the San Francisco Bay block.
Step 6: Shift the Pilarcitos block and Bay block southward by removing 100 km of
dextral offset along the Hayward fault zone (Graymer et al., 2002).
Step 7: Shift the Pilarcitos block, San Francisco Bay block, and East Bay block
southward by removing an additional 75 km of cumulative dextral offset along faults of
the East Bay fault system, excluding the Hayward fault zone (Graymer et al., 2002). Note
that for simplicity, we have treated the East Bay block as a single unit and have
partitioned all 75 km of restored slip along the easternmost fault zone (Green ValleyConcord-Diablo-Greenville-Calaveras fault; Graymer et al., 2002). Apply a 7 degree
counterclockwise rotation to eliminate overlap with the Diablo Range.
Step 8: Shift the Franciscan Coastal Belt southward by removing 175 km of dextral
offset. Note that this reconstruction assumes that all 175 km of slip on the East Bay fault
system is routed to the eastern margin of the Franciscan Coastal Belt (McLaughlin et al.,
1996; Graymer et al., 2002).
Note: We now integrate the above reconstruction with the Jacobson et al. (2011)
palinspastic reconstruction of southern California with the following modifications (see
below). Note that our reconstruction of the Salinian block is approximately 115 km
farther southeast than what Jacobson et al. (2011) show in their Figure 2. This
discrepancy largely results from Grove et al. (2003) restoring the northern Salinian block
290 km southeastward as measured along the length of the San Andreas fault following
Powell (1993). We prefer a reconstruction of ~315 km of Neogene displacement
(Graham et al., 1989) and ~100 km of late Paleogene displacement (Sharman et al.,
2013).
228

Step 9: Restore the San Gabriel block as shown in Figure 2 of Jacobson et al. (2011).
Step 10: Restore the Peninsular Ranges Batholith (including the San Jacinto, Elisinore,
and western PRB blocks) and attached forearc (Santa Catalina Island and Channel
Islands) southward 315 km to account for the opening of the Gulf of California
(Dickinson, 1996). This amount is 15 km greater than the Grove et al. (2003) and
Jacobson et al. (2011) reconstructions.
Step 11: The Peninsular Ranges batholith west of the Elsinore fault is restored 12 km to
the southeast (Dickinson, 1996).
Step 12: The Santa Ynez, Santa Monica Mountains, Simi Hills, Santa Catalina Island,
and Channel Islands are rotated according to Jacobson et al. (2011). Their position is
shifted to reflect the slightly modified position of the western Peninsular Ranges
batholith. Note that the distance between the Santa Maria Mountains and Pine Mountain
block and the Santa Ynez, Santa Monica, and Simi Hills has been reduced due to our
revised reconstruction of the Salinian block.
REFERENCES
Atwater, T., 1970, Implications of plate tectonics for the Cenozoic tectonic evolution of
western North America: Geological Society of America Bulletin, v. 81, p. 3513
3536.
Atwater, T.M., 1989, Plate tectonic history of the northeast Pacific and western North
America, in Winterer, E.L., Hussong, D.M., and Decker, R.W., eds., The Eastern
Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America,
Geology of North America, v. N, p. 2172.
Burnham, K., 2009, Predictive model of San Andreas fault system paleogeography, Late
Cretaceous to early Miocene, derived from detailed multidisciplinary
conglomerate correlations: Tectonophysics, v. 464, p. 195-258.
Dickinson, W.R., 1996, Kinematics of transrotational tectonism in the California
Transverse Ranges and its contribution to cumulative slip along the San Andreas
transform fault system: Geological Society of America Special Paper 305, 46 p.
Dickinson, W.R., Ducea, M., Rosenberg, L.I., Greene, H.G., Graham, S.A., Clark, J.C.,
Weber, G.E., Kidder, S., Ernst, W.G., and Brabb, E.E., 2005, Net dextral slip,
Neogene San GregorioHosgri Fault Zone, Coastal California: Geologic Evidence

229

and Tectonic Implications: Geological Society of America Special Paper 391, 43


p.
Doebbert, A.C., Carroll, A.R., and Johnson, C., 2012, The sandstone-derived provenance
record of the Gualala basin, northern California, U.S.A.: Journal of Sedimentary
Research, v. 82, p. 841-858.
Ducea, M.N., Kidder, S., Chesley, J.T., and Saleeby, J.B., 2009, Tectonic underplating of
trench sediments beneath magmatic arcs, the central California example:
International Geology Review, v. 51, p. 126.
Graham, S.A., 1978, Role of Salinian block in evolution of San Andreas fault system,
California: American Association of Petroleum Geologists Bulletin, v. 62, p.
22142231.
Graham, S.A., and Berry, K.D., 1979, Early Eocene paleogeography of the central San
Joaquin Valley: Origin of the Cantua Sandstone, in Armentrout, J.M., Cole, M.R.,
and Ter Best, H., Jr., eds., Cenozoic paleogeography of the western United States:
Pacific Section, SEPM (Society for Sedimentary Geology), Pacific Coast
Paleogeography Symposium 3, p. 119127.
Graham, S.A., Stanley, R.G., Bent, J.V., and Carter, J.B., 1989, Oligocene and Miocene
paleogeography of central California and displacement along the San Andreas
fault: Geological Society of America Bulletin, v. 101, p. 711730.
Graymer, R.W., Sarna-Wojcicki, A.M., Walker, J.P., McLaughlin, R.J., and Fleck, R.J.,
2002, Controls on timing and amount of right-lateral offset on the East Bay fault
system, San Francisco Bay region, California: Geological Society of America
Bulletin, v. 114, p. 1471-1479.
Grove, M., Jacobson, C.E., Barth, A.P., and Vucic, A., 2003, Temporal and spatial trends
of Late CretaceousEarly Tertiary underplating of Pelona and related schists
beneath southern California and southwestern Arizona, in Johnson, S.E., et al.,
eds., Tectonic Evolution of Northwestern Mexico and the Southwestern USA:
Geological Society of America Special Paper 374, p. 381406.
Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsaglia, K.M., Gehrels, G.E.,
and Nourse, J.A., 2011, Late Cretaceous-early Cenozoic tectonic evolution of the
southern California margin inferred from provenance of trench and forearc
sediments: Geological Society of America Bulletin, v. 123, p. 485506.
Langenheim, V.E., Jachens, R.C., Graymer, R.W., Colgan, J.P., Wentworth, C.M., and
Stanley, R.G., 2012, Fault geometry and cumulative offsets in the central Coast
Ranges, California: Evidence for northward increasing slip along the San
Gregorio-San Simeon-Hosgri fault: Lithosphere, v. 5, p. 29-48.
McLaughlin, R.J., Sliter, W.V., Sorg, D.H., Russell, P.C., and Sarna-Wojcicki, A.M.,
1996, Large-scale right-slip displacement on the East San Francisco Bay Region
fault system, California: Implications for location of late Miocene to Pliocene
Pacific plate boundary: Tectonics, v. 15, p. 1-18.
Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic
paleogeography, southern California: Geologic and kinematic constraints on
evolution of the San Andreas fault system, in Powell, R.E., et al., eds., The San
Andreas fault system: Displacement, palinspastic reconstruction, and geologic
evolution: Geological Society of America Memoir 178, p. 1106.

230

Sharman, G.R., Graham, S.A., Grove, M., and Hourigan, J.K., 2013, A reappraisal of the
early slip history of the San Andreas fault, central California, USA: Geology, v.
41, p. 727-730.
Wernicke, B.P., 1992, Cenozoic extensional tectonics of the U.S. Cordillera, in Burchfiel,
B.C., Lipman, P.W., and Zoback, M.L., eds., The Cordilleran Orogen:
Conterminous U.S.: Boulder, Colorado, Geological Society of America, The
Geology of North America, v. G-3, p. 553581.

231

APPENDIX A-3: Supplementary Description of Sample Locations

INTRODUCTION
Due to space limitations in the main text, we have included this appendix to
provide additional information regarding the geologic context of the detrital zircon
samples analyzed as part of this study. While a comprehensive stratigraphic review of
Upper Cretaceous-Eocene forearc strata is beyond the scope of this study, we do include
a description of the major forearc depocenters that extend from coastal Oregon to Baja
California.
In general, this appendix follows the outline used by Jacobson et al. (2011) in that
we describe the following information for each sample:
1. Stratigraphic name
2. Map references
3. Proximity to other published sample locations, when applicable
4. Geographic descriptions of the sample location, when applicable
5. Inferred depositional environment
6. Age
7. Location coordinates (WGS84 datum)
Note that all age assignments based on the California calcareous nannoplankton, planktic
foraminiferal, or benthic foraminiferal stages are updated relative to absolute time using
McDougall (2007).
OVERVIEW OF FOREARC DEPOCENTERS
Oregon forearc
While Cretaceous-Paleocene forearc strata are generally absent in the Pacific
Northwest (Heller et al., 1987), we include samples from the Albian to lower
Maastrichtian Hornbrook Formation and lower to middle Eocene Tyee Formation in this
study (Figs. 1-2; Dumitru et al., 2012; Surpless and Beverly, 2013). The Hornbrook basin
was located in the vicinity of the Oregon-California border and shares a similar history
232

with the Sacramento basin, although these depocenters were probably only connected
after ca. 85 Ma (Surpless and Beverly, 2013). A well-preserved Eocene forearc basin
(Tyee basin) developed in western Oregon (present day Oregon Coast Range) following
the accretion of an oceanic seamount assemblage to the continental margin during early
Eocene time (Heller et al., 1987). Petrographic and geochemical studies of forearc
assemblages indicate a transition from a local source in the adjacent Klamath Mountains
to an easterly source in the Idaho batholith region (Heller and Ryberg, 1983; Heller et al.,
1985; Dumitru et al., 2012). The lower middle Eocene Tyee Formation comprises a thick
(1.52 km) sequence of deltaic, shelfal, and deep-marine sandstones (Chan and Dott,
1983; Heller and Dickinson, 1985) that were deposited over a short interval of time (ca.
49.4-46.4 Ma) indicating very high average sedimentation rates (>500 m/Ma) that
resulted from rapid exhumation associated with the development of core complexes in the
Idaho region (Dumitru et al., 2012).
Great Valley forearc
The Upper Jurassic(?) to Paleogene Great Valley forearc extends for 675 km
along the continental margin and represents a well-preserved and well-studied example of
an ancient forearc basin (Dickinson and Seely, 1979; Ingersoll, 1979; Bartow, 1991;
Dickinson, 1995a; Williams and Graham, 2013). Great Valley forearc strata were
deposited in an asymmetric trough that was confined to the west by the actively accreting
Franciscan subduction complex (Ernst, 1970; Dumitru et al., 2010). Uplift and erosion of
the western portion of the forearc produced a continuous belt of exposures that form an
east-dipping homocline along the western margin of the basin (Fig. 2; Williams and
Graham, 2013). The eastern margin of the basin laps onto the western margin of the
Sierra Nevada batholith and Foothills belt and is largely buried beneath younger valley
fill (Fig. 2; Ingersoll, 1979).
The Great Valley forearc is subdivided into the Sacramento and San Joaquin
basins based on distinct geologic histories between these depocenters. In general, the
forearc records basin filling and shoaling that progressed longitudinally from north to
south over time (Moxon, 1990). While Upper Jurassic(?) and Cretaceous strata
predominate in the Sacramento basin, the San Joaquin basin contains thick (< 9 km)
accumulations of Campanian-Quaternary strata (Bartow, 1991; Williams and Graham,
233

2013). Sediment delivered to the Great Valley forearc was primarily derived from the
adjacent Klamath Mountains, Sierran Foothills belt, and Sierra Nevada volcano-plutonic
arc (Ingersoll, 1983; DeGraaff-Surpless et al., 2002) with lesser input from the
Franciscan subduction complex that became emergent during latest Cretaceous to
Paleogene time (Dickinson et al., 1979; Schulein, 1993; Mitchell et al., 2010).
Salinian block forearc
Locally thick (up to ~7 km) successions of upper Campanian (?) to Paleogene
strata are discontinuously preserved in the California Coast Ranges west of the San
Andreas fault (Fig. 2; Graham, 1976a,b; Vedder et al., 1982; Grove, 1993). These strata
were primarily deposited within shallow- and deep-marine basins (e.g., Grove, 1993) that
were deposited atop or incised into deeply denuded plutonic rocks of the eastern midCretaceous batholithic belt (i.e., Salinian block; Compton, 1960; Hansen and Stuk, 1993;
Kidder et al., 2003). This configuration is atypical of forearc depocenters to the north and
south (i.e., Great Valley and Peninsular Ranges forearc segments) that are positioned
between the western margin of the volcano-plutonic arc and the subduction complex
(Dickinson, 1995a). For this reason, some authors have argued that these strata should not
be considered true forearc deposits (e.g., Saleeby, 2003). However, these strata may be
more properly considered to be intramassif forearc deposits (sensu Dickinson and
Seely, 1979) and have modern and ancient counterparts in regions that have undergone
extensive subduction erosion (e.g., Laursen et al., 2002; Fildani et al., 2008).
The paleogeographic context under which the Salinian block forearc developed is
controversial. The Salinian block has variously been described as 1) an east-west
extensional basin with steep gradients that formed within a forearc graben (Grove, 1993)
or 2) a continental borderland characterized by locally uplifted structural blocks and
adjacent deep-marine basins that variously resulted from pre-Eocene dextral offset along
a proto-San Andreas fault (Nilsen and Clarke, 1975), northwest-directed thrusting (Hall,
1991), or orogenic collapse along low-angle normal faults (Saleeby, 2003). Forearc strata
are inferred to be derived from locally uplifted exposures of the Salinian block (e.g.,
Critelli and Nilsen, 1996) and/or the adjacent Mojave Desert region (e.g., Grove, 1993).
Many researchers have considered the Salinian block to have been structurally emplaced
northward against the southern San Joaquin basin in Paleogene time (Nilsen and Clarke,
234

1975; Graham, 1978; Dickinson et al., 2005) based on correlated sedimentary bodies
(Graham et al., 1989) and basement features (Ross, 1984). Subsequent research has
suggested that Salinian and San Joaquin depocenters were not likely contiguous (Graham
and Berry, 1979; Seiders and Cox, 1992; Doebbert et al., 2012; Sharman et al., 2013).
Transverse Ranges forearc
Much of the western Transverse Ranges is comprised of thick successions of
uppermost Jurassic to Paleogene forearc strata that were deposited in a wide range of
fluvial, coastal, shelfal, and deep-marine depositional environments (Chipping, 1972;
Dickinson, 1995b). The western Transverse Ranges consist of a series of east-west
oriented crustal blocks that have undergone over 90 degrees of transrotation since ca. 16
Ma associated with microplate capture during the development of the San Andreas
transform plate boundary (Atwater, 1989; Luyendyk, 1991; Nicholson et al., 1994;
Dickinson, 1996). Prior to tectonic rotation, the Transverse Ranges forearc is inferred to
have been contiguous with better preserved forearc segments to the north and south
adjacent to the Sierran and Peninsular Ranges batholiths, respectively (Dickinson,
1995b). Facies relationships suggest a general deepening toward the west (or south prior
to transrotation; Dickinson, 1996) where a relatively complete uppermost Jurassic to
Paleogene succession overlies Coast Range ophiolite and Franciscan subduction complex
(Dickinson, 1995b). More proximal portions of the forearc are preserved where forearc
strata overlie crystalline rocks of the batholithic belt and its associated framework (e.g.,
San Gabriel block; Dickinson, 1995b).
Peninsular Ranges forearc
A relatively thin (~1-2 km) sequence of Cenomanian(?) to Paleogene forearc
strata lap onto the western margin of the Peninsular Ranges batholith and its framework
and include exposures in the northern Santa Ana Mountains, San Diego area, and from
Tijuana to El Rosario in Baja California (Kennedy and Moore, 1971; Schoellhamer et al.,
1981; Kies and Abbott, 1982, 1983; Bottjer and Link, 1984). Forearc strata in the
Channel Islands, Simi Hills, and Santa Monica Mountains are thought to be have been
originally positioned offshore the northern Peninsular Ranges forearc prior to postOligocene transrotation and are considered to represent distal equivalents of the in situ
forearc exposures of the northern Santa Ana Mountains and San Diego area (Fig. 2; Kies
235

and Abbott, 1982; Bottjer and Link, 1984). Forearc sedimentation initiated with
deposition of Cenomanian(?)-Turonian strata in both nonmarine (proximal) and marine
(distal) environments (Bottjer and Link, 1984). An unconformity or condensed section
separates these strata from a pulse of dominantly marine sedimentation that occurred
during middle Campanian to early Maastrichtian time (Kennedy and Moore, 1971;
Bottjer and Link, 1984). Sandstone petrography and conglomerate clast compositions
indicate that the forearc was primarily derived from the adjacent, dissected Peninsular
Ranges batholith during this time period (Nilsen and Abbott, 1981; Kies and Abbott,
1982; Girty, 1986). A regional unconformity separates Cretaceous from overlying upper
Paleocene-Eocene strata that developed along and offshore an extensive coastal plain that
was supplied by rivers that traversed the northern Peninsular Ranges batholith (Peterson
and Nordstrom, 1970; Kies and Abbott, 1982). Evidence for this drainage is provided by
incised valleys that are filled with Eocene fluvial deposits (Ballenas Gravels) that
supplied distinctive clast assemblages to conglomerates deposited along the margin
(Fairbanks, 1893; Kies and Abbott, 1982; Steer and Abbott, 1984). Distinctive rhyodacitic clasts in the Eocene Poway Group have been linked with Jurassic sources in
north-central Sonora, Mexico, suggesting that the trans-Peninsular Ranges drainage
extended at least 300 km eastward prior to the opening of the Gulf of California during
Late Miocene time (Abbott and Smith, 1978, 1989).
SAMPLE LOCATIONS AND DESCRIPTIONS
Central Sierra Nevada
The Ione Formation consists of fluvial non-marine and marginal marine strata that
onlap the western flank of the Sierra Nevada Foothills Belt (Allen, 1929; Creely and
Force, 2007) and is considered to be the basinal equivalent of the fluvial auriferous
gravels present widely in the northern Sierra Nevada (e.g., Cecil et al., 2010; Cassel et
al., 2012). A middle Eocene age assignment is based primarily on macrofossils and
correlation with the lower middle Eocene Domengine Sandstone in the subsurface of the
Sacramento basin (Creely and Force, 2007).
GVG09-08

236

Ione Formation (Allen, 1929). Unit Ec of Ludington et al. (2007). Collected along
Highway 104. Middle Eocene. 38.3459N, -120.9221W
GVG09-07
Ione Formation (Allen, 1929). Unit Ec of Ludington et al. (2007). Collected along
Highway 88. Middle Eocene. 38.3256N, -120.9174W
GVG09-09
Ione Formation (Allen, 1929). Unit Ec of Ludington et al. (2007). Collected near
Camanche Reservoir. Middle Eocene. 38.2020N, -120.9049W
Mount Diablo
A thick succession of Upper Cretaceous and lower Tertiary forearc strata are
exposed along the flanks of Mount Diablo (Colburn, 1961; Cherven, 1983a; Graymer et
al., 1994; Sullivan and Sullivan, 2012). Over 4 km of undifferentiated Upper Cretaceous
Great Valley Group are unconformably overlain by a similar thickness of Paleocenemiddle Eocene strata (Moxon, 1990). While the succession is punctuated by several
prominent unconformities and transgressive-regressive cycles (e.g., Sullivan and
Sullivan, 2012), the succession generally records shoaling from lower bathyal to neritic
and finally non-marine deposition (Moxon, 1990). All but three of our samples come
from exposures along the northeast flank of Mount Diablo where the succession is more
complete than on the southwest flank (Graymer et al., 1994).
GVG09-0.5
Domengine Sandstone. Unit Td of Graymer et al. (1994). Collected from the Black
Diamond Mines Regional Preserve. Fluvial non-marine and marginal marine (deltaic,
paralic, and estuarine; Bodden, 1983; Cherven, 1983b; Sullivan and Sullivan, 2012).
Early to middle Eocene (Moxon, 1990; Kimyai, 1993; Sullivan and Sullivan, 2012).
37.9550N, -121.8655W
GVG09-22
Domengine Sandstone. Unit Tdu of Graymer et al. (1994).Collected from the
southwestern flank of Mount Diablo. Upper to middle bathyal (Moxon, 1990). Early to
middle Eocene (Moxon, 1990; Kimyai, 1993; Sullivan and Sullivan, 2012). 37.8838N, 121.9910W
237

GVG09-23
Domengine Sandstone. Unit Tdu of Graymer et al. (1994). Collected from the
southwestern flank of Mount Diablo. Upper to middle bathyal (Moxon, 1990). Early to
middle Eocene (Moxon, 1990; Kimyai, 1993; Sullivan and Sullivan, 2012). 37.8852N, 121.9919W
GVG09-24
Domengine Sandstone. Unit Tdu of Graymer et al. (1994). Collected from the
southwestern flank of Mount Diablo. Upper to middle bathyal (Moxon, 1990). Early to
middle Eocene (Moxon, 1990; Kimyai, 1993; Sullivan and Sullivan, 2012). 37.8507N, 121.9335W
GVG09-17
Meganos Sandstone, upper D member. Unit Tmd of Graymer et al. (1994). Shallow
marine (mid-neritic; Moxon, 1990). Early Eocene (Cherven, 1983a; Moxon, 1990;
Sullivan and Sullivan, 2012). 37.9241N, -121.7781W
GVG09-16
Meganos Sandstone, lower A member. Unit Tma of Graymer et al. (1994). Bathyal,
submarine canyon fill (Moxon, 1990). Late Paleocene-early Eocene (Cherven, 1983a;
Moxon, 1990; Sullivan and Sullivan, 2012). 37.9092N, -121.7650W
GVG09-15
Deer Valley Formation (Colburn, 1961). Unit Kdv of Graymer et al. (1994). Marginal to
shallow-marine (deltaic to inner neritic; Moxon, 1990). Maastrichtian (Goudkoffs C
zone; Moxon, 1990). 37.9071N, -121.7634W
GVG09-14
Undifferentiated Great Valley Group. Unit Kcm of Graymer et al. (1994). Sampled from
an outcrop along the north side of the road close to biostratigraphic sample locality ICP
575 (Malmblorg et al., 2008).Bathyal (Moxon, 1990). Probably Campanian to
Maastrichtian (Moxon, 1990). 37.8888N, -121.8104W
GVG09-13
Undifferentiated Great Valley Group. Unit Kbsh of Graymer et al. (1994). Bathyal
(Moxon, 1990). Sampled close to biostratigraphic sample locality ICP 576 (Malmblorg et
al., 2008) that indicated Goudkoff (1945) E to G zones (Coniacian to Campanian;
238

Williams and Graham, 2013). Probably Campanian (Moxon, 1990). 37.8851N, 121.8128W
GVG09-12
Undifferentiated Great Valley Group. Unit Kb of Graymer et al. (1994). Lower bathyal
(Moxon, 1990). Turonian to Albian and probably Goudkoff (1945) H zone (Turonian;
Moxon, 1990). 37.8852N, -121.8356W
Del Puerto Canyon
A succession of Albian to Eocene forearc strata is exposed along the western
flank of the northern San Joaquin basin within Del Puerto Canyon (Bartow et al., 1985).
Our sample from the Panoche Formation (GVG09-2) was collected from a conglomeratic
interval present in the lower portion of the unit that is otherwise dominated by shale. A
small grain age population ca. 85-82 Ma suggests that this unit is at least Santonian to
early Campanian in age. This finding is consistent with an inferred Turonian(?)Campanian age of the Panoche Formation in this area (Bartow et al., 1985). A detrital
zircon sample published by Jacobson et al. (2011) from the uppermost Panoche
Formation also has ca. 81 Ma zircon. Taken together, these results suggest that a large
portion of the Panoche Formation in Del Puerto Canyon (including a minimum thickness
of 3.5 kilometers; Bartow et al., 1985) is likely Santonian to Campanian in age.
The Panoche Formation is overlain by the uppermost Cretaceous Moreno
Formation (not sampled) and the Paleogene Tesla Formation. A Paleocene age
assignment for the Tesla Formation is supported by both megafossil and diatom
assemblages collected from exposures in the Del Puerto Canyon area, although the Tesla
Formation ranges into the Eocene farther to the south in its type section (Bartow et al.,
1985). The Tesla Formation is overlain by the lower to middle Eocene Domengine
Sandstone and Kreyenhagen Shale (Bartow et al., 1985), although we did not sample
these units.
GVG09-1
Tesla Formation. Unit Tt of Bartow et al. (1985). Marginal marine to shallow marine
(Dickinson et al., 1979). Paleocene-early Eocene (Bartow et al., 1985). 37.4803N, 121.2111W
239

GVG09-2
Panoche Formation. Unit Kpc of Bartow et al. (1985). Deep-marine (Ingersoll, 1979).
Turonian to Campanian (Bartow et al., 1985), but likely Santonian or Campanian based
on the presence of several 82-85 Ma detrital zircon grains. 37.4527N, -121.2756W
Vallecitos syncline
The Vallecitos syncline is located in western margin of the central San Joaquin
basin, approximately 50 km north of Coalinga. The syncline is flanked on both sides by
uplifted Franciscan Complex exposed in the core of the Diablo Range to the north and in
the Joaquin Ridge to the south. A thick sequence of Upper Cretaceous to Eocene forearc
strata are preserved along both limbs of the Vallecitos syncline (Nilsen, 1981). The
Upper Cretaceous (Cenomanian-Campanian) succession was sampled by DeGraaffSurpless et al. (2002) along the nearby Joaquin Ridge. Overlying Paleocene-Eocene
forearc strata obtain thicknesses up to ~2,300 m and include a range of lithologies and
sedimentary facies deposited in a variety of depositional environments (Nilsen, 1981).
The Maastrichtian-Paleocene Moreno Formation (McGuire, 1988) is overlain by the
deep-marine, Paleocene-lower Eocene Lodo Formation (Anderson, 1998), the neritic to
terrestrial, lower-middle Eocene Domengine Sandstone (Schulein, 1993), and the deepmarine, middle Eocene Kreyenhagen Shale (Schulein, 1993).
The Lodo Formation can further be differentiated into members that include (base
to top) the San Carlos Sandstone Member, Cerros Shale Member, Cantua Sandstone
Member, and Arroyo Hondo Shale Member (Nilsen, 1981; Anderson, 1998). The San
Carlos Sandstone is inferred to be late Paleocene in age based on its stratigraphic position
between shale assigned to the California calcareous nannoplankton zone CP4 (60.5-59.2
Ma) and the Cerros Shale Member that is as old as zone CP8B (56.2-54.5 Ma; Anderson,
1998). The Cantua Sandstone Member is constrained as early Eocene in age by its
position above the Cerros Shale that is as young as CP9B (54.1-53.2 Ma) and below the
Arroyo Hondo Shale Member that has faunas of Laiming (1939) benthic formainiferal
zone C or middle CP10 zone (ca. 51.2 Ma; Almgren et al., 1988; Anderson, 1998). The
Cantua Sandstone Member was originally interpreted to be correlative with the Eocene
German Rancho Formation of the Gualala block (Nilsen and Clarke, 1975) but was
240

subsequently shown to be derived from the east via a submarine canyon cut into the shelf
edge and sourced from the Sierra Nevada batholith (Graham and Berry, 1979; Anderson,
1998). Our detrital zircon results support the latter interpretation.
The Lodo Formation is unconformably overlain by the Domengine Sandstone,
which was deposited during a time of major uplift and erosion of the western margin of
the central San Joaquin basin (Schulein, 1993). The Domengine Sandstone is early to
middle Eocene in age (ca. 49.6-47.9 Ma) based on B-1 benthic foraminiferal assemblages
(Almgren et al., 1988) and a CP12a age assignment (Schulein, 1993). The Domengine
Sandstone was deposited in a variety of depositional environments that include deltaic,
estuarine, and terrestrial (alluvial and fluvial) facies (Schulein, 1993). Sandstone
petrography, conglomerate clast compositions, and paleodispersal directions suggest that
the Domengine Sandstone was derived from multiple sources, including 1) underlying
Great Valley forearc strata, 2) Franciscan subduction complex and overlying Coast Range
ophiolite, and 3) the Sierra Nevada magmatic arc (Schulein, 1993). Our sample of the
Domengine Sandstone (SG-20) was collected from a facies thought to have been derived
from the Sierra Nevada batholith region. Our results corroborate this interpretation.
Deposition of the siliceous Kreyenhagen Shale over the Domengine Sandstone
coincides with a regional flooding event and marks the return of the bathyal conditions in
the Vallecitos Syncline (Nilsen, 1981; Milam, 1985; Bartow, 1991; Schulein, 1993).
SG-20
Domengine Sandstone (Schulein, 1993). Unit Td of Dibblee (2007). Collected along New
Idria Road from the uppermost part of the Domengine Sandstone. Deltaic to estuarine
(Schulein, 1993). Early-middle Eocene (ca. 49.6-47.9 Ma; Schulein, 1993). 36.4441N, 120.6640W
SG-21
Cantua Sandstone Member of the Lodo Formation (Nilsen, 1981; Anderson, 1998). Unit
Tlc of Dibblee (2007). Collected along New Idria Road from the central portion of the
unit. Deep-marine (Anderson, 1998). Early Eocene (ca. 54.1-50.4 Ma; Anderson, 1998).
36.4257N, -120.6693W
SG-22
San Carlos Sandstone Member of the Lodo Formation (Anderson, 1998), or unnamed
241

sandstone member of the Lodo Formation (Dibblee and Nilsen, 1974; Nilsen et al.,
1974). Unit Tlc of Dibblee (2007). Collected along New Idria Road. Deep-marine
(Anderson, 1998). Late Paleocene (ca. 58.3-54.1 Ma; Anderson, 1998). 36.4175N, 120.6704W
San Francisco Bay block
The San Francisco Bay block is bordered by the San Andreas fault on its
southwestern margin and by the Hayward and Calaveras faults on its northeastern margin
(McLaughlin et al., 1996). The San Francisco Bay block is part of the East Bay fault
system and has accumulated as much as ~175 km of right-lateral offset since Miocene
time (Graymer et al., 2002). The majority of inferred displacement has occurred along the
Hayward and Calaveras faults (McLaughlin et al., 1996; Graymer et al., 2002), although
lesser estimates of fault displacement across the East Bay fault system have also been
proposed (~35 10 km; Graham et al., 1984).
Campanian to Eocene forearc strata are discontinuously exposed within the San
Francisco Bay block in several geographic areas, 1) San Bruno Mountain (Snow et al.,
2010), 2) in the vicinity of the cities of Redwood City, Atherton, Woodside, Menlo Park,
and near the campus of Stanford University (Page and Tabor, 1967; Beaulieu, 1970;
Brabb and Pampeyan, 1983; Brabb et al., 2000), 3) in the Santa Theresa Hills south of
San Jose, California (Short, 1986; McLaughlin et al., 2001), and 4) in the Sierra Azul
block (McLaughlin et al., 2001, 2004). Although shallow-water (neritic) deposits are
locally present, the majority of Campanian-Eocene strata were deposited in a deepmarine (bathyal) environment (Beaulieu, 1970; Short, 1986; McLaughlin et al., 2004).
Campanian or Eocene strata occur in structural and depositional contact with underlying
Franciscan Complex and ultramafic rocks equivalent to the Coast Range ophiolite
(McLaughlin et al., 2001, 2004). In general, Campanian strata equivalent to the type
Great Valley Group are restricted to the Sierra Azul block and Santa Theresa Hills, the
latter of which is thought to be a thrust klippe derived from the Sierra Azul block, and are
overlain by lower to middle Eocene strata (Short, 1986; McLaughlin et al., 2001, 2004).
However, the Cretaceous section is usually absent to the northwest where Eocene strata
lie depositionally or structurally on Franciscan Complex in the vicinity of Jasper Ridge
242

(Beaulieu, 1970; Page, 1993) and San Bruno Mountain (Snow et al., 2010). Geologic
relationships suggest that these strata were probably deposited in trench-slope basins
within an actively rising accretionary complex.
Lower to middle Eocene arkosic sandstone and mudstone in the vicinity of
Redwood City, Atherton, Woodside, Menlo Park, and near the campus of Stanford
University are generally referred to as the Whiskey Hill Formation (Beaulieu, 1970;
Pampeyan, 1993). The type section of the Whiskey Hill Formation is located along
Whiskey Hill Road near the town of Woodside, California, and the formation is
approximately 1,300 m thick in this region (Page and Tabor, 1967; Beaulieu, 1970).
Conglomerate and an associated shallow-water fauna are present where the Whiskey Hill
formation lies depositionally on underlying Franciscan Complex in the vicinity of Jasper
Ridge (Beaulieu, 1970; Page, 1993). The majority of the Whiskey Hill formation was
deposited in bathyal water depths, suggesting rapid subsidence and deepening during
Whiskey Hill deposition (Beaulieu, 1970; p. 38). The Whiskey Hill formation likely
ranges from early to middle Eocene in age based on 1) Penutian to Narizian benthic
foraminiferal fauans (ca. 54.6-35.2 Ma; Pampeyan, 1993), 2) coccolith biostratigraphy
that indicate late early Eocene to middle middle Eocene ages (Kheradyar, 1988), 3)
benthic foraminiferal assemblages indicative of the A-2 (Laiming, 1939) zone (middle
Eocene; ca. 47-37 Ma; Graham and Classen, 1955), and 4) occurrence of planktic
foraminieral zones P10-P12 (ca. 48.6-38 Ma; Malmblorg et al., 2008).
SG-4
Whiskey Hill Formation (informal designation; Beaulieu, 1970). Unit Tw of Pampeyan
(1993). Collected from the type section of Whiskey Hill Formation along Whiskey Hill
Road in Woodside, California. Deposited in a deep-marine (bathyal) environment
(Beaulieu, 1970). Probably middle Eocene (ca. 48.6-38 Ma; Beaulieu, 1970; Pampeyan,
1993; Malmblorg et al., 2008). 37.4267N, -122.2487W
SG-23
Greystone Formation (informal designation; Short, 1986) or sandstone and shale of Loma
Chiquita Ridge (McLaughlin et al., 2004). Unit Te2 of McLaughlin et al. (2004).
Collected from fragments of the original Stanford campus buildings (informally called
the Quad Sandstone) that were damaged in the 1906 and 1989 earthquakes and relocated
243

to the Stanford Bone Yard (see Junkerman, 2010; p. 15). The Quad Sandstone was
originally quarried from the Greystone Quarry in the Santa Theresa hills until ca. 1906
when the quarry closed (Junkerman, 2010). Probably middle Eocene (ca. 49-36 Ma)
based on stratigraphic position above the (informal) Bernal Formation that it at least as
young as calcareous nannoplankton zone CP12a (49.3-47.9 Ma; Short, 1986). The
Greystone Formation is poorly fossiliferous, but has a single Turitella cast that is visible
in the Stanford University building stone that indicates an early Eocene age (Short, 1986;
Junkerman, 2010). 37.2233N, -121.8470W
SG-47
Unnamed lower Eocene sandstone of the Sierra Azul block (McLaughlin et al. 2004).
Unit Tcm of McLaughlin et al. (2001) and Te1 of McLaughlin et al. (2004). Collected
along a dirt road on the northeast flank of Mount Chual near a telecommunications tower.
The sample was collected from medium-grained gray sandstone adjacent to a white gate.
Deposited in a deep-marine (lower bathyal-abyssal) environment (McLaughlin et al.,
2004). Early Eocene (Penutian) in age (ca. 54.6-49.6 Ma) based on stratigraphic position
immediately overlying an exposure of mottled maroon and green mudstone with
abundant foraminifera (McLaughlin et al., 2004).
37.1184N, -121.8328W
Pilarcitos block
We collected two samples (SG-6 and SG-46) from the Pilarcitos bock, which is
bounded on the northeast by the modern trace of the San Andreas fault and on the
southwest by the Pilarcitos fault (Brabb and Pampeyan, 1983). The Pilarcitos fault is
generally considered to be an older, abandoned segment of the northern San Andreas
fault (Dickinson et al., 2005; McLaughlin et al., 2007), an interpretation that is supported
by the observation that the Pilarcitos fault locally marks the contact between the Salinian
block and Franciscan subduction complex (Graymer et al., 2006). The active trace of the
Peninsular segment of the San Andreas fault only has 26 4 km of slip (Cummings,
1968; Griscom and Jachens, 1989; Jachens and Zoback, 1999; Wakabayashi, 1999;
Jachens et al., 2002), suggesting that transfer of slip from the Pilarcitos fault to the
modern fault location occurred ca. 1.0-1.75 Ma (Dickinson et al., 2005).
244

Our samples from the Pilarcitos block are from the southern portion of the
mapped fault slice where undifferentiated Eocene sandstone and mudstone are variably
mapped as equivalent to the Butano Sandstone (Brabb and Pampeyan, 1983) or Whiskey
Hill Formation (Brabb et al., 2000). Because only ~26 km of slip are inferred to have
occurred on the present trace of the San Andreas fault (Dickinson et al., 2005), the
Pilarcitos Sandstone should be restored to a position south of the Whiskey Hill Formation
and northwest of the unnamed lower Eocene sandstone of the Sierra Azul block. In this
model, the Pilarcitos Sandstone is correlative with other Eocene sandstones of the San
Francisco Bay block that were deposited atop Franciscan subduction complex in trenchslope basins.
However, detrital zircon U-Pb age distributions indicate that the Pilarcitos
Sandstone closely resembles the Butano Sandstone while being distinct from all other
Eocene sandstones deposited atop the San Francisco Bay block (e.g., Figure 3D). The
Kolgomorov-Smirnov (K-S) statistic (Press et al., 1986) is widely used to test whether
two distributions could have been derived from the same parent population (e.g.,
Dickinson et al., 2012). This approach reveals that the Pilarcitos Sandstone and Eocene
sandstones from the San Francisco Bay block (Whiskey Hill Formation, Greystone
Formation, and unnamed lower Eocene sandstone of the Sierra Azul block) are
statistically distinct from each other at a 95% confidence level (P-value 0). When the
Pilarcitos Sandstone is compared with the Butano Sandstone (Sharman et al., 2013), 10 of
12 sample pairs yield statistically indistinguishable results (i.e., P-value > 0.05).
We interpret these results to indicate that the Pilarcitos Sandstone is actually
correlative with the Butano Sandstone, rather than with Eocene strata of the San
Francisco Bay block. By extension this implies that the location of the Pilarcitos fault is
likely mismapped along the southern extent of the Pilarcitos block (Brabb and Pampeyan,
1983; R. McLaughlin, pers. comm.). We prefer a fault geometry that links the Pilarcitos
fault with the modern trace of the San Andreas fault in the vicinity of the southern end of
the Crystal Springs reservoir. This geometry allows exposures of Franciscan assemblages
to occur east of the Pilarcitos fault while the Pilarcitos Sandstone remains attached to
the Salinian block. For the reasons stated above, we include the Pilarcitos Sandstone
samples with the Butano Sandstone in our reconstructions (e.g., Figure 7).
245

SG-6
Pilarcitos Sandstone (informal designation). Unit Tb of Brabb and Pampeyan (1983).
Collected along California Route 84. Inferred to be deep-marine. Probably early to
middle Eocene (ca. 54-38 Ma) based on correlation with the Butano Sandstone (Poore
and Brabb, 1977; Seiders and Cox, 1992). 37.3995N, -122.2594W
SG-46
Pilarcitos Sandstone (informal designation). Unit Tb of Brabb and Pampeyan (1983).
Collected along Kings Mountain Road near Huddart County Park. Inferred to be deepmarine. Probably early to middle Eocene (ca. 54-38 Ma) based on correlation with the
Butano Sandstone (Poore and Brabb, 1977; Seiders and Cox, 1992). 37.4330N, 122.2944W
San Emigdio Mountains
We collected an additional sample (SG-40) from the upper sandstone member of
the San Emigdio Formation in the San Emigdio Mountains to supplement previously
published data from this region (Lechler and Niemi, 2011; Sharman et al., 2013).
Narizian (late middle Eocene) benthic foraminiferal faunas (ca. 42.5-35.2 Ma) have been
reported from shale from the middle San Emigdio Formation (DeLise, 1967), although
megafossils may indicate a Refugian (late Eocene) age for this formation (Nilsen, 1987).
Thus, the San Emigdio Formation is likely late middle to late Eocene in age (ca. 42.5-34
Ma). Shallow-marine deposition is indicated by both megafossils and sedimentary facies
and structures that may represent intertidal (Nilsen et al., 1973) or nearshore (DeCelles,
1988) deposition.
SG-40
San Emigdio Formation (upper sandstone member). Unit Tss of Dibblee (2005).
Collected along San Emigdio Canyon within the Wind Wolves Preserve. Shallow-marine
(Nilsen et al., 1973; DeCelles, 1988). Late middle to late Eocene (ca. 42.5-34 Ma;
DeLise, 1967, Nilsen, 1987). 34.9145N, -119.1628W

246

Santa Lucia Range Reliz Canyon


Three samples (SG-31, SG-28, and SG-29) were collected from the Rocks
Sandstone (Link and Nilsen, 1979; Mason, 1998; Mason and Graham, 2000) that
outcrops in Reliz Canyon, northern Santa Lucia Range. The Rocks Sandstone is up to 640
m thick and is composed of amalgamated, medium-grained sandstone with lesser
interbedded mudstone and conglomerate (Link and Nilsen, 1979). Sedimentary facies and
structures indicate channelized or braid-plain turbidite deposition in a deep-marine
environment (Link and Nilsen, 1979; Mason, 1998). Ulatisian to Narizian benthic
foraminiferal fauans indicate a middle Eocene age (ca. 49.6-35.2 Ma) for the Rocks
Sandstone (Link and Nilsen, 1979).
SG-31
Rocks Sandstone. Unit Tr of Dibblee (2006). Collected from the uppermost thick sand
package near the contact with the overlying Berry Formation. Deep-marine (Link and
Nilsen, 1979). Middle Eocene (ca. 49.6-35.2 Ma; Link and Nilsen, 1979). 36.1632N, 121.2934W
SG-28
Rocks Sandstone. Unit Tr of Dibblee (2006). Collected from near the middle of the
Rocks Sandstone. Deep-marine (Link and Nilsen, 1979). Middle Eocene (ca. 49.6-35.2
Ma; Link and Nilsen, 1979). 36.1579N, -121.2919W
SG-29
Rocks Sandstone. Unit Tr of Dibblee (2006). Collected from the lowermost thick sand
package near the contact with the underlying Lucia Mudstone. Deep-marine (Link and
Nilsen, 1979). Middle Eocene (ca. 49.6-35.2 Ma; Link and Nilsen, 1979). 36.1558N, 121.2862W
Northern Santa Ana Mountains
Upper Cretaceous, Paleocene, and Eocene samples were collected from the
northern Santa Ana Mountains in and near Silverado Canyon using the excellent geologic
map of Schoellhamer et al. (1981) as a guide. The Upper Cretaceous-Eocene forearc
succession rests depositionally atop Upper Jurassic and Lower Cretaceous rocks of the
Bedford Canyon Formation and Santiago Peak Volcanics (Schoellhamer et al., 1981;
247

Balch et al., 1984). The sample names in parentheses are those used in Jacobson et al.
(2011) and are included for reference.
TRB (Trabuco)
Trabuco Formation (Schoellhamer et al., 1981). Collected in Silverado Canyon from the
west cutbank of the parking lot at the Silverado Post Office, ZIP code 92676. Sample
collected from a sandstone lens within pebble cobble conglomerate. Non-marine alluvial
fan deposits (Bottjer and Link, 1984). The Trabuco Formation is unfossiliferous but its
upper portion interfingers with the overlying Baker Canyon Conglomerate Member thus
indicating an early Late Cretaceous age. In Silverado Canyon, the upper Trabuco
Formation may be as young as Turonian. An estimated depositional age for the sampled
stratum is 93 Ma. 33.7475N, -117.6381W
BAK (Baker Canyon)
Baker Canyon Conglomerate Member of the Ladd Formation (Schoellhamer et al., 1981).
Collected from the type section of the Baker Canyon member of the Ladd Formation
(Popenoe, 1937) immediately west of the paved road at the mouth of Ladd Canyon just
north of Silverado Canyon Road. Collected from medium to conglomeratic sandstone.
Non-marine to shallow marine (Popenoe, 1941; Schoelhamer et al., 1981). The Baker
Canyon Conglomerate Member of the Ladd Formation contains abundant thick-shelled
molluscs of Turonian age (Popenoe, 1941). A late Turonian age is suggested by the
cosmopolitan ammonite genus Sunprionocyclus. An estimated depositional age for the
sample is 90 Ma. 33.7475N, -117.6414W
MUST (Mustang)
Holtz Shale Member of the Ladd Formation (Schoellhamer et al., 1981). Collected from
the Mustang Springs (informal designation) conglomerate lens 1.3 km due north of the
Silverado School in Silverado Canyon from a prominent, well-bedded cliff face (see
geologic map of Schoelhamer et al., 1981). The Holtz Shale Member includes both slope
and deep-marine facies in its lower part and shelfal facies in its upper portion (Bottjer and
Link, 1984). Fossils in the enclosing Holz Shale member suggest a middle Campanian
age for the Mustang Springs Lens, probably between 79 to 75 Ma. An estimated
depositional age for the analyzed sample is 78 Ma. 33.7589N, -117.6425W
WLM (Williams)
248

Shultz Ranch Sandstone Member of the Williams Formation (Schoellhamer et al., 1981).
Collected 0.3 km west-northwest of the old Silverado. Shallow marine and fan delta
facies (Cooper et al., 1982). A reversed magnetic polarity interval measured near this
locality by Fry et al. (1985) suggests a Campanian depositional age of 74.5 Ma.
33.7572N, -117.6558W
SLVL (L Silverado)
Silverado Formation (Schoellhamer, 1981). Collected from the lower Silverado
Formation on the north side of Silverado Canyon Road beneath marker bed
(Schoellhamer et al., 1981). Non-marine (Kies and Abbott, 1982). Late Paleocene
(Popenoe, 1941; Schoellhamer et al., 1981). 33.7515N, -117.6522W
SLVU (U Silverado)
Silverado Formation (Schoellhamer, 1981). Collected from the upper Silverado
Formation on the north side of Silverado Canyon Road above the upper marker bed
(Schoellhamer et al., 1981). Non-marine (Kies and Abbott, 1982). Late Paleocene
(Popenoe, 1941; Schoellhamer et al., 1981). 33.7573N, -117.6693W
SNTGL (L Santiago)
Santiago Formation (Schoellhamer, 1981). Collected from the lower interval of the
Santiago Formation from a roadcut on the south side of Silverado Canyon Road near the
easternmost extent of the Santiago Formation outcrop (see geologic map of Schoellhamer
et al., 1981). Grades from marine in the lower portion to non-marine towards the upper
portion of the formation (Yerkes et al., 1965). Late early to early middle Eocene (Yerkes
et al., 1965; Schoellhamer et al., 1981; Kies and Abbott, 1982). 33.7444N, -117.6619W
SNTGU (U Santiago)
Santiago Formation (Schoellhamer, 1981). Collected from the upper interval of the
Santiago Formation from the southernmost outcrop crossing Santiago Canyon road. The
sample was collected about 12 meters above the roadbed on the northeast side of the
road. Grades from marine in the lower portion to non-marine towards the upper portion
of the formation (Yerkes et al., 1965). Late early to early middle Eocene (Yerkes et al.,
1965; Schoellhamer et al., 1981; Kies and Abbott, 1982). 33.7409N, -117.7409W

249

San Diego Area


A suite of Cretaceous, Paleocene, and Eocene samples were collected in the San Diego
area from coastal outcrops and from the nearby foothills. Sample sites are best located
using the geologic maps of Kennedy (1975).
ROSA-5 (Lusardi-Carlsbad)
Lusardi Formation(?) or Point Loma Formation of the Rosario Group(?) (Kennedy and
Moore, 1971). Collected near the Palomar Airport in Carlsbad, California from massive,
bedded sandstone in a bank off a dirt road on the north side of Hedionda Canyon. This
sample was collected in an area mapped as Lusardi (Kennedy, 1975) but the sedimentary
facies we sampled appeared to be more likely related to the overlying, marine Point
Loma Formation. Thus, possibly as old as Cenomanian-Turonian, but probably middle
Campanian to early Maastrichtian (Kennedy and Moore, 1971; Bottjer and Link, 1984).
33.1435N, -117.2731W
ROSA-4 (Rosario-Carlsbad)
Point Loma Formation of the Rosario Group (Kennedy and Moore, 1971). Collected near
the Palomar Airport in Carlsbad, California from strongly indurated, fossiliferous
sandstone above bedded grey mudstone that outcrops along El Camino Real. Inferred to
be middle Campanian to early Maastrichtian (Kennedy and Moore, 1971; Bottjer and
Link, 1984).
33.1414N, -117.2770W
ROSA-3 (Rosario-La Jolla Bay)
Point Loma Formation of the Rosario Group (Kennedy and Moore, 1971). Collected from
a west-facing cliff about 1 meter above the beach in the first small cove south of the La
Jolla Beach and Tennis Club. This cove lies immediately south of the Mount Soledad
fault on the La Jolla quadrangle map of Kennedy (1975). Dominantly slope and deepmarine facies (Nilsen and Abbott, 1981). Late Campanian in age based on foraminifera
(Sliter, 1968) and coccoliths (Bukry and Kennedy, 1969), and ammonites of the
Metaplacenticeras pacificum zone (Mike Mickey, pers. comm.). The strata are probably
in magnetic polarity Chron 33N. An estimated depositional age of the sampled stratum is
77 Ma. 32.8506N, -117.2631W

250

ROSA-2 (Rosario-Bird Rock)


Point Loma Formation of the Rosario Group (Kennedy and Moore, 1971). Collected from
the northwest side of a small point of Cretaceous sandstone extending into the ocean just
south of Bird Rock Avenue. Site is due east of Bird Rock label on Kennedy map of La
Jolla Quadrangle (1975). Dominantly slope and deep-marine facies (Nilsen and Abbott,
1981). Late Campanian in age based on foraminifera (Sliter, 1968) and coccoliths (Bukry
and Kennedy, 1969), and ammonites of the Metaplacenticeras pacificum zone (Mike
Mickey, pers. comm.). An estimated depositional age is 73 Ma. 32.8147N, 117.2733W
ROSA-1 (Rosario-Tourm Bch)
Cabrillo Formation of the Rosario Group (Kennedy and Moore, 1971). Collected from
Tourmaline Beach in the south-facing sea cliff ~1.5 m above the beach sand east of False
Point. Site lies just northwest of dashed fault shown crossing La Jolla Boulevard on
Kennedy map of La Jolla Quadrangle (1975). Deep-marine inner and middle fan
deposits (Nilsen and Abbott, 1981). Likely uppermost Campanian to early Maastrichtian
in age (Sliter, 1968; Kennedy and Moore, 1971). An estimated depositional age is 72 Ma.
32.8072N, -117.2636W
JOLLA-4 (Mt. Soledad-EPTM)
Mount Soledad Formation of the La Jolla Group (Kennedy and Moore, 1971). Collected
~1.5 m above beach sand from the west-facing beach cliff on the south side of Indian
Trail Canyon just north of the Torrey Pines Glider Port. This site is northwest of the Salk
Institute (see geologic map of the Del Mar Quadrangle by Kennedy, 1975, or photo of
site in Abbott, 1999, p. 76). The sampled strata contain Poway rhyolitic clasts that have
been deeply weathered in the C horizon of a paleosol during the Late Paleocene Thermal
Maximum (Peterson and Abbott, 1979). Paralic (estuarine) facies (Kies and Abbott,
1982). Late Paleocene to early Eocene (Kies and Abbott, 1982). An estimated
depositional age is 56 Ma. 32.8955N, -117.2522W
JOLLA-3 (Mt. Soledad)
Mount Soledad Formation of the La Jolla Group (Kennedy and Moore, 1971). Collected
about 5.5 meters higher stratigraphically relative to sample JOLLA-4 from the paleosol A
horizon (whitish color, kaolinite-rich stratum; see photo in Abbott, 1999, p. 76). Paralic
251

(estuarine) facies (Kies and Abbott, 1982). Late Paleocene to early Eocene (Kies and
Abbott, 1982). An estimated depositional age is 56 Ma. 32.8955N, -117.2522W
JOLLA-2 (Delmar-Bathtub Rock)
Delmar Formation of the La Jolla Group (Kennedy and Moore, 1971). Collected from the
south side of Bathtub Rock (also called Flat Rock) about 0.5 m above beach sand at the
southern end of Torrey Pines State Park (see photo of site in Abbott, 1999, p. 109).
Marginal marine (barrier bar and lagoonal) facies (Kennedy and Moore, 1971; Kies and
Abbott, 1982). Late early Eocene in age based on coccoliths and pollen (Frederiksen,
1991). An estimated depositional age is 50 Ma. 32.9136N, -117.2575W
JOLLA-5 (Scripps-Tourm Bch)
Scripps Formation of the La Jolla Group (Kennedy and Moore, 1971). Collected from a
sand lens in a dominantly conglomeratic sequence immediately north of the parking lot at
Tourmaline Beach. Middle to late Eocene (Kennedy and Moore, 1971). 32.8055N, 117.2622W
JOLLA-1 (Scripps-Pat)
Scripps Formation of the La Jolla Group (Kennedy and Moore, 1971). Collected ~2.5 m
above beach sand from the west-facing sea cliff south of Bathtub Rock and north of the
Torrey Pines Landslide (see Del Mar Quadrangle map in Kennedy, 1975). On p. 111 of
Abbott, 1999, the sample site is just south of the right-hand side of the photo. Deepmarine (submarine canyon fill) facies (Kies and Abbott, 1982). The sample site is within
the Rhabdosphaera inflata coccolith subzone (May, 1982). An estimated depositional
age is 49 Ma. 32.9135N, -117.2580W
POW-2 (Stadium Cgl)
Stadium Conglomerate of the Poway Group (Kennedy and Moore, 1971). Collected from
pebbly sandstone lens within conglomerate of the Poway Group just northwest of the
stadium parking lot. Alluvial fan and fan delta facies (Kies and Abbott, 1982). Inferred to
be middle to late Eocene in age (Kennedy and Moore, 1971). 32.7793N, -117.1161W
POW-1
Pomerado Conglomerate of the Poway Group (Kennedy and Moore, 1971; Kies and
Abbott, 1982). Collected approximately 3 meters up a north-facing roadcut above 1,000
feet elevation about 4.3 km east of the eastern end of Miramar Reservoir (see Poway
252

Quadrangle map in Kennedy, 1975). Alluvial fan facies (Kies and Abbott, 1982).
Microvertebrates collected by Steve Walsh (pers. comm.) at this stratigraphic level
suggest a depositional age of 36 Ma. A single young detrital zircon grain (36.2 1.4 Ma)
supports a late Eocene age assignment. 32.9186N, -117.0582W
San Miguel Island
San Miguel Island Cretaceous samples were collected from the sites adjacent to the
microfossil localities of Miller (1983) that occur along the northwestern shore and cliffs
of the island. Generally agreed upon formation names do not exist for the Upper
Cretaceous and Eocene strata on San Miguel Island.
SMI0401 (SMI1)
Collected at site of Millers (1983) sample 43-35 near the northwestern end of the island.
Miller assigned the coccoliths to the lower part of the Quadrum gartneri zone of the
Turonian. An estimated depositional age is 90 Ma. 34.0366N, -120.4423W
SMI0402 (SMI2)
Collected at site of Millers sample 12-47 about 1.5 km northeast of sample SMI0401.
Miller assigned the coccoliths to the Micula staurophora zone of the ConiacianSantonian. An estimated depositional age is 86 Ma. 34.0470N, -120.4293W
SMI0403 (SMI3)
Collected at site of Millers sample 43-70 about 0.5 km northeast of sample SMI0402.
Miller assigned the coccoliths to the Broinsonia parca zone of the Campanian. An
estimated depositional age is 80 Ma. 34.0501N, -120.4227W
SMI0404 (SMI4)
Collected at site of Millers sample 79-24 about 2.5 km east of sample SMI0403. Miller
assigned the coccoliths to the Quadrum trifidum zone of the early Maastrichtian. An
estimated depositional age is 70 Ma. 34.0500N, -120.3965W
SMI0406 (SMI6)
Collected from Eocene conglomeratic sandstone sandwiched within a mudstone
stratigraphic section about 1.75 km northeast of sample SMI0404. An early Eocene age
is based on forams identified by Robert Arnal and coccoliths identified by David Bukry
(see Kies, 1982). The conglomerates contain abundant Poway rhyolite clasts in
253

percentages similar to the Mount Soledad Formation of San Diego (Abbott, et al., 1983).
An estimated depositional age is 52 Ma. 34.0539N, -120.3863W
Pinyon Mountains
MINE
The Mine Canyon sample was collected from Mine Wash in the Pinyon Mountains of
eastern San Diego county where about 120 m of section is exposed (Lough, 1993). These
deposits appear to the eastern equivalents of the Ballenas Gravels (i.e., were associated
with the large extraregional river system that supplied the rhyo-dacitic clasts to the
Poway Conglomerate on the coastal plain; Fairbanks, 1893; Abbott et al., 1983).
Extraregional clasts in these deposits are all resistant, well rounded, and dominated by
rhyo-dacitic lithologies (up to 48% of clasts) and percussion-marked, whitish quartzites
(up to 33% of clasts; Abbott et al., 1983). The largest rhyodacitic clasts reach long
diameters up to 84 cm. Paleocurrents inferred from imbricated clasts at the outcrop east
of Mine Wash locality indicate generally westward transport, opposite to the present
regional slope and consistent with a mid-Tertiary origin (Axen and Dorsey, unpub. data).
The conglomerate sequence coarsens upward and extraregional clasts become less
common. The upsection decrease in exotic rhyodacitic clasts suggests that the Eocene
steep-walled paleocanyons were reoccupied by younger streams and overland flows.
These later flows of water must have eroded Eocene conglomerate and redeposited
younger gravels containing lesser percentages of Poway rhyolite clasts. Clay-rich
paleosols are present locally in the upper parts of the section showing intervals of time
without major flows of water. Poor age control but inferred to be late Eocene in age
based on a single, young detrital zircon grain (36.3 4.5 Ma). 33.0935N, -116.3463W
Northern Baja California
A suite of Turonian-Eocene samples were collected along the northwestern margin of
Baja California. In this region the Turonian Redonda Formation overlies basement rock
of the Peninsular Ranges batholith (Bottjer and Link, 1984). Overlying strata of late
Campanian to early Maastrichtian age are assigned to the Rosario Group and were
deposited in non-marine, shelfal, slope, and deep-marine settings (Yeo, 1982; Bottjer and
254

Link, 1984). Upper Cretaceous strata are separated by overlying Eocene strata (Las
Palmas Gravels) by a regional unconformity (e.g., Peterson and Nordstrom, 1970). The
Eocene Las Palmas Gravels are sandy conglomerates of fluvial origin that are deposited
within a paleovalley incised into the Peninsular Ranges batholith. The fluvial outcrops
are 65 km long, up to 6 km wide, and 240 m thick (McDonough and Abbott, 1989). The
age of the Las Palmas Gravels is an estimate based on its paleogeographic position as an
incised river valley leading toward the fossiliferous marine Buenos Aires Formation of
middle to late Eocene age in the Tijuana area (~49 34 Ma).
Guad01 (Redonda-Guad-01)
Redonda Formation(?) (Bottjer and Link, 1984). Collected from bedded, fossiliferous
sandstone within a volcanic sequence up canyon from La Mission. Inferred to be
Turonian in age based on stratigraphic position overlying the Santiago Peak Volcanics
and two detrital zircon grains 94.4 4.7 and 88.6 2.8 Ma in age. 32.0980N, 116.7914W
Guad02 (Redonda-Guad-02)
Redonda(?) Formation or Rosario Formation(?) (Bottjer and Link, 1984). Collected in a
canyon about 2.5 kilometers to the north of sample Guad01 from a road cut near a cattle
guard. The sample was collected from sandstone within a bedded, relatively organized
sequence among conglomeratic beds of the Redonda Formation. The inferred age of the
Redonda Formation is Turonian (Bottjer and Link, 1984) but two detrital zircon grains
69.8 1.9 and 72.2 2.6 possibly suggest this sample may be as young as Maastrichtian
(i.e., basal Rosario Formation). 32.1188N, -116.7995W
ROS-1 (Medano)
Rosario Formation (Bottjer and Link, 1984). Collected from the from upper Medano
Valley 5 km east-southeast of the town of Cantamar which lies 40 km south-southeast of
the international border. Conglomeratic sandstone from an amalgamated 2.5m thick
fining upward channel sequence was collected from the basal section of the Rosario
Group in Canon Medano, about 75 meters above the depositional(?) contact with
underlying basement. This unit was interpreted as Redondo Fm of Yeo (1982) which
would be equivalent to the Trabuco north of border (Turonian, Bottjer and Link, 1984),
but detrital zircon grains with ages from 86 to 78 Ma suggest a Campanian or younger
255

age for this sandstone (Rosario Formation; Bottjer and Link, 1984).
32.2599N, -116.8896W
ROS-2 (Medano)
Rosario Formation (Bottjer and Link, 1984). Collected from the from upper Medano
Valley 5 km east-southeast of the town of Cantamar which lies 40 km south-southeast of
the international border. Collected about 50 meters higher stratigraphically than ROS-1
from biotite rich sandstone interbedded with pebble to small cobble conglomerate.
Abundant mollusk shells. Possibly innermost shelf deposits. Inferred to be late
Campanian to early Maastrichtian in age (Bottjer and Link, 1984). 32.2563N, 116.8918W
ROS-3 (Rosario-Salsipuedes)
Rosario Formation (Bottjer and Link, 1984; Yeo, 1984). Collected adjacent to
Salsipuedes Bay along the Highway 1 toll road north of Ensenada and about 100 meters
south of the 91 kilometer sign posting. Middle fan facies (Yeo, 1984). Inferred to be late
Campanian to early Maastrichtian in age (Bottjer and Link, 1984; Yeo, 1984).
31.9614N, -116.7595W
PALM-3 (L Las Palmas)
Las Palmas Gravels (McDonough and Abbott, 1989). Collected from the lowest outcrop
on Mexican Highway 3 at the southernmost outcrop, across the river from El Testerazo,
about 50 kilometers south of Tecate, a large town on the international border. Fluvial
non-marine fill of a paleovalley. Inferred to be late Eocene in age. 32.3129N, 116.5582W
PALM-2 (M Las Palmas)
Las Palmas Gravels (McDonough and Abbott, 1989). Collected midway back up the
highway, but still on the south-facing slope with a view of El Testerazo. Fluvial nonmarine fill of a paleovalley. Inferred to be late Eocene in age. 32.3218N, -116.5645W
PALM-1 (U Las Palmas)
Las Palmas Gravels (McDonough and Abbott, 1989). Collected further back up the
highway near the top of the same south-facing slope. Fluvial non-marine fill of a
paleovalley. Inferred to be late Eocene in age. 32.3298N, -116.5683W

256

El Rosario
A suite of Turonian-lower Maastrichtian samples were collected from the El
Rosario area, Baja California. The La Bocana Roja Formation overlies Peninsular Ranges
basement and is inferred to be Turonian in age (Bottjer and Link, 1984). The Punta Baja
Formation is Campanian in age and was deposited in both shelfal and shallow marine
environments (Bottjer and Link, 1984). The El Gallo and Rosario Formations range in
age from late Campanian to early Maastrichtian and were deposited in fluvial and marine
environments (Bottjer and Link, 1984).
LBR
La Bocana Roja Formation (Bottjer and Link, 1984). Collected from fine-grained
sandstone with shale intraclasts within a predominantly silt-dominated sequence.
Collected from medium-grained, gray, biotite-bearing sandstone. Deposited in a
floodplain environment (Bottjer and Link, 1984). Inferred to be Turonian in age (Yeo,
1982; Bottjer and Link, 1984).
29.9792N, -115.8039W
PB
Punta Baja Formation (Bottjer and Link, 1984). Collected from a medium-grained, tan,
biotite-bearing sandstone bed that is intercalated with conglomeratic facies of the Punta
Baja submarine canyon sequence. Deposited in shelfal and shallow marine environments
(Bottjer and Link, 1984). Campanian (Bottjer and Link, 1984). 29.9806N, -115.8039W
EG-1
El Gallo Formation (Bottjer and Link, 1984). Collected from medium-grained sandstone
from a 10 meter thick sequence of channelized sandstone and minor conglomerate.
Fluvial non-marine (Bottjer and Link, 1984). Late Campanian to early Maastrichtian
(Bottjer and Link, 1984).
30.0542N, -115.7569W
EG-2
El Gallo Formation (Bottjer and Link, 1984). Collected from medium-grained, tan
sandstone from a 2 meter thick tabular bed that outcrops along a dirt road. Fluvial non-

257

marine (Bottjer and Link, 1984). Late Campanian to early Maastrichtian (Bottjer and
Link, 1984).
30.0678N, -115.7769W
RF-1
Rosario Formation (Bottjer and Link, 1984). Collected from medium-grained, biotitebearing sandstone from the Las Caracoles member of the Rosario Formation. Sample was
collected along Highway 1 from an outcrop located across the highway from a Pemex
station in El Rosario. Deposited in shelfal and shallow marine environments (Bottjer and
Link, 1984). Probably Late Campanian (Bottjer and Link, 1984). 30.0589N, 115.7253W
RF-2
Rosario Formation (Bottjer and Link, 1984). Collected from medium-grained, biotitebearing sandstone from the El Viento member of the Rosario Formation. Sample is from
the first resistant sandstone bed beneath the unconformity with the overlying Pliocene,
east of Highway 1 on a spur. Deposited in shelfal and shallow marine environments
(Bottjer and Link, 1984). Probably early Maastrichtian (Bottjer and Link, 1984).
30.0758N, -115.7325W
REFERENCES
Abbott, P.L., 1999, The Rise and Fall of San Diego: Sunbelt Publications, 231 p.
Abbott, P.L., and Smith, T.E., 1978, Trace element comparison of clasts in Eocene
conglomerates, southwestern California and northwestern Mexico: Journal of
Geology, v. 86, p. 753762.
Abbott, P.L., and Smith, T.E., 1989, Sonora, Mexico, source for the Eocene Poway
Conglomerate of southern California: Geology, v. 17, p. 329332.
Abbott, P.L., Kies, R.P., Bachman, W.R. and Natenstedt, C.J., 1983, A tectonic slice of
Eocene strata, northern part of California Continental Borderland, in Larue, D.K.
and Steel, R.J., eds., Cenozoic Marine Sedimentation, Pacific Margin, U.S.A.:
Pacific Section, SEPM (Society for Sedimentary Geologists), p.151-168.
Allen, V.T., 1929, The Ione Formation of California: Berkeley, University of California
Department of Geological Sciences Bulletin, v. 18, p. 347-448.
Almgren, A.A., Filewicz, M.V., and Heitman, H.L., 1988, Lower Tertiary foraminiferal
and calcareous nannofossil zonation of California: An overview and
recommendation, in Filewicz, M.V., and Squires, R.L., eds., Paleogene
Stratigraphy, West Coast of North America: Los Angeles, Pacific Section, SEPM
(Society for Sedimentary Geology), v. 58, p. 83-106.

258

Anderson, K.S., 1998, Facies architecture of two Paleogene structurally-controlled


turbidite systems, central California [Ph.D. thesis]: Stanford University, 391 p.
Atwater, T.M., 1989, Plate tectonic history of the northeast Pacific and western North
America, in Winterer, E.L., Hussong, D.M., and Decker, R.W., eds., The Eastern
Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America,
Geology of North America, v. N, p. 2172.
Balch, D.C., Bartling, S.H., and Abbott, P.L., 1984, Volcaniclastic strata of the Upper
Jurassic Santiago Peak Volcanics, San Diego, California, in Crouch, J.K., and
Bachman, S.B., eds., Tectonics and Sedimentation Along the California Margin,
Pacific Section, SEPM (Society for Sedimentary Geology), v. 38, p. 157-170.
Bartow, J. A., 1991, The Cenozoic evolution of the San Joaquin Valley, California: U.S.
Geological Survey Professional Paper 1501, 40 p.
Bartow, J.A., Lettis, W.R., Sonneman, H.S., and Switzer, J.R., Jr., 1985, Geologic map of
the east flank of the Diablo Range from Hospital Creek to Poverty Flat, San
Joaquin, Stanislaus, and Merced counties, California: U.S. Geological Survey
Miscellaneous Investigation Series, Map I-1656.
Beaulieu, J.D., 1970, Cenozoic stratigraphy of the Santa Cruz Mountains, California and
inferred displacement along the San Andreas fault [Ph.D. thesis]: Stanford
University, 202 p.
Bodden, W.R., 1983, Depositional environments of the Eocene Domengine Formation
outcrop on the north side of Mt. Diablo, California, in Cherven, V.B., and
Graham, S.A., Geology and sedimentology of the southwestern Sacramento basin
and East Bay hills, Pacific Section, SEPM (Society for Sedimentary Geology),
Field Trip Guidebook, p. 43-57.
Bottjer, D.J., and Link, M.H., 1984, A synthesis of Late Cretaceous southern California
and northern Baja California paleogeography, in Crouch, J.K., and Bachman,
S.B., eds., Tectonics and Sedimentation Along the California Margin: Pacific
Section, SEPM (Society for Sedimentary Geology), v. 38, p. 171188.
Brabb, E.E., and Pampeyan, E.H., 1983, Geologic map of San Mateo County, California:
U.S. Geological Survey, scale 1:62000.
Brabb, E.E., Graymer, R.W., and Jones, D.L., 2000, Geologic map and map database of
the Palo Alto 30 x 60 quadrangle, California: U.S. Geological Survey
Miscellaneous Field Studies Map MF-2332b.
Bukry, D., and Kennedy, M.P., 1969, Cretaceous and Eocene coccoliths at San Diego,
California: California Division of Mines and Geology, Special Report 100, p. 3344.
Cassel, E.J., Grove, M., and Graham, S.A., 2012, Eocene drainage evolution and erosion
of the Sierra Nevada batholith across northern California and Nevada: American
Journal of Science, v. 312, p. 117-144, doi: 10.2475/02.2012.03.
Cecil, M.R., Ducea, M.N., Reiners, P., Gehrels, G., Mulch, A., Allen, C., and Campbell,
I., 2010, Provenance of Eocene river sediments from the central northern Sierra
Nevada and implications for paleotopography: Tectonics, v. 29, TC6010,
doi:10.029/2010TC002717, 13 p.
Chan, M.A., and Dott, R.H., Jr., 1983, Shelf and deep-sea sedimentation in Eocene
forearc basin, western Oregon-fan or non-fan?: The American Association of
Petroleum Geologists Bulletin, v. 67, p. 21002116.
259

Cherven, V.B., 1983a, Mesozoic through Paleogene evolution of the Sacramento basin,
California, in Cherven, V.B., and Graham, S.A., Geology and sedimentology of
the southwestern Sacramento basin and East Bay hills, Pacific Section, SEPM
(Society for Sedimentary Geology), Field Trip Guidebook, p. 21-32.
Cherven, V.B., 1983b, Stratigraphy, facies, and depositional provinces of the middle
Eocene Domengine Formation, southern Sacramento basin, in Cherven, V.B., and
Graham, S.A., Geology and sedimentology of the southwestern Sacramento basin
and East Bay hills, Pacific Section, SEPM (Society for Sedimentary Geology),
Field Trip Guidebook, p. 59-72.
Chipping, D.H., 1972, Early Tertiary paleogeography of central California: The
American Association of Petroleum Geologists Bulletin, v. 56, p. 480493.
Colburn, I.P., 1961, The tectonic history of Mount Diablo, California [Ph.D.
Dissertation]: Stanford University, 276 p.
Compton, R.R., 1960, Charnockitic rocks of the Santa Lucia Range, California:
American Journal of Sciences, v. 258, p. 609-636.
Cooper, J.D., Colburn, I.P., and Sundberg, F.A., 1982, Upper Cretaceous environmental
stratigraphy and field trip stops, Silverado Canyon area, in Bottjer, D.J., Colburn,
I.P., and Cooper, J.D., eds., Late Cretaceous Depositional Environments and
Paleogeography, Santa Ana Mountains, Southern California: Pacific Section,
SEPM (Society for Sedimentary Geology), Volume and Guidebook, p. 11-23.
Creely, S., and Force, E.R., 2007, Type region of the Ione Formation (Eocene), central
California: Stratigraphy, paleogeography, and relation to auriferous gravels: U.S.
Geological Survey Open-File Report 2006-1378, 65 p.
Critelli, S., and Nilsen, T.H., 1996, Petrology and diagenesis of the Eocene Butano
Sandstone, La Honda basin, California: Journal of Geology, v. 104, p. 295315,
doi:10.1086/629826.
Cummings, J.C., 1968, The Santa Clara Formation and possible post-Pliocene slip on the
San Andreas fault in central California, in Dickinson, W.R., and Grantz, A., eds.,
Proceedings of conference on geologic problems of San Andreas fault system:
Stanford University Publications in the Geological Sciences, v. 11, p. 191207.
DeCelles, P.G., 1988, Middle Cenozoic depositional, tectonic, and sea level history of
southern San Joaquin basin, California: American Association of Petroleum
Geologists Bulletin, v. 72, p. 1297-1322.
DeGraaff-Surpless, K., Graham, S.A., Wooden, J.L., and McWilliams, M.O., 2002,
Detrital zircon provenance analysis of the Great Valley Group, California:
Evolution of an arc-forearc system: Geological Society of America Bulletin, v.
114, p. 15641580, doi: 10.1130/00167606(2002)114<1564:DZPAOT>2.0.CO;2.
DeLise, K.C., 1967, Biostratigraphy of the San Emigdio Formation, Kern County,
California: California University Publications in Geological Science, v. 68, 66 p.
Dibblee, T.R., Jr., 2005. Geologic map of the Eagle Rest Peak/south of Conner SW
quadrangles Kern County, California. Dibblee Geology Center, v. 172.
Dibblee, T.R., Jr., 2006, Geologic map of the Reliz Canyon quadrangle Monterey
County, California. Dibblee Geology Center, v. 249.
Dibblee, T.R., Jr., 2007, Geologic map of the Idria Quadrangle San Benito County,
California. Dibblee Geology Center, v. 294.
260

Dibblee, T.R., Jr., and Nilsen, T.H., 1974, Lower Tertiary stratigraphy from Panoche
Creek area to Domengine Creek area and Vallecitos area, California, in Payne, M.
ed., The Paleogene of the Panoche Creek-Cantua Creek area: Pacific Section,
SEPM (Society of Economic Paleontologists and Mineralogists), p. 28-37.
Dickinson, W.R., 1995a, Forearc basins, in Busby, C.J., and Ingersoll, R.V., eds.,
Tectonics of sedimentary basins: Cambridge, Massachusetts, Blackwell Science,
p. 221261.
Dickinson, W.R., 1995b, Paleogene depositional systems of the western Transverse
Ranges and adjacent southernmost Coast Ranges, California, in Fritsche, A.E.,
ed., Cenozoic Paleogeography of the Western United StatesII: Los Angeles,
Pacific Section, SEPM (Society for Sedimentary Geology), Book 75, p. 5883.
Dickinson, W.R., 1996, Kinematics of transrotational tectonism in the California
Transverse Ranges and its contribution to cumulative slip along the San Andreas
transform fault system: Geological Society of America Special Paper 305, 46 p.
Dickinson, W.R., and Seely, D.R., 1979, Structure and stratigraphy of forearc regions:
The American Association of Petroleum Geologists Bulletin, v. 63, p. 2-31.
Dickinson, W.R., Ducea, M., Rosenberg, L.I., Greene, H.G., Graham, S.A., Clark, J.C.,
Weber, G.E., Kidder, S., Ernst, W.G., and Brabb, E.E., 2005, Net dextral slip,
Neogene San GregorioHosgri Fault Zone, Coastal California: Geologic Evidence
and Tectonic Implications: Geological Society of America Special Paper 391, 43
p.
Dickinson, W.R., Ingersoll, R.V., and Graham, S.A., 1979, Paleogene sediment dispersal
and paleotectonics in northern California: Geological Society of America
Bulletin, v. 90, pt. I, p. 897898; pt. II, p. 14581528.
Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A.,
2012, Provenance of the Paleogene Colton Formation (Uinta Basin) and
Cretaceous-Paleogene provenance evolution in the Utah foreland: Evidence for
U-Pb ages of detrital zircons, paleocurrent trends, and sandstone petrofacies:
Geosphere, v. 8, p. 854-880, doi:10.1130/GES00763.1.
Doebbert, A.C., Carroll, A.R., and Johnson, C., 2012, The sandstone-derived provenance
record of the Gualala basin, northern California, U.S.A.: Journal of Sedimentary
Research, v. 82, p. 841-858, doi: 10.2110/jsr.2012.72.
Dumitru, T.A., Wakabayashi, J., Wright, J.E., and Wooden, J.L., 2010, Early Cretaceous
transition from nonaccretionary behavior to strongly accretionary behavior within
the Franciscan subduction complex: Tectonics, v. 29, TC5001,
doi:10.1029/2009TC002542.
Dumitru, T.A., Ernst, W.G., Wright, J.E., Wooden, J.L., Wells, R.E., Farmer, L.P., Kent,
A.J.R., and Graham, S.A., 2012, Eocene extension in Idaho generated massive
sediment floods into the Franciscan trench and into the Tyee, Great Valley, and
Green River basins: Geology, v. 41, p. 187-190, doi: 10.1130/G33746.1.
Ernst, W.G., 1970, Tectonic contact between the Franciscan mlange and the Great
Valley sequence: Crustal expression of a late Mesozoic Benioff zone: Journal of
Geophysical Research, v. 75, p. 886901.
Girty, G.H., 1986, Sandstone provenances, Point Loma Formation, San Diego,
California: Evidence for uplift of the Peninsular Ranges during the Laramide
Orogeny, Journal of Sedimentary Petrology, v. 57, p. 839-844.
261

Goudkoff, P.P., 1945, Stratigraphic relations of Upper Cretaceous in Great Valley,


California: American Association of Petroleum Geologists Bulletin, v. 29, p. 9561007.
Graham, J.J., and Classen, W.J., 1955, A lower Eocene foraminiferal faunule from the
Woodside area, Santa Mateo County, California: Contributions from the Cushman
Foundation for Foraminiferal Research, v. 6, p. 1-38.
Graham, S.A., 1976a, Tertiary sedimentary tectonics of the central Salinian block of
California [Ph.D. dissertation]: Stanford, California, Stanford University, 510 p.
Graham, S.A., 1976b, Tertiary stratigraphy and depositional environments near Indians
Ranch, Monterey County, California, in Fritsche, A.E., Ter Best, H., Jr., and
Wornardt, W.W., eds., The Neogene symposium: Los Angeles, Pacific Section,
SEPM (Society for Sedimentary Geology), p. 125136.
Graham, S.A., 1978, Role of Salinian block in evolution of San Andreas fault system,
California: The American Association of Petroleum Geologists Bulletin, v. 62, p.
22142231.
Graham, S.A., and Berry, K.D., 1979, Early Eocene paleogeography of the central San
Joaquin Valley: Origin of the Cantua Sandstone, in Armentrout, J.M., Cole, M.R.,
and Ter Best, H., Jr., eds., Cenozoic paleogeography of the western United States:
Pacific Section, SEPM (Society for Sedimentary Geology), Pacific Coast
Paleogeography Symposium 3, p. 119127.
Graham, S.A., Stanley, R.G., Bent, J.V., and Carter, J.B., 1989, Oligocene and Miocene
paleogeography of central California and displacement along the San Andreas
fault: Geological Society of America Bulletin, v. 101, p. 711730.
Graham, S.A., McCloy, C., Hitzman, M.,Ward, R., and Turner, R., 1984, Basin evolution
during change from convergent to transform margin in central California:
American Association of Petroleum Geologists Bulletin, v. 68, p. 233249.
Graymer, R.W., Jones, D.L., and Brabb, E.E., 1994, Preliminary geologic map
emphasizing bedrock formations in Contra Costa County, California: U.S.
Geological Survey, Open-File Report 94-622.
Graymer, R.W., Moring, B.C., Saucedo, G.J., Wentworth, C.M., Brabb, E.E., and
Knudsen, K.L., 2006, Geologic map of the San Francisco Bay region: U.S.
Geological Survey Scientific Investigations Map 2918, 1 plate.
Graymer, R.W., Sarna-Wojcicki, A.M., Walker, J.P., McLaughlin, R.J., and Fleck, R.J.,
2002, Controls on timing and amount of right-lateral offset on the East Bay fault
system, San Francisco Bay region, California: Geological Society of America
Bulletin, v. 114, p. 1471-1479.
Griscom, A., and Jachens, R.C., 1989, Tectonic history of the north portion of the San
Andreas fault system, California, inferred from gravity and magnetic anomalies:
Journal of Geophysical Research, v. 94, p. 30893099.
Grove, K., 1993, Latest Cretaceous basin formation within the Salinian terrane of westcentral California: Geological Society of America Bulletin, v. 105, p. 447463.
Fairbanks, H.W., 1893, Geology of San Diego; also portions of Orange and San
Bernardino Counties: California State Mining Bureau 11th Annual Report, p. 76120.

262

Fildani, A., Hessler, A.M., and Graham, S.A., 2008, Trench-forearc interactions reflected
in the sedimentary fill of Talara basin, northwest Peru: Basin Research, v. 20, p.
305-331.
Frederiksen, N.O., 1991, Pulses of middle Eocene to earliest Oligocene climatic
deterioration in southern California and Gulf Coast: Palaios, v. 6, p. 564-671.
Fry, J.G., Bottjer, D.J., and Lund, S.P., 1985, Magnetostratigraphy of displaced Upper
Cretaceous strata in southern California: Geology, v. 13, p. 648-651.
Hall, C.A., Jr., 1991, Geology of the Point SurLopez Point Region, Coast Ranges,
California: A Part of the Southern California Allochthon: Geological Society of
America Special Paper 266, 40 p.
Hansen, E., and Stuk, M., 1993, Orthopyroxene-bearing, mafic migmatites at Cone Peak,
California: evidence for the formation of migmatitic granulites by anatexis in an
open system: Journal of Metamorphic Geology, v. 11, p. 291-307.
Heller, P.L., and Dickinson, W.R., 1985, Submarine ramp facies model for delta-fed,
sand-rich turbidite systems: The American Association of Petroleum Geologists
Bulletin, v. 69, p. 960-976.
Heller, P.L., and Ryberg, P.T., 1983, Sedimentary record of subduction to forearc
transition in the rotated Eocene basin of western Oregon: Geology, v. 11, p. 380383.
Heller, P.L., Peterman, Z.E., ONeil, J.R., and Shafiquallah, M., 1985, Isotopic
provenance of sandstones from the Eocene Tyee Formation, Oregon Coast Range:
Geological Society of America Bulletin, v. 96, p. 770-780.
Heller, P.L., Tabor, R.W., and Suczek, C.A., 1987, Paleogeographic evolution of the
United States Pacific Northwest during Paleogene time: Canadian Journal of
Earth Sciences, v. 24, p. 1652-1667.
Ingersoll, R.V., 1979, Evolution of the Late Cretaceous forearc basin, northern and
central California: Geological Society of America Bulletin, v. 90, p. 18131826.
Ingersoll, R.V., 1983, Petrofacies and provenance of late Mesozoic forearc basin,
northern and central California: The American Association of Petroleum
Geologists Bulletin, v. 67, p. 11251142.
Jachens, R.C., and Zoback, M.L., 1999, The San Andreas fault in the San Francisco Bay
region, California: Structure and kinematics of a young plate boundary:
International Geology Review, v. 41, p. 191205.
Jachens, R.C., Wentworth, C.M., Zoback, M.L., Bruns, T.R., and Roberts, C.W., 2002,
Concealed strands of the San Andreas fault system in the central San Francisco
Bay region, as inferred from aeromagnetic anomalies, in Parsons, T., ed., Crustal
structure of the coastal and marine San Francisco Bay region, California: U.S.
Geological Survey Professional Paper 1658, p. 4361.
Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsaglia, K.M., Gehrels, G.E.,
and Nourse, J.A., 2011, Late Cretaceous-early Cenozoic tectonic evolution of the
southern California margin inferred from provenance of trench and forearc
sediments: Geological Society of America Bulletin, v. 123, p. 485-506.
Junkerman, C., 2010, A biography of the Stanford sandstone from Greystone Quarry to
Stone River: Stanford Historical Society, v. 34, p. 3-16.
Kennedy, M.P., 1975, Geology of the San Diego Metropolitan area, California, Section
A: California Division of Mines and Geology Bulletin 200, p. 939.
263

Kennedy, M.P., and Moore, G.W., 1971, Stratigraphic relations of Upper Cretaceous and
Eocene formations, San Diego coastal area, California: The American Association
of Petroleum Geologists Bulletin, v. 55, p. 709722.
Kheradyar, T., 1988, Calcareous nannofossil biostratigraphy of the Butano Sandstone,
Santa Cruz Mountains, California, in Filewicz, M.V., and Squires, R.L., eds.,
Paleogene stratigraphy of the west coast of North America, Pacific Section,
SEPM (Society for Sedimentary Geology) West Coast Paleogene Symposium, v.
58, p. 151-166.
Kidder, S., Ducea, M., Gehrels, G., Patchett, P.J., and Vervoort, J., 2003, Tectonic and
magmatic development of the Salinian Coast Ridge belt, California: Tectonics, v.
22, no. 5, 1058, doi: 10.1029/2002TC0001409, p. 121 to 1220.
Kies, R.P., 1982, Paleogene sedimentology, lithostratigraphic correlations and
paleogeography, San Miguel Island, Santa Cruz Island, and San Diego, California
[M.S. Thesis]: San Diego State University, 577 p.
Kies, R.P., and Abbott, P.L., 1982, Sedimentology and paleogeography of lower
Paleogene conglomerates, southern California Continental Borderland, in Fife,
D.L., and Minch, J.A., eds., Geology and mineral wealth of the California
Transverse Ranges: South Coast Geological Society, p. 337-349.
Kies, R.P., and Abbott, P.L., 1983, Rhyolite clast populations and tectonics in the
California continental borderland: Journal of Sedimentary Petrology, v. 53, p.
461475.
Kimyai, A., 1993, Eocene palynomorphs from the Black Diamond Mines Regional
Preserve, Contra Costa County, California: Palynology, v. 17, p. 101-113.
Laiming, B., 1939, Some foraminiferal correlations in the Eocene of San Joaquin Valley,
California: Sixth Pacific Science Conference Proceedings, v. 2, p. 535-638.
Laursen, J., Scholl, D.W., and von Huene, R., 2002, Neotectonic deformation of the
central Chile margin: Deepwater forearc basin formation in response to hot spot
ridge and seamount subduction: Tectonics, v. 21, 1038,
doi:10.1029/2001TC901023, p. 2-1 to 2-27.
Lechler, A.R., and Niemi, N.A., 2011, Sedimentologic and isotopic constraints on the
Paleogene paleogeographic and paleotopography of the southern Sierra Nevada,
California: Geology, v. 39, p. 379382, doi:10.1130/G31535.1.
Link, M.H., and Nilsen, T.H., 1979, The Rocks Sandstone, an Eocene sand-rich deep-sea
fan deposit northern Santa Lucia Range, California: Journal of Sedimentary
Petrology, v. 50, p. 583-602.
Lough, C. F., 1993, Structural evolution of the Vallecitos Mountains, Colorado Desert
and Salton Trough geology. San Diego Association of Geologists, San Diego,
California, p. 91-109.
Ludington, S., Moring, B.C., Miller, R.J., Stone, P.A., Bookstrom, A.A., Bedford, D.R.,
Evans, J.G., Haxel, G.A., Nutt, C.J., Flyn, K.S., and Hopkins, M.J., 2007,
Preliminary integrated geologic map databases for the United States Western
States: California, Nevada, Arizona, Washington, Oregon, Idaho, and Utah: U.S.
Geological Survey Open-File Report, v. 2005-1305, version 1.3.
Luyendyk, B.P., 1991, A model for Neogene crustal rotations, transtension, and
transpression in southern California: Geological Society of America Bulletin, v.
103, p. 15281536.
264

Malmblorg, W.T., West, W.B., Brabb, E.B., and Parker, J.M., 2008, Digital coordinates
and age of more than 13,000 foraminifers samples collected by Chevron
petroleum geologists in California: U.S. Geological Survey Open-File Report
2008-1187.
Mason, E.L., 1998, Internal facies architecture of a sand-rich, deep-sea depositional
system: The Rocks Sandstone, Reliz Canyon Formation, northern Santa Lucia
Range, Monterey County, California [M.S. thesis]: Stanford University, 100 p.
Mason, E.L., and Graham, S.A., 2000, Structural and tectonic controls on basin
partitioning, sediment dispersal, and stratigraphic architecture of Mesozoic and
Cenozoic deep-marine coarse-clastic deposits, central California, in Appi, C.J.,
dAvila, R.S.F., and Viana, A.R., eds., Deep-water sedimentation: Technological
challenges for the next millennium: 31st International Geological Congress
Workshop Program and Abstract Volume, p. 43-48.
May, J.A., 1982, Basin-margin sedimentation: Eocene La Jolla group, San Diego County,
California [Ph.D. thesis]: Rice University, 364 p.
McDonough, S.D. and Abbott, P.L., 1989, Paleohydrology of the Eocene Las Palmas
Gravels, Baja California, Mexico, in Abbott, P.L., ed., Geologic Studies in Baja
California: Pacific Section, SEPM (Society for Sedimentary Geology), Book 63,
p. 47-62.
McDougall, K., 2007, California Cenozoic biostratigraphy Paleogene, in Hosford
Scheirer, A., ed., Petroleum systems and geologic assessment of oil and gas in the
San Joaquin Basin Province, California: U.S. Geological Survey Professional
Paper 1268, p. 1-56.
McLaughlin, R.J., Clark, J.C., Brabb, E.E., Helley, E.J., and Colon, C.J., 2001, Geologic
maps and structure sections of the southwestern Santa Clara Valley and southern
Santa Cruz Mountains, Santa Clara and Santa Cruz Counties, California: U.S.
Geological Survey Miscellaneous Field Studies Map MF-2373, scale 1:24,000,
maps, cross sections, and map pamphlet.
McLaughlin, R.J., Clark, J.C., Brabb, E.E., Helley, E.J., and Wentworth, C.M., 2004,
Geologic map of the Loma Prieta region, California: U.S. Geological Survey
Professional Paper 1550-E, 3 plates.
McLaughlin, R.J., Powell, C.L., McDougall-Reid, K., and Jachens, R.C., 2007, Cessation
of slip on the Pilarcitos fault and initiation of the San Franciscan Peninsula
Segment of the (Modern) San Andreas fault, California: Eos, Transactions,
American Geophysical Union, v. 88, n. 52, Abstract T34A-1089.
McLaughlin, R.J., Sliter, W.V., Sorg, D.H., Russell, P.C., and Sarna-Wojcicki, A.M.,
1996, Large-scale right-slip displacement on the East San Francisco Bay Region
fault system: Implications for location of late Miocene to Pliocene Pacific plate
boundary: Tectonics, v. 15, p. 118.
McGuire, D.J., 1988, Stratigraphy, depositional history, and hydrocarbon source-rock
potential of the Upper Cretaceous-lower Tertiary Moreno Formation, central San
Joaquin basin, California [Ph.D. thesis]: Stanford University, 309 p.
Milam, R.W., 1985, Biostratigraphy and sedimentation of the Eocene and Oligocene
Kreyenhagen Formation, central California [Ph.D. thesis]: Stanford University,
240 p.

265

Miller, P.L., 1983, Late Cretaceous coccolith biostratigraphy of San Miguel Island,
California: Micropaleontology, v. 29, p. 182-201.
Mitchell, C., Graham, S.A., and Suek, D.H., 2010, Subduction complex uplift and
exhumation and its influence on Maastrichtian forearc stratigraphy in the Great
Valley Basin, northern San Joaquin Valley, California: Geological Society of
America Bulletin, v. 122, p. 2063-2078, doi: 10.1130/B30180.1.
Moxon, I.A.,1990, Stratigraphic and structural architecture of the San Joaquin and
Sacramento valleys [Ph.D. thesis]: Stanford University, 371 p.
Nicholson, C., Sorlien, C.C., Atwater, T., Crowell, J.C., and Luyendyk, B.P., 1994,
Microplate capture, rotation of the western Transverse Ranges, and initiation of
the San Andreas transform as a low-angle fault system: Geology, v. 22, p. 491
459.
Nilsen, T.H., 1981, Early Cenozoic stratigraphy, tectonics and sedimentation of the
central Diablo Range between Hollister and New Idria, in Nilsen, T.H., and
Dibblee, J.W., Jr., eds., Geology of the central Diablo Range between Hollister
and New Idria, California: Cordilleran Section, Geological Society of America,
Field Trip Guidebook, p. 31-55.
Nilsen, T.H., 1987, Stratigraphy and sedimentology of the Eocene Tejon Formation,
western Tehachapi and San Emigdio Mountains, California: U.S. Geological
Survey Professional Paper 1268, 100 p.
Nilsen, T.H., and Abbott, P.L., 1981, Paleogeography and sedimentology of Upper
Cretaceous turbidites, San Diego, California: American Association of Petroleum
Geologists Bulletin, v. 65, p. 12561284.
Nilsen, T.H., and Clarke, S.H., Jr., 1975, Sedimentation and tectonics in the early
Tertiary continental borderland of central California: U.S. Geological Survey
Professional Paper 925, 64 p.
Nilsen, T.H., Dibblee, T.W., Jr., and Addicott, W.O., 1973, Lower and middle Tertiary
stratigraphic units of the San Emigdio and western Tehachapi Mountains,
California: U.S. Geological Survey Bulletin 1372-H, 23 p.
Nilsen, T.H., Dibblee, T.W., Jr., and Simoni, T.R., Jr., 1974, Stratigraphy and
sedimentology of the Cantua Sandstone Member of the Lodo Formation, in
Payne, M., ed., The Paleogene of the Panoche Creek-Cantua Creek area central
California: Pacific Section, SEPM (Society of Economic Paleontologists and
Mineralogists), Field Trip Guidebook, p. 38-68.
Page, B.M., 1993, Geologic maps of Stanford lands and vicinity: Stanford Geological
Survey, one plate prepared by R.G. Coleman.
Page, B.M., and Tabor, L.L., 1967, Chatoic structure and dcollement of Cenozoic rocks
near Stanford University, California: Geological Society of America Bulletin, v.
78, p. 1-12.
Pampeyan, E.H., 1993, Geologic map of the Palo Alto and part of the Redwood Point 71/2 quadrangles, San Mateo and Santa Clara Counties, California: U.S.
Geological Survey Miscellaneous Investigation Series Map, I-2371, scale
1:24,000, map and map pamphlet.
Peterson, G.L., and Abbott, P.L., 1979, Mid-Eocene climate change, southwestern
California and northwestern Baja California: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 26, p. 73-87.
266

Peterson, G.L., and Nordstrom, C.E., 1970, Sub-La Jolla unconformity in the vicinity of
San Diego, California: American Association of Petroleum Geologists Bulletin, v.
9, p. 137-159.
Poore, R.Z., and Brabb, E.E., 1977, Eocene and Oligocene planktonic foraminifera from
the upper Butano Sandstone and type San Lorenzo Formation, Santa Cruz
Mountains, California: Journal of Foraminiferal Research, v. 7, p. 249-272.
Popenoe, W.P., 1937, Upper Cretaceous Mollusca from southern California: Journal of
Paleontology, v. 11, p. 379-752.
Popenoe, W.P., 1941, The Trabuco and Baker conglomerates in the Santa Ana
Mountains: Journal of Geology, v. 49, p. 738-752.
Press, W.H., Flannery, B.P., Teukolsky, S.A., and Vetterling, W.T., 1986, Numerical
recipes, The art of scientific computing: Cambridge, Cambridge University Press,
186 p.
Ross, D.C., 1984, Possible correlations of basement rocks across the San Andreas, San
GregorioHosgri, and RinconadaRelizKing City faults, California: U.S.
Geological Survey Professional Paper 1317, 37 p.
Saleeby, J., 2003, Segmentation of the Laramide slabEvidence from the southern
Sierra Nevada region: Geological Society of America Bulletin, v. 115, p. 655
668.
Schoellhamer, J. E., Vedder, J. G., Yerkes, R. F., and Kinney, D. M., 1981, Geology of
the northern Santa Ana Mountains, California: U.S. Geological Survey
Professional Paper 420-D, 109 p.
Schulein, B.J., 1993, Sedimentation and tectonics of the upper lower to lower middle
Eocene Domengine Formation Vallecitos syncline, California [M.S. thesis]:
Stanford, California, Stanford University, 343 p.
Seiders, V.M., and Cox, B.T., 1992, Place of origin of the Salinian block, California, as
based on clast compositions of Upper Cretaceous and lower Tertiary
conglomerates: U.S. Geological Survey Professional Paper 1526, 80 p.
Sharman, G.R., Graham, S.A., Grove, M., and Hourigan, J.K., 2013, A reappraisal of the
early slip history of the San Andreas fault, central California, USA: Geology, v.
41, p. 727-730, doi: 10.1130/G34214.1.
Short, W.R., Jr., 1986, Geology of the Santa Teresa Hills, Santa Clara County, California
[M.S. thesis]: San Jose, California State University, 112 p.
Sliter, W.V., 1968, Upper Cretaceous foraminifera from southern California and
northwestern Baja California, Mexico: University of Kansas Paleontological
Contributions, serial number 49, article 7, 141 p.
Snow, C.A., Wakabayashi, J., Ernst, W.G., and Wooden, J.L., 2010, Detrital zircon
evidence for progressive underthrusting in Franciscan metagraywackes, westcentral California: Geological Society of America Bulletin, v. 122, p. 282291,
doi:10.1130/B26399.1.
Steer, B.L., and Abbott, P.L., 1984, Paleohydrology of the Eocene Ballena Gravels, San
Diego County, California: Sedimentary Geology, v. 38, p. 181-216.
Sullivan, R., and Sullivan, M.D., 2012, Sequence stratigraphy and incised valley
architecture of the Domengine Formation, Black Diamond Mines Regional
Preserve and the southern Sacramento basin, California, U.S.A.: Journal of
Sedimentary Research, v. 82, p. 781-800, doi: 10.2110/jsr.2012.66.
267

Surpless, K.D., and Beverly, E.M., 2013, Understanding a critical basinal link in
Cretaceous Cordilleran paleogeography: Detailed provenance of the Hornbrook
Formation, Oregon and California: Geological Society of America Bulletin, v.
125, p. 709-727, doi: 10.1130/B30690.1.
Vedder, J.G., Howell, D.G., and McLean, H., 1982, Stratigraphy, sedimentation, and
tectonic accretion of exotic terranes, southern Coast Ranges, California, in
Watkins, J.S., and Darke, C.L., eds., Studies in continental geology: The
American Association of Petroleum Geologists Memoir No. 34, p. 471-496.
Wakabayashi, J., 1999, Distribution of displacement on and evolution of a young
transform fault system: The northern San Andreas fault system, California:
Tectonics, v. 18, p. 12451274, doi: 10.1029/1999TC900049.
Williams, T.A., and Graham, S.A., 2013, Controls on forearc basin architecture from
seismic and sequence stratigraphy of the Upper Cretaceous Great Valley Group,
central Sacramento Basin, California: International Geology Review, v. 55, p.
2030-2059.
Yeo, R.K., 1982, The stratigraphy and sedimentology of Upper Cretaceous sediments of
southwestern California and Baja California, Mexico [Ph.D. thesis]: Houston,
Rice University, 604 p.
Yeo, R.K., 1984, Sedimentology of Upper Cretaceous strata, northern Baja California,
Mexico, in Abbott, P.L., Upper Cretaceous depositional systems southern
California-northern Baja California: Pacific Section, SEPM (Society of Economic
Paleontologists and Mineralogists), Field Trip Volume and Guidebook 36, p. 109120.
Yerkes, R.F., McCulloh, T.H., Schoellhamer, J.E., and Vedder, J.G., 1965, Geology of
the Los Angeles basin, California an introduction: U.S. Geological Survey
Professional Paper 420-A, 57 p.

268

APPENDIX A-4: Cumulative Detrital Zircon U-Pb Age Distributions (available


electronically on Stanford Universitys library website)

269

APPENDIX A-5: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the


University of Arizona Laserchron Center (available electronically on Stanford
Universitys library website)

270

APPENDIX A-6: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the


University of California Santa Cruz (available electronically on Stanford
Universitys library website)

271

APPENDIX A-7: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the


University of California Los Angeles (available electronically on Stanford
Universitys library website)

272

APPENDIX A-8: U-Th-Pb Isotopic Composition of Detrital Zircons Analyzed at the


Stanford-USGS SHRIMP (available electronically on Stanford Universitys library
website)

273

APPENDIX A-9. Detrital Zircon Age Population Proportions


Sample

Eocene forearc sandstones


Tyee
48
LF020
49
LF024
24
LF023
26
LF021
9
JV395
53
JV437
49
JW290
38
Average
296

Age (Ma)
135100

200135

300200

>300

Location

<65

85-65

10085

Tyee basin
Tyee basin
Tyee basin
Tyee basin
Tyee basin
Tyee basin
Tyee basin
Tyee basin

60.4%
24.5%
58.3%
26.9%
22.2%
34.0%
26.5%
23.7%
35.1%

0.0%
16.3%
8.3%
19.2%
11.1%
1.9%
2.0%
7.9%
7.1%

10.4%
10.2%
8.3%
11.5%
11.1%
5.7%
12.2%
5.3%
9.1%

8.3%
14.3%
4.2%
15.4%
33.3%
20.8%
8.2%
7.9%
12.5%

2.1%
8.2%
0.0%
7.7%
0.0%
3.8%
8.2%
13.2%
6.1%

2.1%
2.0%
0.0%
11.5%
0.0%
0.0%
10.2%
2.6%
3.7%

16.7%
24.5%
20.8%
7.7%
22.2%
34.0%
32.7%
39.5%
26.4%

CR-2
CR-25
CR-85
CR-90
CR-55
Average

93
89
69
96
92
439

Franciscan Coastal Belt


Franciscan Coastal Belt
Franciscan Coastal Belt
Franciscan Coastal Belt
Franciscan Coastal Belt

11.8%
13.5%
23.2%
25.0%
18.5%
18.2%

30.1%
16.9%
29.0%
17.7%
15.2%
21.4%

21.5%
22.5%
13.0%
13.5%
19.6%
18.2%

6.5%
5.6%
7.2%
5.2%
9.8%
6.8%

18.3%
27.0%
23.2%
29.2%
17.4%
23.0%

3.2%
3.4%
1.4%
0.0%
3.3%
2.3%

8.6%
11.2%
2.9%
9.4%
16.3%
10.0%

SN07-060NC
SN09-001DF
SN09-006YB
SN09-007RD
SN09-070HP
CEC-3a
CEC-4
CEC-5
Average

97
92
106
97
95
102
99
100
788

Northern Sierra Nevada


Northern Sierra Nevada
Northern Sierra Nevada
Northern Sierra Nevada
Northern Sierra Nevada
Northern Sierra Nevada
Northern Sierra Nevada
Northern Sierra Nevada

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
1.0%
0.1%

0.0%
1.1%
0.0%
1.0%
0.0%
0.0%
0.0%
0.0%
0.3%

53.6%
54.3%
34.9%
37.1%
11.6%
26.5%
49.5%
33.0%
37.4%

33.0%
30.4%
52.8%
46.4%
57.9%
43.1%
44.4%
58.0%
45.9%

8.2%
9.8%
6.6%
12.4%
29.5%
21.6%
4.0%
2.0%
11.7%

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

5.2%
4.3%
5.7%
3.1%
1.1%
8.8%
2.0%
6.0%
4.6%

CEC-1
CEC-2
GVG09-8
GVG09-7
GVG09-9
Average

90
96
95
92
83
456

Central Sierra Nevada


Central Sierra Nevada
Central Sierra Nevada
Central Sierra Nevada
Central Sierra Nevada

0.0%
0.0%
0.0%
0.0%
2.4%
0.4%

1.1%
0.0%
1.1%
0.0%
1.2%
0.7%

33.3%
24.0%
30.5%
31.5%
31.3%
30.0%

43.3%
59.4%
44.2%
42.4%
33.7%
45.0%

14.4%
16.7%
21.1%
22.8%
30.1%
20.8%

5.6%
0.0%
0.0%
0.0%
1.2%
1.3%

2.2%
0.0%
3.2%
3.3%
0.0%
1.8%

TD1F-105
GVG09-18
GVG09-0.5
GVG09-22
GVG09-23
GVG09-24
Average

90
87
89
85
94
96
541

Mount Diablo
Mount Diablo
Mount Diablo
Mount Diablo
Mount Diablo
Mount Diablo

6.7%
17.2%
0.0%
1.2%
0.0%
0.0%
4.1%

3.3%
4.6%
1.1%
2.4%
2.1%
0.0%
2.2%

12.2%
5.7%
40.4%
21.2%
26.6%
30.2%
22.9%

21.1%
18.4%
43.8%
44.7%
42.6%
36.5%
34.6%

24.4%
18.4%
12.4%
22.4%
21.3%
26.0%
20.9%

0.0%
1.1%
2.2%
1.2%
0.0%
3.1%
1.3%

32.2%
34.5%
0.0%
7.1%
7.4%
4.2%
14.0%

SG-20
SG-21
Average

99
97
196

Vallecitos syncline
Vallecitos syncline

0.0%
0.0%
0.0%

1.0%
1.0%
1.0%

36.4%
39.2%
37.8%

49.5%
40.2%
44.9%

10.1%
15.5%
12.8%

0.0%
0.0%
0.0%

3.0%
4.1%
3.6%

SG-4
SG-23
SG-47
SBM

98
99
100
55

San Francisco Bay block


San Francisco Bay block
San Francisco Bay block
San Francisco Bay block

0.0%
0.0%
0.0%
21.8%

1.0%
1.0%
4.0%
25.5%

42.9%
31.3%
42.0%
14.5%

34.7%
45.5%
44.0%
7.3%

20.4%
17.2%
9.0%
21.8%

0.0%
0.0%
0.0%
0.0%

1.0%
5.1%
1.0%
9.1%

274

Average

352

3.4%

5.7%

34.9%

36.1%

16.5%

0.0%

3.4%

POR-1
POR-2
POR-3
Average

100
97
101
298

Temblor Range
Temblor Range
Temblor Range

1.0%
9.3%
1.0%
3.7%

2.0%
3.1%
0.0%
1.7%

41.0%
24.7%
38.6%
34.9%

36.0%
39.2%
41.6%
38.9%

13.0%
10.3%
12.9%
12.1%

1.0%
1.0%
2.0%
1.3%

6.0%
12.4%
4.0%
7.4%

SG-40
SEF-1
TEJ-2
TEJ-1
SEM-1
Average

97
99
97
100
104
497

San Emigdio Mountains


San Emigdio Mountains
San Emigdio Mountains
San Emigdio Mountains
San Emigdio Mountains

1.0%
0.0%
1.0%
0.0%
0.0%
0.4%

1.0%
4.0%
5.2%
0.0%
0.0%
2.0%

16.5%
26.3%
24.7%
23.0%
15.4%
21.1%

22.7%
15.2%
10.3%
17.0%
71.2%
27.8%

39.2%
45.5%
51.5%
45.0%
11.5%
38.2%

7.2%
5.1%
2.1%
7.0%
0.0%
4.2%

12.4%
4.0%
5.2%
8.0%
1.9%
6.2%

ST08
GCV2
GC1B
Average

103
161
178
442

Gualala block
Gualala block
Gualala block

0.0%
0.0%
0.0%
0.0%

1.0%
0.6%
1.1%
0.9%

30.1%
13.7%
28.1%
23.3%

19.4%
11.8%
15.2%
14.9%

33.0%
48.4%
27.5%
36.4%

15.5%
22.4%
21.3%
20.4%

1.0%
3.1%
6.7%
4.1%

SG-46
SG-6
BUT-1
BUT-2
BUT-3
BUT-4
BUT-5
BB1
SJB-1
Average

95
97
100
99
100
99
96
28
96
810

Pilarcitos block
Pilarcitos block
La Honda basin
La Honda basin
La Honda basin
La Honda basin
La Honda basin
La Honda basin
La Honda basin

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

5.3%
4.1%
3.0%
3.0%
7.0%
0.0%
3.1%
0.0%
1.0%
3.2%

18.9%
17.5%
20.0%
29.3%
25.0%
18.2%
27.1%
25.0%
32.3%
23.6%

10.5%
13.4%
9.0%
10.1%
10.0%
5.1%
13.5%
10.7%
10.4%
10.2%

56.8%
55.7%
43.0%
39.4%
44.0%
61.6%
37.5%
28.6%
40.6%
46.7%

4.2%
0.0%
17.0%
10.1%
8.0%
7.1%
6.3%
28.6%
5.2%
8.0%

4.2%
9.3%
8.0%
8.1%
6.0%
8.1%
12.5%
7.1%
10.4%
8.3%

IR3
IR4
SG-31
SG-28
SG-29
Average

22
28
100
99
98
347

Santa Lucia Range


Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

13.6%
25.0%
3.0%
4.0%
2.0%
5.5%

9.1%
7.1%
9.0%
2.0%
2.0%
4.9%

18.2%
3.6%
14.0%
2.0%
0.0%
6.1%

31.8%
14.3%
70.0%
82.8%
91.8%
72.9%

0.0%
14.3%
1.0%
6.1%
2.0%
3.7%

27.3%
35.7%
3.0%
3.0%
2.0%
6.9%

OR432
OR433
MP1
MP2
MP3
LV2
LV3
Average

70
59
15
12
12
12
14
194

Orocopia Mountains
Orocopia Mountains
Sierra Madre Mountains
Sierra Madre Mountains
Sierra Madre Mountains
Pine Mountain block
Pine Mountain block

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

2.9%
0.0%
20.0%
0.0%
16.7%
8.3%
0.0%
4.1%

0.0%
0.0%
0.0%
0.0%
8.3%
0.0%
0.0%
0.5%

0.0%
1.7%
0.0%
0.0%
0.0%
8.3%
0.0%
1.0%

24.3%
20.3%
0.0%
50.0%
16.7%
16.7%
35.7%
22.7%

4.3%
3.4%
13.3%
0.0%
0.0%
0.0%
7.1%
4.1%

68.6%
74.6%
66.7%
50.0%
58.3%
66.7%
57.1%
67.5%

LP3
LP7
LP13
LP14
LP15
Average

15
16
14
15
14
74

Santa Ynez Mountains


Santa Ynez Mountains
Santa Ynez Mountains
Santa Ynez Mountains
Santa Ynez Mountains

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

20.0%
18.8%
14.3%
33.3%
14.3%
20.3%

13.3%
18.8%
7.1%
0.0%
7.1%
9.5%

6.7%
6.3%
7.1%
6.7%
0.0%
5.4%

26.7%
6.3%
21.4%
33.3%
28.6%
23.0%

0.0%
6.3%
0.0%
0.0%
0.0%
1.4%

33.3%
43.8%
50.0%
26.7%
50.0%
40.5%

SH6
SM6
SM6A

48
15
15

Simi Hills
Santa Monica Mountains
Santa Monica Mountains

0.0%
0.0%
0.0%

8.3%
20.0%
20.0%

4.2%
13.3%
13.3%

2.1%
0.0%
6.7%

39.6%
6.7%
0.0%

2.1%
0.0%
0.0%

43.8%
60.0%
60.0%

275

Average

78

0.0%

12.8%

7.7%

2.6%

25.6%

1.3%

50.0%

SNTGU
SNTGL
Average

88
98
186

Santa Ana Mountains


Santa Ana Mountains

1.1%
1.0%
1.1%

42.0%
26.5%
33.9%

1.1%
2.0%
1.6%

0.0%
1.0%
0.5%

5.7%
2.0%
3.8%

5.7%
5.1%
5.4%

44.3%
62.2%
53.8%

SMI0406
POW-1
POW-2
JOLLA-5
JOLLA-1
JOLLA-2
Average

53
74
53
64
45
50
339

San Miguel Island


San Diego
San Diego
San Diego
San Diego
San Diego

0.0%
1.4%
0.0%
0.0%
4.4%
0.0%
0.9%

11.3%
12.2%
15.1%
28.1%
28.9%
34.0%
20.9%

17.0%
14.9%
17.0%
9.4%
4.4%
8.0%
12.1%

35.8%
32.4%
17.0%
10.9%
4.4%
12.0%
19.8%

9.4%
13.5%
20.8%
23.4%
4.4%
0.0%
12.7%

1.9%
4.1%
5.7%
1.6%
4.4%
6.0%
3.8%

24.5%
21.6%
24.5%
26.6%
48.9%
40.0%
29.8%

PALM-1
PALM-2
PALM-3
Average

49
52
45
146

Northern Baja California


Northern Baja California
Northern Baja California

0.0%
3.8%
2.2%
2.1%

20.4%
26.9%
26.7%
24.7%

36.7%
30.8%
35.6%
34.2%

12.2%
19.2%
15.6%
15.8%

26.5%
11.5%
11.1%
16.4%

2.0%
1.9%
0.0%
1.4%

2.0%
5.8%
8.9%
5.5%

0.0%
0.0%
0.0%
0.0%

1.1%
0.0%
2.0%
1.0%

47.9%
29.8%
33.0%
36.8%

33.0%
48.9%
33.0%
38.2%

13.8%
19.1%
21.0%
18.1%

0.0%
0.0%
2.0%
0.7%

4.3%
2.1%
9.0%
5.2%

Maastrichtian-Paleocene forearc samples


GVG09-17
94
Mount Diablo
GVG09-16
94
Mount Diablo
GVG09-15
100
Mount Diablo
Average
288
GVG09-1

97

Del Puerto Canyon

0.0%

2.1%

24.7%

33.0%

34.0%

2.1%

4.1%

SG-22

99

Vallecitos syncline

2.0%

5.1%

55.6%

29.3%

8.1%

0.0%

0.0%

GP01
MM1
MM1A
MM2
MM3
Average

125
12
14
16
20
187

Gualala block
La Honda basin
La Honda basin
La Honda basin
La Honda basin

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

9.6%
0.0%
14.3%
6.3%
0.0%
8.0%

39.2%
8.3%
28.6%
31.3%
40.0%
35.8%

42.4%
25.0%
21.4%
6.3%
35.0%
35.8%

8.0%
41.7%
21.4%
31.3%
15.0%
13.9%

0.0%
0.0%
14.3%
6.3%
5.0%
2.1%

0.8%
25.0%
0.0%
18.8%
5.0%
4.3%

03-04
MO1
IR5
IR1
ES1
PS1
PT1
ROJO
ESALEN
ROCKY
POINT
Average

16
13
21
19
13
9
13
79
104

Santa Lucia Range


Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range
Santa Lucia Range

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

12.5%
15.4%
42.9%
0.0%
15.4%
0.0%
30.8%
1.3%
1.0%

50.0%
30.8%
38.1%
57.9%
30.8%
44.4%
53.8%
3.8%
16.3%

18.8%
7.7%
9.5%
15.8%
7.7%
11.1%
7.7%
2.5%
17.3%

0.0%
7.7%
4.8%
10.5%
7.7%
33.3%
7.7%
17.7%
9.6%

18.8%
38.5%
4.8%
0.0%
38.5%
11.1%
0.0%
8.9%
50.0%

0.0%
0.0%
0.0%
15.8%
0.0%
0.0%
0.0%
65.8%
5.8%

89
376

Santa Lucia Range

0.0%
0.0%

0.0%
5.6%

2.2%
18.1%

3.4%
9.3%

15.7%
12.5%

13.5%
22.9%

65.2%
31.6%

PZ1
PZ3
PZ4
SML1
SF1
SF2
SF4
SF5

9
14
12
15
20
13
13
15

La Panza Range
La Panza Range
La Panza Range
La Panza Range
San Gabriel bock
San Gabriel bock
San Gabriel bock
San Gabriel bock

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

88.9%
21.4%
25.0%
26.7%
20.0%
0.0%
7.7%
0.0%

0.0%
0.0%
8.3%
0.0%
0.0%
7.7%
0.0%
6.7%

0.0%
7.1%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
7.1%
0.0%
0.0%
10.0%
0.0%
7.7%
13.3%

0.0%
7.1%
8.3%
33.3%
0.0%
0.0%
7.7%
6.7%

11.1%
57.1%
58.3%
40.0%
70.0%
92.3%
76.9%
73.3%

276

VA1
Average

12
123

San Gabriel bock

0.0%
0.0%

33.3%
22.0%

0.0%
2.4%

0.0%
0.8%

8.3%
5.7%

0.0%
7.3%

58.3%
61.8%

LP12
LP21
Average

29
72
101

Santa Ynez Mountains


Santa Ynez Mountains

0.0%
0.0%
0.0%

37.9%
31.9%
33.7%

27.6%
20.8%
22.8%

6.9%
9.7%
8.9%

3.4%
1.4%
2.0%

6.9%
1.4%
3.0%

17.2%
34.7%
29.7%

SH2
SH8
SM2
SM4
Average

15
72
17
15
119

Simi Hills
Simi Hills
Santa Monica Mountains
Santa Monica Mountains

0.0%
0.0%
0.0%
0.0%
0.0%

13.3%
19.4%
47.1%
0.0%
20.2%

0.0%
44.4%
11.8%
0.0%
28.6%

6.7%
23.6%
11.8%
93.3%
28.6%

6.7%
6.9%
11.8%
6.7%
7.6%

0.0%
1.4%
0.0%
0.0%
0.8%

73.3%
4.2%
17.6%
0.0%
14.3%

SLVU
SLVL
Average

51
31
82

Santa Ana Mountains


Santa Ana Mountains

0.0%
0.0%
0.0%

35.3%
0.0%
22.0%

19.6%
54.8%
32.9%

9.8%
41.9%
22.0%

7.8%
0.0%
4.9%

9.8%
0.0%
6.1%

17.6%
3.2%
12.2%

SMI0404
JOLLA-3
JOLLA-4
Average

30
45
48
123

San Miguel Island


San Diego
San Diego

0.0%
0.0%
0.0%
0.0%

0.0%
2.2%
0.0%
0.8%

30.0%
0.0%
8.3%
10.6%

56.7%
95.6%
81.3%
80.5%

13.3%
2.2%
6.3%
6.5%

0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
4.2%
1.6%

El Rosario

0.0%

7.0%

54.4%

24.6%

10.5%

1.8%

1.8%

0.0%
0.0%
0.0%
0.0%
0.0%

1.2%
0.0%
0.0%
3.1%
1.1%

31.7%
7.1%
0.0%
33.0%
17.3%

14.6%
19.4%
35.7%
18.6%
22.4%

18.3%
52.0%
48.0%
39.2%
40.3%

4.9%
5.1%
7.1%
3.1%
5.1%

29.3%
16.3%
9.2%
3.1%
13.9%

RF-2

Santonian-Campanian forearc samples


07-HB-05
82
Hornbrook basin
07-HB-26
98
Hornbrook basin
07-HB-21
98
Hornbrook basin
06-KS-11
97
Hornbrook basin
Average
375
KDS11

56

Redding

0.0%

0.0%

10.7%

50.0%

28.6%

0.0%

10.7%

KDS23
KDS21
Average

55
60
115

Chico
Chico

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

3.6%
1.7%
2.6%

20.0%
8.3%
13.9%

32.7%
61.7%
47.8%

1.8%
1.7%
1.7%

41.8%
26.7%
33.9%

KDS3

56

Cache Creek

0.0%

0.0%

19.6%

8.9%

39.3%

5.4%

26.8%

GVG09-14
GVG09-13
Average

89
93
182

Mount Diablo
Mount Diablo

0.0%
0.0%
0.0%

1.1%
3.2%
2.2%

22.5%
18.9%
20.9%

48.3%
23.2%
35.7%

22.5%
37.9%
30.8%

1.1%
4.2%
2.7%

4.5%
12.6%
7.7%

GVG09-2
99-65
Average

86
36
122

Del Puerto Canyon


Del Puerto Canyon

0.0%
0.0%
0.0%

2.3%
2.8%
2.5%

11.6%
11.1%
11.5%

16.3%
50.0%
26.2%

17.4%
36.1%
23.0%

4.7%
0.0%
3.3%

47.7%
0.0%
33.6%

GV13
GV8
GV21
Average

56
56
54
166

Coalinga
Coalinga
Coalinga

0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
0.0%
0.0%

16.1%
10.7%
0.0%
9.0%

60.7%
44.6%
66.7%
57.2%

17.9%
35.7%
33.3%
28.9%

0.0%
0.0%
0.0%
0.0%

5.4%
8.9%
0.0%
4.8%

SBSS
WB03
SM19
SMSS

136
111
105
86

Gualala block
Gualala block
Gualala block
Gualala block

0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
1.9%
0.0%

11.8%
0.9%
65.7%
45.3%

56.6%
37.8%
26.7%
50.0%

30.9%
60.4%
1.0%
2.3%

0.7%
0.0%
0.0%
1.2%

0.0%
0.9%
4.8%
1.2%

277

Average

438

0.0%

0.5%

28.5%

43.4%

25.6%

0.5%

1.6%

PP1
PP2
PP3
03-01
03-03
99-66A
99-68
YM1
YM3
YM5
Average

12
14
13
14
18
32
14
14
13
14
158

Pigeon Point block


Pigeon Point block
Pigeon Point block
Cambria slab
Cambria slab
Cambria slab
Atascadero
Atascadero
Atascadero
Atascadero

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
7.1%
0.0%
0.0%
0.6%

8.3%
0.0%
7.7%
7.1%
0.0%
6.3%
21.4%
0.0%
15.4%
0.0%
6.3%

33.3%
35.7%
84.6%
14.3%
22.2%
43.8%
35.7%
42.9%
30.8%
42.9%
38.6%

58.3%
50.0%
7.7%
28.6%
27.8%
40.6%
21.4%
35.7%
53.8%
42.9%
36.7%

0.0%
7.1%
0.0%
28.6%
16.7%
6.3%
14.3%
0.0%
0.0%
14.3%
8.9%

0.0%
7.1%
0.0%
7.1%
5.6%
0.0%
0.0%
7.1%
0.0%
0.0%
2.5%

0.0%
0.0%
0.0%
14.3%
27.8%
3.1%
7.1%
7.1%
0.0%
0.0%
6.3%

LP1
LP9
TP1
BS1
BS1A
PS2
Average

15
15
25
13
16
14
98

San Rafael Mountains


San Rafael Mountains
San Rafael Mountains
Point Sur block
Point Sur block
Point Sur block

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

73.3%
26.7%
40.0%
23.1%
25.0%
35.7%
37.8%

20.0%
46.7%
48.0%
61.5%
56.3%
14.3%
41.8%

6.7%
13.3%
12.0%
0.0%
6.3%
7.1%
8.2%

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
13.3%
0.0%
15.4%
12.5%
42.9%
12.2%

SH1
SH7
SM3
SM8
Average

15
74
22
67
178

Simi Hills
Simi Hills
Santa Monica Mountains
Santa Monica Mountains

0.0%
0.0%
0.0%
0.0%
0.0%

20.0%
13.5%
18.2%
1.5%
10.1%

33.3%
27.0%
31.8%
55.2%
38.8%

40.0%
31.1%
22.7%
23.9%
28.1%

6.7%
10.8%
0.0%
0.0%
5.1%

0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
17.6%
27.3%
19.4%
18.0%

WLM
MUST
Average

30
26
56

Santa Ana Mountains


Santa Ana Mountains

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

36.7%
65.4%
50.0%

63.3%
34.6%
50.0%

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

SMI0403
ROSA-1
ROSA-2
ROSA-3
ROSA-4
ROSA-5
Average

34
30
55
30
58
54
261

San Miguel Island


San Diego
San Diego
San Diego
San Diego
San Diego

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
1.8%
0.0%
0.0%
0.0%
0.4%

20.6%
26.7%
29.1%
20.0%
0.0%
1.9%
14.6%

73.5%
53.3%
60.0%
66.7%
98.3%
98.1%
78.2%

5.9%
20.0%
7.3%
13.3%
1.7%
0.0%
6.5%

0.0%
0.0%
1.8%
0.0%
0.0%
0.0%
0.4%

0.0%
0.0%
0.0%
0.0%
0.0%
0.0%
0.0%

ROS-1
ROS-2
ROS-3
Average

33
65
60
158

Northern Baja California


Northern Baja California
Northern Baja California

0.0%
0.0%
0.0%
0.0%

6.1%
9.2%
0.0%
5.1%

36.4%
61.5%
56.7%
54.4%

48.5%
26.2%
21.7%
29.1%

9.1%
0.0%
16.7%
8.2%

0.0%
1.5%
0.0%
0.6%

0.0%
1.5%
5.0%
2.5%

RF-1
EG-2
EG-1
PB
Average

54
60
58
60
232

El Rosario
El Rosario
El Rosario
El Rosario

0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
3.3%
5.2%
1.7%
2.6%

35.2%
43.3%
56.9%
61.7%
49.6%

55.6%
41.7%
27.6%
30.0%
38.4%

5.6%
10.0%
8.6%
5.0%
7.3%

0.0%
0.0%
1.7%
1.7%
0.9%

3.7%
1.7%
0.0%
0.0%
1.3%

0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
0.0%
0.0%

0.0%
3.1%
0.0%
0.0%

10.1%
13.3%
4.9%
2.2%

85.9%
80.6%
90.2%
97.8%

1.0%
2.0%
1.6%
0.0%

3.0%
1.0%
3.3%
0.0%

Cenomanian-Coniacian forearc samples


07-HB-11
99
Hornbrook basin
07-HB-13
98
Hornbrook basin
07-HB-17
61
Hornbrook basin
06-KS-06
74
Hornbrook basin

278

06-KS-04
Average

93
425

Hornbrook basin

0.0%
0.0%

0.0%
0.0%

0.0%
0.7%

12.2%
8.7%

73.0%
85.6%

0.0%
0.9%

14.9%
4.0%

KDS13
KDS10
Average

55
56
111

Redding
Redding

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

14.5%
3.6%
9.0%

34.5%
53.6%
44.1%

49.1%
41.1%
45.0%

0.0%
0.0%
0.0%

1.8%
1.8%
1.8%

KDS22

57

Chico

0.0%

0.0%

1.8%

5.3%

75.4%

1.8%

15.8%

GV44
GV45
GV42
GV40
Average

56
56
56
56
224

Cache Creek
Cache Creek
Cache Creek
Cache Creek

0.0%
0.0%
0.0%
0.0%
0.0%

0.0%
0.0%
0.0%
0.0%
0.0%

1.8%
0.0%
0.0%
0.0%
0.4%

25.0%
1.8%
7.1%
0.0%
8.5%

66.1%
96.4%
92.9%
100.0%
88.8%

0.0%
0.0%
0.0%
0.0%
0.0%

7.1%
1.8%
0.0%
0.0%
2.2%

GVG09-12

93

Mount Diablo

0.0%

0.0%

0.0%

15.1%

81.7%

0.0%

3.2%

GV33
GV24
Average

56
56
112

Coalinga
Coalinga

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

1.8%
0.0%
0.9%

66.1%
69.6%
67.9%

28.6%
17.9%
23.2%

0.0%
0.0%
0.0%

3.6%
12.5%
8.0%

YM4

26

Atascadero

0.0%

0.0%

15.4%

50.0%

26.9%

0.0%

7.7%

LP10
LP11
Average

16
15
31

San Rafael Mountains


San Rafael Mountains

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

25.0%
20.0%
22.6%

43.8%
53.3%
48.4%

31.3%
13.3%
22.6%

0.0%
6.7%
3.2%

0.0%
6.7%
3.2%

BAK
TRB
Average

29
32
61

Santa Ana Mountains


Santa Ana Mountains

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

17.2%
0.0%
8.2%

82.8%
84.4%
83.6%

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

0.0%
15.6%
8.2%

SMI0402
SMI0401
Average

34
35
69

San Miguel Island


San Miguel Island

0.0%
0.0%
0.0%

0.0%
0.0%
0.0%

17.6%
17.1%
17.4%

73.5%
57.1%
65.2%

5.9%
14.3%
10.1%

0.0%
2.9%
1.4%

2.9%
8.6%
5.8%

Guad01
Guad02
Average

30
60
90

Northern Baja California


Northern Baja California

0.0%
0.0%
0.0%

0.0%
3.3%
2.2%

6.7%
6.7%
6.7%

83.3%
88.3%
86.7%

6.7%
1.7%
3.3%

0.0%
0.0%
0.0%

3.3%
0.0%
1.1%

LBR

59

El Rosario

0.0%

0.0%

61.0%

39.0%

0.0%

0.0%

0.0%

Notes:
See Table 1 for additional information on individual samples
Data in bold used to construct Figure 7A

279

APPENDIX B: SUPPLEMENTARY MATERIAL FOR CHAPTER 2

Summary of Contents
Appendix B-1: A New Paleogeographic Model for the Point of Rocks Sandstone, San
Joaquin Basin, California
Appendix B-2: Analytical Methods
Appendix B-3: Kolgomorov-Smirnov Statistics
Appendix B-4: Discussion of Piercing Points between the Northern Gabilan Range and
the Western San Emigdio Mountains
Appendix B-5: U-Th-Pb Isotopic Compositions of Detrital Zircons Analyzed at the
University of Arizona Laserchron Center
Appendix B-6: U-Th-Pb Isotopic Compositions of Detrital Zircons Analyzed at the
University of California Santa Cruz
Appendix B-7: Explanation of Tectonic and Depositional Events Shown in Figure 4
Appendix B-8: Explanation of the San Andreas Fault Piercing Points Shown in Figure 4

Appendices B-5 and B-6 are available electronically as supplemental files on Stanford
Universitys library website

280

APPENDXI B-1: A New Paleogeographic Model for the Point of Rocks Sandstone,
San Joaquin Basin, California
Glenn R. Sharman1, Stephan A. Graham1, and Christopher Sine2
1. Department of Geological and Environmental Sciences, Stanford University 94305
2. Occidental Petroleum, Bakersfield, California 93311

INTRODUCTION
The Point of Rocks Sandstone Member of the Kreyenhagen Formation is a middle
Eocene sand-dominated unit that constitutes a historically important petroleum reservoir
in the southwestern San Joaquin basin. The Point of Rocks Sandstone attains great
thicknesses (likely over 2000 meters) and is present over a wide area, both in outcrop and
in the subsurface (Fig. 1). Both sedimentary facies and recovered benthic foraminifera
indicate that the Point of Rocks Sandstone was predominantly deposited in a deep-marine
(lower bathyal to abyssal) environment as part of a submarine fan system (Clarke, 1973;
see also Carter, this volume, for additional description of the Point of Rocks Sandstone).
In addition to its economic value, the Point of Rocks Sandstone has played an
important role by influencing the historical development of paleotectonic and
paleogeographic models in California. For example, the Point of Rocks Sandstone helped
early workers to recognize and quantify strike-slip displacement along the San Andreas
fault. In a seminal paper, Hill and Dibblee (1953) argued that the San Andreas fault had
accumulated 100s of kilometers of dextral strike-slip motion since Cretaceous time,
based in part upon correlation of Eocene deep-marine sequences of the Santa Cruz
Mountains (Butano Sandstone) and Temblor Range (Point of Rocks Sandstone; Hill and
Dibblee, 1953, p. 449). The recognition that large-scale, horizontal fault offsets could
occur was a revolutionary and controversial idea that played an important role in
reconciling continental and offshore geologic relationships in the emerging theory of
plate tectonics (e.g., Atwater, 1970).
While the Butano-Point of Rocks correlation was proposed in the early 1950s, it
was not until twenty years later that additional work done by Samuel Clarke and Tor
281

Nilsen solidified the correlation within a coherent paleogeographic model (Clarke, 1973;
Clarke and Nilsen, 1973; Nilsen and Clarke, 1975; see also Fig. 2 in Carter, this volume).
In this model, the Butano Sandstone and Point of Rocks Sandstone formed two halves of
a large, northeast-to-northwest flowing submarine fan that was derived from erosion of
Salinian granitic rocks in the vicinity of the northern Gabilan Range and Monterey Bay
area (Clarke, 1973; see also Fig. 2 in Carter, this volume). Portions of this submarine fan
were deposited in the La Honda basin (modern-day Santa Cruz Mountains) and
southwestern San Joaquin basin and constitute the Butano Sandstone and Point of Rocks
Sandstone, respectively. Evidence for the Butano-Point of Rocks correlation includes
many similarities between these units: 1) abrupt truncation of thick sand depocenters
across the San Andreas fault, 2) great thickness (> 1,500 m) and widespread distribution,
3) age-equivalence, 4) similar depositional facies and paleobathymetry (outer neritic to
bathyal/abyssal), 5) generally compatible paleocurrent distributions, 6) very similar
sandstone composition and inferred provenance from a silicic-intermediate granitic
terrane, and 7) shared stratigraphic positions overlying Paleocene mudstone and
underlying middle-late Eocene mudstone (see discussion in Clarke, 1973, p. 225-230).
In addition to providing a paleogeographic model for the central California
region, the Butano-Point of Rocks correlation provided a key constraint on the preNeogene offset history of the San Andreas fault (e.g., Nilsen and Clarke, 1975). For
example, the middle Eocene Butano-Point of Rocks submarine fan (ca. 48-38 Ma) has
approximately the same amount of offset (~315 km) as the ca. 23 Ma Pinnacles-Neenach
volcanic field (Matthews, 1976; Graham et al., 1989), suggesting that the central San
Andreas fault was inactive during late Paleogene time. Researchers have long recognized
that restoration of ~315 km of Neogene San Andreas displacement does not fully account
for the apparent offset of Salinian granitic rocks with respect to the main Cretaceous
batholithic belt that runs through the Sierra Nevadan and Peninsular Ranges batholiths
(e.g., Suppe, 1970). By one estimate, the amount of unaccounted slip is at least 100 km
(Dickinson et al., 2005). This discrepancy has been accounted for by a dextral, proto-San
Andreas fault that was hypothesized to have been active during latest Cretaceous-early
Paleogene time (e.g., Nilsen and Clarke, 1975). However, the proto-San Andreas fault
model has not been widely accepted and remains controversial (e.g., Atwater, 1989).
282

Since its proposal in the early 1950s, the Butano-Point of Rocks correlation has
been widely accepted in the literature and has been used to constrain both
paleogeographic and paleotectonic models (e.g., Graham et al., 1989; Dickinson et al.,
2005). Perhaps a single exception was a regional study of conglomerate clast and
sandstone compositions by Seiders and Cox (1992) that noted a marked dissimilarity
between conglomerate clast types within the Butano Sandstone and Point of Rocks
Sandstone. In particular, the Butano Sandstone is enriched in felsic volcanic and granitic
clasts compared to the Point of Rocks Sandstone that is enriched in chert (Seiders and
Cox 1992, p. 28). Seiders and Cox (1992) also noted that Point of Rocks sandstone is
enriched in quartz grains relative to the Butano Sandstone (Fig. 2). While the Butano
Sandstone is compositionally similar to the German Rancho Formation, sandstone and
conglomerate clast compositions of the Point of Rocks Sandstone closely resemble the
lower Eocene Cantua Sandstone that is interpreted to have been derived from the Sierra
Nevada (Graham and Berry, 1979; Seiders and Cox, 1992). In their conclusions, Seiders
and Cox (1992) suggested that the Butano Sandstone and Point of Rocks Sandstone were
derived from different source regions, in contrast to the prevailing paleogeographic
model (Clarke, 1973).
The main purpose of this article is to further evaluate the Butano-Point of Rocks
correlation using two complimentary methods: 1) detrital zircon U-Pb geochronology and
2) subsurface mapping of middle Eocene strata in the southern Joaquin basin using well
log, biostratigraphic, and seismic datasets. Our results will be used to reassess current
thinking regarding Eocene paleogeography of the central California margin (see also
Sharman et al., 2013, 2014).
DETRITAL ZIRCON U-Pb GEOCHRONOLOGY
Detrital zircon U-Pb geochronology offers a method of determining sandstone
provenance that can have distinct advantages over traditional methods (e.g., sandstone
petrography, conglomerate clast assemblages, etc.). Individual crystallization ages of
detrital zircon grains in sandstone can be directly linked with the ages of crystalline rocks
in potential source regions. In California, igneous rocks are widespread due to the
presence of a volcanic arc that extended north-south across the margin during Mesozoic
283

time (Dickinson, 2008). Today, the remnants of this volcanic arc are primarily preserved
as plutonic rocks in the Sierra Nevada Mountains, Salinian block, Transverse Ranges,
and Peninsular Ranges. Although many of the plutonic rocks in California have similar
compositions (e.g., quartz monzonite), detrital zircon geochronology is able to readily
distinguish between compositionally-similar plutons on the basis of their crystallization
age.
Sedimentary Provenance Analysis
Normalized and cumulative detrital zircon age distributions are presented in
Figure 3 for twelve sandstone samples (early to middle Eocene) from the southern San
Joaquin and La Honda basins (Sharman et al., 2013). All samples are dominated by late
Permian-Cretaceous (ca. 280-80 Ma) zircon with lesser abundances of Paleogene and
pre-Permian zircon (Fig. 3). In particular, all samples have variable proportions of midCretaceous (ca. 125-80 Ma) and Jurassic-earliest Cretaceous (ca. 175-140 Ma) zircon
(Fig. 3).
The dominance of ca. 280-80 Ma zircon in Eocene sandstone in the La Honda and
southern San Joaquin basins suggests that these strata were derived in large part from
igneous rocks of the Mesozoic arc (Fig. 3; Sharman et al., 2013). Although all samples
share similar age populations, the details of the age distributions suggest that there are
two distinct detrital zircon age assemblages that were derived from distinct source
regions (Fig. 3).
The first age assemblage is represented by the Point of Rocks Sandstone and is
characterized by a dominance of mid-Cretaceous zircon (ca. 125-80 Ma; 67%-80%) with
lesser amounts of Jurassic zircon (10%-13%; Fig. 3). Eocene zircon (ca. 55-43 Ma) is
also found in low abundances (1%-9%; Fig. 3). Strikingly similar detrital zircon age
distributions are found in other Eocene fluvial and marine strata that extend ~400 km to
the north along the San Joaquin and Sacramento basins (Fig. 4). We interpret this
provenance signature to indicate that these sands were derived from the ancestral Sierra
Nevada Mountains. The same age populations found in the forearc are present in the midCretaceous Sierra Nevada batholith (125-85 Ma) and in local Jurassic rocks (Irwin and
Wooden, 2001).
284

The second age assemblage is represented by Eocene strata from the La Honda
basin (Butano Sandstone) and from the southernmost San Joaquin basin (Tejon
Formation; Fig. 3). These samples are characterized by 1) approximately equal
abundances of Cretaceous and Jurassic zircon, 2) a small population of Permian-Triassic
zircon (5%-17%), and 3) a lack of Eocene zircon (Fig. 3). Very similar age assemblages
are found in other Eocene sandstone units from the Salinian block, including samples
from the Gualala block (German Rancho Formation; Doebbert et al., 2012) and Santa
Lucia Range (Carmelo Formation and Rocks Sandstone; Sharman et al., in press) that
also have abundant Permian-Jurassic zircon (Fig. 4). Because Permian-Jurassic rocks are
very uncommon in the Salinian block (Kistler and Champion, 2001), the abundance of
Jurassic (38%-62%) and Permian-Triassic (6%-17%) zircon in these units rules out a
local Salinian source, as traditionally interpreted (e.g., Nilsen and Clarke, 1975).
Kolgomorov-Smirnov Statistics
The Butano-Point of Rocks correlation can be quantitatively evaluated using the
Kolgomorov-Smirnov (K-S) statistic that tests the null hypothesis that two distributions
are derived from the same population (Press et al., 1986). This approach reveals that the
two age assemblages are statistically distinct from each other at the 95% confidence level
(i.e., P-value < 0.05; Table 1). In other words, differences in the detrital zircon age
distributions between the Butano Sandstone and Point of Rocks Sandstone cannot be
explained by random sampling of the same parent population, and these units were not
likely derived from the same source region. In addition, the Point of Rocks samples,
when compared to each other, are unable to be statistically distinguished by the K-S test
(Table 1). Similarly, 31 of 36 sample pairs from the La Honda basin and San Emigdio
Mountains are statistically indistinguishable at the 95% confidence level (Table 1).
Summary
Although the Butano-Point of Rocks correlation has been widely accepted and
used to constrain paleogeographic and tectonic reconstructions (Clarke, 1973; Graham et
al., 1989; Dickinson et al., 2005), our results demonstrate that the Butano and Point of
Rocks Sandstones do not share a common provenance. These results are in agreement
with the work by Seiders and Cox (1992) that found significant differences between
285

sandstone and conglomerate clast compositions (Fig. 2). This interpretation casts doubt
on the existence of the Butano-Point of Rocks submarine fan with a single source from
the Salinian block.
SUBSURFACE INVESTIGATION
Isopach Mapping
Point of Rocks Sandstone
The Point of Rocks Sandstone is present over a wide area in the subsurface of the
southwestern San Joaquin Basin and has been penetrated by numerous oil wells (Fig. 1).
We mapped the thickness of the Point of Rocks Sandstone using a database of 43 wells of
which 16 are interpreted to have penetrated the full thickness of the formation (see
Appendix A). Our Point of Rocks isopach map closely resembles the excellent mapping
of Clarke (1973), demonstrating that additional well penetrations in the last 40 years have
not significantly altered our knowledge of the subsurface distribution of the Point of
Rocks. However, we have extended the zero-thickness line farther east than Clarkes
(1973) map, based on the presence of probable Eocene sand in the Tupman USL 1-10
(API: 02947380) and 31X-10 Great Basins (API: 02947361) wells (Malmblorg, 2008).
The Point of Rocks Sandstone is thickest in the vicinity of Cymric oil field where
it likely attains a thickness of over 2 kilometers in a deep well (Cymric Unit 1; API:
02929580) that reached total depth without exiting Eocene sandstone (CDOGGR, 1998).
The Point of Rocks Sandstone systematically thins to the north and east (Fig. 1), where it
pinches out within the encasing Kreyenhagen shale. The Point of Rocks is variably
erosionally truncated along its western margin in outcrop and in the nearby subsurface
(Clarke, 1973). The southern extent of the Point of Rocks Sandstone is poorly known due
to deep burial and a lack of well penetrations, although the Point of Rocks is absent in
Eocene exposures and well penetrations in the vicinity of the San Emidgio Mountains
(Nilsen, 1987; Fig. 1).
Famoso Sandstone
The Famoso Sandstone is a poorly understood unit that occurs exclusively in the
subsurface along the southeastern margin of the San Joaquin basin (Fig. 1). The Famoso
Sandstone is inferred to have been deposited within a shallow-marine shelfal
286

environment and grades into non-marine facies (Walker Formation) to the east (Reid,
1988; Fig. 1). The age of the Famoso Sandstone is poorly constrained but likely ranges
from middle to late Eocene in age based on stratigraphic correlation with the underlying
Domengine Sandstone (Reid, 1988) and by the presence of late Eocene (Refugian)
benthic foraminifera within the formation (Bartow and McDougall, 1984; Fig. 5). Thus,
the Point of Rocks Sandstone and Famoso Sandstone were likely deposited
synchronously, in part (Fig. 5).
We mapped the thickness of the Famoso Sandstone using a database of 40 wells,
all of which penetrated the full thickness of the formation (see Appendix A). Although
the Famoso Sandstone reaches thicknesses up to 160 m, its average thickness is about 85
m over its mapped distribution (Fig. 1). The Famoso Sandstone forms a highly linear
thickness trend that parallels the inferred position of the Eocene shoreline (Dickinson et
al., 1979), an observation that is consistent with an inferred shallow-marine (shelfal)
origin of the Famoso sands (Reid, 1988). The thickness distribution of the Famoso
Sandstone is poorly constrained south of the Bakersfield Arch where the unit becomes
deeply buried beneath younger valley fill (Fig. 1).
Metralla Sandstone
The middle Eocene Metralla Sandstone Member of the Tejon Formation is
present both in outcrop in the San Emigido Mountains (Nilsen, 1987) and the nearby
subsurface (Weber, 1973). The Metralla Sandstone shares many similarities with the
Famoso Sandstone including a middle Eocene age assignment (Fig. 5) and a shelfal
depositional environment that grades into non-marine facies to the east (Tecuya
Formation) and shale-dominated facies to the west (Nilsen, 1987). For these reasons, the
Metralla Sandstone likely represents a southerly equivalent of the Famoso Sandstone.
The Point of Rocks-Famoso Depositional System
Although the deep-marine Point of Rocks Sandstone and shelfal Famoso
Sandstone have previously been interpreted as depositionally unrelated (Clarke, 1973),
several lines of evidence suggests that these units may have formed a genetically-linked,
shelf-to-basin depositional system (Fig. 6). 1) Detrital zircon U-Pb age distributions
indicate that the Point of Rocks Sandstone has a Sierran provenance signature that is
287

distinct from coeval Eocene sands deposited atop or adjacent to the northern Salinian
block (Figs. 3-4). Sierran-derived sand must have come from the east and would have
passed through the shelf prior to entering the deep-marine basin. 2) The Famoso and
Point of Rocks Sandstones overlap in age (Fig. 5), and thus could have been part of the
same depositional system. 3) The thickness distributions of both units indicate that
Famoso sands were located very closely to the Point of Rocks zero-thickness line (Fig.
1). 4) Paleocurrent measurements from the Point of Rocks Sandstone (n = 446; Clarke,
1973) indicate a mean west-northwesterly direction (296) that is consistent with
derivation from the east with lesser northward deflection.
Together, these observations are consistent with the interpretation that Famoso
shelfal sands were routed into the deep-marine Point of Rocks basin (Fig. 6). We
speculate that a submarine canyon linked the Famoso and Point of Rocks depocenters
(Point of Rocks canyon, Fig. 6), based in part on analogy to other well-documented
Late Cretaceous-Paleogene submarine canyons in the Sacramento basin (e.g., Williams,
Markley, Meganos, and Martinez canyons; Almgren et al., 1978; Williams et al., 1998)
and central San Joaquin basin (Cantua canyon, Graham and Berry, 1979; Anderson,
1998). Any such canyon (if it existed) would be located on the Eocene shelf edge that
was positioned a short distance west of Belleview, Greely, and Wasco oil fields
(Clarke, 1973, p. 203) and also between the mapped distributions of the Point of Rocks
and Famoso Sandstones (Figs. 1). Based on this reasoning, we depict the hypothetical
Point of Rocks canyon to be located between the Tupman USL 1-10 and Mushrush 5
wells in our east-west cross section (Fig. 6).
MIDDLE EOCENE PALEOGEOGRAPHY OF CENTRAL CALIFORNIA
Both sandstone provenance (i.e., detrital zircon U-Pb geochronology) and
subsurface mapping of the southern San Joaquin basin suggest that the prevailing
paleogeographic model of central California during Eocene time requires significant
modification (Fig. 7). Because the Butano-Point of Rocks correlation also constitutes an
important piercing point used to constrain the offset history of the San Andreas fault
(Nilsen and Clarke, 1975), our results have important implications for the tectonic
development of the California margin (see discussion in Sharman et al., 2013).
288

Southern San Joaquin Basin


We argue that the Point of Rocks Sandstone was derived from the east and is the
deep-marine equivalent of the shallow-marine Famoso Sandstone (Figs. 6, 7). This
interpretation is primarily based upon detrital zircon U-Pb ages that are similar to other
Eocene Great Valley sands that were derived from the Sierra Nevada batholith (Figs. 3-4)
and proximity to age-equivalent shelfal sands (Fig. 5). We speculate that sand transported
by alongshore currents along the paleo-Sierran shelf encountered a submarine canyon on
the northwest flank of the Bakersfield arch and was routed into the deep-marine Point of
Rocks basin (Fig. 7). This process was very effective in transporting a large volume of
sand (~1,200-1,500 km3; Fig. 1) into the southern San Joaquin deep-marine basin despite
occurring during widespread middle Eocene marine transgression (Bartow, 1991). A
potential analog for this system is the modern California coastline where similar
processes transport sand offshore despite the present-day sea level highstand (e.g.,
Covault et al., 2009).
By analogy to the Cantua Sandstone to the north (Anderson, 1998), the Point of
Rocks basin may have been structurally controlled by an actively rising accretionary
complex that allowed great thicknesses of sand to accumulate in ponded depocenters
(Fig. 7). We speculate that some amount of Point of Rocks sand may have bypassed the
forearc to reach trench-slope basins through an outlet that may have existed in the trenchslope break (Fig. 7). Some evidence for this process is provided by early-middle Eocene
sandstone present in trench-slope basins in the San Francisco Bay region that have a
similar detrital zircon provenance character to the Point of Rocks Sandstone (Fig. 4;
Sharman et al., in press).
Northern Salinian Block
An important, and surprising, observation is that the majority of zircon found in
Paleogene sandstone of the Salinian block (e.g., Butano Sandstone) was not derived
locally as previously thought (e.g., Nilsen and Clarke, 1975). This observation is at odds
with the continental borderland model which holds that local basement uplifts within
the Salinian block supplied detritus to adjacent, bathyal basins (Nilsen and Clarke, 1975).
Although the abundant Permian-Jurassic zircon found in the Butano Sandstone (and
related units) could not have been derived from the Salinian block, plutonic and volcanic
289

rocks of this age are abundant in the southeastern Sierra Nevada and western Mojave
Desert region (Walker et al., 2002; Chapman et al., 2012). The presence of such source
rocks is supported by locally preserved non-marine Paleogene basins in the southernmost
Sierra Nevada region (Goler and Witnet basins) that contain significant amounts of
Triassic-Jurassic zircon (Lechler and Niemi, 2011). We speculate that an extraregional
fluvial system delivered sediment to marine basins atop the Salinian block from the
southeastern Sierra and/or western Mojave regions (Fig. 7; Sharman et al., 2013).
The presence of western Mojave detritus in the Butano Sandstone (Sharman et al.,
2013) and German Rancho Formation (Doebbert et al., 2012) is at odds with current
paleogeographic models that depict the northern Salinian block juxtaposed northward
against the southern San Joaquin basin during Eocene time (Fig. 7). Sharman et al. (2013)
suggested that the Salinian block was positioned 50-100 km farther south in middle
Eocene time than typically depicted in paleotectonic reconstructions (Fig. 7). This
reconstruction has the advantage of 1) allowing sediment to be readily delivered from the
western Mojave region to the northern Salinian block, and 2) negating the need for a Late
Cretaceous-early Paleogene proto-San Andreas fault by restoring the Salinian block to a
position within the Cretaceous batholithic belt (Sharman et al., 2013)
CONCLUSIONS

The Butano Sandstone and Point of Rocks Sandstone have quantitatively distinct

detrital zircon U-Pb age distributions. As a result, these units were not likely part of a
contiguous submarine fan system with a common source region from Salinian granitic
rocks.

The Point of Rocks Sandstone was derived from erosion of the ancestral Sierra

Nevada Mountains. This interpretation is supported by 1) detrital zircon U-Pb ages in the
Point of Rocks that closely resemble other Eocene Great Valley sands, and 2) the
abundance of ca. 125-85 Ma detrital zircon that matches the age range of the midCretaceous Sierra Nevada batholith.

The Point of Rocks Sandstone is the deep-marine equivalent of the shelfal

Famoso Sandstone and the non-marine Walker Formation. We speculate that a submarine

290

canyon was positioned on the Eocene shelf-slope break and delivered shelfal sands to the
adjacent deep-marine basin over much of middle Eocene time.

The Butano Sandstone, and related Paleogene sandstone from the Salinian block,

was derived from the southeastern Sierra Nevada and/or western Mojave region. This
interpretation is supported by the abundance of Late Permian-Jurassic zircon in these
sands that lack a local source in the Salinian block.
ACKNOWLEDGEMENTS
This work benefited from discussion with I. Alark, M. Johns, B. Stupp, and the
rest of the Oxy California Exploration group.
REFERENCES
Almgren, A.A., 1978, Timing of Tertiary submarine canyons and marine cycles of
deposition in the southern Sacramento Valley, California, in Stanley, D.J., and
Kelling, G., eds., Sedimentation in submarine canyons, fans, and trenches:
Dowden, Hutchinson & Ross, Stroudsburg, p. 276-291.
Almgren, A.A., Filewicz, M.V., and Heitman, H.L., 1988, Lower Tertiary foraminiferal
and calcareous nannofossil zonation of California: An overview and
recommendation, in Filewicz, M.V., and Squires, R.L., eds., Paleogene
Stratigraphy, West Coast of North America: Los Angeles, Pacific Section, SEPM
(Society for Sedimentary Geology), v. 58, p. 83-106.
Anderson, K.S., Graham, S.A., and Hubbard, S.M., 2006, Facies, architecture, and origin
of a reservoir-scale sand-rich succession within submarine canyon fill: Insights
from Wagon Caves Rock (Paleocene), Santa Lucia Range, California, U.S.A.:
Journal of Sedimentary Research, v. 76, p. 819-838.
Anderson, K.S., 1998, Facies architecture of two Paleogene structurally-controlled
turbidite systems, central California [Ph.D. thesis]: Stanford University, 391 p.
Atwater, T., 1970, Implications of plate tectonics for the Cenozoic tectonic evolution of
western North America: Geological Society of America Bulletin, v. 81, p. 3513
3536.
Atwater, T.M., 1989, Plate tectonic history of the northeast Pacific and western North
America, in Winterer, E.L., Hussong, D.M., and Decker, R.W., eds., The Eastern
Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America,
Geology of North America, v. N, p. 2172.
Bartow, J.A., and McDougall, K., 1984, Tertiary stratigraphy of the southeastern San
Joaquin Valley, California: U.S. Geological Survey Bulletin 1529-J, 41 p.
Bartow, J.A., 1991, The Cenozoic evolution of the San Joaquin Valley, California: U.S.
Geological Survey Professional Paper 1501, 40 p.

291

Carter, J.B., this volume, The Point of Rocks Sandstone, Temblor Range, California, in
Clark, M., Sharman, G.R., and Carter, J.B., Westside Turbidite Field Trip, San
Joaquin Valley California, PS-AAPG Field Trip Guidebook.
Cassel, E.J., Grove, M., and Graham, S.A., 2012, Eocene drainage evolution and erosion
of the Sierra Nevada batholith across northern California and Nevada: American
Journal of Science, v. 312, p. 117-144.
CDOGGR, 1998, California Oil and Gas Fields, Volume 1 Central California: California
Department of Conservation, Division of Oil, Gas, and Geothermal Resources,
Publication No TR11.
Cecil, M.R., Ducea, M.N., Reiners, P., Gehrels, G., Mulch, A., Allen, C., and Campbell,
I., 2010, Provenance of Eocene river sediments from the central northern Sierra
Nevada and implications for paleotopography: Tectonics, v. 29, TC6010, 13 p.
Chapman, A.D., Saleeby, J.B., Wood, D.J., Piasecki, A., Kidder, S., Ducea, M.N., and
Farley, K.A., 2012, Late Cretaceous gravitational collapse of the southern Sierra
Nevada batholith, California: Geosphere, v. 8, p. 314-341.
Clarke, S.H., Jr., and Nilsen, T.H., 1973, Displacement of Eocene strata and implications
for the history of offset along the San Andreas fault, central and northern
California, in Kovach, R.L., and Nur, A., eds., Proceedings of the conference on
tectonic problems of the San Andreas fault system: Stanford University
Publications in the Geological Sciences, v. 13, p. 358-367.
Clarke, S.H., Jr., 1973, The Eocene Point of Rocks Sandstone: Provenance, mode of
deposition and implications for the history of offset along the San Andreas fault in
central California [Ph.D. thesis]: Berkeley, University of California, 302 p.
Covault, J.A., Normark, W.R., Romans, B.W., and Graham, S.A., 2009, Highstand fans
in the California borderland: The overlooked deep-water depositional systems:
Geology, v. 35, p. 783-786.
Critelli, S., and Nilsen, T.H., 1996, Petrology and diagenesis of the Eocene Butano
Sandstone, La Honda basin, California: The Journal of Geology, v. 104, p. 295
315.
Dickinson, W.R., 2008, Accretionary Mesozoic-Cenozoic expansion of the Cordilleran
continental margin in California and adjacent Oregon: Geosphere, v. 4, p. 329353.
Dickinson, W.R., Ingersoll, R.V., and Graham, S.A., 1979, Paleogene sediment dispersal
and paleotectonics in northern California: Geological Society of America
Bulletin, v. 90, pt. I, p. 897898; pt. II, p. 14581528.
Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Inman,
K.F., Knepp, R.A., Lindberg, F.A., and Ryberg, P.T., 1983, Provenance of North
American Phanerozoic sandstones in relation to tectonic setting: Geological
Society of America Bulletin, v. 94, p. 222235.
Dickinson, W.R., Ducea, M., Rosenberg, L.I., Greene, H.G., Graham, S.A., Clark, J.C.,
Weber, G.E., Kidder, S., Ernst, W.G., and Brabb, E.E., 2005, Net dextral slip,
Neogene San GregorioHosgri Fault Zone, Coastal California: Geologic Evidence
and Tectonic Implications: Geological Society of America Special Paper 391, 43
p.

292

Doebbert, A.C., Carroll, A.R., and Johnson, C., 2012, The sandstone-derived provenance
record of the Gualala basin, northern California, U.S.A.: Journal of Sedimentary
Research, v. 82, p. 841-858.
Dumitru, T.A., Ernst, W.G., Wright, J.E., Wooden, J.L., Wells, R.E., Farmer, L.P., Kent,
A.J.R., and Graham, S.A., 2012, Eocene extension in Idaho generated massive
sediment floods into the Franciscan trench and into the Tyee, Great Valley, and
Green River basins: Geology, v. 41, p. 187-190.
Graham, S.A., and Berry, K.D., 1979, Early Eocene paleogeography of the central San
Joaquin Valley: Origin of the Cantua Sandstone, in Armentrout, J.M., Cole, M.R.,
and Ter Best, H., Jr., eds., Cenozoic paleogeography of the western United States:
Pacific Section, SEPM (Society for Sedimentary Geology), Pacific Coast
Paleogeography Symposium 3, p. 119127.
Graham, S.A., Stanley, R.G., Bent, J.V., and Carter, J.B., 1989, Oligocene and Miocene
paleogeography of central California and displacement along the San Andreas
fault: Geological Society of America Bulletin, v. 101, p. 711730.
Harun, H., 1984, Distribution and deposition of lower to middle Eocene strata in central
San Joaquin Valley, California: Stanford, Calif., Stanford University, M.S. thesis,
100 p.
Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults,
California: Geological Society of America Bulletin, v. 64, p. 443458.
Irwin, W.P., and Wooden, J.L., 2001, Map showing plutons and accreted terranes of the
Sierra Nevada, California with a tabulation of U/Pb isotopic ages: U.S. Geological
Survey Open-File Report 2001-229.
Johnson, C.L., and Graham, S.A., 2005, Middle Tertiary stratigraphic sequences of the
San Joaquin basin, California, in Scheirer, ed., A.H., Petroleum Systems and
Geologic Assessment of Oil and Gas in the San Joaquin Basin Province,
California: U.S. Geological Survey Professional Paper 1713, 18 p.
Kistler, R.W., and Champion, D.E., 2001, Rb-Sr whole-rock and mineral ages, K-Ar,
40Ar/39Ar, and U-Pb mineral ages, and strontium, lead, neodymium, and oxygen
isotopic compositions for granitic rocks from the Salinian composite terrane,
California: U.S. Geological Survey Open-File Report 01-453, 84 p.
Lechler, A.R., and Niemi, N.A., 2011, Sedimentologic and isotopic constraints on the
Paleogene paleogeographic and paleotopography of the southern Sierra Nevada,
California: Geology, v. 39, p. 379382.
Ludington, S., Moring, B.C., Miller, R.J., Stone, P.A., Bookstrom, A.A., Bedford, D.R.,
Evans, J.G., Haxel, G.A., Nutt, C.J., Flyn, K.S., and Hopkins, M.J., 2007,
Preliminary integrated geologic map databases for the United States Western
States: California, Nevada, Arizona, Washington, Oregon, Idaho, and Utah: U.S.
Geological Survey Open-File Report, v. 2005-1305, version 1.3.
Malmblorg, W.T., West, W.B., Brabb, E.B., and Parker, J.M., 2008, Digital coordinates
and age of more than 13,000 foraminifers samples collected by Chevron
petroleum geologists in California: U.S. Geological Survey Open-File Report
2008-1187.
Matthews, V., III, 1976, Correlation of Pinnacles and Neenach volcanic fields and their
bearing on San Andreas fault problem: The American Association of Petroleum
Geologists Bulletin, v. 60, p. 21282141.
293

McDougall, K., 2007, California Cenozoic biostratigraphy Paleogene, in Hosford


Scheirer, A., ed., Petroleum systems and geologic assessment of oil and gas in the
San Joaquin Basin Province, California: U.S. Geological Survey Professional
Paper 1268, p. 1-56.
Milam, R.W., 1985, Biostratigraphy and sedimentation of the Eocene and Oligocene
Kreyenhagen Formation, central California [Ph.D. thesis]: Stanford University,
240 p.
Moxon, I.A.,1990, Stratigraphic and structural architecture of the San Joaquin and
Sacramento valleys [Ph.D. thesis]: Stanford University, 371 p.
Nilsen, T.H., and Clarke, S.H., Jr., 1975, Sedimentation and tectonics in the early
Tertiary continental borderland of central California: U.S. Geological Survey
Professional Paper 925, 64 p.
Nilsen, T.H., 1987, Stratigraphy and sedimentology of the Eocene Tejon Formation,
western Tehachapi and San Emigdio Mountains, California: U.S. Geological
Survey Professional Paper 1268, 110 p.
Press, W.H., Flannery, B.P., Teukolsky, S.A., and Vetterling, W.T., 1986, Numerical
recipes, The art of scientific computing: Cambridge, Cambridge University Press,
186 p.
PS-AAPG, 1957, Cenozoic correlation section across South San Joaquin Valley the San
Andreas fault to Sierra Nevada foot hills, California: prepared by the San Joaquin
Valley Sub-committee on the Cenozoic of the Geologic Names and Correlation
Committee of the American Association of Petroleum Geologists, 1 sheet.
PS-AAPG, 1958, Correlation section across central San Joaquin Valley from Riverdale
thru Tejon Ranch Area, California: prepared by the San Joaquin Valley Subcommittee of the Geologic Names and Correlation Committee of the American
Association of Petroleum Geologists, Section 10S, 1 sheet.
PS-AAPG, 1959, Correlation section across westside San Joaquin Valley from Coalinga
to Midway Sunset and across the San Andreas fault into southeast Cuyama
Valley, California: prepared by the San Joaquin Valley Sub-committee of the
Geologic Names and Correlation Committee of the American Association of
Petroleum Geologists, Section 11, 1 sheet.
Reid, S.A., 1988, Late Cretaceous and Paleogene sedimentation along the east side of the
San Joaquin Basin, in Graham, S.A., and Olson, H.C., eds., Studies of the geology
of the San Joaquin Basin: Los Angeles, Pacific Section, Society of Economic
Paleontologists and Mineralogists, book 60, p. 157-171.
Scheirer, A.H., and Magoon, L.B., 2007, Age, Distribution, and Stratigraphic
Relationship of Rock Units in the San Joaquin Basin Province, California, in
Scheirer, A.H., ed., Petroleum Systems and Geologic Assessment of Oil and Gas
in the San Joaquin Basin Province, California: U.S. Geological Survey
Professional Paper 1713, 38 p.
Seiders, V.M., and Cox, B.T., 1992, Place of origin of the Salinian block, California, as
based on clast compositions of Upper Cretaceous and lower Tertiary
conglomerates: U.S. Geological Survey Professional Paper 1526, 80 p.
Sharman, G.R., Graham, S.A., Grove, M., and Hourigan, J.K., 2013, A reappraisal of the
early slip history of the San Andreas fault, central California, USA: Geology, v.
41, p. 727-730.
294

Sharman, G.R., Graham, S.A., Grove, M., Kimbrough, D.L., and Wright, J.E., in press,
Detrital zircon provenance of the Late Cretaceous-Eocene California forearc:
Influence of Laramide low-angle subduction on sediment dispersal and
paleogeography: Geological Society of America Bulletin.
Suppe, J., 1970, Offset of late Mesozoic basement terrains by the San Andreas fault
system: Geological Society of America Bulletin, v. 81, p. 32533258.
Taylor, B.L., 2004, Petrography and diagenesis of the Eocene Point of Rocks Sandstone,
McKittrick oil field, Kern County, California [M.S. thesis]: California State
University, Bakersfield, 273 p.
Weber, G.E., 1973, Subsurface facies variations in the Metralla Sandstone Member of the
Tejon Formation in the Wheeler Ridge and North Tejon oil fields, Kern County,
California, in Vedder, J.G., chm., Sedimentary facies changes in Tertiary rocks
California Transverse and southern Coast Ranges: Society of Economic
Paleontologists and Mineralogists Guidebook to Field Trip 2, Anaheim,
California, p. 34-38.
Williams, T.A., Graham, S.A., and Constenius, K.N., 1998, Recognition of a Santonian
submarine canyon, Great Valley Group, Sacramento Basin, California:
Implications for petroleum exploration and sequence stratigraphy of deep-marine
strata: American Association of Petroleum Geologists Bulletin, v. 82, p. 15751595.

295

Table 1. Wells used to constrain isopach maps shown in Figure 1


Top
Well Name
API
(MD)
Point of Rocks Sandstone Well Penetrations
Hand 35-28
03100157
1150
2-32
03100193
2485
Pyramid Hills 1-9
03120423 19109
Levis 1
02911381
2465
Ocean Orchard 1
02935470
800
Maybury 3
02937117
1160
Texaco Beer 66-17
02919874
4115
OLC 10
02929571
1110
General Williamson 33
02917540
9000
Hopkins B 35X-23
02913514
2200
31X-10 "Great Basins"
02947361 20600
MST Theta 2
02936550
1000
AML 83-18
02952784
1807
AML 65X-18
02966363
1200
934H-29R
02974651 21650
9-363-31S
03026488 16680
Texaco-2
02919869
6940
OLC 4
02929567 11330
55-26
02935057
9450
Cahn 58
02903956 10385
Layman 23
02935762
3150
CWOD 75-20
02937643
6950
51X-33
02963779 12900
Tupman USL 1-10
02947380 19875
Victory J-12
02936235
1700
Superior Cymric Unit 1
02929580
5000
Bergen USL 1
02960381
8700
PML 8-14
02913551
6660
CWOD 1
02938138
6150
Seabord-Anderson 1
02919894
6760
555-15Z
02909446 10800
534-16Z
02905802
9252
733-17Z
03003315
8770
572-18Z
02905307
9120
Midway-McKittrick 'A' 33-30 02901077
5100
Midway-McKittrick 'A' 22-30 02925785
5350
52-33Z
02905341
6400
Rheem-Magee 36-X
02941289
8330
Dodds-Thomas 3
07900242
3239
Oceanic-Strickland 1
02935471
1140
Corehole 17-13
02913620
50
Oil Explorers 31
02903649
0
Seabord-Bandini Gov't 41-10
02919887
900
38-19V
03100231
-Bravo 1-31
03120135
-Morris H C 1
10700112
--

296

Base
(MD)

Thickness
(ft)

Status

2020
3430
19709
3800
1940
3800
6720
3460
9800
5000
20750
3845
6400
6025
23880
17760
7240
12880
10800
10400
6700
7190
14565
20753
2050
12022
10106
7450
8552
7510
12700
10522
13682
10800
7000
10867
6935
8572
3889
1935
562
1862
2330
----

870
945
600
1335
1140
2640
2605
2350
800 (?)
2800 (?)
150
2845
4593
4825
2230 (?)
950
>300
>1550
>1350
>15
>3550
>240 (?)
>1665
>878 (?)
>350
>7022
>1406 (?)
>790 (?)
>2402
>750
>1900
>1270
>4912
>1680
>1900
>5517
>535
>242
>650 (?)
>795
>512
>1862
>1430
0
0
0

Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Partial
Truncated
Truncated
Truncated
Truncated
Absent
Absent
Absent

S.F. & F.L. 4-2


Twisselman 1
Coles Levee A 26-29
Shell Posuncula 1
KCL 15X-24

02903645
02917530
02960650
02918565
02920544

Famoso Sandstone Well Penetrations


Churchhill 1
10700218
Hoffman 24
10700245
R.E.S. Hesse, et al 1
10720026
Transamerica 1
10720207
Moran 42-30
10700283
Pixley Comm. 1
10700374
Deer Creek 81
10700384
G.R.I. 65-20
10720029
Brunner B1
10700439
Beane 81-28
10700441
Curry 1
02930522
Mobile-Pan 86-35
02941957
Cities-Tenneco 35X
02948583
Mushrush 5
02909487
KCL 31-15
02940286
Famoso 12-1
02930718
Kuhn 81
02930721
KCL-A 58-8
02930725
D-L-K 1
02914763
Shell Fuhrman 1
02926316
Fee C 74
02924112
Kern County Land Lease 31
12-1
02916795
KCL 67 21-10
02940003
115
02908536
33
02930973
KCL A 57-13
02930978
KCL-B 45
02906949
Kramer 1
02908933
Section 5 34
02916269
Gow 1
02930728
Russell 73-17
02930755
McCulloch Camp et al 1-36
02956763
McKevitt-DiGiorgio Co. 1
10700444
Guinee 1
02942852
Occidental-KCL 18X-13
02930764
KCL 33-34
02900132
Lucy 1
10720142
KCL-A 85-35
02930606
Tenneco-Sun 11X-31
02970053
Parsons 1
03003222
Merritt Estate 1
10720008
Salyer 1-16
03120162

------

------

0
0 (?)
0
0
0

Absent
Absent
Absent
Absent
Absent

4999
4879
7073
4853
4263
6107
4543
10101
6997
7051
6276
14600
15221
14926
11324
6146
7046
8749
4390
4641
5653

5217
5062
7604
5075
4460
6569
4587
10246
7415
7523
6648
14960
15371
15095
11573
6489
7420
9112
4520
4682
5921

218
183
530
222
197
463
44
146
418
471
373
360
150
170
249
343
374
363
131
41
268

Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete

10413
12491
13336
6810
7888
10112
7157
6707
7720
13961
5607
5353
7318
9011
11171
8772
9529
8434
7239
---

10746
12670
13545
6971
8143
10502
7572
7107
8029
14155
5945
5478
7665
9278
11510
9033
9791
8742
7606
---

334
179
209
161
255
389
414
400
309
194
338
125
347
267 (?)
339 (?)
261
262
308
367
0
0

Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Complete
Absent
Absent

297

Boswell-Richardson 72-10
Hansen 1
Bravo 1-31
Elmer C. Von Glahn 1
S.F. & F.L. 4-2
Hahesy 36-1
Richgrove Community 1
Coles Levee A 26-29
Shell Posuncula 1
KCL 15X-24
Morris H C 1

03100634
03100004
03120135
03100739
02903645
03120005
02930540
02960650
02918565
02920544
10700112

------------

------------

Notes:
MD-measured depth in feet
Thicknesses queried when uncertain

298

0
0
0
0
0
0
0
0
0
0
0

Absent
Absent
Absent
Absent
Absent
Absent
Absent
Absent
Absent
Absent
Absent

Figure 1. Isopach (thickness) maps of major sand-dominated units within the


southern San Joaquin basin during middle Eocene time. Isopach maps of the Point
of Rocks Sandstone, Famoso Sandstone, and Metralla Sandstone are modified
from Clarke (1973), Weber (1973), Nilsen (1987), and Reid (1988). Region in
gray shading is poorly constrained due to lack of well penetrations of Eocene
strata. See Appendix A for oil wells used to constrain isopach mapping. Geologic
map modified from Ludington et al. (2007). Fault locations follow Chapman et al.
(2012). Oil Fields: C-Coalinga; KMD-Kettleman Middle Dome; LH-Lost Hills;
EH-Elk Hills; MS-Midway-Sunset; MV-Mountain View; KR-Kern River; TTrico. Other Abbreviations: B-Bakersfield; GF-Garlock fault; Fm-Formation;
SEM-San Emigdio Mountains; Ss - Sandstone; WWF-White wolf fault.

299

Figure 2. QFL diagram showing the sandstone composition of the Point of Rocks
Sandstone (blue) and Butano Sandstone (red). Mean values are shown as larger squares.
Data from Clarke (1973), Seiders and Cox (1992), Critelli and Nilsen (1996), and Taylor
(2004). Tectonic setting overlay from Dickinson et al. (1983). Qt-total quartz; F-feldspar;
L-lithic grains; K-potassium feldspar; P-plagioclase.

300

Figure 3. Cumulative (above) and normalized (below) detrital zircon U-Pb age
distributions of Eocene sandstone from the southern San Joaquin basin, San Emigdio
Mountains, and northern Salinian block. Data from Sharman et al. (2013). The gray
shading in the cumulative distributions encompass all samples within each group.
Northern and central Sierra Nevada samples from Cecil et al., 2010 and Cassel et al.,
2012. Southernmost Sierra Nevada samples from Lechler and Niemi (2011). NgNeogene; Pg-Paleogene; K-Cretaceous; Fm-Formation; Mbr-Member.

301

302

Figure 4. Regional compilation of detrital zircon U-Pb ages from Eocene sandstone
along the southern Oregon-California-northern Baja forearc (modified from Sharman et
al., in press). A) bar graphs of major detrital zircon age populations. B) cumulative
(above) and normalized (below) U-Pb age distributions for groups of forearc sandstone
samples.

Gray bar indicates depositional age range of the samples. Number of

samples/grains in parentheses.

303

304

Figure 5. Chronostratigraphic diagram of the southern San Joaquin basin. Paleogene time
scale from McDougall et al. (2007). Depositional age constraints from Bartow and
McDougall (1984), Milam (1985), Nilsen (1987), Almgren et al. (1988), Reid (1988),
Moxon (1990), Johnson and Graham (2005), Scheirer et al. (2007). Paleobathymetry
(water depth) from Clarke (1973), Harun (1984), Nilsen (1987), and Moxon (1990).

305

Figure 6. Well log cross-section across the southern San Joaquin basin emphasizing
middle Eocene stratigraphic relationships (see Fig. 1 for location). Spontaneous potential
and deep resistivity logs are shown on the left and right, respectively. The section is hung
on the top of the Tumey-Kreyenhagen Shale. Vertical exaggeration is ~16X.
constraints are from PS-AAPG (1957, 1958, 1959) and Malmblorg et al. (2008).

306

Age

307

Figure 7. Cartoon of Paleogene paleogeography of the Great Valley forearc and northern Salinian block. A) Traditional

paleogeographic model. Modified from Nilsen and Clarke (1975), Graham and Berry (1979), and Anderson et al. (2006). B)

Revised paleogeographic model (this study). Modified from Dumitru et al. (2012) and Sharman et al. (2013). BS-Butano

Sandstone; CF-Carmelo Formation; CS-Cantua Sandstone; FCB-Franciscan Coastal Belt; FS-Famoso Sandstone; MC-Markely

Canyon; MS-Markley Sandstone; PC-Princeton Canyon; PRS-Point of Rocks Sandstone; RS-Rocks Sandstone; SFBB-San
Francisco Bay Block.

308

APPENDIX B-2: Analytical Methods


Overview
Zircons were extracted from 3-5 kg sandstone samples using standard size,
hydrodynamic, density, and magnetic mineral separation techniques at Stanford
University. Sample locations were selected to ensure stratigraphic coverage of each
formation represented by more than one sample. U-Pb analyses of detrital zircons were
conducted using laser ablation-inductively coupled plasma-mass spectrometry (LA-ICPMS) at the University of Arizona Laserchron Center and at the University of California
Santa Cruz (Appendices B-5 and B-6). The results were conservatively filtered (< 3%
total data exclusion) to exclude less interpretable analyses based on the following criteria:
discordance (>30%), reverse discordance (>5%), common 204Pb content (>200-400
counts per second), and age precision (>10-15%).
University of Arizona Laserchron Center
Data from the University of Arizona Laserchron Center were analyzed using a
multicollector LA-ICP-MS (Nu HR ICPMS) with a 30 micron spot diameter (Appendix
B-5). An age cutoff of 800 Ma was used between common 204Pb-corrected 206Pb/207Pb
and 206Pb/238U ages. For an additional description of methods, refer to Gehrels et al.
(2008) and Cassel et al. (2012).
University of California Santa Cruz LA-ICP-MS Laboratory
Overview
Data from the University of California Santa Cruz LA-ICP-MS laboratory were
analyzed using a single-collector Element XR high-resolution magnetic-sector ICP-MS
and a Photon Machines Analyte.H 193 nm ArF excimer laser equipped with a Helex 2volume laser ablation cell. A 26 micron spot diameter was used for all analyses. Ablated
aerosol is transported through 4 mm OD Teflon tubing and through an all-Teflon Squid
signal smoothing manifold after addition of argon sample gas. The ATLEX 300i laser is
energy stabilized at 4.5 mJ. Downstream, a user settable beam attenuator provides energy
density control. An age cutoff of 800 Ma was used between common 207Pb-corrected
206

Pb/238U ages and common 204Pb-corrected 207Pb/206Pb ages.


309

Sample Preparation
As many as eight detrital and/or igneous zircon samples are mounted in rows on doublesided sticky tape using a mask cut from the tape backing film. Several fragments of
zircon standard are mounted in smaller rows at the center of the mount. Sri Lankan zircon
(SL2) (563 Ma; Gehrels et al., 2008) is used as the primary standard and Plesovice (337
Ma; Slma et al., 2008) is used as a secondary standard. The grains are then potted in a
2.5 cm ring form using Struers Epofix epoxy. Cured mounts are removed from their ring
forms and the upper meniscus is cut off with a parting tool and lathe. The back of the
mount is lightly polished to permit transmitted light imaging. The mount surface is lightly
polished with 1,500 grit sandpaper followed by 9 m and then 3 m Struers polishing
compounds on a LaboPol 5 lap wheel. Optical and/or cathodoluminescene (CL) imaging
is conducted at the Stanford-USGS Microanalytical Center. If applied, the gold coat for
CL imaging is removed with a brief 1 m polish. All mounts are washed in 1% HNO3
and rinsed in MilliQ water prior to installation in the Helex-2 volume cell.
Tuning Parameters
Argon and helium gas flows and torch XY position are tuned to achieve Th/U 1, with
minimal (<0.2%) ThO. A well-tuned instrument with 5.76 J/cm2 laser energy density
yields typical 238U sensitivity on the order of 15,000 cps/ppm on NIST 612 glass (10
m/s raster, 43 m spot) and 9,000 cps/ppm on SL2 zircon (26 m spot). Higher U
sensitivity is possible although this comes at the expense of increased oxides and
dramatically decreased Th/U ratios.
Analytical Process
A series of four primary (SL2) and secondary standard (Plesovice) analyses are run at the
beginning and end of each session. Primary standards are run after every fifth unknown
analysis and are paired with a secondary standard after every tenth unknown analysis. A
series of two primary and two secondary standards are run between samples on the same
mount. This protocol allows the secondary standard to provide an accuracy and precision
monitor (approximately 15 secondary standard analyses are collected per 100 unknown
analyses). Each analysis consists of 30 seconds of background measurement (laser off),
310

30 seconds of sample measurement (laser firing), and 20 seconds of delay before the next
analysis to purge the previous sample.
Data Reduction
Data are reduced using the Iolite (version 2.2; www.iolite.org.au/Iolite.html) (Paton et al.,
2010) and VisualAge (version 2012-05-29; http://www.japetrus.net/va/) add-ons for Igor
Pro. We maintain that the exponential detrending algorithm based on the down-hole
fractionation of standards is a more robust approach than linear regression, ratio-ofmeans, and mean-of-ratios data reduction methodologies. Specifically, we favor the
detrending approach (Paton et al., 2010) because unknowns that exhibit different
fractionation behavior from the primary standards maintain a temporal trend after downhole correction and result in a higher standard deviation when the resultant ratios are
averaged. Thus, the estimate of internal precision accounts for any differences between
the ablation behavior of the standards and the unknowns.
A combination of triggered acquisition and the reproducible sample washout of
the Helex-2 allow for automatic integration based on fixed time windows without
modification. Integration regions are resized if: (1) drill-through is observed based on a
rapid decrease in total beam prior to the end of lasing, or (2) spikes of 204Pb are observed
in the background corrected 204Pb signal. Total 204Pb backgrounds (Pb +Hg) are typically
~300 cps +/- 10 cps. Other than 204Pb spikes related to inclusions or correlated with high
U zones (possibly reflecting high radiation damage), average background-subtracted
signals are typically less than a conservatively estimated limit of detection of three times
the standard error of the background signal. For this reason, we do not apply a 204Pb
correction and instead utilize the 207Pb-corrected 206Pb/238U age. The 207Pb correction is
calculated using Isoplot (version 4.15) assuming an initial Pb composition based on a
two-stage model of terrestrial lead isotope evolution (Stacey and Kramers, 1975;
Appendix B-5). A 204Pb-corrected 207Pb/206Pb age is applied to grains with 206Pb/238U
ages greater than 800 Ma (Appendix B-5).
Typical Laser Parameters
Laser Energy: 4.5 mJ
311

Spot size: 26 m or 34 m diameter spot


Rep. Rate: 8 Hz
Shot count: 240 (30 s ablation)
Attenuator setting: 30% 5.76 J/cm2
Average Ablation Rate: 0.75 nm/pulse 20 m pit depth
Gas Flow Parameters
Nebulizer (argon) ~0.700 L/min
Add 1 (sampling cone; helium) ~0.700 L/min
Add 2 (cell pressure; helium) ~0.500 L/min)
Method Parameters
202

Hg (15 ms)

204

Pb (15 ms)

206

Pb (15 ms)

207

Pb (30 ms)

208

Pb (15 ms)

232

Th (3 ms)

235

U (6 ms)

238

U (3 ms)

Total Dwell time (102 ms)


Cycle period (120 ms; 8.3 Hz)
Duty cycle (84%)

312

APPENDIX B-3: Kolgomorov-Smirnov Statistics


The Butano-Point of Rocks correlation can be quantitatively evaluated using the
Kolgomorov-Smirnov (K-S) statistic. The K-S test is a nonparametric assessment of
whether two distributions are different from each other, and it tests the null hypothesis
that two distributions are derived from the same parent population (Press et al., 1986).
The K-S test is applicable to unbinned distributions that are functions of a single
independent variable and is based on the maximum separation (D-value) of the
cumulative distribution functions (CDFs) between two detrital zircon age distributions
(Press et al., 1986). Sample error was used in the CDFs which effectively smooths an
otherwise step-like distribution and yields a more conservative result (i.e., difficult to
reject the null hypothesis). All statistical calculations use the software of J. Guynn
(https://sites.google.com/a/laserchron.org/laserchron/).
This approach reveals that the Butano Sandstone and Point of Rocks Sandstone
are statistically distinct from each other at a 95% confidence level (i.e., P-value < 0.05;
Table 2). In addition, the K-S test is unable to distinguish between samples from the same
geographic areas at a 95% confidence level (i.e., northern Salinian block, Great Valley
forearc, and San Emigdio Mountains), with the exception of one Butano Sandstone
sample (BUT-4) that is statistically distinct from three other northern Salinian block
samples (BUT-2, BUT-3, and SJB-1) at 95% confidence (Table 2). In addition, the
northern Salinian block samples are indistinguishable from the San Emigdio Mountain
samples with the exception of one Butano Sandstone sample (BUT-4) that is distinct
from SEF-1 and TEJ-2 at 95% confidence (Table 2).

313

APPENDIX B-4: Discussion of northern Gabilan Range western San Emigdio


Mountain piercing points

Correlation of Logan quartz gabbro and anorthositic gabbro with the Western San
Emigdio Mafic Complex (WSEMC)
Isolated exposures of Middle- to Late-Jurassic gabbroic rocks are present in the
central Coast Ranges adjacent to the San Andreas fault at several locations: near the town
of Logan in the northern Gabilan Range, Gold Hill in the southern Diablo Range, and in
the western San Emigdio Mountains (Ross, 1970; Ross, 1984). Ross (1970) noted the
petrographic, geochronological, and geochemical similarity of these units and inferred
that their distribution is a result of about 320 km of displacement and slivering along the
San Andreas fault.
The isolated exposures at Logan and Gold Hill comprise virtually identical
hornblende quartz gabbro and anorthositic gabbro that were interpreted to be of oceanic
affinity (Ross, 1970). The Gold Hill exposures were interpreted to represent an isolated
sliver of gabbroic rocks left behind during strike-slip faulting (Ross, 1970). The western
San Emigdio mafic complex (WSEMC; Chapman et al., 2012) was identified as a likely
candidate for the parent mass of the Logan and Gold Hill gabbroic rocks based on
somewhat similar petrographic and geochemical characteristics (Ross, 1970). Ross
(1970) noted that this correlation was less certain, as a result of imperfect petrographic
similarity with the Logan-Gold Hill gabbroic rocks, but still plausible, especially since
there are no known alternative occurrences of gabbroic rocks that could provide the
distinctive mafic lithologies found at Logan and Gold Hill.
Correlation of the San Juan Bautista Formation and Tejon Formation
Both the Logan gabbroic rocks and the WSEMC are overlain by Eocene to
Miocene sedimentary successions (Nilsen, 1984; Nilsen, 1987). Nilsen (1984) suggested
that Eocene strata that overlie the Logan gabbroic rocks (San Juan Bautista Formation)
and the WSEMC (Tejon Formation) are correlative based on the similarity of the
stratigraphic successions present at these locations. Both formations include a basal
conglomerate that rests directly on crystalline basement, a middle shale unit, and an
314

upper sandstone unit (Nilsen, 1984; Nilsen, 1987). These units are unnamed for the San
Juan Bautista Formation (Nilsen, 1984), but are formal members of the Tejon Formation
(Uvas Conglomerate member, Liveoak Shale member, and Metralla Sandstone member;
Nilsen, 1987).
Discussion of the northern Gabilan Range western San Emigdio Mountain
piercing points
While the correlation of gabbroic rocks and overlying Eocene strata in the vicinity
of the northern Gabilan Range and western San Emigdio Mountains has been supported
in the literature (Ross, 1984; Dickinson, 1997; Dickinson et al., 2005; Chapman et al.,
2012), we consider these piercing points to be tenuous for the following reasons:
1)

The exposures of gabbroic rocks at Logan and in the western San Emigdio

Mountains are relatively small and isolated (~5 km2 and 18 km2, respectively). In both
cases they are onlapped by sedimentary cover that obscures their true extent. In the case
of the WSEMC, present-day outcrops are located ~4 km from the trace of the San
Andreas fault, requiring projection of the outcropping lithologies to the San Andreas
fault. The combination of these factors results in a relatively high spatial uncertainty for
these piercing points.
2)

The origin and structural history of the Logan-Gold Hill-WSEMC body is

controversial. Originally interpreted as having oceanic affinity akin to the Coast Range
ophiolite (Ross, 1970), more recent work has suggested similarity with lithologies in the
southwestern Foothills Belt of the Sierra Nevada (Chapman et al., 2012). Chapman et al.
(2012) conclude that the Logan-Gold Hill-WSEMC body is an allochthon that was
structurally emplaced by gravitational collapse from ca. 85-80 Ma along a detachment
bounded to the southeast by the proto-White Wolf fault. This event is broadly
synchronous with, or even predates, the structural emplacement of the northern Salinian
block during Late Cretaceous to Paleocene time (Suppe, 1970; Page, 1981; Dickinson,
1983; Hall, 1991; Chapman et al., 2012; Fig. 4). Offsets of the Logan-WSEMC body are
~100 km less than offsets between Late Cretaceous plutonic rocks of the northern
Salinian block and the western edge of Sierran basement beneath the Great Valley
(Dickinson et al., 2005). Such a large discrepancy between offset basement features that
315

were apparently emplaced during a similar time frame is difficult to reconcile. Given
these uncertainties, additional research is needed to: (1) establish the structural and
geochronological relationship between the Jurassic WSEMC and the Cretaceous granitic
and gneissic lithologies of the southernmost Sierra Nevada, and (2) corroborate the
Logan-WSEMC correlation using modern methodologies.
3)

Poor outcrop quality, limited exposure, and revised interpretations of field

relationships of the San Juan Bautista Formation hamper its definitive correlation with
the Tejon Formation. For example, the basal conglomerate unit of the San Juan Bautista
Formation is reported to overlie anorthositic gabbro about 2 km northwest of the town of
San Juan Bautista in limited exposures (<0.04 km2) within an abandoned quarry (Nilsen,
1984). However, attempts to locate this unit in its mapped location have proven
unsuccessful; instead Eocene shale that is in probable fault contact with the crystalline
basement is present locally (R. McLaughlin, personal correspondence). We consider the
absence of overlying conglomerate at Logan or San Juan Bautista to suggest that the San
Juan Bautista Formation and Tejon Formation cannot be uniquely correlated to each
other. In addition, the lack of conspicuous conglomerate overlying the Logan gabbroic
rocks suggests that these mafic lithologies cannot be precisely correlated to those in the
WSEMC where sandstone and conglomerate of the Uvas Conglomerate member form
prominent ridges (Nilsen, 1987).
4)

Correlation of the Logan gabbroic rocks with the WSEMC juxtaposes the

incompatible detrital zircon age distributions of the Butano Sandstone with the Point of
Rocks Sandstone across the San Andreas fault (Figs. 1, 2). See discussion in main text for
more details.

316

APPENDIX B-5: U-Th-Pb isotope composition of detrital zircons analyzed at the


University of Arizona Laserchron Center (available electronically on Stanford
Universitys library website)

317

APPENDIX B-6: U-Th-Pb Isotopic Compositions of Detrital Zircons Analyzed at


the University of California Santa Cruz (available electronically on Stanford
Universitys library website)

318

APPENDIX B-7: Explanation of tectonic and depositional events shown in Figure 4


Age Range
(Ma)

Reference(s)

Neogene displacement on the San


Andreas fault

18-0

Atwater (1989)

Eruption of the Pinnacles-Neenach


volcanic field

24.5-22

Matthews (1976), Dickinson


(1997)

Initiation of Pacific-North American


plate interaction

29-24

Atwater (1989), Wilson et al.


(2005)

Deposition of the Point of Rocks


Sandstone

48-37

Clarke (1973), Milam (1985),


Almgren et al. (1988)

Deposition of the Butano Sandstone

54-38

Poore and Brabb (1977),


Seiders and Cox (1992)

Structural emplacement of the


northern Salinian block

90-58

Suppe (1970), Page (1981),


Dickinson (1983), Hall
(1991), Saleeby (2003),
Chapman et al. (2012)

Cooling ages of northern Salinian


block plutons

120-80

Mattinson (1990), Kistler and


Champion (2001), Barth et al.
(2003)

Tectonic or Depositional Event

Biostratigraphic age ranges are correlated to absolute time using McDougall (2007)

319

APPENDIX B-8: Explanation of San Andreas fault piercing points used in Figure 4

Cross-fault tie
a
Temblor Range conglomerates
derived from Gabilan Range

Age
Range
(Ma)
13.56.5

Displacement
Range (km)
260-225

Huffman (1972)

Reference(s)

Relizian-Luisian basinal facies

17.513.5

320-260

Graham et al. (1989)

Saucesian paleobathymetry

22.917.5

325-320

Stanley (1987)

Early Miocene shorelines and


volcanic rocks

23-20.8

305-280

Turner (1968),
Huffman (1972),
Graham et al. (1989)

Pinnacles-Neenach volcanic centers

24.5-22

305-325

Matthews (1976),
Dickinson (1997)

Castle Rock-Recruit Pass submarine


fan

2817.5

320-315

Graham et al. (1989)

Zemmorian paleobathymetry

30.322.9

330-325

Stanley (1987)

Eocene forearc strata of the northern


Gabilan Range and San Emigdio
Mountains
Butano-Point of Rocks submarine
fan

54.635.2

320-305

Nilsen (1984)

48-38

330-300

Clarke (1973), Poore


and Brabb (1977),
Milam (1985),
Almgren et al.
(1988), Seiders and
Cox (1992)

54-38

>365

This study

120-80

>415

Mattinson (1990),
Kistler and
Champion (2001),
Barth et al. (2003),
Dickinson et al.
(2005)

Provenance of northern Salinian


block and southern Sierra Nevada
forearc strata
Cretaceous granitic rocks of the
northern Salinian block and the
southern Sierra Nevada

Biostratigraphic age ranges are correlated to absolute time using McDougall (2007)

320

REFERENES (APPENDICES B-2 to B-8)


Almgren, A.A., Filewicz, M.V., and Heitman, H.L., 1988, Lower Tertiary foraminiferal
and calcareous nannofossil zonation of California: An overview and
recommendation, in Filewicz, M.V., and Squires, R.L., eds., Paleogene
Stratigraphy, West Coast of North America: Los Angeles, Pacific Section, SEPM
(Society for Sedimentary Geology), v. 58, p. 83-106.
Barth, A.P., Wooden, J.L., Grove, M., Jacobson, C.E., and Pedrick, J.N., 2003, U-Pb
zircon geochronology of rocks in the Salinas Valley region of California: A
reevaluation of the crustal structure and origin of the Salinian block: Geology, v.
31, p. 517-520.
Dickinson, W.R., 1983, Cretaceous sinistral strike slip along Nacimiento fault in coastal
California: American Association of Petroleum Geologists Bulletin, v. 67, p. 624
645.
Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy,
efficiency, and spatial resolution of U-Pb ages by laser ablationmulticollector
inductively coupled plasmamass spectrometry: Geochemistry, Geophysics,
Geosystems, v. 9, p. Q03017, doi: 10.1029/ 2007GC001805.
Huffman, O.F., 1972, Lateral displacement of upper Miocene rocks and the Neogene
history of offset along the San Andreas fault in central California: Geological
Society of America Bulletin, v. 83, p. 2913-2946.
Kistler, R.W., and Champion, D.E., 2001, Rb-Sr whole-rock and mineral ages, K-Ar,
40
Ar/39Ar, and U-Pb mineral ages, and strontium, lead, neodymium, and oxygen
isotopic compositions for granitic rocks from the Salinian composite terrane,
California: U.S. Geological Survey Open-File Report 01-453, 84 p.
Mattinson, J.M., 1990, Petrogenesis and evolution of the Salinian magmatic arc, in
Anderson, J.L., ed., The nature and origin of Cordilleran magmatism: Geological
Society of America Memoir 174, p. 237250.
McDougall, K., 2007, California Cenozoic biostratigraphy Paleogene, in Hosford
Scheirer, A., ed., Petroleum systems and geologic assessment of oil and gas in the
San Joaquin Basin Province, California: U.S. Geological Survey Professional
Paper 1268, p. 1-56.
Milam, R.W., 1985, Biostratigraphy and sedimentation of the Eocene and Oligocene
Kreyenhagen Formation, central California [Ph.D. thesis]: Stanford University,
240 p.
Nilsen, T.H., 1987, Stratigraphy and sedimentology of the Eocene Tejon Formation,
western Tehachapi and San Emigdio Mountains, California: U.S. Geological
Survey Professional Paper 1268, 110 p.
Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., and Mass, R., 2010,
Improved laser ablation U-Pb zircon geochronology through robust downhole
fractionation correction: Geochemistry, Geophysics, Geosystems, v. 11, p.
Q0AA06, doi:10.1029/2009GC002618.
Poore, R.Z., and Brabb, E.E., 1977, Eocene and Oligocene planktonic foraminifera from
the upper Butano Sandstone and type San Lorenzo Formation, Santa Cruz
Mountains, California: Journal of Foraminiferal Research, v. 7, p. 249-272.

321

Press, W.H., Flannery, B.P., Teukolsky, S.A., and Vetterling, W.T., 1986, Numerical
recipes, The art of scientific computing: Cambridge, Cambridge University Press,
186 p.
Ross, D.C., 1970, Quartz gabbro and anorthositic gabbro: Markers of offset along the San
Andreas fault in the California Coast Ranges: Geological Society of America
Bulletin, v. 81, p. 3647-3662.
Seiders, V.M., and Cox, B.T., 1992, Place of origin of the Salinian block, California, as
based on clast compositions of Upper Cretaceous and lower Tertiary
conglomerates: U.S. Geological Survey Professional Paper 1526, 80 p.
Slma, J., Koler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood,
M.S.A., Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U., Schoene, B.,
Tubrett, M.N., and Whitehouse, M.J., 2008, Pleovice zircon A new natural
reference material for U-Pb and Hf isotopic microanalysis: Chemical Geology, v.
249, p. 1-35.
Stacey, J.S., and Kramers, J.D., 1975, Approximation of terrestrial lead isotope evolution
by a two-stage model: Earth and Planetary Science Letters, v. 26, p. 207-221.
Stanely, R.G., 1987, New estimates of displacement along the San Andreas fault in
central California based on paleobathymetry and paleogeography: Geology, v. 15,
p. 171-174.
Turner, D.L., 1968, Potassium-argon dates concerning the Tertiary foraminiferal time
scale and San Andreas fault displacement [Ph.D. thesis]: Berkeley, University of
California, 99 p.
Wilson, D.S., McCrory, P.A., and Stanley, R.G., 2005, Implications of volcanism in
coastal California for the Neogene deformation history of western North America:
Tectonics, v. 24, p. TC3008.

322

APPENDIX C: SUPPLEMENTARY MATERIAL FOR CHAPTER 3

Summary of Contents
Appendix C-1: Detailed measured sections
Appendix C-2: Slump fold orientations
Appendix C-3: Slump fault orientations
Appendix C-4: Unrestored versus restored slump fault orientations
Appendix C-5: Full-size cliff photomosaics
Appendix C-6: Wave-cut platform orthophotos
Appendix C-7: Biostratigraphy
Appendix C-8: Paleocurrent measurements

Appendices C-5 and C-6 are available electronically as supplemental files on Stanford
Universitys library website

323

APPENDIX C-1: Detailed measured sections

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

Fold
ID
FD067
FD145
FD047
FD048
FD066
FD155
FD154
FD156
FD096
FD094
FD119
FD118
FD120
FD121
FD122
FD008
FD149
FD009
FD148
FD124
FD147
FD125
FD126
FD159
FD093
FD158
FD127
FD128
FD129
FD131
FD117
FD134
Location
Stm 19 South
Stm 19 South
Stm 19 South
Stm 19 South
Stm 19 South
Transition
Transition
Transition
Transition
Transition
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South

R
R1
R1
R1
R1
R1
R2
R2
R2
R2
R2
R3
R3
R3
R3
R3
R3
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4

Type
Outcrop
Outcrop
WCP
Outcrop
WCP
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
WCP
WCP
WCP
WCP
WCP
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
Outcrop
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
Outcrop
WCP

Easting
174.6348
174.6345
174.6330
174.6337
174.6326
174.6332
174.6325
174.6332
174.6332
174.6333
174.6321
174.6315
174.6316
174.6318
174.6325
174.6329
174.6319
174.6303
174.6309
174.6311
174.6308
174.6289
174.6291
174.6296
174.6300
174.6298
174.6297
174.6293
174.6296
174.6291
174.6304
174.6295

Conf.
Good
Good
Good
Good
Good
Fair
Fair
Fair
Good
Fair
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good

Trd
284
-293
285
273
313
334
195
13
13
42
198
22
30
203
61
156
3
159
153
--317
167
147
164
177
356
324
346
330
6

Plng
4
-8
6
10
9
4
0
0
1
16
0
6
0
2
5
6
21
1
11
--43
7
20
1
36
6
45
6
11
2

Fold Axis
Stk
276
294
281
103
90
313
333
-7
6
222
208
202
209
24
30
343
340
343
4
342
182
161
349
329
164
156
174
180
179
318
6

Dip
23
15
32
83
87
70
83
-6
8
85
72
83
80
73
17
44
45
26
20
9.5
48
65
78
88
79
62
41
60
25
46
88

Dir
N
N
N
S
S
NE
E
-E
E
NW
W
W
NW
SE
SE
E
E
E
E
E
W
W
E
E
W
SW
W
W
W
NE
E

Axial Plane

APPENDIX C-2: Slump fold orientations

Southing
-38.5361
-38.5364
-38.5386
-38.5416
-38.5422
-38.5443
-38.5449
-38.5459
-38.5476
-38.5484
-38.5512
-38.5518
-38.5519
-38.5520
-38.5523
-38.5531
-38.5561
-38.5562
-38.5567
-38.5590
-38.5596
-38.5599
-38.5599
-38.5600
-38.5600
-38.5604
-38.5613
-38.5616
-38.5622
-38.5628
-38.5629
-38.5632

Vergence*
186
204
191
--223
--277
276
132
118
112
119
294
300
253
250
253
274
252.2
92
71
-239
-66
84
90
89
228
--

Interlimb
Angle
16
13
22
122
130
133
60
-38
28
116
101
80
74
116
14
6
56
15
25
59
15
34
117
47
115
28
43
32
16
42
153

Symmetry
S
-S
--S
--S
S
Z
S
Z
Z
Z
S
Z
S
Z
Z
--Z
-Z
-S
Z
Z
Z
S
--

N
2
3
16
4
5
7
1
1
2
2
2
2
2
2
3
7
4
3
2
2
2
4
3
2
8
2
2
4
2
7
1
2

342

FD144
FD133
FD132
FD025
FD136
FD137
FD143
FD116
FD139
FD141
FD115
FD138
FD140
FD142
FD011
FD013
FD014
FD111
FD016
FD017
FD110
FD018
FD019
FD020
FD021
FD022
FD109
FD023
FD108
FD113
FD107
FD106
FD146
FD097
FD105
FD084
FD150

Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Ounutae
Poporotaupo
Waiongaro
South
Waiongaro
South
R6

R6

R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
--

WCP

Outcrop

WCP
WCP
WCP
Outcrop
WCP
WCP
WCP
Outcrop
WCP
WCP
Outcrop
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
Outcrop
WCP
WCP
Outcrop
Outcrop
WCP

-38.6015

-38.6011

-38.5633
-38.5633
-38.5635
-38.5644
-38.5649
-38.5659
-38.5669
-38.5672
-38.5675
-38.5678
-38.5680
-38.5682
-38.5685
-38.5689
-38.5725
-38.5727
-38.5728
-38.5731
-38.5731
-38.5732
-38.5733
-38.5735
-38.5738
-38.5745
-38.5750
-38.5751
-38.5751
-38.5753
-38.5753
-38.5763
-38.5764
-38.5764
-38.5765
-38.5768
-38.5863

174.6265

174.6281

174.6298
174.6293
174.6291
174.6304
174.6285
174.6289
174.6290
174.6304
174.6290
174.6288
174.6304
174.6294
174.6291
174.6296
174.6297
174.6298
174.6296
174.6296
174.6296
174.6296
174.6295
174.6298
174.6296
174.6292
174.6292
174.6292
174.6291
174.6293
174.6292
174.6302
174.6295
174.6295
174.6303
174.6302
174.6279

Good

Fair

Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Fair
Fair
Fair
Good
Good
Fair
Fair
Good
Good
Good
Fair
Fair
Good
Good
Good
Good
Good
Fair
Fair
Fair
Good

173

231

18
340
332
335
313
321
342
43
340
167
-156
166
139
282
293
218
215
190
209
15
196
57
243
256
227
344
242
233
140
247
290
-280
--

64

14

27
11
10
19
23
3
4
18
4
3
-7
9
3
10
38
24
3
17
17
2
11
16
3
16
22
40
22
16
1
24
21
-25
--

47

138

196
162
331
-139
144
342
4
161
356
143
145
152
137
-141
-37
28
197
196
21
57
256
78
50
208
58
56
333
71
283
128
-218

60

11

90
75
82
-77
47
51
24
73
81
45
32
39
74
-61
-74
36
56
59
75
79
89
84
86
50
53
76
6
82
72
30
-39

SE

SW

NA
W
E
-W
W
E
E
W
E
SW
W
W
W
-SW
-SE
SE
W
W
SE
SE
N
S
SE
NW
SE
SE
NE
S
N
SW
-NW

317

--

-72
---54
252
274
71
-53
55
62
47
---307
298
-106
-----118
----193
--128

56

43

82
124
113
-111
13
55
10
20
131
5
30
4
106
-53
168
115
32
45
20
135
158
139
152
160
32
157
156
66
149
153
0
64
4

--

-Z
---Z
S
S
Z
--S
S
S
-Z
-Z
Z
S
Z
-----Z
----S
----

3
4
2
3
2
5
5
5
7
4
2
6
7
5
4
3
4
2
4
4
2
3
4
3
3
3
2
3
2
3
3
2
0
3
2

343

FD085
FD151
FD153
FD086
FD152
FD087
FD005
FD077
FD080
FD078
FD006
FD088
FD130
FD059
FD074
FD089
FD090
FD091
FD060
FD058
FD073
FD083
FD092
FD071
FD082
FD064
FD069
FD062
FD063
FD003
FD057
FD002

Waiongaro
South
Waiongaro
South
Waiongaro
South
Waiongaro
South
Waiongaro
South
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Ngatupaku
Waiakapua
North
Waiakapua
North
Waiakapua
North
Waiakapua
North
Waiakapua
North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
R6
R6
R6
R6
R6
R6
R6
R6

R6

R6

R6

R6

R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6

R6

R6

R6

R6

Outcrop
Outcrop
WCP
Outcrop
Outcrop
Outcrop
WCP
Outcrop

WCP

Outcrop

Outcrop

WCP

WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP

Outcrop

WCP

WCP

Outcrop

-38.6144
-38.6179
-38.6189
-38.6204
-38.6204
-38.6205
-38.6211
-38.6215

-38.6144

-38.6133

-38.6131

-38.6131

-38.6052
-38.6073
-38.6089
-38.6089
-38.6089
-38.6090
-38.6091
-38.6094
-38.6098
-38.6099
-38.6100
-38.6100
-38.6102
-38.6102
-38.6102
-38.6103

-38.6051

-38.6051

-38.6051

-38.6032

174.6252
174.6248
174.6233
174.6241
174.6241
174.6239
174.6235
174.6239

174.6245

174.6251

174.6250

174.6239

174.6268
174.6260
174.6258
174.6258
174.6259
174.6254
174.6255
174.6251
174.6255
174.6251
174.6249
174.6253
174.6253
174.6250
174.6257
174.6253

174.6274

174.6269

174.6267

174.6278

Good
Good
Good
Good
Good
Fair
Good
Good

Good

Fair

Good

Fair

Good
Good
Good
Fair
Fair
Fair
Good
Fair
Good
Good
Good
Good
Good
Good
Good
Fair

Fair

Good

Fair

Fair

264
37
255
32
23
18
229
311

70

217

47

288

189
18
204
14
230
85
129
21
345
78
274
81
167
146
41
152

--

194

336

202

1
10
12
0
1
7
8
19

14

8
25
28
16
21
11
25
6
18
6
6
5
13
14
27
7

--

258
237
188
34
36
-50
278

255

179

242

--

160
246
74
202
75
257
143
220
337
259
100
258
159
329
224
146

194

162

172

201

8
28
13
42
31
-85
32

89

12

20

--

13
24
41
64
45
82
85
30
64
85
74
85
61
72
81
42

10

25

47

N
NW
W
SE
SE
-SE
N

NW

--

W
NW
S
W
S
N
SW
NW
NE
N
S
N
SW
NE
NW
SW

168
147
98
304
306
-320
188

--

--

152

--

70
156
344
112
345
--130
247
---69
--56

--

72

82

111

18
65
6
102
110
56
152
78

148

22

41

134
25
52
14
56
148
163
56
56
161
142
154
105
124
142
54

139

-7

109

S
Z
-S
S
-Z
S

--

--

--

S
Z
Z
Z
Z
--Z
S
---S
--S

--

2
3
4
3
2
3
3
7

4
4
5
3
2
3
3
3
6
3
6
3
3
3
3
2

344

FD061
FD001
FD099
FD042
FD041
FD056
FD040
FD039
FD055
FD100
FD053
FD052
FD104
FD051
FD037
FD038
FD050
FD103
FD034
FD033
FD035
FD102
FD036
FD046
FD032
FD027
FD031
FD028
FD043
FD101
FD114
FD029
FD030

Paritutu North
Paritutu North
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South
Paritutu South

R6
R6
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7

Outcrop
Outcrop
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
WCP
Outcrop
WCP
Outcrop
WCP
WCP
WCP
Outcrop
Outcrop

-38.6215
-38.6218
-38.6269
-38.6272
-38.6273
-38.6274
-38.6275
-38.6276
-38.6276
-38.6277
-38.6278
-38.6279
-38.6282
-38.6283
-38.6284
-38.6285
-38.6285
-38.6285
-38.6287
-38.6289
-38.6289
-38.6289
-38.6289
-38.6290
-38.6291
-38.6292
-38.6298
-38.6299
-38.6300
-38.6301
-38.6303
-38.6306
-38.6308

174.6239
174.6238
174.6221
174.6217
174.6224
174.6221
174.6221
174.6222
174.6219
174.6216
174.6215
174.6220
174.6219
174.6219
174.6218
174.6218
174.6221
174.6220
174.6211
174.6214
174.6210
174.6217
174.6210
174.6220
174.6216
174.6223
174.6218
174.6223
174.6219
174.6219
174.6219
174.6223
174.6224

Good
Fair
Good
Fair
Good
Good
Good
Good
Fair
Good
Fair
Fair
Good
Fair
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Fair
Good
Fair
Good
Fair
Good
Good

320
322
207
252
256
238
249
271
102
303
-285
264
267
273
296
271
327
291
253
273
283
274
278
285
264
299
245
230
315
251
86
30

19
10
3
14
7
9
15
13
1
15
-11
6
17
8
12
13
40
25
4
19
9
19
12
12
9
17
22
8
33
11
3
9

265
288
27
250
76
56
62
100
-283
284
248
82
89
270
126
259
265
291
252
271
282
272
276
283
88
312
58
228
136
242
273
30

Notes:
* Vergence is calculated as the up-dip direction of the axial plane
S-clockwise fold; Z-counter-clockwise fold (asymmetry is measured looking down-plunge of the fold axis)
N is the number of bedding measurements used to constrain the orientation and asymmetry of the fold

24
33
84
75
83
86
83
42
-38
34
40
81
84
74
51
50
43
89
83
89
81
87
83
86
62
84
82
86
81
46
65
89

N
N
SE
NW
SE
SE
SE
S
-N
N
NW
S
S
N
SW
N
N
NE
NW
N
N
N
N
N
S
NE
SE
NW
SW
NW
N
SE

175
198
--346
326
332
10
-193
194
158
352
-180
36
169
175
-162
-192
-186
-358
----152
183
--

52
39
160
166
138
150
137
42
161
16
14
65
129
60
118
25
40
26
110
159
154
141
148
136
149
111
163
158
160
158
120
94
132

S
S
-S
Z
Z
Z
Z
-S
-S
Z
-S
Z
S
S
-S
-S
-S
-Z
----S
Z
--

3
10
2
4
6
3
5
6
2
4
2
4
4
2
3
3
3
5
3
4
3
3
3
3
3
3
3
3
3
3
3
4
3

345

Fault_ID
Group
Normal faults
SF029
1
SF034
1
SF035
1
SF019
3
SF154
5
SF155
5
SF156
5
SF157
5
SF158
5
SF159
5
SF160
5
SF166
5
SF167
5
SF168
5
SF106
6
SF107
6
SF108
6
SF109
6
SF086
7
SF089
7
SF090
7
SF094
7
SF097
7
SF101
7
SF084
7
SF085
7
SF087
7
SF088
7
SF093
7
SF099
7
SF100
7

APPENDIX C-3: Slump fault orientations


Apparent
Disp.
(cm)*
5-20
100
5-20
5-20
0-5
5-20
5-20
0-5
0-5
0-5
0-5
5-20
5-20
5-20
20-100
20-100
5-20
5-20
5-20
0-5
0-5
5-20
5-20
5-20
0-5
0-5
0-5
5-20
20-100
0-5
0-5

Apparent
Disp.
(cm)*
----Fair
Fair
Fair
Fair
Fair
Fair
Good
Good
Good
Good
Fair
Good
Good
Good
Fair
Fair
Fair
Fair
Fair
Fair
Good
Good
Good
Good
Good
Good
Good

Confidence
280
240
15
296
90
92
90
288
277
278
253
7
355
356
22
25
27
203
337
10
265
2
18
19
335
345
325
24
331
151
2

Stk
70
19
38
14
88
89
29
64
78
43
43
70
75
64
77
64
54
31
35
29
75
68
70
55
11
8
16
43
46
35
67

Dip
N
NW
E
N
S
S
S
N
N
N
N
E
E
E
E
E
E
W
NE
E
N
E
E
E
NE
NE
NE
E
NE
SW
E

Dir
-259
196
--------20
20
20
----59
-----59
59
59
-----

Stk

-39
12
--------18
18
18
----23
-----23
23
23
-----

Dip

-N
W
--------E
E
E
----SE
-----SE
SE
SE
-----

Dir

280
95
15
296
90
92
90
288
277
278
253
5
352
350
22
25
27
203
306
10
265
2
18
19
266
260
274
24
331
151
2

Stk

70
22
50
14
88
89
29
64
78
43
43
53
59
48
77
64
54
31
38
29
75
68
70
55
24
22
29
43
46
35
67

Dip

N
S
E
N
S
S
S
N
N
N
N
E
E
E
E
E
E
W
NE
E
N
E
E
E
N
N
N
E
NE
SW
E

Dir

10
185
105
26
180
182
180
18
7
8
343
95
82
80
112
115
117
293
36
100
355
92
108
109
356
350
4
114
61
241
92

Vergence#

Fault (corr.)

R
20
100
20
15
5
10
19
1
4
4
1
10
10
10
30
30
15
10
20
3
2
10
10
10
3
3
5
10
30
5
3

Bedding

Location
R1
R1
R1
R1
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4

Fault

Kopuai North
Kopuai North
Kopuai North
Kopuai North
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
Waioroko North
Waioroko North
Waioroko North
Waioroko North
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South

346

SF102
SF141
SF219
SF220
SF221
SF222
SF223
SF224
SF225
SF226
SF188
SF187
SF189
SF190
SF191
SF192
SF193
SF194
SF197
SF198
SF199
SF200
SF201
SF202
SF212
SF205
SF206
SF207
SF209
SF210
SF213
SF214
SF230
SF249
SF242
SF246
SF247
SF248

7
8
8
8
8
8
8
8
8
8
9
9
9
9
9
9
9
9
9
9
9
9
9
9
10
10
10
10
10
10
10
11
12
13
13
13
13
13

Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Poporotaupo North
Poporotaupo North
Poporotaupo North
Poporotaupo North
Poporotaupo North

R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5

50
10
2
2
2
10
5
5
15
10
3
3
3
7
7
5
1
2
1
1
1
3
10
2
5
15
3
10
5
2
5
20
10
5
5
5
20
3

20-100
5-20
0-5
0-5
0-5
5-20
0-5
0-5
5-20
5-20
0-5
0-5
0-5
5-20
5-20
0-5
0-5
0-5
0-5
0-5
0-5
0-5
5-20
0-5
0-5
5-20
0-5
5-20
0-5
0-5
0-5
5-20
5-20
0-5
0-5
0-5
5-20
0-5

Good
Fair
Fair
Fair
Fair
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Fair
Fair
Fair
Good
Good
Good
Good

143
195
205
210
60
6
347
1
55
68
37
343
18
7
14
20
148
330
264
244
222
331
312
306
303
98
98
92
90
100
247
168
336
196
260
142
38
341

33
80
6
11
1
15
38
30
74
76
61
41
61
59
40
63
76
28
51
50
41
31
36
55
51
69
71
62
61
58
82
85
70
41
55
49
43
59

SW
W
W
NW
E
E
NE
E
SE
SE
SE
E
E
E
E
E
SW
NE
N
NW
NW
NE
NE
NE
NE
S
S
S
S
S
NW
W
NE
W
N
SW
SE
E

--157
157
157
145
145
145
11
11
185
185
185
185
185
---145
145
145
194
194
194
-------128
89
--134
134
--

--44
44
44
46
46
46
23
23
20
20
20
20
20
---21
21
21
31
31
31
-------33
23
--15
15
--

--SW
SW
SW
W
W
W
E
E
W
W
W
W
W
---SW
SW
SW
W
W
W
-------SW
S
--SW
SW
--

143
195
330
323
338
337
338
343
63
75
33
349
17
7
12
20
148
330
275
259
247
349
336
320
303
98
98
92
90
100
247
175
346
196
260
145
23
341

33
80
40
38
44
58
82
72
58
65
78
60
81
79
60
63
76
28
63
56
41
57
57
71
51
69
71
62
61
58
82
61
63
41
55
34
47
59

SW
W
NE
NE
E
NE
E
E
SE
E
SE
E
E
E
E
E
SW
NE
N
N
N
E
NE
NE
NE
S
S
S
S
S
NW
W
E
W
N
SW
SE
E

233
285
60
53
68
67
68
73
153
165
123
79
107
97
102
110
238
60
5
349
337
79
66
50
33
188
188
182
180
190
337
265
76
286
350
235
113
71

347

SF250
SF236
SF237
SF238
SF239
SF240
SF081
SF082
SF037
SF038
SF039
SF040
SF041
SF042
SF043
SF044
SF045
SF046
SF047
SF048
SF049
SF050
SF051
SF070
SF069
SF071
SF072
SF073
SF074
SF053
SF061
SF062
SF063
SF064
SF065
SF066
SF067
SF068

13
14
14
14
14
14
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
16
17
17
17
17
17
17
17
17
17

Poporotaupo North
Poporotaupo Stream
Poporotaupo Stream
Poporotaupo Stream
Poporotaupo Stream
Poporotaupo Stream
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Ngatupaku South
Waiakapua North
Waiakapua North
Waiakapua North
Waiakapua North
Waiakapua North
Waiakapua North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North

R5
R5
R5
R5
R5
R5
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6

39
10
3
3
2
30
3
3
70
5
3
3
2
2
70
3
10
2
2
2
2
1
100
5
30
100
100
100
150
3
5
20
30
10
3
150
10
70

20-100
5-20
0-5
0-5
0-5
20-100
0-5
0-5
20-100
0-5
0-5
0-5
0-5
0-5
20-100
0-5
5-20
0-5
0-5
0-5
0-5
0-5
100
0-5
20-100
100
100
100
100
0-5
0-5
5-20
20-100
5-20
0-5
100
5-20
20-100

Good
Good
Good
Good
Good
Good
Good
Good
---------------Fair
Good
Good
---Fair
Good
Good
Good
Good
Good
Good
Good
Good

5
267
262
284
285
258
122
87
31
255
256
64
271
282
87
103
76
124
134
139
2
288
40
264
294
336
273
2
24
321
165
234
343
11
266
27
315
116

20
59
55
69
48
51
42
46
53
35
56
61
61
66
38
45
70
62
64
46
80
83
67
43
47
22
29
40
46
76
54
76
27
35
56
42
18
18

E
N
N
N
N
N
SW
S
E
N
N
SE
N
N
S
S
S
S
S
S
E
N
SE
N
NE
NE
N
E
E
NE
W
NW
E
E
N
E
NE
SW

-358
358
358
358
358
-------------------254
----------87
--

-15
15
15
15
15
-------------------43
----------34
--

-E
E
E
E
E
-------------------N
----------S
--

5
258
252
278
271
247
122
87
31
255
256
64
271
282
87
103
76
124
134
139
2
288
40
264
294
42
273
2
24
321
165
234
343
11
266
27
285
116

20
60
58
66
45
55
42
46
53
35
56
61
61
66
38
45
70
62
64
46
80
83
67
43
47
44
29
40
46
76
54
76
27
35
56
42
48
18

E
N
N
N
N
NW
SW
S
E
N
N
SE
N
N
S
S
S
S
S
S
E
N
SE
N
NE
SE
N
E
E
NE
W
NW
E
E
N
E
N
SW

95
348
342
8
1
337
212
177
121
345
346
154
1
12
177
193
166
214
224
229
92
18
130
354
24
132
3
92
114
51
255
324
73
101
356
117
15
206

348

SF001
SF002
SF004
SF005
SF006
SF007
SF008
SF009
SF060
SF056
SF057
SF058
SF059
SF124
SF127
SF128
SF137
SF138
SF116
SF117
SF118
SF119
SF120
SF129
SF130
SF135
SF136
Reverse faults
SF027
SF028
SF030
SF031
SF032
SF033
SF036
SF020
SF021
1
1
1
1
1
1
1
2
2

17
17
17
17
17
17
17
17
17
17
17
17
17
18
18
18
18
18
18
18
18
18
18
18
18
18
18

Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North

Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North

R1
R1
R1
R1
R1
R1
R1
R1
R1

R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7
R7

10
200
20
3
3
3
3
---

NA
NA
NA
NA
NA
NA
NA
NA
15
5
2
3
4
5
5
2
3
3
3
10
15
20
50
10
2
3
10

--------5-20
0-5
0-5
0-5
0-5
0-5
0-5
0-5
0-5
0-5
0-5
5-20
5-20
5-20
20-100
5-20
0-5
0-5
5-20

----------

---------Very good
Very good
Very good
Very good
Fair
Fair
Fair
Fair
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Good

118
157
135
271
281
220
139
295
270

235
284
226
221
287
189
12
12
13
183
192
189
175
223
6
32
245
241
285
343
75
40
15
43
335
253
257

84
17
20
14
51
59
21
16
19

31
33
58
66
53
55
42
57
32
66
66
59
63
61
64
54
62
87
5
40
32
28
41
25
87
54
28

SW
W
SW
N
N
N
SW
N
N

NW
N
NW
NW
NE
W
E
E
E
W
W
W
W
NW
E
SE
NW
N
N
E
S
SE
E
SE
NE
N
N

281
--261
261
261
----

---------------140
--86
----192
171
---

79
--40
40
40
----

---------------24
--43
----28
28
---

N
--N
N
N
----

---------------SW
--S
----W
W
---

174
157
135
76
321
187
139
295
270

235
284
226
221
287
189
12
12
13
183
192
189
175
223
6
19
245
241
268
343
75
40
15
18
153
253
257

17
17
20
26
18
36
21
16
19

31
33
58
66
53
55
42
57
32
66
66
59
63
61
64
64
62
87
48
40
32
28
41
47
66
54
28

W
W
SW
S
NE
W
SW
N
N

NW
N
NW
NW
NE
W
E
E
E
W
W
W
W
NW
E
E
NW
N
N
E
S
SE
E
E
W
N
N

84
67
45
346
231
97
29
205
180

325
14
316
311
17
279
102
102
103
273
282
279
265
313
96
109
335
331
358
73
165
130
105
108
243
343
347

349

SF022
SF023
SF024
SF025
SF216
SF217
SF218
SF010
SF011
SF012
SF013
SF014
SF015
SF016
SF017
SF018
SF174
SF183
SF175
SF176
SF177
SF178
SF179
SF180
SF181
SF182
SF142
SF144
SF145
SF148
SF151
SF161
SF162
SF163
SF164
SF171
SF173
SF184

2
2
2
2
3
3
3
3
3
3
3
3
3
3
3
3
4
4
4
4
4
4
4
4
4
4
5
5
5
5
5
5
5
5
5
5
5
5

Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
Kopuai North
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach

R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R1
R2
R2
R2
R2
R2
R2
R2
R2
R2
R2
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3

---20
250
100
70
10
10
10
10
5
20
20
5
5
20
3
100
100
100
10
5
10
20
10
3
30
20
10
200
15
15
100
20
15
20
10

----Good
Good
Good
---------Fair
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Fair
Fair
Fair
Fair
Fair
Fair
Fair
Fair
Fair
Fair
Fair
Fair

229
143
155
108
286
295
292
325
214
326
310
190
272
280
133
287
154
107
288
285
216
144
120
326
346
124
12
201
192
83
205
74
258
58
22
34
270
214

11
21
28
19
31
22
28
27
33
21
20
43
34
25
51
19
9
12
15
18
11
28
13
17
14
19
30
19
30
10
22
23
17
58
33
49
77
36

NW
W
SW
S
N
NE
NE
NE
NW
NE
N
W
N
N
SW
N
SW
S
N
N
NW
SW
SW
NE
NE
SW
E
W
W
S
W
E
N
SE
E
SE
N
NW

---115
----------------------341
-----242
25
-26
70
194

---22
----------------------15
-----49
17
-24
56
44

---S
----------------------E
-----NW
E
-E
SE
W

229
143
155
331
286
295
292
325
214
326
310
190
272
280
133
287
154
107
288
285
216
144
120
326
346
124
35
201
192
83
205
74
53
66
22
40
96
323

11
21
28
4
31
22
28
27
33
21
20
43
34
25
51
19
9
12
15
18
11
28
13
17
14
19
19
19
30
10
22
23
33
44
33
32
51
15

NW
W
SW
NE
N
NE
NE
NE
NW
NE
N
W
N
N
SW
N
SW
S
N
N
NW
SW
SW
NE
NE
SW
SE
W
W
S
W
E
SE
SE
E
SE
S
NE

139
53
65
241
196
205
202
235
124
236
220
100
182
190
43
197
64
17
198
195
126
54
30
236
256
34
305
111
102
353
115
344
323
336
292
310
6
233

350

SF185
SF186
SF143
SF146
SF147
SF149
SF150
SF152
SF153
SF165
SF169
SF170
SF172
SF110
SF111
SF112
SF113
SF096
SF092
SF095
SF098
SF103
SF104
SF140
SF195
SF196
SF203
SF204
SF208
SF211
SF215
SF228
SF227
SF229
SF245
SF243
SF244
SF233

5
5
5
5
5
5
5
5
5
5
5
5
5
6
6
6
6
7
7
7
7
7
7
8
9
9
10
10
10
10
11
12
12
12
13
13
13
14

North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
North Beach
Waioroko North
Waioroko North
Waioroko North
Waioroko North
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Waioroko South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Ounutae South
Poporotaupo North
Poporotaupo North
Poporotaupo North
Poporotaupo Stream

R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R3
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R4
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5
R5

5
5
20
30
15
15
300
300
50
10
30
15
15
30
20
150
40
5
20
20
40
20
30
5
5
10
10
10
5
10
3
5
5
10
3
5
2
3

Fair
Fair
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Fair
Good
Good
Good
Good
Good
Good
Fair
Fair
Good
Good
Fair
Good
Good
Fair

202
184
283
191
84
343
192
210
190
24
0
30
9
260
19
355
248
333
20
16
0
184
1
359
289
143
139
108
28
128
33
185
236
119
84
65
322
152

23
57
10
29
12
26
21
27
35
25
35
30
48
20
15
16
12
31
15
24
23
77
62
47
69
38
35
11
66
24
10
28
52
46
22
18
15
64

W
W
N
W
S
E
W
W
W
E
E
SE
E
N
E
E
NW
NE
E
E
E
W
E
E
N
SW
SW
S
SE
SW
SE
W
NW
SW
S
SE
NE
W

----------20
20
26
--------350
350
86
220
141
-122
-98
75
198
198
127
159
--143

----------18
18
24
--------37
37
14
24
25
-20
-10
35
21
21
23
21
--37

----------E
E
E
--------E
E
S
NW
SW
-SW
-S
SE
W
W
SW
W
--SW

202
184
283
191
84
343
192
210
190
24
343
43
4
260
19
355
248
333
20
16
0
5
12
346
300
147
139
318
28
145
269
154
251
112
35
65
322
161

23
57
10
29
12
26
21
27
35
25
19
13
31
20
15
16
12
31
15
24
23
67
26
48
62
13
35
10
66
16
28
9
37
23
26
18
15
28

W
W
N
W
S
E
W
W
W
E
E
SE
E
N
E
E
NW
NE
E
E
E
E
E
E
NE
SW
SW
NE
SE
SW
N
SW
N
S
SE
SE
NE
W

112
94
193
101
354
253
102
120
100
294
253
313
274
170
289
265
158
243
290
286
270
275
282
76
210
57
49
228
298
55
179
64
161
22
305
335
232
71

351

SF235
SF241
SF234
SF078
SF080
SF079
SF083
SF076
SF231
SF232
SF075
SF052
SF055
SF000
SF003
SF121
SF122
SF123
SF131
SF132
SF133
SF134
SF139

14
14
14
15
15
15
16
16
16
16
16
17
17
17
17
18
18
18
18
18
18
18
18

Poporotaupo Stream
Poporotaupo Stream
Poporotaupo Stream
Waiongaro South
Waiongaro South
Waiongaro South
Ngatupaku South
Waiakapua North
Waiakapua North
Waiakapua North
Waiakapua North
Paritutu North
Paritutu North
Paritutu North
Paritutu North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North
Pitone North

R5
R5
R5
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R6
R7
R7
R7
R7
R7
R7
R7
R7

3
3
30
10
20
10
10
50
15
15
10
3
15
25
-1
1
10
5
3
3
3
20

Fair
Fair
Good
Fair
Fair
Good
Good
Good
Good
Good
-Fair
Fair
--Good
Good
Good
Good
Good
Good
Good
Good

58
75
20
264
306
296
91
152
245
96
241
20
44
265
294
99
97
103
272
120
236
263
41

Notes:
* Apparent displacement of faults is approximate
Bedding orientation used to restore fault orientation (if applicable, see text)
Fault orientation following back-rotation of bedding (if applicable, see text)
# Vergence direction is assumed to be up-dip and down-dip for reverse and normal faults, respectively

11
20
19
43
33
35
30
19
10
30
25
16
16
52
34
81
76
64
35
26
15
14
41

SE
SE
SE
N
NE
N
S
SW
N
S
NW
E
SE
N
NE
S
S
S
N
SW
NW
N
E

358
358
-----54
91
91
-----100
100
100
------

15
15
-----22
7
7
-----41
41
41
------

E
E
-----SE
S
S
-----S
S
S
------

132
115
20
264
306
296
91
195
256
97
241
20
44
265
294
98
95
107
272
120
236
263
41

13
22
19
43
33
35
30
31
17
23
25
16
16
52
34
40
35
23
35
26
15
14
41

SW
SW
SE
N
NE
N
S
W
N
S
NW
E
SE
N
NE
S
S
S
N
SW
NW
N
E

42
25
290
174
216
206
1
105
166
7
151
290
314
175
204
8
5
17
182
30
146
173
311

352

APPENDIX C-4: Unrestored versus restored slump fault orientations

353

APPENDIX C-5: Full-size cliff photomosaics (available electronically on Stanford


Universitys library website)

354

APPENDIX C-6: Wave-cut platform orthophotos (available electronically on


Stanford Universitys library website)

355

APPENDIX C-7: Biostratigraphy


Samples analyzed by Martin Crundwell, GNS Science
7-8 March, 2011
PS-01B: 03-07-11 (Road cutting on farm track)
Unwashed weight: 360.86 g
Washed weight (>75 um): 31.59 g
Coordinates (WGS84): -38.52042, 174.67878
Map Grid (coordinates): BF31 463351
Photo: P1020986
Locality Description: Paparahia Station farm track cut. Mohakatino (?) overlies a 2-4 meter thick
bioclastic unit (Mangarara Ss?), which overlies massive gray mudstone (Manganui Fm?). PS01B-030711 collected from the mudstone about 1.5 meters beneath the bioclastic unit.

Comment on sample: Washed residue consists almost entirely of forams.

Comment on fauna: Very rich, well preserved microfauna. 56% Planktics (extra-neritic). Lower
bathyal (1000-1500 m water depth) based on Cibicides robertsonianus.

Comment on age: Lower Lillburnian (14.62-15.1 Ma) based on the range zone of Orbulina
suturalis. The lower Sl age is consistent with the presence of Globoconella miotumida,
Globoconella cf. miozea, Menardella praemenardii, Zealandella conica? and Paragloborotalia cf.
partimlabiata.

Age and environmentally significant species:


Globoconella miotumida
Globoconella cf. miozea
Menardella praemenardii
Zealandella conica?
Paragloborotalia cf. partimlabiata
Globoquadrina dehiscens
Cibicides molestus
Cibicides neoperforatus
Karreriella cylindrica
Sigmoilopsis schlumbergeri
Vulvulina pennatula

356

Cibicides robertsonianus
__________________________________________________________________________

PS-02B: 03-07-11 (Road cutting on farm track)


Unwashed weight: 250.88 g
Washed weight (>75 um):
Coordinates (WGS84): -38.52042, 174.67878
Map Grid (coordinates): BF31 463351
Photo: P1020986
Locality Description: Paparahia Station farm track cut. Mohakatino (?) overlies a 2-4 meter thick
bioclastic unit (Mangarara Ss?), which overlies massive gray mudstone (Manganui Fm?). PS02B-030711 collected from the bioclastic unit about 2 meters above contact with Manganui Fm
(?) mudstone.

Comment on sample: Sandy, poor disaggregation.

Comment on fauna: Moderately rich, poorly preserved microfauna. ~35% Planktics (extraneritic). Deep lower bathyal (>1500 m water depth) based on Tritaxilina zealandica.

Comment on age: Lillburnian-upper Clifdenian

Age and environmentally significant species:


Globoconella miotumida
Orbulina sp. (or advanced Praeorbulina sp.?)
Globoquadrina dehiscens
Cibicides molestus
Cibicides neoperforatus
Karreriella cylindrica
Vulvulina pennatula
Tritaxilina zealandica
Amphistegina sp. (shelfal species transported downslope)
__________________________________________________________________________

PS-03B: 03-07-11 (Inside cave)


Unwashed weight: 221.10 g
Washed weight (>75 um): 202.53 g
Coordinates (WGS84): -38.518 174.676 (approx. entrance to cave)

357

Map Grid (coordinates): BF31 462353


Locality Description: Collected from mudstone (?) or sandy mudstone (?) at base of cave
entrance (See diagram for schematic location).

Comment on sample: Cemented. Very poor disaggregation.

Comment on fauna: Very sparse poorly preserved microfauna. Outermost shelf or deeper?
(>150 m water depth) based on Cibicides molestus.

Comment on age: non-determinate (nd)

Age and environmentally significant species:


Globoquadrina dehiscens?
Cibicides molestus
__________________________________________________________________________

PS-04B: 03-07-11 (Inside cave)


Unwashed weight: 204.43 g
Washed weight (>75 um): 188.68 g
Coordinates (WGS84): -38.518 174.676 (approx. entrance to cave)
Map Grid (coordinates): BF31 462353
Locality Description: Collected from bioclastic limestone above the conglomeratic limestone unit
(See diagram for schematic location).

Comment on sample: Limestone. Recrystallized.

Comment on fauna: Very poor disaggregation. Very poorly preserved microfauna. Forams,
heavily recrystallized and barely recognizable, except for porecellaneous forms. Bryozoan and
shell fragments. Shallow shelf environment?

Comment on age: Non-determinate (ND)

Age and environmentally significant species:


Pyrgo sp.
__________________________________________________________________________

PS-05B: 03-07-11 (Inside cave)


358

Unwashed weight: 290.80 g


Washed weight (>75 um): 185.21 g
Coordinates (WGS84): -38.518 174.676 (approx. entrance to cave)
Map Grid (coordinates): BF31 462353
Locality Description: Collected from sandy mudstone above the bioclastic unit sampled in PS04B-030711 (See diagram for schematic location).

Comment on sample: Weathered, calcareous.

Comment on fauna: Rare, very poorly preserved forams and poorly preserved shell (bivalve)
and bryozoan fragments.

Comment on age: Non-determinate (ND)

Age and environmentally significant species:


Cibicides sp.
Anomalinoides subnonionoides
Globocassidulina subglobosa
__________________________________________________________________________

PS-06B: 03-07-11 (Stream near cave)


Unwashed weight:
Washed weight (>75 um):
Coordinates (WGS84): -38.517 174.680 (approx.)
Map Grid (coordinates): BF31 465355 (approx.)
Locality Description: Collected from mudstone about 500 meters upstream from the cave
entrance (See diagram for schematic location).

Comment on sample: Poorly disaggregated.

Comment on fauna: Common, poorly preserved microfauna. Common bryozoan fragments, few
ostracods and echinoid spines. Neritic, shelfal fauna.

Comment on age: Poorly constrained, lower Tongaporutuan to Waitakian (lTt-Lw), Miocene?

Age and environmentally significant species:


Globoquadrina dehiscens

359

Paragloborotalia cf. mayeri s.l.? (rare)


Zeaglobigerina woodi
Cribrorotalia ornatissima?
Amphistegina sp.
__________________________________________________________________________

PRS-07B: 03-07-11 (Waterfall about 400 meters south of Paparahia Stream)


Unwashed weight: 256.67 g
Washed weight (>75 um): 65.22 g
Coordinates (WGS84): -38.5348 174.6351
Map Grid (coordinates): BF31 425336 (approx.)
Photo: P1020988
Locality Description: Collected under waterfall, about 1.5 meters above top of North Awakino
mass-transport deposit.

Comment on sample: Weathered, poorly disaggregated. Volcaniclastic.

Comment on fauna: Lower bathyal (600-1000 m water depth). Sub-oceanic (ca. 85% planktics).

Comment on age: Lower Tongaporutuan to Lillburnian (lTt-Sl), probably lower Tt (10.50-10.91


Ma, excluding 10.65-10.68 Ma) based on the presence of dextral Hirsutella panda, sinistral
Globoconella miotumida, and the absence of Neogloboquadrina and Paragloborotalia.

Age and environmentally significant species:


Globoquadrina dehiscens
Hirsutella panda (0S:7D)
Globoconella miotumida (64S:3D)
Orbulina sp.
Sigmoilopsis schlumbergeri
Eggerella bradyi
Hopkinsina mioindex
Cibicides robertsonianus
__________________________________________________________________________

PRS-08B: 03-07-11 (Moist region in cliff about 750 meters south of Paparahia
Stream)
360

Unwashed weight: 749.04 g


Washed weight (>75 um): 121.74 g
Coordinates (WGS84): -38.53824 174.63446
Map Grid (coordinates): BF31 425332 (approx.)
Photo: P1020989
Locality Description: Collected on cliff in moist area, within the North Awakino mass-transport
deposit, about 750 meters south of Paparahia Stream.

Comment on sample: Siltstone. Poor disaggregation. Traces of carbonaceous material and


pyrite.

Comment on fauna: Sparse microfauna dominated by Chilostomella (unusual stressed benthic


environment, anoxic or dysoxic infauna).

Comment on age: Non-determinate (ND)

Age and environmentally significant species:


Chilostomella ovoidea (dominant, most specimens crushed)
__________________________________________________________________________

TP-01B: 03-08-11 (northern end Middle Bay beach: Tirua Formation)


Unwashed weight: 604.05 g
Washed weight (>75 um):
Coordinates (WGS84): -38.405788 174.639847
Map Grid (coordinates): BF31 432479 (approx.)
Photo: P1030049
Locality Description: Collected mudstone sample below Tirua Formation, above 2 meters above
first thin (20 cm) sandstone bed encountered. Northern end of Ngararahae (Middle) Bay.

Comment on sample: Poor disaggregation.

Comment on fauna: Abundant, moderately preserved microfauna. Deep mid bathyal or deeper
(>800 m water depth). Sub-oceanic (ca. 70% planktics).

Comment on age: Lillburnian (Sl) based on the presence of Orbulina, an early population of
Globoconella miotumida including weakly forms, and Globoconella cf. miozea, and the absence
of Paragloborotalia mayeri s.l.

361

Age and environmentally significant species:


Globoconella miotumida (early population including weakly forms)
Globoconella cf. miozea
Orbulina sp.
Globoquadrina dehiscens
Sphaeroidinellopsis sp.
Vulvulina pennatula
__________________________________________________________________________

WW-01B: 03-08-11 (Mohakatino Fm in Waikawau Beach)


Unwashed weight: 273.76 g
Washed weight (>75 um): 248.73 g
Coordinates (WGS84): -38.47402 174.641085
Map Grid (coordinates): BF31 432403 (approx.)
Photo: P1030087
Locality Description: Collected sample above thick volcaniclastic sandstone bed about 200
meters north of Waikawau beach entrance.

Comment on sample: Siltstone. Weathered, non-calcareous. Poor disaggregation.

Comment on fauna: Non-fossilfereous (nf)


__________________________________________________________________________

WW-02B: 03-08-11 (Mohakatino Fm on Ngarupupu Point, south of Waikawau


Beach)
Unwashed weight: 640.16 g
Washed weight (>75 um): 2.36 g
Coordinates (WGS84): -38.476604 174.636229
Map Grid (coordinates): BF31 427400 (approx.)
Photo: P1030100
Locality Description: Collected sample above in Mohakatino at Ngarupupu Point, south end of
Waikawau Beach.

Comment on sample: Mostly pyrite, no calcareous material.

362

Comment on fauna: Very rare pyritized foram casts.

Comment on age: Non-determinate (ND)

Age and environmentally significant species:


Chilostomella ovoidea

363

APPENDIX C-8: Paleocurrent measurements


Paleocurrent
Measured section
type
ID
depth (m)
Opito Point (-38.51156S, -174.63175E)
Flute marks
18
26
Ripple cross-sets
18
31
Ripple cross-sets
18
31
Ripple cross-sets
18
31
Ripple cross-sets
18
31
Ripple cross-sets
18
31
Ripple cross-sets
18
36.5
Ripple cross-sets
18
36.5
Ripple cross-sets
18
36.5
Ripple cross-sets
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5
Ripple tops
18
36.5

Paleocurrent
Strike/
Dip/
Trend
Plunge

Dip
Dir

Bedding*
Strike Dip Dir

Flow
direction

-78
81
105
93
66
214
180
181
135
----------

125
31
15
32
27
24
13
7
17
20
12
21
28
13
18
33
35
36
33

-S
S
S
S
S
NW
W
W
SW
----------

-106
106
106
106
106
102
102
102
102
102
102
102
102
102
102
102
102
102

-13
13
13
13
13
9
9
9
9
9
9
9
9
9
9
9
9
9

-S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S
S

125
152
112
194
122
127
330
330
300
246
12
21
28
13
18
33
35
36
33

South of Paparahia stream (-38.53464S, -174.63503E)


Ripple x-laminations
1
6-11.5
60
Ripple x-laminations
1
6-11.5
100
Ripple x-laminations
1
6-11.5
81
Ripple x-laminations
1
6-11.5
90
Ripple x-laminations
1
6-11.5
76
Ripple x-laminations
1
6-11.5
77

16
13
14
15
14
9

SE
S
S
S
S
S

236
236
-----

6
6
-----

NW
NW
-----

149
177
171
180
166
167

South of Paparahia stream - within MTD (-38.53585S, 174.63474E)


Flute marks
2
21.5
38
Flute marks
2
21.5
20
Flute marks
2
21.5
350
Flute marks
2
21.5
58
Flute marks
2
21.5
33
Flute marks
2
21.5
33
Flute marks
2
21.5
54
Flute marks
2
21.5
65
Flute marks
2
21.5
28
Flute marks
2
21.5
65
Flute marks
2
21.5
75
Flute marks
2
21.5
54
Flute marks
2
21.5
55

--------------

--------------

271
271
271
271
271
271
271
271
271
271
271
271
271

16
16
16
16
16
16
16
16
16
16
16
16
16

N
N
N
N
N
N
N
N
N
N
N
N
N

37
20
350
57
32
32
53
64
27
64
74
53
54

South of Paparahia stream - within MTD (-38.53883S, 174.63389E)


Flute marks
2
21.5
-Flute marks
2
21.5
-Flute marks
2
21.5
-Flute marks
2
21.5
-Flute marks
2
21.5
-Flute marks
2
15.25
-Flute marks
2
15.25
--

49
38
37
31
29
19
14

--------

100
100
100
100
100
276
276

81
81
81
81
81
83
83

S
S
S
S
S
N
N

42
63
64
70
72
53
57

364

Flute marks
Flute marks
Flute marks
Flute marks
Flute marks

2
2
2
2
2

15.25
15.25
15.25
15.25
15.25

------

33
25
36
19
17

------

276
276
276
276
276

83
83
83
83
83

N
N
N
N
N

65
70
70
72
75

South of Kopuai stream - sea stack (-38.54819S, -174.63278E)


Flute marks
4
1.75
82
Flute marks
4
1.75
91
Flute marks
4
1.75
89
Flute marks
4
1.75
83
Flute marks
4
1.75
79
Flute marks
4
1.75
76
Flute marks
4
1.75
74
Flute marks
4
1.75
77
Flute marks
4
1.75
61
Flute marks
4
1.75
77
Flute marks
4
1.75
80
Flute marks
4
1.75
95
Flute marks
4
1.75
77
Flute marks
4
1.75
84

---------------

---------------

---------------

---------------

---------------

82
91
89
83
79
76
74
77
61
77
80
95
77
84

South of Kopuai stream (-38.55018S, -174.63308E)


Flute marks#
5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5
Flute marks#
5
5.5

------------------

------------------

10
10
10
10
10
10
10
10
10
10
10
10
10
10
10
10
10

3
3
3
3
3
3
3
3
3
3
3
3
3
3
3
3
3

E
E
E
E
E
E
E
E
E
E
E
E
E
E
E
E
E

119
117
94
137
137
101
104
106
134
142
133
153
127
64
108
127
139

---

---

106
106

29
29

N
N

310
316.5

NE
NE
NE
NE
NE
NE
NE

--------

--------

--------

37
73
37
59
44
39
61

NE

--

--

--

46 or 226

261
263
286
243
243
279
276
274
246
238
247
227
253
316
272
253
241

South of Ounutae stream - within MTD (-38.57682S, -174.63019E)


Flute marks**
--313
Flute marks**
--320

North of Waiakapua stream - above MTD (-38.61454S, -174.62620E)


Ripple x-laminations
13
8-9
307
26
Ripple x-laminations
13
8-9
343
16
Ripple x-laminations
13
8-9
307
17
Ripple x-laminations
13
8-9
329
23
Ripple x-laminations
13
8-9
314
20
Ripple x-laminations
13
8-9
309
19
Ripple x-laminations
13
8-9
331
28
Channel margin
orientation
13
8
136
14
Pitone stream north bank (-38.63575S, 174.62192E)

365

Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple x-laminations
Ripple crests
Ripple crests
Ripple crests
Ripple crests
Ripple crests
Ripple crests
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Cross-bedding
Ripple x-lamination
Ripple x-lamination

17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17
17

6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
6.15
??
??
??
??
??
??
12.0-12.5
12.0-12.5
12.0-12.5
14.4-15.0
14.4-15.0
14.4-15.0
14.4-15.0
14.4-15.0
14.4-15.0
14.4-15.0
15.75-16.0
15.75-16.0
15.75-16.0
15.75-16.0
15.75-16.0
17.0-17.25
17.0-17.25
17.0-17.25
17.0-17.25
17.0-17.25
17.5-17.75
17.5-17.75
17.5-17.75
17.5-17.75
18.75-19.25
18.75-19.25
18.75-19.25
18.75-19.25
18.75-19.25
18.75-19.25
18.75-19.25
20.25-20.75
20.25-20.75
20.25-20.75
27-?
27-?

125
125
175
170
190
193
93
129
199
150
130
139
140
189
195
126
122
125
124
112
105
185
185
220
205
216
234
220
220
230
235
234
215
230
233
233
225
225
224
219
245
210
208
225
209
216
228
203
185
221
210
185
245
245
230
81
90

25
25
31
31
29
21
25
25
27
26
25
22
26
21
29
------19
30
40
29
27
26
27
27
16
16
19
30
20
25
20
19
17
25
21
18
22
34
21
31
24
20
20
19
20
21
17
21
30
34
14
15

366

SW
SW
W
W
W
W
S
SW
W
SW
SW
SW
SW
W
W
------W
W
NW
W
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
W
NW
NW
W
NW
NW
N
S
S

236
236
236
236
236
236
236
236
236
236
236
236
236
236
236
------242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
242
---

6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
------15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
15
---

NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
------NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
NW
---

204
204
255
250
271
270
177
208
280
228
209
215
218
265
277
216
212
215
214
202
195
227
247
299
269
281
314
288
288
253
267
298
284
290
311
299
272
257
292
271
349
261
279
282
277
275
285
246
227
270
257
221
342
338
312
171
180

Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination
Ripple x-lamination

17
17
17
17
17
17
17
17
17
17
17
17

27-?
27-?
27-?
27-?
27-?
27-?
27-?
27-?
27-?
27-?
27-?
27-?

20
76
27
77
7
180
204
195
200
181
184
204

16
14
18
9
20
30
32
24
23
19
19
22

SE
S
SE
S
E
W
NW
W
NW
W
W
NW

-----204
204
204
204
204
204
204

-----8
8
8
8
8
8
8

-----NW
NW
NW
NW
NW
NW
NW

* Paleocurrent measurements rotated to bedding horizontal


Paleocurrents collected ~400 meters north of section 2
Measured rake of flute
# Flutes measured on an overturned bedding plane. Results should be considered approximate.
** Flutes measured on two nearby bedding planes. Numerous flutes in each case with similar orientations.
General flow direction collected from a bedding base with numerous flute marks
Dip/plunge is given in the direction of ripple movement, as observed on the top of a bedding surface

367

110
166
117
167
97
262.6
294.2
281
288.2
256.7
261.3
294.3

You might also like