<<

UCLA UCLA Electronic Theses and Dissertations

Title Tectonic Reconstruction of the Southern System Using Segments of the Chocolate Mountains Anticlinorium in the , Southern , U.S.A.

Permalink https://escholarship.org/uc/item/6sb262md

Author Coffey, Kevin Thomas

Publication Date 2019

Supplemental Material https://escholarship.org/uc/item/6sb262md#supplemental

Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital Library University of California

UNIVERSITY OF CALIFORNIA

Los Angeles

Tectonic Reconstruction of the Southern San

Andreas Fault System Using Segments of the

Chocolate Mountains Anticlinorium in the San

Gabriel Mountains, , U.S.A.

A dissertation submitted in partial satisfaction of the

requirements for the degree Doctor of Philosophy

in Geology

by

Kevin Coffey

2019

© Copyright by

Kevin Coffey

2019 ABSTRACT OF THE DISSERTATION

Tectonic Reconstruction of the Southern San

Andreas Fault System Using Segments of the

Chocolate Mountains Anticlinorium in the San

Gabriel Mountains, Southern California, U.S.A.

by

Kevin Coffey

Doctor of Philosophy in Geology

University of California, , 2019

Professor Raymond V. Ingersoll, Chair

The San Gabriel Mountains of southern California contain anticlinoria of schist thought to correlate with one another, and with the Chocolate Mountains anticlinorium to the east. These correlations provide cumulative-slip estimates for the Punchbowl and southern San Andreas faults, respectively; however, the validity of these correlations, as well as the original orientations of these anticlinoria, is debated. This dissertation presents four related studies, each addressing a different aspect of this problem. Chapter 1 presents sedimentologic and structural data verifying that the two anticlinoria of the San Gabriel Mountains were continuous prior to

40-50 km of dextral slip on the Punchbowl fault, and casting doubt on the 80-110 km of dextral

ii slip on the San Francisquito-Fenner-Clemens Well fault suggested by some prior studies.

Chapter 2 presents structural data from the schist of the larger anticlinorium of the San Gabriel

Mountains, which supports its correlation with the Chocolate Mountains anticlinorium, and thus estimates of approximately 240 km of dextral slip on the southern San Andreas fault. New data are combined with previously published structural and paleomagnetic data, which have been interpreted as implying different architectures for the anticlinorium, suggesting a model reconciling the datasets, and implying an originally linear, approximately northeast/southwest trend of the anticlinorium. Chapter 3 proposes a strain-partitioning model for the anomalously oriented Mojave segment of the southern San Andreas fault, in which slip south of this segment is partitioned within this segment into a strike-slip component along the San Andreas fault itself, and a shortening component beside it. This model, together with a previous study’s insight, brings otherwise contradictory cumulative-slip estimates of approximately 240 km and 160 km for the southern San Andreas fault into agreement. Chapter 4 presents new fault-kinematic data, and sequential reconstructions of the San Gabriel block as maps and cross sections, suggesting that the anticlinorium is in the hanging wall of a detachment fault, rather than the exhumed footwall as previously suggested. New geochronologic and geochemical data support correlation of the anticlinoria and constrain the onset of extension to ca. 26-25 Ma.

iii

The dissertation of Kevin Coffey is approved.

An Yin

Kevin D. McKeegan

Gilles F. Peltzer

Raymond V. Ingersoll, Committee Chair

University of California, Los Angeles

2019

iv

DEDICATION

I dedicate this dissertation to my parents, Paula Sullivan Coffey and Lewis Coffey.

Throughout my life, they encouraged and supported me in my pursuit of anything that I found interesting. Geology was one of those things, and, although it was not itself important to either of them, once I began to develop a passion for it, they made sure I had every opportunity to explore it. When administrators at the local community college tried to deny me entry to a geology course because of my young age, my mother researched and wrote a counterargument grounded in education law, and my father took the class with me. This was only the first of countless examples, big and small, from then through the present. I would never have achieved this without them. Thank you, mom and dad.

v

TABLE OF CONTENTS: Page:

ABSTRACT ii

APPROVAL PAGE iv

DEDICATION v

LIST OF FIGURES x

LIST OF TABLES xiv

LIST OF APPENDICES xv

LIST OF SUPPLEMENTARY MATERIALS xvi

ACKNOWLEDGMENTS xvii

VITA xix

FOREWORD 1

REFERENCES CITED 3

CHAPTER 1: STRATIGRAPHY, PROVENANCE, AND TECTONIC 6

SIGNIFICANCE OF THE PUNCHBOWL BLOCK, SAN GABRIEL

MOUNTAINS, CALIFORNIA, U.S.A.

ABSTRACT 6

INTRODUCTION 8

GEOLOGIC BACKGROUND 8

Oligocene-Miocene Basins 8

Potential Source Rocks 13

Tectonic Reconstructions of the Southern San Andreas Fault System 14

GEOLOGY OF THE PUNCHBOWL BLOCK 15

vi

Stratigraphy 15

Sediment Composition 20

Structure 30

DISCUSSION 34

Basin Development 34

Regional Stratigraphic Correlations 36

Regional Tectonic and Paleogeographic Reconstructions 41

CONCLUSIONS 47

REFERENCES CITED 51

CHAPTER 2: RECONCILING STRUCTURAL AND PALEO MAGNETIC 73

DATA TO CONSTRAIN EVOLUTION OF THE CHOCOLATE MOUNTAINS

ANTICLINORIUM ALONG THE SAN ANDREAS FAULT SYSTEM,

SOUTHERN CALIFORNIA, U.S.A.

ABSTRACT 73

INTRODUCTION 74

GEOLOGIC BACKGROUND 76

The Chocolate Mountains Anticlinorium 76

Penetrative Lineations 77

Paleomagnetic Declinations 78

METHODS 79

RESULTS 81

Mount Pinos 81

Sierra Pelona 83

vii

Crafton Hills 86

DISCUSSION 89

Decoupling-Surface Model 90

Origin of Penetrative Lineations Within the Schist 92

NNW-Lineations Model 93

Comparison of the Two Models 94

Formation of NNW/SSE-Oriented Penetrative Lineations 104

Possible Relationship to Nacimiento Fault 106

Origin of Neogene-Quaternary Clockwise Vertical-Axis Rotations 107

Implications of the NNW-Lineations Model 109

CONCLUSIONS 110

REFERENCES CITED 113

CHAPTER 3: STRAIN PARTITIONING EXPLAINS DISPARATE CROSS- 134

FAULT CORRELATIONS ALONG THE SAN ANDREAS FAULT, SOUTHERN

CALIFORNIA, U.S.A.

ABSTRACT 134

INTRODUCTION 135

STRAIN-PARTITIONING MODEL 138

CONCLUSIONS 144

REFERENCES CITED 146

CHAPTER 4: CONSTRAINING THE GEOMETRY AND TIMING OF 155

NORMAL FAULTING AND ASSOCIATED VOLCANISM IN THE SAN

GABRIEL MOUNTAINS AND ITS CROSS-FAULT CORRELATIVES,

viii

SOUTHERN CALIFORNIA, U.S.A.

ABSTRACT 155

INTRODUCTION 156

SIERRA PELONA AND ASSOCIATED NORMAL-FAULT SYSTEM 158

Kinematics of the Pelona Fault 160

SEQUENTIAL RECONSTRUCTION OF THE SAN GABRIEL BLOCK 164

Ca. 5 Ma Reconstruction 164

Ca. 14 Ma Reconstruction 165

Ca. 18 Ma Reconstruction 165

Ca. 26 Ma Reconstruction 171

TIMING OF NORMAL FAULTING AND ANTICLINORIUM 172

DEVELOPMENT

K-Ar Dating of Volcanic Flows (Previous Work) 172

Zircon U-Pb Dating of Volcanic Flows and Domes 173

Geochemistry of Volcanic Flows and Domes 176

CONCLUSIONS 182

REFERENCES CITED 183

AFTERWORD 196

REFERENCES CITED 198

APPENDICES 200

ix

LIST OF FIGURES Page:

Chapter 1

Figure 1-1: Regional map showing Punchbowl block, and Tejon, Soledad, and Orocopia 7

regions.

Figure 1-2. Time-stratigraphic chart showing approximate ages of deposition of 9

Oligocene-Miocene strata of Punchbowl block, and of Tejon, Soledad, and

Orocopia regions.

Figure 1-3. Geologic map of central Punchbowl block. 17

Figure 1-4: Cross sections along section lines of Figure 1-3. 18

Figure 1-5: Schematic composite stratigraphic sections of Oligocene-Miocene strata (A) 21

south and (B) north of Blue Ridge in Punchbowl block.

Figure 1-6: Sandstone composition of Oligocene-Miocene strata of Punchbowl block. 24

Figure 1-7: Relative-probability distributions of detrital-zircon U-Pb ages from 27

Oligocene-Miocene strata of Punchbowl block (A) south and (B) north of Blue

Ridge.

Figure 1-8: Schematic cross section showing hypothesized positive flower structure 33

along Punchbowl fault: (A) in its initial geometry, and (B) in its present,

deformed geometry.

Figure 1-9: Comparison of sandstone composition of Vasquez Formation of Punchbowl 37

block and Vasquez Formation of Soledad region, which are likely correlative.

Figure 1-10: Schematic composite stratigraphic sections for Oligocene-Miocene strata of 38

Punchbowl block, and Tejon, Soledad, and Orocopia regions.

Figure 1-11: Comparison of sandstone composition of Paradise Springs formation of 40

x

Punchbowl block and Tick Canyon strata of Soledad region, which are likely

correlative.

Figure 1-12: Schematic representation of interpreted paleogeography, source rocks, and 46

depositional systems of Tejon-Soledad-Punchbowl-Orocopia regions.

Chapter 2

Figure 2-1: Regional map and schematic restorations, showing present and restored 74

locations and schematic extents of exposures of Pelona, Orocopia, Rand, and

related schists, and relevant faults.

Figure 2-2: Field photos showing lineations on foliation surfaces exposed at Sierra 79

Pelona.

Figure 2-3: Map of area, showing measurements of A: foliation within 81

schist and mylonitic gneiss, and B: tilt-corrected penetrative and crenulation

lineations within schist, and penetrative lineations within mylonitic gneiss.

Figure 2-4: Foliation and lineation data from Mount Pinos area 82

Figure 2-5: Map of Sierra Pelona, showing measurements of A: foliation within schist 83

and mylonitic gneiss, and B: tilt-corrected penetrative and crenulation lineations

within schist, and penetrative lineations within mylonitic gneiss.

Figure 2-6: Foliation and lineation data from schist of Sierra Pelona. 84

Figure 2-7: Map of northeastern Crafton Hills, showing measurements of A: foliation 86

within schist and mylonitic gneiss, and B: tilt-corrected penetrative and

crenulation lineations within schist, and penetrative lineations within mylonitic

gneiss.

Figure 2-8: Foliation and lineation data from northeastern Crafton Hills. 87

xi

Figure 2-9: Paleotectonic reconstruction of exposures of Pelona, Orocopia, Rand, and 97

related schists, including those along the Chocolate Mountains anticlinorium, ca.

25 Ma, according to the decoupling-surface model.

Figure 2-10: Paleotectonic reconstruction of exposures of Pelona, Orocopia, Rand, and 98

related schists, including those along the Chocolate Mountains anticlinorium, ca.

25 Ma, according to the NNW-lineations model.

Figure 2-11: Schematic block diagram illustrating hypothesized geometry and kinematics 107

of the Nacimiento-Vincent fault, ca. 70 Ma, along which top-to-NNW shear may

have occurred in the NNW-lineations model.

Chapter 3

Figure 3-1: Map of southern California, showing Mojave and segments 134

of the San Andreas fault, and related strike-slip faults, as well as approximate

extent of zone of north-south shortening, and three cross-fault correlations

discussed in text.

Figure 3-2: Diagrams showing vector-sum relationship of differential motion within the 137

restraining double bend of San Andreas fault.

Figure 3-3: Diagrams showing vector-sum relationships of differential motion within the 142

restraining double bend of San Andreas fault, as in Fig. 3-2C, except with

addition of slip on related to eastward extrusion of Mojave block

added as an additional component.

Chapter 4 155

Figure 4-1: Regional map, showing southern San Andreas fault system, insets of three

correlated regions separated by it, and extent of Fig. 4-5A.

xii

Figure 4-2: Schematic cross section across Sierra Pelona, illustrating hypothesis that 158

Pelona and San Francisquito faults are an exhumed detachment fault exposed on

either limb of the Sierra Pelona anticlinorium.

Figure 4-3: Field photos from Vasquez Formation within damage zone of Pelona fault. 159

Figure 4-4: Kinematic data measured along Pelona fault and their interpretation. 160

Figure 4-5: Sequential palinspastic reconstructions of the San Gabriel block, as paired 164

maps and cross -sections, from present (A) to ca. 26 Ma (E).

Figure 4-6: Palinspastic reconstruction of Sierra Pelona, Blue Ridge, and Orocopia 174

Mountains anticlinoria and surrounding areas, ca. 18 Ma, showing ages of

volcanic units determined by U-Pb dating of zircon.

Figure 4-7: Classification of volcanic samples according to Le Bas et al. (1986) 177

classification.

Figure 4-8: Plots of A: Th, B: U, and C: Hf vs. Ta for volcanic samples. 178

Figure 4-9: Chondrite-normalized A: rare-earth-element and B: trace-element 179

abundances for volcanic samples.

xiii

LIST OF TABLES Page:

Chapter 1

Table 1-1: Definition of original and recalculated sandstone point-count grain categories. 23

Chapter 2

Table 2-1: Summary of paleomagnetic data from rocks in the vicinity of schist exposures 95

Table 2-2: Summary of structural data from exposures of Pelona, Orocopia, Rand, and 96

related schists

xiv

LIST OF APPENDICES Page:

Chapter 2

Appendix A: Chapter 2 – Basis of representative orientations of penetrative lineations 198

used in Figure 2-9 and Table 2-2

Chapter 3

Appendix B: Chapter 3 – Relationship between rate of total differential motion, dextral 200

slip rate, and north-south shortening rate in the restraining double bend of the San

Andreas fault

Appendix C: Chapter 3 – Geometric model for evolution of the restraining double bend 204

of the San Andreas fault

Appendix D: Chapter 3 – Calculation of components of dextral slip and north-south 210

shortening in a restraining double bend of time-varying geometry

Appendix E: Chapter 3 – Calculation of components of dextral slip and north-south 212

shortening incorporating eastward extrusion of the Mojave block along the

Garlock fault

Chapter 4

Appendix F: Chapter 4 – Details of palinspastic reconstructions (Fig. 4-5) 218

Appendix G: Chapter 4 – Inverse-concordia and weighted-mean age results and plots 247

xv

LIST OF SUPPLEMENTARY MATERIALS (AVAILABLE ONLINE)

Chapter 1

Figure S1-1: Full-size version of Figure 1-3

Table S1-1: Igneous-zircon data and methods

Table S1-2: Conglomerate-clast-count data and lithologic descriptions

Table S1-3: Raw and recalculated sandstone point-count data

Table S1-4: Detrital-zircon data

Chapter 2

Table S2-1: All Mount Pinos raw measurements

Table S2-2: All Sierra Pelona raw measurements

Table S2-3: All Crafton Hills raw measurements

Chapter 4

Figure S4-1 (parts A-E): Full-size versions of Figure 4-5 (parts A-E).

Table S4-1: Volcanic-zircon U-Pb data

Table S4-2: X-ray fluorescence volcanic geochemistry data

Table S4-3: Inductively coupled plasma mass spectrometry volcanic geochemistry data

xvi

ACKNOWLEDGMENTS

I thank my advisor, Raymond V. Ingersoll, for advice, support, mentoring, discussion, insights, and edits throughout my time at UCLA, and sandstone point-count data contributed to

Chapter 1. I thank An Yin for in-the-field mentoring on kinematic measurements, and advice and feedback throughout my research. I thank my other committee members, Kevin McKeegan and

Gilles Peltzer, and Axel K. Schmitt (now at Heidelberg University) of my master’s committee, for advice, discussion, and edits. I thank Axel K. Schmitt, Jeremy Hourigan of UCSC, and Mark

Pecha, Nicky Giesler, and Chelsi White of the UA LaserChron Center for assistance generating and interpreting zircon age data. I thank Clinton Colasanti, Johanna Hoyt, and Dallon Stang of

UCLA (the latter two now at Aera Energy LLC) for assistance with sampling and field reconnaissance for Chapter 1. I thank Juliet Ryan-Davis (now at USGS), Jonathan Harris, and

Jade Star Lackey of Pomona College for assistance with mineral separation and use of their water shaking table, and Winnie Wu (now at CSULA), Matthew Wielicki (now at University of

Alabama), and Melanie Barboni (now at ASU) of UCLA for help with sample preparation and mineral separation. I thank Gordon Haxel and L. Sue Beard of USGS, and Carl Jacobson of Iowa

State University and West Chester University for discussion and productive debate relating to

Chapter 2. I thank Bryan Murray of Cal Poly Pomona and Kim Bishop of CSULA for helpful reviews of a previous version of Chapter 1, G. Peter Bird and Nathan Brown (now at UCB) of

UCLA and Jonathan Matti of USGS for helpful feedback on Chapter 3, Abijah Simon of UCLA for assistance drafting Figure 2-11, Peter Haproff (now at UNCW), Randon Flores, and Drew

Gomberg of UCLA for discussion of map relationships for Chapter 1, Kim Bishop for discussion of Chapter 4, and Carl Jacobson, Ryan Crow of the USGS, and Jason Ricketts of UT El Paso for general discussion. I thank Pine Mountain Club and Paradise Springs for property access.

xvii

Fieldwork, U-Pb dating of zircon, microscope thin sections, and geochemical analysis were paid for by two Graduate Student Research Grants from the Geological Society of America and by UCLA Academic Senate research grants awarded to Raymond V. Ingersoll. I was supported financially during the last year of my doctoral program by a Dissertation Year

Fellowship from UCLA. The ion-microprobe facility at UCLA is partly supported by a grant from the Instrumentation and Facilities Program, Division of Earth Sciences, National Science

Foundation. The Arizona LaserChron Center is partly supported by National Science Foundation grant NSF-EAR 1032156. Jeremy Hourigan’s laboratory at UCSC is partly supported by

National Science Foundation grant NSF OCE 0821303. Purchase of the XRF facility at

Washington State University was supported by the Murdoch Foundation and the National

Science Foundation.

xviii

VITA

Education:

 M.S. in Geology, University of California, Los Angeles (UCLA): 2015

 B.S. in Geology and Biology (double major), UCLA (Summa Cum Laude): 2012

 Los Angeles Harbor College: 2002-2008

 Reedley College: 2001

Employment:

 Teaching Assistant/Associate/Fellow, UCLA Department of Earth, Planetary, and

Space Sciences: 2013-2015 & 2017-2018

 Instructor, El Camino College Department of Earth Science: 2015-2017

 Geology Intern, Occidental Petroleum Corporation (now CRC): 2013 Summer

 Tutor (math and science), self-employed: 2007-2009

 Tutor (math and science), Los Angeles Harbor College: 2006-2008

Publications:

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019, Stratigraphy, provenance and tectonic

significance of the Punchbowl block, San Gabriel Mountains, California, U.S.A.:

Geosphere, v. 15, p. 479-501, https://doi.org/10.1130/GES02025.1

Hoyt, J.F., Coffey, K.T., Ingersoll, R.V., and Jacobson, C.E., 2018, Paleogeographic and

paleotectonic setting of the middle Miocene Mint Canyon and Caliente formations,

southern California: An integrated provenance study, in Ingersoll, R.V., Lawton, T.F, and

Graham, S.A., eds., Tectonics, sedimentary basins, and provenance: A celebration of the

career of William R. Dickinson: Geological Society of America Special Paper 540, p.

463-480, https://doi.org/10.1130/2018.2540(21) xix

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Coffey, K.T., Schmitt, A.K., Ford, A., Spera, F.J., Christensen, C., and Garrison, J., 2014,

Volcanic ash provenance from zircon dust with an application to Maya pottery: Geology,

v. 42, p. 595-598, https://doi.org/10.1130/G35376.1

Ingersoll, R.V., Pratt, M.J., Davis, P.M., Caracciolo, L., Day, P.P., Hayne, P.O., Petrizzo, D.A.,

Gingrich, D.A., Cavazza, W., Critelli, S., Diamond, D.S., Coffey, K.T., Stang, D.M.,

Hoyt, J.F., Reith, R.C., and Hendrix, E.D., 2014, Paleotectonics of a complex Miocene

half graben formed above a detachment fault: the Diligencia basin, Orocopia Mountains,

southern California: Lithosphere, v. 6, p. 157-176, https://doi.org/10.1130/L334.1

Selected Presentations:

Coffey, K.T., 2018, Can structural and paleomagnetic data be reconciled to determine the

original geometry of the Chocolate Mountains anticlinorium of southern California and

southwestern Arizona?: GSA Abstracts with Programs, v. 50, no. 5,

https://doi.org/10.1130/abs/2018RM-313789

Coffey, K.T., Ingersoll, R.V., and Hoyt, J.F., 2016, Implications of the Vincent Gap region of

the eastern San Gabriel Mountains for the Neogene history of southern California: GSA

Abstracts with Programs, v. 48, no. 4, https://doi.org/10.1130/abs/2016CD-274120

Coffey, K.T., and Ingersoll, R.V., 2014, Offset equivalents along the San Andreas fault system

were originally linked by a fault-bounded sliver of Miocene strata in the Vincent Gap

region, San Gabriel Mountains, southern California USA: GSA Abstracts with Programs,

v. 46, no. 6, p. 156.

xx

FOREWORD

The dextral San Andreas fault system in southern California is a complex arrangement of faults, active and inactive, with distinct slip histories (e.g., Wallace, 1990). Understanding the

San Andreas fault system is important, both because it is a well studied and often cited example of a transform plate boundary, and because it poses a serious seismic hazard to the tens of millions of people inhabiting southern California (e.g., Wallace, 1990). Despite several decades of research, a question as fundamental as the distribution of cumulative slip among the system’s individual faults is still debated.

A similar suite of distinctive lithologies and structures, including anticlinoria of schist, is present in three locations in southern California, on opposite sides of the San Gabriel/Canton and

San Andreas faults (e.g., Crowell, 1975). Correlation of these suites, including correlation of the schist anticlinoria with the Chocolate Mountains anticlinorium, suggests approximately 240 km of cumulative dextral slip on the southern San Andreas fault, and 60-70 km of cumulative dextral slip on the San Gabriel/Canton fault (Hill and Dibblee, 1953; Crowell, 1962, 1975). The sum of these estimates is approximately 300-310 km, close to the 315 km of dextral slip north of where the San Andreas and San Gabriel/Canton faults merge, implied by correlation of the Pinnacles and Neenach volcanic fields (Matthews, 1976). However, correlation of other units has been used to suggest cumulative dextral slip of only 160-185 km on the southern San Andreas fault, and 40-45 km on the San Gabriel/Canton fault (Powell, 1981, 1993; Frizzell et al., 1986; Matti and Morton, 1993; Weldon et al., 1993). Adding these estimates leaves an approximately 80-110 km deficit relative to the Pinnacles/Neenach correlation, which it has been proposed occurred along an older strand of the San Andreas, now dismembered into the San Francisquito, Fenner, and Clemens Well faults (Smith, 1977; Powell, 1981, 1993). Uncertainty in the cumulative slip 1 implied by cross-fault correlations is amplified by uncertainty in the extent to which crustal blocks along the San Andreas fault system have undergone vertical-axis rotations (e.g., Terres and Luyendyk, 1985; Bennett et al., 2016; Ingersoll and Coffey, 2017; Haxel et al., 2018).

The geology of the San Gabriel Mountains is central to the above debates. It includes the schist anticlinorium at Sierra Pelona, and surrounding lithologies and structures, proposed as equivalent to the Chocolate Mountains anticlinorium to the east, across the San Andreas fault, and schist exposed to the west, across the San Gabriel/Canton fault (Hill and Dibblee, 1953;

Crowell, 1962, 1975). It also contains the megaporphyritic monzogranite whose correlation with a matching body to the east, across the San Andreas fault, implies only 160 km of cumulative dextral slip (Frizzell et al., 1986; Matti and Morton, 1993). The San Francisquito and Fenner faults, components of the hypothesized proto-San Andreas fault with 80-110 km of cumulative dextral slip (Powell, 1981, 1993), are also located in the San Gabriel Mountains, as are structures and paleomagnetic sampling sites which have been used to argue for different amounts of

Neogene-Quaternary vertical-axis rotation (Harvill, 1969; Terres and Luyendyk, 1985; Haxel et al., 2018). My research, presented as four distinct but related studies, examines these critical components of the geology of the San Gabriel Mountains, including the schist anticlinorium at

Sierra Pelona, and their implications for the evolution of the Chocolate Mountains anticlinorium and the southern San Andreas fault system.

2

REFERENCES CITED

Bennett, S.E.K., Darin, M.H., Dorsey, R.J., Skinner, L.A., Umhoefer, P.J., and Oskin, M.E.,

2016, Animated tectonic reconstruction of the Lower Colorado River region: implications

for late Miocene to present deformation: in Reynolds, R.E., ed., Going LOCO:

investigations along the Lower Colorado River: 2016 Desert Symposium, Desert Studies

Center, California, p. 73-86.

Crowell, J.C., 1962, Displacement along the San Andreas fault, California: Geological Society of

America Special Paper 71, 61 p., https://doi.org/10.1130/SPE71

Crowell, J.C., 1975, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Frizzell, V.A., Jr., Mattison, J.M., and Matti, J.C., 1986, Distinctive Triassic megaporphyritic

monzogranite: Evidence for only 160 km offset along the San Andreas fault, southern

California: Journal of Geophysical Research, v. 91, p. 14,080-14,088,

https://doi.org/10.1029/JB091iB14p14080

Harvill, L.L., 1969, Deformational history of the Pelona Schist, northwestern Los Angeles

County, California [Ph.D. thesis]: Los Angeles, University of California, 117 p.

Haxel, G.B., Beard, L.S., and Jacobson, C.E., 2018, Significance of northeast-southwest

orientation of prograde lineation in the Pelona-Orocopia-Rand-Schist low-angle

subduction complex, southern California and southwest Arizona: Geological Society of

America Abstracts with Programs. v. 50, no. 5, https://doi.org/10.1130/abs/2018RM-

313363

3

Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults, California: a

study of the character, history, and tectonic significance of their displacements:

Geological Society of America Bulletin v. 64, p. 443-458, https://doi.org/10.1130/0016-

7606(1953)64[443:SAGABP]2.0.CO;2

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Matthews, V., III, 1976, Correlation of the Pinnacles and Neenach volcanic formations and their

bearing on San Andreas fault problem: American Association of Petroleum Geologists

Bulletin, v. 60, p. 2,128-2,141.

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

Powell, R.E., 1981, Geology of the crystalline basement complex, eastern ,

southern California: constraints on regional tectonic interpretation [Ph.D. dissertation]:

Pasadena, California Institute of Technology, 441 p.

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

4

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

Smith, D.P., 1977, San Juan-St Francis fault – Hypothesized major middle tertiary right-lateral

fault in central and southern California: California Division of Mines and Geology

Special Report 129, p. 41-50.

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Wallace, R.E., 1990, 1. General features, in Wallace, R.E., ed., The San Andreas fault system,

California: U.S. Geological Survey Professional Paper 1515, p. 3-12.

Weldon, R.J., II, Meisling, K.E., and Alexander, J., 1993, A speculative history of the San

Andreas fault in the central Transverse Ranges, California, in Powell, R.E., Weldon, R.J.,

II, and Matti, J.C., eds., The San Andreas fault system: displacement, palinspastic

reconstruction, and geologic evolution: Geological Society of America Memoir 178, p.

161-198, https://doi.org/10.1130/MEM178-p161

5

CHAPTER 1: STRATIGRAPHY, PROVENANCE, AND

TECTONIC SIGNIFICANCE OF THE PUNCHBOWL

BLOCK, SAN GABRIEL MOUNTAINS, CALIFORNIA,

U.S.A.

This chapter has been published as a journal article:

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019, Stratigraphy, provenance and tectonic significance of the Punchbowl block, San Gabriel Mountains, California, USA: Geosphere, v. 15, p. 479-501, https://doi.org/10.1130/GES02025.1

ABSTRACT

The Punchbowl block is a fault-bounded crustal sliver in the eastern San Gabriel

Mountains of southern California with important implications for conflicting reconstructions of the San Andreas fault system. Detailed mapping, determination of conglomerate-clast and sandstone compositions, and dating of detrital and igneous zircon of Oligocene-Miocene strata define two distinct subbasins and document initiation of extension and volcanism ca. 25-24 Ma, followed by local exhumation of Pelona Schist, and transition from alluvial-fan to braided-fluvial deposition. Strata of the Punchbowl block correlate with those of other regions in southern

California, confirming 40-50 km of dextral slip on the Punchbowl fault, and supporting reconstructions with 60-70 km of dextral slip on the San Gabriel/Canton fault and approximately

240 km of dextral slip on the southern San Andreas fault. Provenance and probable correlations of Punchbowl-block strata argue against 80-110 km of dextral slip on the San Francisquito-

6

Fenner-Clemens Well fault, and limit the time interval in which such slip could have occurred.

Synthesis of these findings with previous work produces paleogeographic reconstructions of the

Punchbowl block and its probable correlatives through time.

Figure 1-1: Regional map showing Punchbowl block (including location of Figure 1-3), and Tejon,

Soledad and Orocopia regions. Schematically shown in gray is the areal extent of upper Oligocene- lower Miocene strata in these regions, which delineate Simmler, Plush Ranch and Diligencia basins and Charlie Canyon, Texas Canyon and subbasins of Soledad basin. Relevant faults

also shown. Sierra Pelona, site of paleodrainage divide and anticlinorium of Pelona Schist, also

shown. f. = fault; BP/LVf = Big Pine/Lockwood Valley fault; CW/OMf = Clemens Well/Orocopia

Mountains fault; Pf = Punchbowl fault; SFf = San Francisquito fault; SGP = San Gorgonio Pass;

WTR = western Transverse Ranges. Figure after Frizzell and Weigand (1993) and Law et al.

(2001).

7

INTRODUCTION

The Punchbowl block, a crustal sliver between the Punchbowl and San Andreas faults

(e.g., Dibblee, 1987; Fig. 1-1), contains Oligocene-Miocene strata, the older parts of which had not been thoroughly investigated prior to this study. Similar Oligocene-Miocene strata are present in the Tejon, Soledad and Orocopia regions of southern California, which lie on different sides of the San Gabriel/Canton and San Andreas faults (e.g., Crowell, 1975a; Fig. 1-1). Some palinspastic reconstructions show the Tejon, Soledad and Orocopia regions as correlated and originally adjacent (e.g., Hill and Dibblee, 1953; Crowell, 1962, 1975a; Carman, 1964; Ehlig and

Ehlert, 1972; Bohannon, 1975), whereas others do not correlate some of these regions (e.g.,

Powell, 1981, 1993; Spittler and Arthur, 1982; Frizzell et al., 1986). The central Punchbowl block is generally accepted as an offset equivalent of the Soledad region; therefore, its previously understudied strata provide important constraints on these palinspastic reconstructions.

Furthermore, Oligocene-Miocene strata of the Punchbowl block straddle the Fenner fault, a component of a proposed early trace of the San Andreas fault, whose existence is debated (e.g.,

Powell, 1981, 1993; Richard, 1993). For these reasons, I conducted a detailed study of these strata. My findings support original alignment of the Tejon, Soledad, Punchbowl and Orocopia regions, and the slip estimates implied thereby, and argue against an early trace of the San

Andreas fault system along the Fenner fault.

GEOLOGIC BACKGROUND

Oligocene-Miocene Basins

Punchbowl block

8

The Punchbowl block contains the Punchbowl Formation (Noble, 1953, 1954), approximately 1500 m of fluvial/alluvial conglomerate, sandstone and minor mudstone, which accumulated during the middle-late Miocene (Tedford and Downs, 1965; Woodburne and Golz,

1972; Woodburne, 1975; Dibblee, 1987; Liu, 1990; Fig. 1-2). A distinct basal member is middle

Figure 1-2: Time-stratigraphic chart showing approximate ages of deposition of Oligocene-Miocene strata of Punchbowl block, and of Tejon, Soledad and Orocopia regions, arranged from west (left)

to east (right). Straight and wavy lines indicate conformable and unconformable relationships, respectively; queried contacts indicate uncertainty in nature of contact. Details based on data and references discussed in text, data of Chapter 4, and Stirton (1933), Woodburne and Whistler (1973),

Woodburne (1975), Ensley and Verosub (1982), McDougall (1982), Lander (1985), Frizzell and

Weigand (1993) and Coffey (2015). Age ranges of North American Land Mammal Ages from

Woodburne (1987) and Alroy (2000). Figure after Hoyt et al. (2018). 9

Miocene (Noble, 1954; Liu, 1990; personal communications of Allen and Whistler in Liu, 1990).

The older units documented in this study (the Paradise Springs and Vasquez formations; Fig. 1-

2) have been interpreted in previous studies as either part of the basal Punchbowl Formation

(e.g., Noble, 1954; Dibblee, 2002a, b, c), or deposits in a fault-bound sliver along the Punchbowl fault that originated in a separate basin (Weldon et al., 1993).

Tejon region

The oldest non-marine strata of the Tejon region belong to the Plush Ranch Formation

(Carman, 1954, 1964; Fig. 1-2), composed of alluvial and lacustrine conglomerate, sandstone, siltstone, shale, limestone and evaporites (Carman, 1964; Cole and Stanley, 1995; Hendrix et al.,

2010). Interbedded basalt has been dated by whole-rock and plagioclase K-Ar methods as ca. 26-

23 Ma (Crowell, 1973; recalculated after Dalrymple, 1979; Frizzell and Weigand, 1993).

Northwest of Plush Ranch basin, on the opposite side of Mount Pinos (which includes exposures of Pelona Schist) Oligocene-Miocene strata are generally mapped as Simmler Formation (e.g.,

Kellogg and Miggins, 2002; Dibblee, 2005a, b, 2006b; Fig. 1-1), but are considered equivalent to the Plush Ranch Formation (personal communications of Hill and Dibblee in Carman, 1964).

These strata coarsen upward, from mostly sandstone at the base to coarse conglomerate at the top

(Dibblee, 2005a, b).

Atop the Plush Ranch Formation in angular unconformity is the non-marine Caliente

Formation (named by T. W. Dibblee, Jr. in Stock, 1947; Schwade, 1954), which is composed of fluvial and lacustrine conglomerate, sandstone and mudstone, and minor tuffaceous and limestone beds (Carman, 1964; Ehlert, 2003; Fig. 1-2).

10

Soledad region

The oldest non-marine strata of the Soledad region belong to the Vasquez Formation

(Sharp, 1935; Jahns and Muehlberger, 1954; Muehlberger, 1958; Fig. 1-2), which is dominated by alluvial conglomerate and sandstone (Jahns and Muehlberger, 1954; Hendrix and Ingersoll,

1987). The Vasquez Formation is preserved in three subbasins (Jahns and Muehlberger, 1954;

Muehlberger, 1958; Hendrix and Ingersoll, 1987; Fig. 1-1). The Vasquez Rocks and Texas

Canyon subbasins, south of Sierra Pelona (where Pelona Schist is exposed; Fig. 1-1), are interpreted to have been depositionally distinct but kinematically linked for most of their history

(Bohannon, 1976; Hendrix and Ingersoll, 1987; Hendrix, 1993; Hendrix et al., 2010). The

Vasquez Rocks subbasin contains interbedded volcanic rocks, primarily basaltic andesite with some rhyodacite and rhyolite (Hendrix and Ingersoll, 1987; Frizzell and Weigand, 1993), dated by whole-rock and plagioclase K-Ar methods as ca. 26-23 Ma (Crowell, 1973; Spittler, 1974;

Woodburne, 1975; recalculated after Dalrymple, 1979; Frizzell and Weigand, 1993). Strata of the Charlie Canyon subbasin, north of Sierra Pelona (Fig. 1-1), coarsen upward, from siltstone and fine sandstone at the base to coarse conglomerate near the top (Sams, 1964; Hendrix and

Ingersoll, 1987).

Atop the Vasquez Formation in angular unconformity are strata designated as the Tick

Canyon Formation by Jahns (1939, 1940; Fig. 1-2). They consist of alluvial, fluvial and lacustrine conglomerate, sandstone, siltstone and claystone (Jahns, 1940; Woodburne, 1975).

These strata are overlain by, and were originally considered part of, the

(Kew, 1923, 1924), but were subsequently distinguished on the basis of an inferred disconformity (Jahns, 1940). Subsequent work discounted the presence of a disconformity

(Ehlert, 1982, 2003; Lander, 1985; Bishop, 1990). In this study, I use the term “Tick Canyon

11 strata” to distinguish these deposits from overlying strata of the Mint Canyon Formation and underlying strata of the Vasquez Formation.

Tick Canyon strata contain an unroofing sequence, culminating up-section in clasts of

Pelona Schist (Ehlert, 1982, 2003; Hendrix, 1993). The Tick Canyon strata also contain abundant volcanic clasts, most of which resemble volcanic rocks of the Vasquez Formation (Hendrix,

1993). The Charlie Canyon subbasin of the Soledad region also contains an unroofing sequence, in the form of a Pelona Schist-bearing, poorly sorted breccia, stratigraphically above the Vasquez

Formation but below the Mint Canyon Formation (e.g., Sams, 1964; Weber, 1994; Dibblee,

1997; Coffey, 2015). This breccia has previously been considered basal Mint Canyon Formation

(Dibblee, 1997) and a separate formation (part of the breccia: Sams,

1964; the Powerhouse breccia-conglomerate: Weber, 1994). In this study, these strata are referred to as Tick Canyon strata.

The Mint Canyon Formation consists primarily of fluvial, alluvial and lacustrine conglomerate, sandstone and mudstone (Kew, 1923, 1924; Ehlert, 1982, 2003; Fig. 1-2). The

Mint Canyon Formation is overlain by the dominantly marine Castaic Formation (Crowell,

1954), which consists of shale, sandstone and minor conglomerate (Crowell, 1954; Ehlert, 1982).

The contact between the Mint Canyon and Castaic formations is an angular unconformity in some places, and apparently conformable and gradational in others (Ehlert, 1982; Fig. 1-2).

Orocopia region

The only Oligocene-Miocene strata of the Orocopia region belong to the non-marine

Diligencia Formation (Crowell, 1975b; Fig. 1-2), which is composed of alluvial, fluvial and lacustrine conglomerate, sandstone, siltstone and limestone (Spittler and Arthur, 1982; Law et al., 2001; Ingersoll et al., 2014). It contains interbedded basalt and andesite flows, dated by

12 whole-rock and plagioclase K-Ar methods as ca. 24-21 Ma (Crowell, 1973; Spittler, 1974; recalculated after Dalrymple, 1979; Frizzell and Weigand, 1993), and andesitic sills and dikes

(Spittler and Arthur, 1982; Terres, 1984).

Potential Source Rocks

The oldest rocks of the San Gabriel and Punchbowl blocks are Paleoproterozoic gneisses, formed 1800-1660 Ma (Silver, 1966; Ehlig, 1981; Barth et al., 2001; Premo et al., 2007; Nourse and Premo, 2016), which were intruded by a complex of anorthosite, gabbro, syenite and norite

(e.g., Crowell, 1975a; Ehlig, 1981) ca. 1200 Ma (Barth et al., 1995, 2001). Perturbation of the

Paleoproterozoic gneisses during intrusion produced discordant zircon with 207Pb/206Pb ages of

1760-1300 Ma (Silver et al., 1963; Barth et al., 1995, 2001). “Anorogenic” plutons (1460-1400

Ma) are present throughout southern California and surrounding regions (Anderson and Bender,

1989), including minor exposure in the eastern San Gabriel Mountains (Premo et al., 2007; personal communication of J. Nourse, 2017 in Hoyt et al., 2018).

The Proterozoic basement of southern California is intruded by numerous, overlapping

Mesozoic plutons (e.g., Ehlig, 1981). In the San Gabriel Mountains, the oldest of these is the compositionally zoned Mount Lowe intrusion, emplaced 218-207 Ma (Barth and Wooden,

2006); similar-age plutons are present in the southern Mojave region (e.g., Barth et al., 1997).

Most plutons in southern California are substantially younger than the Mount Lowe intrusion

(e.g., Barth et al., 1997). In the San Gabriel Mountains, distinct magmatic episodes occurred

170-149 Ma and 90-75 Ma (e.g., Silver, 1971; May and Walker, 1989; Barth et al., 2008), producing quartz diorite to quartz monzonite (e.g., Ehlig, 1981).

The Paleogene San Francisquito Formation is exposed north of Blue Ridge in the

Punchbowl block (Dibblee, 1967, 1987). It consists of almost entirely marine shale, mudstone,

13 sandstone, conglomerate and minor carbonate, both in the Punchbowl block and in its type area in the Soledad region (Dibblee, 1967; Kooser, 1982); coeval marine deposits are present in the

Tejon (e.g., Kellogg et al., 2008) and Orocopia (Crowell and Susuki, 1959; Advocate et al.,

1988) regions.

The gneissic, granitic and sedimentary rocks of the San Gabriel Mountains lie within the upper plate of the Vincent thrust (e.g., Ehlig, 1981). The lower plate is composed of Pelona

Schist, which is predominantly meta-arkose (Haxel and Dillon, 1978; Ehlig, 1981; Jacobson et al., 2011). This relationship is exposed in a structural window in the southern San Gabriel

Mountains. Pelona Schist is also exposed in the core of an anticlinorium along Blue Ridge in the

Punchbowl block (e.g., Dibblee, 1968), and in the core of anticlinoria in the Tejon and Soledad regions (e.g., Ehlig, 1968; Crowell, 1975a). The correlated Orocopia Schist is exposed in an anticlinorium in the Orocopia region (e.g., Crowell, 1962; Ehlig, 1968; Haxel and Dillon, 1978;

Jacobson et al., 2007; Ingersoll et al., 2014).

Tectonic Reconstructions of the Southern San Andreas Fault System

The central Punchbowl block has been correlated with the Soledad region, implying 40-

50 km of dextral slip on the Punchbowl fault (Dibblee, 1967, 1968; Ehlig, 1968, 1981; Powell,

1993). Correlation has also been suggested between the central Punchbowl block and either the northwestern Orocopia region (Ehlert and Ehlig, 1977; Ehlig and Joseph, 1977) or the northern

Little San Bernardino Mountains (Ehlig and Joseph, 1977; Matti and Morton, 1993).

Correlation of the Tejon, Soledad and Orocopia regions (Fig. 1-1) has been proposed and refined based on similarities in both basement rocks, and sedimentary and volcanic strata (e.g.,

Hill and Dibblee, 1953; Crowell, 1962, 1975a; Carman, 1964; Ehlig and Ehlert, 1972;

Bohannon, 1975; Ehlert, 1982, 2003; Weigand, 1982; Frizzell and Weigand, 1993; Ingersoll et

14 al., 2014; Hoyt et al., 2018). Such correlations imply 60-70 km and approximately 240 km of dextral slip along the San Gabriel/Canton and San Andreas faults, respectively (e.g., Crowell,

1975a; Ingersoll et al., 2014). These correlations have been widely accepted, but also challenged by alternate reconstructions (e.g., Smith, 1977; Powell, 1981, 1993; Spittler and Arthur, 1982;

Frizzell et al., 1986; Matti and Morton, 1993; Weldon et al., 1993), most of which suggest that

80-110 km of dextral slip occurred along the San Francisquito-Fenner-Clemens Well fault, and only 42 km and 160-185 km along the San Gabriel/Canton and southern San Andreas faults, respectively (e.g., Powell, 1993). Some of the differences among these reconstructions may be reconcilable (Darin and Dorsey, 2013; Chapter 3).

GEOLOGY OF THE PUNCHBOWL BLOCK

I mapped the study area in the central Punchbowl block at 1:12,000 scale, documented sediment composition via conglomerate-clast counts, sandstone point counts and U-Pb dating of detrital zircon, and determined ages of igneous rocks via U-Pb dating of zircon.

Stratigraphy

Vasquez Formation

The stratigraphically lowest non-marine strata in the study area are compositionally distinct from overlying strata and contain interbedded volcanic rocks. Accordingly, I consider them a separate formation, which I refer to as the Vasquez Formation based on probable correlation with the Vasquez Formation of the Soledad region, as discussed below. The Vasquez

Formation of the study area is composed primarily of bright-red conglomerate and sandstone.

The conglomerate exhibits low degrees of rounding and sorting, a muddy matrix and commonly, reverse grading. South of Blue Ridge, the Vasquez Formation nonconformably overlies

15 granitoid, although the base of the section is excised by the Blue Ridge fault along much of its length (Fig. 1-3). Interbedded trachyandesite, previously undescribed, is present near the base of the section (Fig. 1-3), in places capped by thinly bedded tan limestone (too small to map). North of Blue Ridge, the Vasquez Formation also lies depositionally atop granitoid, possibly with some intervening San Francisquito Formation in places (Fig. 1-3). The Vasquez Formation north of

Blue Ridge contains lenses of very poorly sorted, very angular, matrix-poor megabreccias (map unit PεNvg in Figs. 1-3 and 1-4) interbedded with Vasquez Formation conglomerate and sandstone. These deposits fit the definitions of “crackle breccia facies” and “jigsaw breccia facies” (i.e., Yarnold and Lombard, 1989, p. 12).

The sorting, rounding, grading and matrix within Vasquez conglomerate suggest deposition by debris-flow and hyperconcentrated-flow mechanisms (e.g., Blackwelder, 1928;

Fisher, 1971). I interpret the Vasquez Formation to represent primarily proximal alluvial-fan deposits. Using the criteria of Yarnold and Lombard (1989) and Yarnold (1993), I interpret lenses of granitoid crackle and jigsaw breccia in the Vasquez Formation north of Blue Ridge as rock-avalanche deposits. I interpret the thin interval of thinly bedded limestone atop the interbedded trachyandesite as lacustrine, likely the result of ponding against the volcanic flows.

The high relative abundance of trachyandesite and rhyolite clasts immediately up section of the trachyandesite flows (see below) suggests that the rhyolite clasts are from deposits closely related to, and thus broadly coeval with, the trachyandesite flows. If so, then the 24.4 ± 0.9 Ma age I determined for the rhyolite clasts (Table S1-1) should approximate the age of the trachyandesite flows, and thus of the interbedded strata. Therefore, deposition of the Vasquez

Formation in the Punchbowl block probably began ca. 25-24 Ma or earlier.

16

Figure 1-3: Geologic map of central Punchbowl block. Note

distinct units of Oligocene-Miocene strata both north and

south of Blue Ridge, along which an anticlinorium of Pelona

Schist is exposed. Location shown in Figure 1-1, cross sections in Figure 1-4 and schematic stratigraphic columns in Figure 1-

5. For full-size figure, see Fig. S1-1 in Supplementary

Materials.

17

Figure 1-4: Cross sections along section lines of Figure 1-3. No vertical exaggeration.

18

Paradise Springs formation

Overlying the Vasquez Formation are compositionally distinct strata here informally dubbed the “Paradise Springs formation.” The Paradise Springs formation is texturally similar to the Vasquez Formation: it is composed of bright-red conglomerate and sandstone, with the conglomerate exhibiting low to moderate degrees of rounding and sorting, and a sandy to muddy matrix. Paradise Springs formation is present both north and south of Blue Ridge; in both cases, basal strata and the contact with the Vasquez Formation are covered by Quaternary deposits (Fig.

1-3). South of Blue Ridge, along the Blue Ridge fault, is a spatially restricted breccia (map unit

Npss in Fig. 1-3). This breccia is very angular and very poorly sorted, with a deep-maroon sandy matrix.

I interpret the Paradise Springs formation to represent alluvial-fan deposits. The textural differences compared to the Vasquez Formation suggest less proximal deposition, except for the breccia unit.

Punchbowl Formation

North of Blue Ridge, the Paradise Springs formation is overlain by the Punchbowl

Formation (Noble, 1953, 1954), with no discernable angular discordance (Fig. 1-3). The main member of the Punchbowl Formation is gray to white to tan conglomerate, sandstone and minor mudrock. The conglomerate is typically well sorted and sub-rounded to rounded, with sandy matrix. Channels are abundant, and exhibit normal grading. I recognized basal strata of the

Punchbowl Formation as a distinct member (Fig. 1-3). This basal member is composed of pink to red to buff conglomerate and sandstone. It is texturally intermediate between the overlying main member of the Punchbowl Formation and the underlying Paradise Springs formation, though

19 more similar to the former. The contact between the basal and main members of the Punchbowl

Formation is transitional.

The Punchbowl Formation has been interpreted as dominantly fluvial (e.g., Dibblee,

1987). The basal Punchbowl Formation’s intermediate texture and color compared to the

Paradise Springs and Punchbowl formations suggest that it represents the transition from alluvial to fluvial deposition.

Sediment Composition

Conglomerate-clast composition

Conglomerate-clast composition was determined at locations across the study area. A flexible grid was affixed to conglomeratic beds and the lithology of the clast at each crosshair determined until 100 counts were reached; grid spacing was varied between locations such that it always exceeded the average clast size.

Conglomerate-clast composition data are shown in Figure 1-5 (raw data are in Table S1-

2). Clasts of Pelona Schist are only present in the upper part of the Miocene strata (the Paradise

Springs and Punchbowl formations). Throughout the study area, granitoid makes up the majority of the conglomerate-clast population of the Vasquez Formation. Adjacent to and immediately upsection from the interbedded trachyandesite flows, trachyandesite and rhyolite clasts make up

30-39% of the conglomerate clasts; the more stable rhyolite clasts are found in low abundance throughout most of the section. The Paradise Springs formation conglomerate is dominated by sandstone clasts derived from the San Francisquito Formation; south of Blue Ridge, granitoid clasts are also abundant. Pelona Schist clasts are absent in the lower part of the Paradise Springs formation, but constitute up to 7% of the conglomerate clasts higher in the section; the areally restricted breccia unit is composed entirely of Pelona Schist clasts. The Punchbowl Formation,

20

21

Figure 1-5 (previous page): Schematic composite stratigraphic sections of Oliogocene-Miocene

strata (A) south and (B) north of Blue Ridge in Punchbowl block. Units, unit symbols (e.g., Nps) and unit colors same as in Figures 1-3 and 1-4; more detailed lithologic descriptions of units in text

and Figure 1-3. Paradise Springs formation and upper Vasquez Formation occur on both sides of

Blue Ridge, whereas lower Vasquez Formation is only present south of Blue Ridge, and Punchbowl

Formation is only north of Blue Ridge. Conglomerate-clast compositions are shown at approximate

stratigraphic positions at which they were measured (measurements at similar stratigraphic

positions have been combined; raw data, locations and details of clast lithologies in Table S1-2).

Note differences in composition between units, and occurrence of Pelona Schist clasts upsection in

both columns. Stratigraphic thicknesses were estimated using Figure 1-3. N = number of

measurement locations; n = total number of clasts counted; cgl. = conglomerate; ss. = sandstone.

including the basal member, is dominated by granitoid and gneiss clasts; clasts of sandstone from the San Francisquito Formation are present, but much less abundant than in the underlying

Paradise Springs formation.

Sandstone composition

Sandstone composition was determined for samples collected throughout the study area.

Sandstone samples were mounted as standard 30-μm-thick thin sections by R.A. Petrographic, then etched with concentrated hydrofluoric acid (HF) and stained with a saturated solution of sodium hexanitrocobaltate(III) (Na3Co(NO2)6). Etching and staining distinguish quartz

(unetched; unstained), potassium feldspar (stained with yellow dots) and plagioclase feldspar

(heavily etched; unstained; Gabriel and Cox, 1929; Reeder and McAllister, 1957; Ingersoll and

Cavazza, 1991). For each thin section, 500 points were analyzed using the Gazzi-Dickinson method of point-counting (Gazzi, 1966; Dickinson, 1970; Ingersoll et al., 1984). For each

22

Table 1-1. Definition of original and recalculated sandstone point-count grain categories Original Categories Category Symbol Description quartz, monocrystalline Qm quartz crystal with a maximum diameter >0.0625 mm quartz, polycrystalline Qp interlocking quartz crystals with maximum diameters <0.0625 mm; no other associated mineral phases feldspar, plagioclase Fp plagioclase-feldspar crystal with a maximum diameter >0.0625 mm feldspar, potassium Fk potassium-feldspar crystal with a maximum diameter >0.0625 mm mica, monocrystalline M mica (or chlorite) crystal with a maximum diameter >0.0625 mm dense mineral, monocrystalline D crystal of any mineral phase not listed above with a maximum diameter >0.0625 mm lithic fragment, metamorphic, Lma interlocking crystals with maximum diameters <0.0625 mm and little or no aggregate preferred orientation (typically quartz, feldspar and/or micas) lithic fragment, metamorphic, Lmt interlocking crystals with maximum diameters <0.0625 mm and distinct preferred tectonite orientation (typically quartz, feldspar and/or micas) lithic fragment, metamorphic, Lmm interlocking crystals of mica with maximum diameters <0.0625 mm; no other micaceous associated mineral phases lithic fragment, metamorphic, Lmv interlocking crystals with maximum diameters <0.0625 mm recrystallized from a metavolcanic volcanic rock (commonly plagioclase and chlorite) lithic fragment, volcanic, Lvl fragment of volcanic rock containing plagioclase laths lathwork lithic fragment, volcanic, Lvm fragment of volcanic rock with groundmass containing plagioclase microlites microlitic lithic fragment, volcanic, felsitic, Lvfs fragment of volcanic rock with groundmass of variable crystal size seriate lithic fragment, volcanic, felsitic, Lvfg fragment of volcanic rock with equigranular, microcrystalline groundmass granular lithic fragment, volcanic, vitric Lvv fragment of glassy volcanic rock lithic fragment, sedimentary, Lss fragment of mudrock, i.e., grains have maximum diameters <0.0625 mm siliciclastic (siltstone, shale, mudstone) lithic fragment, sedimentary, Lsc fragment of clastic or chemical sedimentary rock of dominantly carbonate carbonate composition and grain/crystal size <0.0625 mm miscellaneous and unknown M/U any sand-sized grain which does not fall within one of the above categories or which could not be confidently identified interstitial material Int. Interstitial material: cement, primary pore space, and detrital grains with maximum diameters <0.0625 mm

Recalculated Categories: Category Symbol Relationship to original categories total quartz Q Q = Qm + Qp total feldspar F F = Fk + Fp total (nonquartzose) lithics L L = Lma + Lmt + Lmm + Lmv + Lvm +Lvl +Lvfg +Lvfs + Lvv +Lss + Lsc total metamorphic lithics Lm Lm = Lma + Lmt + Lmm + Lmv total volcanic lithics Lv Lv = Lvm + Lvl + Lvfg + Lvfs + Lvv total (nonquartzose) Ls Ls = Lss + Lsc sedimentary lithics percent total quartz QFL%Q 100% ∙ Q / (Q + F + L) percent total feldspar QFL%F 100% ∙ F / (Q + F + L) percent total lithics QFL%L 100% ∙ L / (Q + F + L) percent monocrystalline quartz QmFkFp%Qm 100% ∙ Qm / (Qm + Fk + Fp) percent potassium feldspar QmFkFp%Fk 100% ∙ Fk / (Qm + Fk + Fp) percent plagioclase feldspar QmFkFp%Fp 100% ∙ Fp / (Qm + Fk + Fp) percent metamorphic lithics LmLvLs%Lm 100% ∙ Lm / (Lm + Lv + Ls) percent volcanic lithics LmLvLs%Lv 100% ∙ Lv / (Lm + Lv + Ls) percent sedimentary lithics LmLvLs%Ls 100% ∙ Ls / (Lm + Lv + Ls)

23 sample, grid spacing was greater than the average grain size. Categories of framework grains

(>0.0625 mm) are defined in Table 1-1; points interstitial to framework grains were also counted.

Figure 1-6: Sandstone composition of Oligocene-Miocene strata of Punchbowl block. Composition determined using Gazzi-Dickinson point-count method and plotted on QFL, QmFkFp and LmLvLs ternary diagrams. Q = quartz; F = feldspar; L = nonquartzose lithics; Qm = moncrystalline quartz;

Fk = potassium feldspar; Fp = plagioclase feldspar; Lm = metamorphic lithics; Lv = volcanic and

hypabyssal lithics; Ls = sedimentary lithics (full definitions of grain categories in Table 1-1). Note

differences in composition and compositional variability among units. Fm. = Formation; fm. =

formation. Raw data in Table S1-3.

24

Where diagenetic alteration had occurred, framework grains were counted as their original grain type. Point counts were also performed on samples collected and prepared by others in the same manner, and points counts of R. V. Ingersoll were added to my dataset; details are given in Table

S1-3.

Most of the Punchbowl-block Oligocene-Miocene strata show considerable variation in sandstone composition (Fig. 1-6). Sandstone from the main member of the Punchbowl

Formation, however, clusters very tightly on total quartz-total feldspar-total (nonquartzose) lithics (QFL) and monocrystalline quartz-potassium feldspar-plagioclase feldspar (QmFkFp) ternaries; the large spread on the metamorphic-volcanic-sedimentary lithic (LmLvLs) ternary is presumably caused by the extremely low abundance of lithic grains (grain categories defined in

Table 1-1; Hoyt et al., 2018). Sandstone from the basal Punchbowl Formation shows more variation than the overlying main Punchbowl Formation, but less variation than the underlying

Vasquez and Paradise Springs formations (Fig. 1-6). The Vasquez Formation, which is dominated by granitoid and volcanic conglomerate clasts, is dominated by feldspar, especially plagioclase, in the sandstone fraction. The Paradise Springs formation, which is dominated by sandstone conglomerate clasts from the San Francisquito Formation, contains significantly more quartz and sedimentary lithics in the sandstone fraction.

Detrital-zircon ages

Nine >1-kg sandstone samples were collected for detrital-zircon analysis (Fig. 1-3).

Samples were crushed and sieved, then separated by density and magnetism using: 1. a Mineral

Technologies MD Gemini shaking table at Pomona College or tetrabromoethane (ρ = 2.97 g/cm3), 2. a neodymium hand magnet and a model L-1 Frantz Isodynamic magnetic separator and 3. methylene iodide (ρ = 3.32 g/cm3). At the Arizona LaserChron Center at the University of

25

Arizona, large splits of these separates were mounted, together with grains of zircon references, on epoxy plugs, which were polished to a depth of approximately 20 μm to expose crystal interiors.

Detrital zircon was dated by U-Pb methods using laser-ablation multicollector inductively coupled plasma mass spectrometry (LA MC-ICPMS) at the Arizona LaserChron Center, using protocols described by Gehrels et al. (2006, 2008). Material was ablated using a Photon

Machines Analyte G2 excimer laser, then carried in helium into the plasma source of a Nu HR

ICPMS. A 204Pb-based common-lead correction was performed, using the common Pb composition from Stacey and Kramers (1975). U/Pb, Th/U and Pb-isotope relative sensitivities were determined using zircon reference Sri Lanka (563.5 ± 1.6 Ma; approximately 518 ppm U and 68 ppm Th; Gehrels et al., 2008). The accuracy of final ages was verified using zircon reference R33 (419.3 ± 0.4 Ma; Black et al., 2004). Analytical data are reported in Table S1-4.

Additional details of preparation and analysis are given in Coffey (2015).

Relative-probability distributions of detrital-zircon ages are given in Figure 1-7 (raw data are in Table S1-4). Those for all samples south of Blue Ridge (Fig. 1-7A) are dominated by peaks at ca. 1200 Ma and 1660-1800 Ma, and exhibit low, broad peaks at ca. 1400 Ma.

Distributions for all samples north of Blue Ridge (Fig. 1-7B) lack ca. 1200 Ma peaks, and exhibit smaller peaks at ca. 1400 Ma and 1660-1800 Ma. Many samples contain peaks at ca. 150 Ma, and some samples contain peaks at ca. 220 Ma, ca. 75 Ma, and/or ca. 25 Ma.

Provenance interpretations

All Oligocene-Miocene deposits south of Blue Ridge and the Fenner fault contain substantial ca. 1200 Ma zircon (Fig. 1-7A), whereas zircon of this age is entirely absent from all

26

Figure 1-7: Relative-probability distributions of detrital-zircon U-Pb ages from Oligocene-Miocene

strata of Punchbowl block (A) south and (B) north of Blue Ridge. Within A and B, samples are arranged in approximate stratigraphic order, with youngest on top. Note change of horizontal scale

at 300 Ma; vertical scale also differs to left and right of this scale break, such that equal areas

represent equal probability across graph. Labels indicate sample number and name of geologic

unit. n = number of analyses. Sample locations in Figure 1-3; raw data in Table S1-4. 27 such deposits north of Blue Ridge and the Fenner fault (Fig. 1-7B). This is consistent with a drainage divide along Blue Ridge during the Miocene (Sadler, 1993; Hoyt et al., 2018), and indicates that this divide was established prior to deposition of the lower Paradise Springs formation, and probably prior to deposition of the Vasquez Formation. Strata north and south of

Blue Ridge would thus have formed in distinct basins, hereafter referred to as the “northern” and

“southern” subbasins, respectively.

The granitoid depositionally beneath the Vasquez Formation (compositionally near the quadruple point of granite, quartz monzonite, granodiorite and quartz monzodiorite on a QAPF diagram; Streckeisen, 1974; and dated as 146 ± 3 Ma 1σ; Table S1-1) is presumably the primary proximal source of the Vasquez Formation. This interpretation accounts for the dominance of granitoid clasts in the conglomerate fraction (Fig. 1-5), the consistency of sandstone composition with that of both the granitoid and a granitoid conglomerate clast (Fig. 1-6), and the presence of ca. 150 Ma detrital-zircon ages (Fig. 1-7). South of Blue Ridge, these proximal alluvial-fan deposits likely interfingered with more distal deposits shed approximately northward from western and central parts of the San Gabriel block (presumably adjacent prior to Punchbowl-fault slip), where the anorthosite-gabbro-syenite-norite complex and Mount Lowe intrusion are exposed. This would account for the ca. 1200 Ma and ca. 220 Ma detrital-zircon ages, respectively, in the Vasquez Formation samples south of Blue Ridge (Fig. 1-7A), as well as the variability in sandstone composition (Fig. 1-6).

The Paradise Springs formation, in contrast to the Vasquez Formation, was derived primarily from the San Francisquito Formation, as documented by the abundance of sandstone clasts within Paradise Springs conglomerate (Fig. 1-5), and the greater relative abundance of quartz and sedimentary lithic fragments and lower relative abundance of plagioclase feldspar

28 within Paradise Springs sandstone (Fig. 1-6). Because present exposures of San Francisquito

Formation are abundant in the vicinity of Paradise Springs formation exposures, but absent in most of the area of Vasquez Formation exposure (Fig. 1-3), this contrast in source rock may be more spatial than temporal. Prominent peaks in the detrital-zircon age distributions at ca. 75 Ma, ca. 150 Ma, and 1660-1800 Ma imply diverse ultimate source rocks. These are also the dominant zircon ages for the San Francisquito Formation of the Soledad region (Jacobson et al., 2011), and so most of this diversity probably results from recycling of San Francisquito zircon, rather than direct input from multiple sources.

North of Blue Ridge, the presence of Pelona Schist clasts in the upper part of the Paradise

Springs formation, but absence in the lower part and in the underlying Vasquez Formation, represents an unroofing sequence (Fig. 1-5; Ingersoll and Colasanti, 2004; Colasanti and

Ingersoll, 2006). A similar unroofing sequence is present south of Blue Ridge (Fig. 1-5). The areally restricted schist breccia within the Paradise Springs formation, interpreted as a proximal alluvial-fan deposit, is located along the Blue Ridge fault, which bounds the anticlinorium of

Pelona Schist along Blue Ridge. This suggests that the Pelona Schist of Blue Ridge is the source of this breccia, and likely also of the Pelona Schist clasts within the unroofing sequence of the

Paradise Springs formation.

Basal Punchbowl Formation sandstone is compositionally much more variable than that of the overlying Punchbowl Formation, but less variable than that of the underlying Paradise

Springs and Vasquez formations (Fig. 1-6). The basal Punchbowl Formation contains a smaller proportion of gneissic clasts than overlying Punchbowl conglomerate, but a larger proportion than the Paradise Springs and Vasquez formations (Fig. 1-5). These observations support my interpretation of the basal Punchbowl Formation as documenting the shift in depositional

29 environment from alluvial fans to an integrated braided-fluvial system. In the northern subbasin, a ca. 235 Ma peak is prominent in the upper Paradise Springs formation (samples KTC-14-dz7 and dz10; Fig. 1-7B) and in the basal Punchbowl Formation (sample KTC-14-dz8), but rare or absent down section (sample KTC-14-dz6) and in the San Francisquito Formation of the Soledad region and the Pelona-Orocopia schist (Jacobson et al., 2011). Therefore, the presence of this age peak may document the beginning of this change in drainage patterns. The detrital-zircon age distribution for the basal Punchbowl Formation sample is similar to those of the overlying main member of the Punchbowl Formation (Hoyt et al., 2018), but with more ca. 220 Ma zircon and less ca. 245 Ma zircon. This may indicate that the source of the ca. 245 Ma zircon, which Hoyt et al. (2018) suggested is likely Middle Triassic plutons of the southern Mojave region (e.g., Barth et al., 1997), was one of the last source areas to become integrated into the developing

Punchbowl Formation drainage system. The provenance of the main member of the Punchbowl

Formation is discussed in Hoyt et al. (2018).

Structure

Blue Ridge fault

South of Blue Ridge, the contact between the Pelona Schist and the Vasquez and Paradise

Springs formations, mapped as depositional by Dibblee (2002a, c), is a fault, here termed the

“Blue Ridge fault,” along its entire length. The fault is poorly exposed, but in places slickensides and a narrow zone of hydrothermal alteration are present; slickenlines indicate almost pure dip slip (Fig. 1-3). The Blue Ridge fault terminates to the northwest against the Fenner fault.

Accordingly, it predates the Fenner and Punchbowl faults, and is likely a normal fault associated with the phase of extension that much of southern California underwent beginning ca. 25 Ma

(e.g., Tennyson, 1989), in which case it would be coeval with deposition of the Vasquez

30

Formation. Truncation of Paradise Springs strata indicates that the Blue Ridge fault either remained active until this time or was reactivated.

Fenner fault

North of Blue Ridge, the contact between the Pelona Schist and the San Francisquito and

Vasquez formations is the Fenner fault. Noble (1954) and this study (Fig. 1-3) inferred continuation of the Fenner fault westward through the Paradise Springs formation to the

Punchbowl fault, whereas other studies (e.g., Liu, 1990; Weldon et al., 1993; Dibblee, 2002a) interpreted the Fenner fault as predating, and being overlain by, the Paradise Springs formation; differing interpretations are possible because of poor, discontinuous exposure of the Fenner fault in this region.

Punchbowl fault

The Punchbowl fault is a regionally significant strand of the San Andreas fault (e.g.,

Dibblee, 1967), with documented reverse-dextral slip (e.g., Chester and Chester, 1998). In the southeastern part of the study area, a single fault trace is present; to the northwest, the fault splits into two diverging subparallel traces (Fig. 1-3; Dibblee, 2002a, c). The northeastern branch dips southwestward, and the dip at the surface is shallower farther northwest (Fig. 1-3; orientations were not measured on the southwestern branch). Uplift of the San Gabriel Mountains has generally been greater in the eastern half, as indicated by higher elevations and exposure of deeper structural levels (e.g., Bull, 1987). Accordingly, the present surface exposure of the

Punchbowl fault within the study area is an oblique view, with progressively deeper structural levels exposed progressively southeastward. Splitting of the fault and shallowing of the dip of the northeastern branch to the northwest suggests a positive flower structure, in which a nearly

31 vertical strike-slip fault at depth splits into two branches, which shallow and exhibit more reverse slip upward (i.e., Wilcox et al., 1973; Sylvester, 1988).

Smaller faults and folds

Faults and folds subparallel to the Punchbowl fault in the Punchbowl Formation in the northwestern part of the study area (in the region of A-A’ in Figs. 1-3 and 1-4) are presumably transpressional features associated with slip and shortening along the Punchbowl fault. The fault southeast of the Punchbowl syncline (near the middle of A-A’ and B-B’ in Figs. 1-3 and 1-4) does not terminate in the basal Punchbowl Formation, as indicated by previous studies (e.g.,

Dibblee, 2002a), but rather cuts, and thus post-dates, the lower part of the main member of the

Punchbowl Formation. Additionally, shear-sense indicators along exposure of this fault zone in a road-cut immediately west of B-B’ (Figs. 1-3 and 1-4) suggest oblique, reverse-dextral slip.

Accordingly, this fault is here interpreted as a transpressional feature related to the Punchbowl fault, rather than as an extensional feature from earlier in the Miocene. This fault presumably originally dipped southwest (Fig. 1-8A), and underwent horizontal-axis rotation to achieve its present, steep northeast dip (Fig. 1-8B); this rotation is likely also responsible for the steeply southwest-dipping Punchbowl strata in this area (Fig. 1-3). This fault was likely kinematically linked with the Punchbowl fault, and may splay off the Punchbowl fault at depth, as part of the positive flower structure proposed above (Figs. 1-4 and 1-8). Reverse motion on the Fenner fault could also be part of this flower structure.

Faults and folds within the San Francisquito Formation and granitoid basement are subparallel to both the Punchbowl fault and the (presumably older) Blue Ridge fault (Fig. 1-3), and could be either high-angle reverse faults and associated folds related to transpression, potentially comprising part of the proposed flower structure, or listric normal faults and

32 associated folds related to Oligocene-Miocene extension. It is also possible that these faults began as normal faults during Oligocene-Miocene extension, and have subsequently been

Figure 1-8: Schematic cross section showing hypothesized positive flower structure along

Punchbowl fault, A: in its initial geometry, and B: in its present, deformed geometry. This

hypothesized structure is composed of two strands of the Punchbowl fault, the Fenner fault

(possibly representing reactivation of an older, normal fault), and potentially, the fault south of

Punchbowl syncline (Fig. 1-3). Boxes in B indicate how cross sections of Figure 1-4 fit this

schematic. See text for discussion.

33

reactivated as reverse faults because of their favorable orientation; such reactivation at a larger scale could potentially explain why the trace of the Punchbowl fault so closely follows that of the

Blue Ridge fault. These possibilities predict different ages, geometries and kinematics for these faults; additional detailed work might clarify these relations.

Vincent thrust

Southwest of the southeastern part of the Punchbowl fault, the Vincent thrust separates

Pelona Schist from mylonitic gneiss (Ehlig, 1981; Jacobson, 1983). Previous mapping (e.g.,

Dibblee, 2002c) inferred a fault subparallel to the Punchbowl fault along the southwestern edge of an intrusive body (map unit Pε?gd in Fig. 1-3) of probable late Oligocene age (May and

Walker, 1989; Nourse, 2002), against which the Vincent thrust terminates. I did not find evidence for this fault; rather, my mapping suggests that the Vincent thrust was intruded by this intrusive body (Fig. 1-3). The Vincent thrust may have acted as a conduit along which magma could more easily flow (H-H’ and I-I’ in Fig. 1-4); this would explain the anomalously large size of the body intruding the Vincent thrust compared to the surrounding, coeval sills and dikes intruding the Pelona Schist and mylonitic basement (Fig. 1-3; Dibblee, 2002c).

DISCUSSION

Basin Development

Mid-Cenozoic basin development likely began with initiation of normal faulting in the latest Oligocene. Distinct subbasins formed north and south of Blue Ridge, separated by an ancestral Blue Ridge topographic high, which acted as a drainage divide. Relief and erosion along this topographic high was sufficient to produce the Vasquez Formation alluvial-fan

34 deposits that dominate the margins of the two subbasins. These proximal deposits interfingered with more distal, finer-grained alluvial deposits derived from the opposite margins of these subbasins. The Pelona Schist along Blue Ridge was entirely in the subsurface throughout deposition of the Vasquez Formation, presumably covered by the granitoid presently exposed on either side of Blue Ridge (Fig. 1-3), clasts of which dominate the conglomerate fraction of the

Vasquez Formation in both subbasins. Extension via normal faulting was accompanied by bimodal volcanism, which produced the trachyandesite flows and the rhyolite and trachyandesite conglomerate clasts found in the Vasquez Formation of the southern subbasin (Figs. 1-3 and 1-

5). The southern subbasin may have been primarily a half-graben controlled by a normal fault on its southern margin, with the Blue Ridge fault forming later as an antithetic normal fault. This would explain both the substantial fine-grained sediment input from the south implied by detrital-zircon data and tilting of Vasquez strata away from, rather than toward, the Blue Ridge fault (Fig. 1-3).

The northern and southern subbasins persisted until (or were regenerated during) deposition of the Paradise Springs formation, separated as before by an ancestral Blue Ridge drainage divide. Alluvial-fan deposits in these subbasins were sourced from the San Francisquito

Formation and, in the southern subbasin, granitoid and Vasquez volcanic rocks. During deposition of the Paradise Springs formation, Pelona Schist was first exposed along Blue Ridge, and began contributing detritus to both subbasins.

Following deposition of the Paradise Springs formation, drainage patterns north of Blue

Ridge gradually changed, resulting in transition from alluvial-fan to braided-fluvial deposition as the basal member of the Punchbowl Formation accumulated. The main member of the

Punchbowl Formation accumulated in a well integrated fluvial system with its headwaters

35 outside the Punchbowl block (Hoyt et al., 2018). A similar transition may have occurred south of

Blue Ridge (Hoyt et al., 2018), but no post-Paradise Springs strata are preserved in this part of the Punchbowl block.

The southern San Andreas fault became active ca. 5 Ma (Nicholson et al., 1994; Ingersoll and Rumelhart, 1999; Oskin et al., 2001; Crowell, 2003; Oskin and Stock, 2003), with the

Punchbowl fault probably representing the initial main trace (Sharp and Silver, 1971). The San

Andreas fault has been transpressional throughout this part of southern California as a result of a restraining double bend (term of McClay and Bonora, 2001) of regional scale, extending from

San Gorgonio Pass (Fig. 1-1) in the southeast to the Tejon region just north of the Garlock fault in the northwest (e.g., Hill and Dibblee, 1953; Ingersoll and Coffey, 2017). This transpression shut down the Punchbowl drainage system, and caused uplift, deformation and erosion of

Oligocene-Miocene strata.

Regional Stratigraphic Correlations

Vasquez Formation

The Vasquez Formation of the Punchbowl block generally resembles the Plush Ranch,

Vasquez and Diligencia formations of the Tejon, Soledad and Orocopia regions, respectively, in terms of age, lithology, depositional mechanisms and relationships with surrounding units. The

Vasquez Formation of the Punchbowl block partly overlaps in sandstone composition with the

Vasquez Formation of the Soledad region (Fig. 1-9). In all four regions, lacustrine deposits are rare, but occur immediately above volcanic strata. The basin geometry I propose for the southern sub-basin is analogous to that described for the Plush Ranch basin of the Tejon region and the

Texas Canyon sub-basin of the Soledad region: a major fault along the southern margin which

36

Figure 1-9: Comparison of sandstone composition of Vasquez Formation of Punchbowl block and

Vasquez Formation of Soledad region, which are likely correlative. Details as in Figure 1-6.

Soledad-region data from Hendrix (1986); Punchbowl-block data of this study (Table S1-3).

exposed Proterozoic basement and generated large alluvial-fan systems, the fine-grained, distal parts of which interfingered and mixed with proximal, coarse-grained deposits shed from granitoid along the northern margin of the basin, which was bounded by a smaller, antithetic fault (Hendrix, 1993; Cole and Stanley, 1995; Hendrix et al., 2010). In both the southern subbasin and the Texas Canyon subbasin, Vasquez Formation and underlying granitoid along the northern margin is in fault contact with an anticlinorium of Pelona Schist (e.g., Hendrix, 1993).

37

38

Figure 1-10 (previous page): Schematic composite stratigraphic sections for Oligocene-Miocene strata of Punchbowl block, and Tejon, Soledad and Orocopia regions. Sections are arranged from west (left) to east (right). Approximate present locations of these sections are shown in Figure 1-1.

Tick marks along left margins of columns are spaced every 1000 meters and begin at base of

Oligocene-Miocene strata. Units of matching color are likely correlative. A: Basins south of Blue

Ridge, Sierra Pelona and equivalent bodies of Pelona-Orocopia schist. Sections are based on following sources: Tejon region: Carman (1964), Dibblee (2006a); Soledad region: Hendrix and

Ingersoll (1987), Dibblee (1996a, b); Punchbowl block: this study. B: Basins north of Blue Ridge,

Sierra Pelona and equivalent bodies of Pelona-Orocopia schist. Sections are based on following sources: Tejon region: Dibblee (2005a, b); Soledad region: Sams (1964), Hendrix and Ingersoll

(1987), Dibblee (1997), Coffey (2015); Punchbowl block: Dibblee (1987), this study; Orocopia region: Spittler and Arthur (1982).

Both the southern subbasin and the Vasquez Rocks subbasin contain interbedded, intermediate volcanic rocks and detritus of the anorthosite-gabbro-syenite-norite-complex derived from the south (Fig. 1-10A; Hendrix and Ingersoll, 1987). In light of these similarities, the southern subbasin of the central Punchbowl block is likely a close equivalent of the Plush Ranch basin of the Tejon region and the Texas Canyon and Vasquez Rocks subbasins of the Soledad region.

The northern subbasin of the Punchbowl block lies north of the Fenner fault and Blue

Ridge anticlinorium, a position analogous to that of the Charlie Canyon subbasin of the Soledad region (e.g., Dibblee, 1997), with which it is likely equivalent. Similar rockslide megabreccias near the top of the Vasquez Formation in each subbasin (Fig. 1-10B; Sams, 1964; Weber, 1994;

Dibblee, 1997) support this correlation. These subbasins may also correlate with the Simmler

Formation north of Pelona Schist exposures in the Tejon region. Consistent with this correlation

39 is a general increase in clast size upsection in both the Charlie Canyon subbasin (Sams, 1964;

Hendrix and Ingersoll, 1987) and the Simmler Formation (Fig. 1-10B; Dibblee, 2005a, b). The

Diligencia Formation occupies an analogous position north of the Orocopia Mountains anticlinorium in the Orocopia region, and thus may correlate with these basins as well (Fig. 1-

10B; Ingersoll et al., 2014).

Figure 1-11: Comparison of sandstone composition of Paradise Springs formation of Punchbowl block and Tick Canyon strata of Soledad region, which are likely correlative. Details as in Figure 1-

6. All data of this study (Table S1-3).

40

Paradise Springs formation

The Paradise Springs formation resembles the Tick Canyon strata of the Soledad region: both represent primarily alluvial-fan deposits with highly variable and partly overlapping sandstone composition (Fig. 1-11), and both contain unroofing sequences documenting exhumation of Pelona Schist (Sams, 1964; Ehlert, 1982, 2003; Hendrix, 1993; Weber, 1994;

Dibblee, 1997; Coffey, 2015). Additionally, both units overlie sandstone, conglomerate and coeval volcanic strata of the Vasquez Formation, and both are overlain by middle-upper Miocene sandstone and conglomerate with no significant angular discordance (Fig. 1-10). The Paradise

Springs formation of the southern subbasin presumably correlates with the type Tick Canyon strata, south of Sierra Pelona, whereas that of the northern subbasin presumably correlates with the Tick Canyon strata of Charlie Canyon subbasin, north of Sierra Pelona.

Punchbowl Formation

The Punchbowl Formation is largely coeval with the Caliente and Mint Canyon formations of the Tejon and Soledad regions, respectively, but represents a distinct drainage system (Hoyt et al., 2018). Strata mapped as Punchbowl Formation are present southeast of

Sierra Pelona, at the eastern edge of the Soledad region (e.g., Dibblee, 2001). A sample of these strata analyzed by Coffey (2015) closely matched those of the Punchbowl Formation of the

Punchbowl block (Hoyt et al., 2018) in both sandstone composition and detrital-zircon age distributions, suggesting that these strata are indeed part of the Punchbowl Formation. Additional sampling and study of these strata would be required to determine their paleogeographic significance.

Regional Tectonic and Paleogeographic Reconstructions

Punchbowl fault

41

Previous correlation of the central Punchbowl block and the Soledad region has been based on correlation of 1. the anticlinoria of Pelona Schist along Blue Ridge and Sierra Pelona,

2. the Fenner and San Francisquito faults along the northern margins of these anticlinoria, and 3. the presence of San Francisquito Formation north of these anticlinoria (Dibblee, 1967, 1968;

Ehlig, 1968, 1981; Powell, 1993). My correlation of Oligocene-Miocene strata on either side these anticlinoria confirms these correlations, and the 40-50 km of dextral slip on the Punchbowl fault that they imply.

San Gabriel/Canton and San Andreas faults

My findings support reconstructions of the San Gabriel/Canton and San Andreas faults that closely align crustal blocks of the Tejon, Soledad and Orocopia regions (e.g., Hill and

Dibblee, 1953; Crowell, 1962, 1975a; Carman, 1964; Ehlig and Ehlert, 1972; Bohannon, 1975;

Ehlert, 1982, 2003; Weigand, 1982; Frizzell and Weigand, 1993; Ingersoll et al., 2014; Hoyt et al., 2018). In particular, the presence of Oligocene-Miocene strata on both flanks of the Blue

Ridge anticlinorium helps link the Soledad region, where strata south of the Sierra Pelona anticlinorium have been given greater emphasis (e.g., Crowell, 1975a; Hendrix and Ingersoll,

1987), with the Orocopia region, where Oligocene-Miocene strata are presently preserved only on the north side of its anticlinorium (e.g., Crowell, 1975b). Correlating the Diligencia basin of the Orocopia region with the Charlie Canyon subbasin of the Soledad region (e.g., Bohannon,

1975; Ingersoll et al., 2014), with the northern subbasin of the central Punchbowl block as the link, eliminates problems with correlation of the Soledad and Orocopia regions (e.g., those discussed by Law et al., 2001).

San Francisquito-Fenner-Clemens Well fault

42

The similarity of the stratigraphic sequences in the northern and southern subbasins of the central Punchbowl block argues against the proposed 80-110 km of dextral slip along the Fenner fault (and the correlated San Francisquito and Clemens Well faults; Powell, 1981, 1993), which separates the two subbasins. In particular, the presence of clasts of San Francisquito Formation sandstone in Paradise Springs formation conglomerate of the southern subbasin (Fig. 1-5) seems incompatible with this magnitude of slip. These clasts link the southern subbasin, which is south of the Fenner fault, with the San Francisquito Formation, which is north of the Fenner fault (Fig.

1-3). Even if deposition of the Paradise Springs formation occurred after this proposed Fenner fault slip, the presence of a drainage divide along Blue Ridge throughout deposition of the

Paradise Springs formation, discussed above, would have prevented transport of these clasts across the Fenner fault into the southern subbasin. Rather, these clasts are presumably derived from San Francisquito Formation outcrop originally present south of the Fenner fault.

Furthermore, whereas restoration of 60-70 km of dextral slip on the San Gabriel/Canton fault aligns the Simmler Formation of the Tejon region with its likely correlative, the Vasquez

Formation of the Charlie Canyon subbasin of the Soledad region (e.g., Ingersoll et al., 2014), restoration of 80-110 km of dextral slip along the San Francisquito-Fenner-Clemens Well fault, as proposed by Powell (1981, 1993), would eliminate this cross-fault match.

If dextral slip of 80-110 km accumulated along the San Francisquito-Fenner-Clemens

Well fault, then most or all of it would have done so in the few million years after deposition of the Vasquez and Diligencia formations, and before deposition of the Paradise Springs formation and Tick Canyon strata (Fig. 1-2). Unroofing sequences document exhumation of the Pelona

Schist in the northern subbasin of the Punchbowl block (Paradise Springs formation; Ingersoll and Colasanti, 2004; Colasanti and Ingersoll, 2006; this study) and in the Charlie Canyon

43 subbasin of the Soledad region (Tick Canyon strata; Sams, 1964; Weber, 1994; Dibblee, 1997;

Coffey, 2015). These unroofing sequences link these strata, which are north of the San

Francisquito-Fenner-Clemens Well fault, with the Pelona Schist anticlinoria south of this fault.

Any significant dextral slip must have post-dated deposition of the Vasquez and Diligencia formations, which are truncated by the San Francisquito-Fenner-Clemens Well fault (e.g., Jahns and Muehlberger, 1954; Muehlberger, 1958; Crowell, 1975b).

Whereas the San Francisquito-Fenner-Clemens Well fault is largely high angle and bears evidence of some dextral slip (e.g., Crowell, 1962; Stanley, 1966; Konigsberg, 1967; Spittler and

Arthur, 1982; Terres, 1984; Ebert, 2004; Yan et al., 2005), it has been suggested that this represents minor (i.e., ≤10 km; Crowell, 1962; Spittler and Arthur, 1982; Terres, 1984; Ebert,

2004) reactivation of a what was originally an Oligocene-Miocene normal fault (e.g., Spittler and

Arthur, 1982; Hendrix and Ingersoll, 1987; Goodmacher et al., 1989; Robinson and Frost, 1996;

Bunker and Bishop, 2001; Yan et al., 2005; Jacobson et al., 2007; Ingersoll et al., 2014). I suggest that down-to-north normal faulting along the Fenner fault likely contributed to subsidence of the northern subbasin of the central Punchbowl block.

Synthesis

Correlation of Oligocene-Miocene strata of the central Punchbowl block with those of the

Tejon, Soledad and Orocopia regions supports the hypothesis that all four crustal blocks were aligned prior to slip along the San Gabriel/Canton and San Andreas/Punchbowl faults (e.g., Hill and Dibblee, 1953; Crowell, 1962, 1975a; Carman, 1964; Dibblee, 1967, 1968; Ehlig, 1968,

1981; Ehlig and Ehlert, 1972; Bohannon, 1975; Ehlert, 1982, 2003; Weigand, 1982; Frizzell and

Weigand, 1993; Ingersoll et al., 2014; Hoyt et al., 2018). Combining these correlations with paleomagnetic data (Terres, 1984; Terres and Luyendyk, 1985; Hornafius et al., 1986; Carter et

44 al., 1987; Ellis et al., 1993), and stratigraphic and provenance data (see references above and in

Fig. 1-12 caption), I propose the following sequence of paleogeographic reconstructions: 1.

Normal faulting initiated ca. 25 Ma, forming two parallel belts of basins separated by a topographic high along Sierra Pelona (Hendrix and Ingersoll, 1987) and Blue Ridge. Alluvial-fan deposits, partly sourced from this topographic high, and bimodal volcanic flows accumulated in these basins (Fig. 1-12A). 2. The Canton fault (precursor to the ; e.g., Crowell,

2003) initiated ca. 18 Ma. East of the Canton fault, Pelona-Orocopia schist was exhumed along the ancestral Sierra Pelona/Blue Ridge/Orocopia Mountains topographic high (e.g., Sams, 1964;

Konigsberg, 1967) as these crustal blocks underwent clockwise vertical-axis rotation (Fig. 1-

12B). Exposed Pelona Schist contributed detritus to alluvial-fan deposits on either side of this topographic high. 3. Beginning ca. 15 Ma, braided-fluvial systems began to develop on either side of the ancestral Sierra Pelona/Blue Ridge/Orocopia Mountains topographic high. By ca. 13

Ma, these fluvial systems were established, with the intervening topographic high continuing to act as a drainage divide (Fig. 1-12C). 4. Continued dextral slip along the Canton fault and its successor, the San Gabriel fault (e.g., Crowell, 2003), displaced the Tejon block from its eastern equivalents. This disrupted the Caliente/Mint Canyon drainage system, and allowed a marine incursion into the western Soledad region, where the Castaic Formation began to accumulate

(e.g., Crowell, 1954; Ehlert, 1982; Fig. 1-12D). The Punchbowl drainage system may have supplied sediment to the Castaic Formation, and/or breached the ancestral Sierra Pelona/Blue

Ridge drainage divide to deliver sediment to the eastern Soledad region. The northern Caliente

Formation, which is compositionally distinct from the rest of the Caliente Formation (Hoyt et al.,

2018), may have formed during this time as well. 5. Slip began along the southern San Andreas

45

Figure 1-12 (following pages): Schematic representation of interpreted paleogeography, source

rocks and depositional systems of Tejon-Soledad-Punchbowl-Orocopia regions. A: ca. 23 Ma,

during deposition of Vasquez, Plush Ranch and Diligencia formations. An ancestral Sierra

Pelona/Blue Ridge drainage divide separates two parallel belts of basins, in which alluvial-fan

deposits and volcanic flows accumulate. Paleogeography: ABR = ancestral Blue Ridge; ASP = ancestral Sierra Pelona; MCR = Mint Canyon Ridge; Mt. = Mount. Active faults (red): BRf = Blue

Ridge fault; CWf = Clemens Well fault; Df = Diligencia fault; Ff = Fenner fault; OMf = Orocopia

Mountains fault; SFf = San Francisquito fault; LVf = Lockwood Valley (or Big Pine) fault; Pf =

Pelona fault; Sf = Soledad fault; VCf = Vasquez Canyon fault. Sedimentary deposits: CC = Vasquez

Formation, Charlie Canyon subbasin; D = Diligencia Formation; PR = Plush Ranch Formation; S =

Simmler Formation; TC = Vasquez Formation, Texas Canyon subbasin; VPS, VPN = Vasquez

Formation of Punchbowl block, southern and northern subbasins, respectively; VR = Vasquez

Formation, Vasquez Rocks subbasin. B: ca. 18 Ma, during deposition of Paradise Springs

formation and Tick Canyon strata. Pelona-Orocopia schist has been exhumed along ancestral

Sierra Pelona/Blue Ridge drainage divide, and contributes detritus to alluvial-fan deposits accumulating on either side of this divide. Dextral Canton fault (precursor to San Gabriel fault) is active, and crustal blocks on its eastern side are undergoing vertical-axis clockwise rotation. PSN,

PSS = Paradise Springs formation, north of and south of, respectively, Blue Ridge; TCC = Tick

Canyon strata, Charlie Canyon; TCT = Tick Canyon strata, type locality. C: ca. 13 Ma, during

deposition of older parts of Punchbowl, Mint Canyon and Caliente formations. Alluvial-fan deposition has given way to braided-fluvial deposition. Dextral slip continues to accumulate along

Canton fault. Clockwise rotation already underway in B is now complete (though subsequent

rotations which bring these blocks into their modern orientations have yet to occur). P =

Punchbowl Formation. D: ca. 9 Ma, during deposition of younger parts of Punchbowl and Caliente formations, and of Castaic Formation. Continued dextral slip along Canton fault and its successor,

46

San Gabriel fault, has displaced Tejon block northwest. Marine deposition has begun in western

Soledad region, possibly supplied with sediment by Punchbowl drainage system. Details of A-D

after Crowell (1975a), Hendrix and Ingersoll (1987), Link (2003), Ingersoll et al. (2014), Ingersoll

and Coffey (2017), Hoyt et al. (2018), and other ideas and references presented in text.

fault ca. 5 Ma, inducing transpression, which shut down the Punchbowl drainage system and deformed and uplifted Oligocene-Miocene strata. Slip along the Punchbowl fault, probably the initial trace of the southern San Andreas fault, separated the Soledad block from the Punchbowl block. Following abandonment of the Punchbowl fault in favor of the present trace of the southern San Andreas fault, the Punchbowl block was separated from the Orocopia block.

Shortening between the southern and northern Peninsular batholiths likely induced additional vertical-axis rotations along the San Andreas fault system, bringing the crustal blocks in Figure 1-12D into their present orientations (e.g., Ingersoll and Coffey, 2017).

CONCLUSIONS

Oligocene-Miocene strata of the central Punchbowl block are composed of three distinct units: the Vasquez, Paradise Springs and Punchbowl formations. These strata were deposited in distinct subbasins north and south of a drainage divide along what is now Blue Ridge. The alluvial-fan deposits and interbedded volcanic rocks of the Vasquez Formation document extension via normal faulting initiating ca. 25-24 Ma. The alluvial-fan deposits of the Paradise

Springs formation contain an unroofing sequence that documents final exhumation of the Pelona

Schist during the middle Miocene. A distinct basal member of the Punchbowl Formation records

47

48

49 transition to the braided-fluvial deposits that comprise the main member of the Punchbowl

Formation.

Correlation of the Vasquez and Paradise Springs formations with equivalent strata of the

Soledad region confirms previous estimates of 40-50 km of dextral slip along the Punchbowl fault. Probable correlation with strata of the Tejon and Orocopia regions supports estimates of

60-70 km of dextral slip along the San Gabriel/Canton fault and approximately 240 km of dextral slip along the southern San Andreas fault. Conversely, similarities between the northern and southern subbasins of the central Punchbowl block and probable correlations between strata of the Tejon and Soledad regions argue against dextral slip of 80-110 km for the San Francisquito-

Fenner-Clemens Well fault.

50

REFERENCES CITED

Advocate, D.M., Link, M.H., and Squires, R.L., 1988. Anatomy and history of an Eocene

submarine canyon: the Maniobra Formation, southern California, in Filewicz, M.V., and

Squires, R.L., eds., Paleogene stratigraphy, west coast of North America [Book 58]:

Pacific Section, Society of Economic Paleontologists and Mineralogists (SEPM), Los

Angeles, p. 45-58.

Alroy, J., 2000, New methods for quantifying macroevolutionary patterns and processes:

Paleobiology, v. 26, p. 707-733.

Anderson, J.L., and Bender, E.E., 1989, Nature and origin of Proterozoic A-type granitic

magmatism in the southwestern United States of America: Lithos, v. 23, p. 19–52,

https://doi.org/10.1016/0024-4937(89)90021-2

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, https://doi.org/10.1086/499573

Barth, A.P., Wooden, J.L., Tosdal, R.M., Morrison, J., Dawson, D.L., and Hernly, B.M., 1995,

Origin of gneisses in the aureole of the San Gabriel anorthosite complex and implications

for Proterozoic crustal evolution of southern California: Tectonics, v. 14, p. 736-752,

https://doi.org/10.1029/94TC02901

Barth, A.P., Tosdal, R.M., Wooden, J.L., and Howard, K.A., 1997, Triassic plutonism in

southern California: Southward younging of arc initiation along a truncated continental

margin: Tectonics, v. 16, p. 290-304, https://doi.org/10.1029/96TC03596

51

Barth, A.P., Wooden, J.L., Coleman, D.S., 2001, SHRIMP-RG U-Pb zircon geochronology of

Mesoproterozoic metamorphism and plutonism in the southwesternmost United States:

Journal of Geology, v. 109, p. 319-327, https://doi.org/10.1086/319975

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, https://doi.org/10.1130/2008.2438(13)

Bishop, K.M., 1990, Tectonic model for the evolution of the Texas Canyon basin and Pelona

fault, Soledad basin, southern California, in Buckley, C.I., and Larson, R.A., eds.,

Geology and engineering geology of the western Soledad basin, Los Angeles County,

California: Association of Engineering Geologists, Southern California Section, Annual

Field Trip Guidebook, p. 35-56.

Black, L., Kamo, S., Allen, C., Davis, D., Aleinikoff, J., Valley, J., Mundil, R., Campbell, I.,

Korsch, R., Williams, I., and Foudoulis, C., 2004, Improved 206Pb/238U microprobe

geochronology by the monitoring of a trace-element-related matrix effect; SHRIMP, ID-

TIMS, ELA-ICP-MS and oxygen isotope documentation for a series of zircon standards:

Chemical Geology, v. 205, p. 115-140, https://doi.org/10.1016/j.chemgeo.2004.01.003

Blackwelder, E., 1928, Mudflow as a geologic agent in semiarid mountains: Geological Society

of America Bulletin, v. 39, p. 465-484, https://doi.org/10.1130/GSAB-39-465

Bohannon, R.G., 1975, Mid-Tertiary conglomerates and their bearing on Transverse Range

tectonics, southern California, in Crowell, J.C., ed., The San Andreas fault in southern

California: A guide to the San Andreas fault from Mexico to Carrizo Plain: California

Division of Mines and Geology Special Report, v. 118, p. 75-82.

52

Bohannon, R.G., 1976, Mid-Tertiary nonmarine rocks along the San Andreas fault in southern

California [Ph.D. thesis]: Santa Barbara, University of California, 311 p.

Bull, W.B., 1987, Relative rates of long-term uplift of mountain fronts in Crone, A.J., and

Omdahl, E.M., eds., Proceedings of Conference XXXIX – Directions in

Paleoseismology: U.S. Geological Survey Open-file Report 87-673, p. 192-202.

Bunker, J.V., and Bishop, K.M., 2001, The San Francisquito fault, southern California: a normal

fault related to uplift of the Sierra Pelona anticline?: Geological Society of America

Abstracts with Programs, v. 33, no. 3, p. 73.

Carman, M.F., Jr., 1954, Geology of the Lockwood Valley area, Kern and Ventura Counties,

California [Ph.D. thesis]: Los Angeles, University of California, 194 p.

Carman, M.F., Jr., 1964, Geology of the Lockwood Valley area, Kern and Ventura Counties,

California: California Division of Mines and Geology Special Report 81, 62 p.

Carter, J.N., Luyendyk, B.P., and Terres, R.R., 1987, Neogene clockwise tectonic rotation of the

eastern Transverse Ranges, California, suggested by paleomagnetic vectors: Geological

Society of America Bulletin, v. 98, p. 199-206, https://doi.org/10.1130/0016-

7606(1987)98<199:NCTROT>2.0.CO;2

Chester, F.M., and Chester, J.S., 1998, Ultracataclasite structure and friction processes of the

Punchbowl fault, San Andreas system, California: Tectonophysics, v. 295, p. 199-221,

https://doi.org/10.1016/S0040-1951(98)00121-8

Coffey, K.T., 2015, Oligocene-Miocene sedimentary and volcanic strata of the Vincent Gap

region, eastern San Gabriel Mountains, southern California, USA, and their tectonic

significance [M.S. thesis]: Los Angeles, University of California, 174 p.

53

Colasanti, C.V., and Ingersoll, R.V., 2006, Unroofing of the Pelona-Orocopia Schist and

palinspastic reconstruction of southern California: Geological Society of America

Abstracts with Programs, v. 38, no. 5, p. 19.

Cole, R.B., and Stanley, R.G., 1995, Middle Tertiary extension recorded by lacustrine fan-delta

deposits, Plush Ranch basin, Western Transverse Ranges, California: Journal of

Sedimentary Research, v. 65B, p. 455-468.

Crowell, J.C., 1954, Geology of the Ridge basin area, Los Angeles and Ventura Counties:

California Division of Mines Bulletin 170, Map Sheet 7, ~1:95000

Crowell, J.C., 1962, Displacement along the San Andreas fault, California: Geological Society of

America Special Paper 71, 61 p., https://doi.org/10.1130/SPE71-p1

Crowell, J.C., 1973, Problems concerning the San Andreas fault system in southern 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 Geological Science, v.

13, p. 125-135.

Crowell, J.C., 1975a, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Crowell, J.C., 1975b, Geologic sketch of the Orocopia Mountains, southeastern California, in

Crowell, J.C., ed., San Andreas fault in southern California: A guide to the San Andreas

fault from Mexico to Carrizo Plain: California Division of Mines and Geology Special

Report, v. 118, p. 99–110.

Crowell, J.C., 2003, Introduction to geology of Ridge Basin, southern California, in Crowell,

J.C., ed., Evolution of Ridge Basin, southern California: An interplay of sedimentation

54

and tectonics: Geological Society of America Special Paper 367, p. 1-15,

https://doi.org/10.1130/SPE367-p1

Crowell, J.C., and Susuki, T., 1959, Eocene stratigraphy and paleontology, Orocopia Mountains,

southeastern California: Geological Society of America Bulletin, v. 70, p. 581-592,

https://doi.org/10.1130/0016-7606(1959)70[581:ESAPOM]2.0.CO;2

Dalrymple, G.G., 1979, Critical tables for conversion of K-Ar ages from old to new constants:

Geology, v. 7, p. 558-560, https://doi.org/10.1130/0091-

7613(1979)7<558:CTFCOK>2.0.CO;2

Darin, M.H., and Dorsey, R.J., 2013, Reconciling disparate estimates of total offset on the

southern San Andreas fault: Geology, v. 41, p. 975-978,

https://doi.org/10.1130/G34276.1

Dibblee, T.W., Jr., 1967, Areal geology of the western , California: U.S.

Geological Survey Professional Paper, v. 522, 153 p., 2 plates, 1:125,000.

Dibblee, T.W., Jr., 1968, Displacements on the San Andreas fault system in the San Gabriel, San

Bernardino, and , Southern California in Dickinson, W.R., and

Grantz, A., eds., Proceedings of conference on geological problems of San Andreas fault

system: Stanford University Publications in Geological Sciences, v. 11, p. 260-278.

Dibblee, T.W., Jr., 1987, Geology of the Devil’s Punchbowl, Los Angeles County, California, in,

Hill, M.L., ed., Cordilleran Section of the Geological Society of America Centennial

Field Guide: Geological Society of America Centennial Field Guide, v. 1: p. 207-210.

Dibblee, T.W., Jr., 1996a, Geologic Map of the Mint Canyon quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-57, 1:24,000.

55

Dibblee, T.W., Jr., 1996b, Geologic Map of the Agua Dulce quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-58, 1:24,000.

Dibblee, T.W., Jr., 1997, Geologic Map of the Warm Springs Mountain quadrangle, Los Angeles

County, California: Dibblee Geological Foundation, Map DF-64, 1:24,000.

Dibblee, T.W., Jr., 2001, Geologic Map of the Pacifico Mountain and Palmdale quadrangles, Los

Angeles County, California: Dibblee Geological Foundation, Map DF-76, 1:24,000.

Dibblee, T.W., Jr., 2002a, Geologic Map of the Valyermo quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-80, 1:24,000.

Dibblee, T.W., Jr., 2002b, Geologic Map of the Crystal Lake quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-87, 1:24,000.

Dibblee, T.W., Jr., 2002c, Geologic Map of the Mount San Antonio quadrangle, Los Angeles

and San Bernardino Counties, California: Dibblee Geological Foundation, Map DF-88,

1:24,000.

Dibblee, T.W., Jr., 2005a, Geologic Map of the Ballinger Canyon quadrangle, San Luis Obispo,

Santa Barbara, Ventura and Kern Counties, California: Dibblee Geological Foundation,

Map DF-170, 1:24,000.

Dibblee, T.W., Jr., 2005b, Geologic Map of the Santiago Creek quadrangle, Kern and Ventura

Counties, California: Dibblee Geological Foundation, Map DF-171, 1:24,000.

Dibblee, T.W., Jr., 2006a, Geologic Map of the Cuddy Valley quadrangle, Ventura and Kern

Counties, California: Dibblee Geological Foundation, Map DF-199, 1:24,000.

Dibblee, T.W., Jr., 2006b, Geologic Map of the Apache Canyon quadrangle, Ventura and Kern

Counties, California: Dibblee Geological Foundation, Map DF-201, 1:24,000.

56

Dickinson, W.R., 1970, Interpreting detrital modes of graywacke and arkose: Journal of

Sedimentary Petrology, v. 40, p. 695-707.

Ebert, K.A., 2004, Exhumational history of the Orocopia Schist and development of hanging-

wall structures of the Orocopia fault, southeastern California [M.S. thesis]: Los Angeles,

University of California, 124 p.

Ehlert, K.W., 1982, Basin analysis of the Miocene Mint Canyon Formation, southern California,

in Ingersoll, R.V., and Woodburne, M.O., eds., Cenozoic nonmarine deposits of

California and Arizona: Los Angeles, Pacific Section, Society of Economic

Paleontologists and Mineralogists (SEPM), p. 51-64.

Ehlert, K.W., 2003, Tectonic significance of the middle Miocene Mint Canyon and Caliente

Formations, southern California, in Crowell, J.C., ed., Evolution of Ridge Basin, southern

California: An interplay of sedimentation and tectonics: Geological Society of America

Special Paper 367, p. 113-130, https://doi.org/10.1130/0-8137-2367-1.113

Ehlert, K.W., and Ehlig, P.L., 1977, The “Polka-dot” granite and the rate of displacement of the

San Andreas fault in southern California: Geological Society of America Abstracts with

Programs, v. 9, p. 415-416.

Ehlig, P.L., 1968, Causes of distribution of Pelona, Rand, and Orocopia Schists along the San

Andreas and Garlock faults, in Dickinson, W.R., and Grantz, A., eds., Proceedings of

conference on geological problems of San Andreas fault system: Stanford University

Publications in Geological Sciences, v. 11, p. 294-305.

Ehlig, P.L., 1981, Origin and tectonic history of the basement terrane of the San Gabriel

Mountains, central Transverse Ranges, in Ernst, W.G., ed., The Geotectonic

57

Development of California (Rubey Volume I): Englewood Cliffs, New Jersey, Prentice-

Hall, p. 253-283.

Ehlig, P.L., and Ehlert, K.W., 1972, Offset of Miocene Mint Canyon Formation from volcanic

source along San Andreas fault, Southern California: Geological Society of America

Abstracts with Programs, v. 4, p. 154.

Ehlig, P.L., and Joseph, S.E., 1977, Polka dot granite and correlation of La Panza quartz

monzonite with Cretaceous batholithic rocks north of , in Howell, D.G.,

Vedder, J.C., and McDougal, K.A., eds., Cretaceous geology of the California Coast

Ranges west of the San Andreas fault: Pacific Section, Society of Economic

Paleontologists and Mineralogists (SEPM) Field Guide 2, p. 91-96.

Ellis, B.J., Levi, S., and Yeats, R.S., 1993, Magnetic stratigraphy of the Morales Formation: late

Neogene clockwise rotation and compression in the Cuyama basin, California Coast

Ranges: Tectonics, v. 12, p. 1170-1179, https://doi.org/10.1029/93TC00314

Ensley, R.A., and Verosub, K.L., 1982, A magnetostratigraphic study of the sediments of the

Ridge Basin, southern California and its tectonic and sedimentologic implications: Earth

and Planetary Science Letters, v. 59, p. 192-207, https://doi.org/10.1016/0012-

821X(82)90125-X

Fisher, R.V., 1971, Features of coarse-grained, high-concentration fluids and their deposits:

Journal of Sedimentary Petrology, v. 41, p. 916-927.

Frizzell, V.A., Jr., and Weigand, P.W., 1993, Whole-rock K-Ar ages and geochemical data from

middle Cenozoic volcanic rocks, southern California: A test of correlations across the

San Andreas fault, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San

Andreas fault system: displacement, palinspastic reconstruction, and geologic evolution:

58

Geological Society of America Memoir 178, p. 273-288,

https://doi.org/10.1130/MEM178-p273

Frizzell, V.A., Jr., Mattison, J.M., and Matti, J.C., 1986, Distinctive Triassic megaporphyritic

monzogranite: Evidence for only 160 km offset along the San Andreas fault, southern

California: Journal of Geophysical Research, v. 91, p. 14,080-14,088,

https://doi.org/10.1029/JB091iB14p14080

Gabriel, A., and Cox, E.P., 1929, A staining method for the quantitative determination of certain

rock minerals: American Mineralogist, v. 14, p. 290-292.

Gazzi, P., 1966, Le arenarie del flysch sopracretaceo dell’Appennino modenese: correlazioni con

il flysch di Monghidoro: Mineralogica e Petrografia Acta, v. 12, p. 69-97.

Gehrels, G.E., Valencia, V., and Pullen, A., 2006, Detrital zircon geochronology by Laser-

Ablation Multicollector ICPMS at the Arizona LaserChron Center, in Olszewski, T., ed.,

Geochronology: Emerging Opportunities, Paleontological Society Short Course:

Paleontological Society Papers, v. 12, p. 67-76.

Gehrels, G.E., Valencia, V., and Ruiz, J., 2008, Enhanced precision, accuracy, efficiency, and

spatial resolution of U-Pb ages by laser ablation-multicollector-inductively coupled

plasma-mass spectrometry: Geochemistry, Geophysics, Geosystems, v. 9, p. 1-13,

https://doi.org/10.1029/2007GC001805

Goodmacher, J., Barnett, L., Buckner, G., Ouachrif, L., Vidigal, A., and Frost, E., 1989, The

Clemens Well fault in the Orocopia Mountains of southern California: a strike-slip or

normal fault structure?: Geological Society of America Abstracts with Programs, v. 21,

no. 5, p. 85.

59

Haxel, G., and Dillon, J., 1978, The Pelona-Orocopia Schist and Vincent-Chocolate Mountain

thrust system, southern California, in Howell, D.G., and McDougall, K.A., eds.,

Mesozoic paleogeography of the western United States: Pacific Section, Society of

Economic Paleontologists and Mineralogists (SEPM) Pacific Coast Paleogeography

Symposium 2, p. 453-469.

Hendrix, E.D., 1986, Sedimentology and basin analysis of the upper Oligocene Vasquez

Formation, Soledad basin, southern California [M.S. thesis]: Los Angeles, University of

California, 296 p.

Hendrix, E.D., 1993, Soledad Basin, Central Transverse Ranges, California: U.S. Geological

Survey Bulletin 2053, p. 243-350.

Hendrix, E.D., and Ingersoll, R.V., 1987, Tectonics and alluvial sedimentation of the upper

Oligocene/lower Miocene Vasquez Formation, Soledad Basin, southern California:

Geological Society of America Bulletin, v. 98, p. 647-663, https://doi.org/10.1130/0016-

7606(1987)98<647:TAASOT>2.0.CO;2

Hendrix, E.D., Cole, R.B., and Ingersoll, R.V., 2010, Soledad and Plush Ranch basins: mid-

Tertiary extensional terrane dismembered by the San Andreas fault system, in Clifton,

H.E., and Ingersoll, R.V., eds., Geologic excursions in California and Nevada: tectonics,

stratigraphy and hydrogeology: Pacific Section, SEPM (Society for Sedimentary Geology)

Book 108, p. 103-171.

Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults, California: a

study of the character, history, and tectonic significance of their displacements:

Geological Society of America Bulletin v. 64, p. 443-458, https://doi.org/10.1130/0016-

7606(1953)64[443:SAGABP]2.0.CO;2

60

Hornafius, J.S., Luyendyk, B.P., Terres, R.R., and Kamerling, M.J., 1986, Timing and extent of

Neogene tectonic rotation in the western Transverse Ranges, California: Geological

Society of America Bulletin, v. 97, p. 1476-1487, https://doi.org/10.1130/0016-

7606(1986)97<1476:TAEONT>2.0.CO;2

Hoyt, J.F., Coffey, K.T., Ingersoll, R.V., and Jacobson, C.E., 2018, Paleogeographic and

paleotectonic setting of the middle Miocene Mint Canyon and Caliente formations,

southern California: An integrated provenance study, in Ingersoll, R.V., Lawton, T.F, and

Graham, S.A., eds., Tectonics, sedimentary basins, and provenance: A celebration of the

career of William R. Dickinson: Geological Society of America Special Paper 540, p.

463-480, https://doi.org/10.1130/2018.2540(21)

Ingersoll, R.V., and Cavazza, W., 1991, Reconstruction of Oligo-Miocene volcaniclastic

dispersal patterns in north-central New Mexico using sandstone petrofacies, in Fisher,

R.V., and Smith, G.A., eds., Sedimentation in volcanic settings: SEPM (Society for

Sedimentary Geology) Special Publication 45, p. 227-236.

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Ingersoll, R.V., and Colasanti, C.V., 2004, Unroofing of the Pelona Schist, as documented by

sedimentologic evidence along the San Andreas and Punchbowl faults: Programs and

Abstracts, 2004 Annual Meeting, Pacific Sections, AAPG, SEPM and SEG, Bakersfield,

p. A13-A14.

61

Ingersoll, R.V., and Rumelhart, P.E., 1999, Three-stage evolution of the ,

southern California: Geology, v. 27, p. 593-596, https://doi.org/10.1130/0091-

7613(1999)027<0593:TSEOTL>2.3.CO;2

Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., and Sares, S.W., 1984, The

effect of grain size on detrital modes: a test of the Gazzi-Dickinson point-counting

method: Journal of Sedimentary Research, v. 54, p. 103-116,

https://doi.org/10.1306/212F83B9-2B24-11D7-8648000102C1865D

Ingersoll, R.V., Pratt, M.J., Davis, P.M., Caracciolo, L., Day, P.P., Hayne, P.O., Petrizzo, D.A.,

Gingrich, D.A., Cavazza, W., Critelli, S., Diamond, D.S., Coffey, K.T., Stang, D.M.,

Hoyt, J.F., Reith, R.C., and Hendrix, E.D., 2014, Paleotectonics of a complex Miocene

half graben formed above a detachment fault: the Diligencia basin, Orocopia Mountains,

southern California: Lithosphere, v. 6, p. 157-176, https://doi.org/10.1130/L334.1

Jacobson, C.E., 1983, Structural geology of the Pelona Schist and Vincent thrust, San Gabriel

Mountains, California: Geological Society of America Bulletin, v. 94, p. 753–767,

https://doi.org/10.1130/0016-7606(1983)94<753:SGOTPS>2.0.CO;2

Jacobson, C.E., Grove, M., Vućić, A., Pedrick, J.N., and Ebert, K.A., 2007, Exhumation of the

Orocopia Schist and associated rocks of southeastern California: Relative rates of

erosion, synsubduction tectonic denudation, and middle Cenozoic extension: Geological

Society of America Special Paper 419, p. 1-37, https://doi.org/10.1130/SPE419-p1

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, https://doi.org/10.1130/B30238.1

62

Jahns, R.H., 1939, Miocene stratigraphy of the easternmost Ventura Basin, California: a

preliminary statement: American Journal of Science, v. 237, p. 818-825,

https://doi.org/10.2475/ajs.237.11.818

Jahns, R.H., 1940, Stratigraphy of the easternmost Ventura Basin, California, with a description

of a new lower Miocene mammalian fauna from the Tick Canyon Formation, in

Henshaw, P.C., Wilson, R.W., Howard, H., Miller, A.H., Dougherty, J.F., and Jahns,

R.H., eds., Studies of Cenozoic vertebrates and stratigraphy of western North America:

Washington, D.C., Carnegie Institute of Washington Publication No. 514, p. 147-194.

Jahns, R.H., and Muehlberger, W.P., 1954, Geology of the Soledad Basin, Los Angeles County:

California Division of Mines Bulletin 170, v. 2, Map Sheet 6, 1:85,000.

Kellogg, K.S., and Miggins, D.P., 2002, Geologic map of the Sawmill Mountain quadrangle,

Kern and Ventura Counties, California: U.S. Geological Survey Open-File Report 02-

406, 1:24,000.

Kellogg, K.S., Minor, S.A., and Cossette, P.M., 2008, Geologic map of the eastern three-quarters

of the Cuyama 30’x60’ quadrangle, California: U.S. Geological Survey Scientific

Investigations Map 3002, 1:100,000.

Kew, W.S.W., 1923, Geologic formations of a part of southern California, and their correlation:

American Association of Petroleum Geologists Bulletin, v. 7, p. 411-420.

Kew, W.S.W., 1924, Geology and oil resources of a part of Los Angeles and Ventura Counties,

California: U.S. Geological Survey Bulletin 753, 202 p.

Konigsberg, R.L., 1967, Geology along the San Francisquito fault, Los Angeles County,

California [M.A. thesis]: Los Angeles, University of California, 84 p.

63

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, Society of Economic

Paleontologists and Mineralogists (SEPM), p. 53-61.

Lander, E.B., 1985, Early and middle Miocene continental vertebrate assemblages, central

Mojave desert, San Bernardino County, California, in Reynolds, R.E., compiler,

Geological investigations along Interstate 15, Cajon Pass to Manix Lake, California:

Redlands, San Bernardino County Museum, p. 127-144.

Law, R.D., Eriksson, K., and Davisson, C., 2001, Formation, evolution, and inversion of the

middle Tertiary Diligencia basin, Orocopia Mountains, southern California: Geological

Society of America Bulletin, v. 113, p. 196-221, https://doi.org/10.1130/0016-

7606(2001)113<0196:FEAIOT>2.0.CO;2

Link, M.H., 2003, Depositional systems and sedimentary facies of the Miocene-Pliocene Ridge

Basin Group, Ridge Basin, southern California, in Crowell, J.C., ed., Evolution of Ridge

Basin, southern California: An interplay of sedimentation and tectonics: Geological

Society of America Special Paper 367, p. 17-87, https://doi.org/10.1130/0-8137-2367-

1.17

Liu, W., 1990, Paleomagnetism of Miocene sedimentary rocks in the Transverse Ranges: the

implications for tectonic history [Ph.D. thesis]: Pasadena, California Institute of

Technology, 233 p.

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

64

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

May, D.J., and Walker, N.W., 1989, Late Cretaceous juxtaposition of metamorphic terranes in

the southeastern San Gabriel Mountains, California: Geological Society of America

Bulletin, v. 101, p. 1246-1267, https://doi.org/10.1130/0016-

7606(1989)101<1246:LCJOMT>2.3.CO;2

McClay, K., and Bonora, M., 2001, Analog models of restraining stepovers in strike-slip fault

systems: American Association of Petroleum Geologists Bulletin 85: 233-260,

https://doi.org/10.1306/8626C7AD-173B-11D7-8645000102C1865D

McDougall, K., 1982, Microfossil assemblages from the Castaic Formation, Ridge Basin,

southern California, in Crowell, J.C., and Link, M.H., eds., Geologic history of Ridge

Basin, southern California: Los Angeles, Pacific Section, Society of Economic

Paleontologists and Mineralogists (SEPM), p. 219-228.

Miller, W.S., 1934, Geology of the western San Gabriel Mountains of California: University of

California, Los Angeles, Publications in Math and Physical Sciences, v. 1, 113 p.,

1:84,480.

Muehlberger, W.R., 1958, Geology of northern Soledad basin, Los Angeles County, California:

American Association of Petroleum Geologists Bulletin, v. 42, p. 1,812-1,844.

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. 491-495,

https://doi.org/10.1130/0091-7613(1994)022<0491:MCROTW>2.3.CO;2

65

Noble, L.F., 1953, Geology of the Pearland quadrangle, California: U.S. Geological Survey

Geologic Quadrangle Map GQ-24, 1:24,000.

Noble, L.F., 1954, Geologic map of the Valyermo quadrangle and vicinity, California: U.S.

Geological Survey Geologic Quadrangle Map GQ-50, 1:24,000.

Nourse, J.A., 2002, Middle Miocene reconstruction of the central and eastern San Gabriel

Mountains, southern California, with implications for evolution of the San Gabriel fault

and Los Angeles basin: Geological Society of America Special Paper 365, p. 161-185,

https://doi.org/10.1130/0-8137-2365-5.161

Nourse, J.A., and Premo, W.R., 2016, New SHRIMP-RG U-Pb zircon and Sr analysis from

Jurassic and Late Cretaceous plutonic sheets in the east-central San Gabriel Mountains,

California: Geological Society of America Abstracts with Programs, v. 48, no. 4,

https://doi.org/10.1130/abs/2016CD-274654

Oskin, M., and Stock, J., 2003, Marine incursion synchronous with plate-boundary localization in

the Gulf of California: Geology, v. 31, p. 23-26, https://doi.org/10.1130/0091-

7613(2003)031<0023:MISWPB>2.0.CO;2

Oskin, M., Stock, J., and Martin-Barajas, A., 2001, Rapid localization of Pacific-North America

plate motion in the Gulf of California: Geology, v. 29, p. 459-462,

https://doi.org/10.1130/0091-7613(2001)029<0459:RLOPNA>2.0.CO;2

Powell, R.E., 1981, Geology of the crystalline basement complex, eastern Transverse Ranges,

southern California: constraints on regional tectonic interpretation [Ph.D. dissertation]:

Pasadena, California Institute of Technology, 441 p.

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

66

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

Premo, W.R., Nourse, J.A., Castineiras, P., and Kellogg, K., 2007, New SHRIMP-RG U-Pb

zircon ages and Sm-Nd analyses of Proterozoic metamorphic rocks of the San Gabriel

basement terrane: Keys for Laurentian crustal reconstruction: Ores and Orogenesis: A

Symposium Honoring the Career of William R. Dickinson, Program with Abstracts:

Tucson, Arizona, Arizona Geological Society, p. 150-151.

Reeder, S.W., and McAllister, A.L., 1957, A staining method for the quantitative determination

of feldspars in rocks and sands from soils: Canadian Journal of Soil Science, v. 37, p. 57-

59, https://doi.org/10.4141/cjss57-008

Richard, S.M., 1993, Palinspastic reconstruction of southeastern California and southwestern

Arizona for the middle Miocene: Tectonics, v. 12, p. 830-854.

Robinson, K.L., and Frost, E.G., 1996, Orocopia Mountains detachment system: progressive

development of a tilted crustal slab and half-graben sedimentary basin during regional

extension, in Abbott, P.L., and Cooper, J.D., eds., Field Conference Guide 1996: Pacific

Section, AAPG GS 73 and Pacific Section, SEPM Book 80, p. 277-284.

Sadler, P.M., 1993, The Santa Ana basin of the central San Bernardino Mountains: Evidence of

the timing of uplift and strike slip relative to the San Gabriel Mountains, in Powell, R.E.,

Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 307-321, https://doi.org/10.1130/MEM178-p307

67

Sams, R.H., 1964, Geology of the Charlie Canyon area, northwest Los Angeles County,

California [M.A. thesis]: Los Angeles, University of California, 101 p.

Schwade, I.T., 1954, Geology of and adjacent ranges, San Luis Obispo, Santa

Barbara, Kern and Ventura Counties: California Division of Mines Bulletin 170, Map

Sheet 1, ~1:194,000.

Sharp, R.P., 1935, Geology of the Ravenna quadrangle, Los Angeles County, California [M.S.

thesis]: Pasadena, California Institute of Technology, 81 p.

Sharp, R.V., and Silver, L.T., 1971, Quaternary displacement on the San Andreas and

Punchbowl faults at the San Gabriel Mountains, southern California: Geological Society

of America Abstracts with Programs, v. 3, p. 191.

Silver, L.T., 1966, Preliminary history of the crystalline complex of the central Transverse

Ranges, Los Angeles County, California, in Abstracts of papers submitted for six

meetings with which the Society was associated: Geological Society of America Special

Paper 101, p. 201-202.

Silver, L.T., 1971, Problems of crystalline rocks of the Transverse Ranges: Geological Society of

America Abstracts with Programs, v. 3, p. 193-194.

Silver, L.T., McKinney, C.R., Deutsch, S., and Bolinger, J., 1963, Precambrian age

determination in the western San Gabriel Mountains, California: Journal of Geology, v.

71, p. 196-214, https://doi.org/10.1086/626893

Smith, D.P., 1977, San Juan–St Francis fault—Hypothesized major Middle Tertiary right-lateral

fault in central and southern California, in Short contributions to California geology:

California Division of Mines and Geology Special Report 129, p. 41–50.

68

Spittler, T.E., 1974, Volcanic petrology and stratigraphy of non-marine strata, Orocopia

Mountains: their bearing on Neogene slip on the San Andreas fault, southern California

[M.S. thesis]: Riverside, University of California, 115 p.

Spittler, T.E., and Arthur, M.A., 1982, The lower Miocene Diligencia Formation of the Orocopia

Mountains, southern California: Stratigraphy, petrology, sedimentology, and structure, in

Ingersoll, R.V., and Woodburne, M.O., eds., Cenozoic nonmarine deposits of California

and Arizona: Los Angeles, Pacific Section, Society of Economic Paleontologists and

Mineralogists (SEPM), p. 83-89.

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. 207-211,

https://doi.org/10.1016/0012-821X(75)90088-6

Stanley, K.O., 1966, The structural history of the Clearwater fault, northwest Los Angeles

County, California [M.A. thesis]: Los Angeles, University of California, 73 p.

Stirton, R.A., 1933, Critical review of the Mint Canyon mammalian fauna and its correlative

significance: American Journal of Science, v. 26, p. 569-576.

Stock, C., 1947, A peculiar new carnivore from the Cuyama Miocene, California: Southern

California Academy of Science Bulletin, v. 46, p. 84-89.

Streckeisen, A., 1974, Classification and nomenclature of plutonic rocks: recommendations of

the IUGS subcommission on the systematics of igneous rocks: Geologische Rundschau,

v. 63, p. 773-786, https://doi.org/10.1007/BF01820841

Sylvester, A.G., 1988, Strike-slip faults: Geological Society of America Bulletin, v. 100, p.

1666-1703, https://doi.org/10.1130/0016-7606(1988)100<1666:SSF>2.3.CO;2

69

Tedford, R.H., and Downs, T., 1965, Age of the Punchbowl Formation, Los Angeles and San

Bernardino Counties, California: Geological Society of America Special Paper 87, p.

234, https://doi.org/10.1130/SPE87

Tennyson, M.E., 1989, Pre-transform early Miocene extension in western California: Geology, v.

17, p. 792-796, https://doi.org/10.1130/0091-7613(1989)017<0792:PTEMEI>2.3.CO;2

Terres, R.R., 1984, Paleomagnetism and tectonics of the central and eastern Transverse Ranges,

southern California [Ph.D. thesis]: Santa Barbara, University of California, 323 p.

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Weber, F.H., Jr., 1994, Geologic map of southeast-central Warm Springs Mountain 7.5-minute

quadrangle, Los Angeles County, California: California Department of Conservation,

Division of Mines and Geology, Open File Report 93-04, 23 p., 1:12,000.

Weigand, P.W., 1982, Middle Cenozoic volcanism of the Western Transverse Ranges, in Fife,

D.L., and Minch, J.A., eds., Geology and mineral wealth of the California Transverse

Ranges: Santa Ana, South Coast Geological Society Annual Symposium and Guidebook

10, p. 170-188.

Weldon, R.J., II, Meisling, K.E., and Alexander, J., 1993, A speculative history of the San

Andreas fault in the central Transverse Ranges, California, in Powell, R.E., Weldon, R.J.,

II, and Matti, J.C., eds., The San Andreas fault system: displacement, palinspastic

reconstruction, and geologic evolution: Geological Society of America Memoir 178, p.

161-198, https://doi.org/10.1130/MEM178-p161

70

Wilcox, R.E., Harding, T.P, and Seely, D.R., 1973, Basic wrench tectonics: Bulletin of the

American Association of Petroleum Geologists, v. 57, p. 74-96.

Woodburne, M.O., 1975, Cenozoic stratigraphy of the Transverse Ranges and adjacent areas,

southern California: Geological Society of America Special Paper 162, 91 p.,

https://doi.org/10.1130/SPE162

Woodburne, M.O., 1987, A prospectus of the North American Mammal Ages, in Woodburne,

M.O., ed., Cenozoic Mammals of North America: Berkeley, University of California

Press, p. 285-290.

Woodburne, M.O., and Golz, D.J., 1972, Stratigraphy of the Punchbowl Formation, Cajon

Valley, Southern California: University of California Publications in Geological

Sciences, v. 92, 73 p.

Woodburne, M.O., and Whistler, D.P., 1973, An early Miocene oreodont (Merychyinae,

Mammalia) from the Orocopia Mountains, Southern California: Journal of Paleontology,

v. 47, p. 908-912.

Yan, Z., Clayton, R.W., and Saleeby, J., 2005, Seismic refraction evidence for steep faults

cutting highly attenuated continental basement in the central Transverse Ranges,

California: Geophysical Journal International, v. 160, p. 651-666,

https://doi.org/10.1111/j.1365-246X.2005.02506.x

Yarnold, J.C., 1993, Rock-avalanche characteristics in dry climates and the effect of flow into

lakes: Insights from mid-Tertiary sedimentary breccias near Artillery Peak, Arizona:

Geological Society of America Bulletin, v. 105, p. 345-360, https://doi.org/10.1130/0016-

7606(1993)105<0345:RACIDC>2.3.CO;2

71

Yarnold, J.C., and Lombard, J.P., 1989, Facies model for large rock-avalanche deposits formed

in dry climates, in Colburn, I.P., Abbott, P.L., and Minch, J., eds., Conglomerates in

basin analysis: a symposium dedicated to A.O. Woodford: Bakersfield, Pacific Section,

Society of Economic Paleontologists and Mineralogists (SEPM), p. 9-31.

72

CHAPTER 2: RECONCILING STRUCTURAL AND

PALEOMAGNETIC DATA TO CONSTRAIN EVOLUTION

OF THE CHOCOLATE MOUNTAINS ANTICLINORIUM

ALONG THE SAN ANDREAS FAULT SYSTEM,

SOUTHERN CALIFORNIA, U.S.A.

This chapter has been submitted to the journal Tectonics and, as of 2019 July, is in review:

Coffey, K.T., in review, Reconciling structural and paleomagnetic data to constrain evolution of the Chocolate Mountains anticlinorium along the San Andreas fault system, southern California,

U.S.A.: Tectonics.

ABSTRACT

The Chocolate Mountains anticlinorium, cored by schist emplaced beneath North

America via subduction, extends from southern California into southwestern Arizona, and provides important constraints on regional tectonic events of the Cenozoic. Along the anticlinorium, the orientations of penetrative lineations within the schist, if interpreted as having formed parallel to northeastward subduction, preclude significant vertical-axis rotations, whereas paleomagnetic declinations in nearby volcanic rocks imply them. These two possibilities suggest substantially different original architectures of the anticlinorium and subsequent histories of deformation. Sierra Pelona is a probable offset segment of the Chocolate Mountains anticlinorium, with nearby volcanic rocks from which paleomagnetic data have been reported.

The spatial distribution of types and orientations of lineations at Sierra Pelona matches that of 73 schist exposures in the western Chocolate Mountains anticlinorium, supporting their correlation.

As in other segments of the anticlinorium, the orientation of lineations appears to preclude substantial vertical-axis rotation, whereas the nearby paleomagnetic data imply it. A recent attempt at reconciling these datasets proposed that rotation of volcanic rocks was decoupled from underlying schist. However, the relative timing of these rotations and exhumation of the underlying schist is problematic for this model, and recent kinematic measurements call into question the assumption that prograde lineations formed via subduction underthrusting. Adding my results to existing structural and paleomagnetic data suggests an alternative way to reconcile the two datasets: prograde lineations in the schist along the western part of the anticlinorium formed or were reoriented in a NNW/SSE orientation, and were then rotated clockwise to achieve their present orientations.

INTRODUCTION

The Chocolate Mountains anticlinorium (herein, “the anticlinorium”; Fig. 2-1A) is a structural feature with important implications for regional tectonic evolution through the

Cenozoic. Structures both within and bounding the schist at its core document deformation interpreted as associated with early Cenozoic denudation (e.g., Jacobson et al., 2002, 2007; Yin,

2002). Normal faults and sedimentary and volcanic deposits along the flanks of the anticlinorium record details of latest Oligocene to middle Miocene extension and associated magmatism (e.g.,

Spittler and Arthur, 1982; Hendrix and Ingersoll, 1987; Sherrod and Tosdal, 1991; Ingersoll et al., 2014; Coffey et al., 2019). The anticlinorium itself likely formed during this extension, in the early to middle Miocene (e.g., Hendrix and Ingersoll, 1987; Sherrod and Tosdale, 1991;

Richards, 1993; Jacobson et al., 2007; Ricketts et al., 2011; but see Yin, 2002). Lastly, the

74

Figure 2-1: A: Regional map, showing

present locations and schematic extents of

exposures of Pelona, Orocopia, Rand, and

related schists (Sierra de Salinas is outside

map area), and relevant faults. Areas of

Figs. 2-3, 2-5 and 2-7 indicated. Figure

modified from Haxel et al. (2002). B:

Schematic restoration of A, after reversing

slip on strands of San Andreas fault system

and Garlock fault. No vertical-axis rotations

have been restored, except 16° of

counterclockwise rotation of San Gabriel

block (between San Gabriel and San

Andreas faults), and 60° of clockwise

rotation of eastern Mountains

(see discussion in text). This is the approximate architecture of the Chocolate Mountains anticlinorium implied by the orientations of

penetrative lineations within schist, assuming that these lineations formed parallel to the former subduction direction. Restoration modified from Haxel et al. (2018). Areas of Fig. 2-9 indicated. C:

Schematic restoration, as in B, but with schist exposures corrected for all vertical-axis rotations

implied by paleomagnetic data, and schematic restoration of hypothesized related strike-slip faulting and extension caused by Basin and Range normal faulting east of the Colorado River. This

is the approximate architecture of the Chocolate Mountains anticlinorium implied by paleomagnetic data. Areas of Fig. 2-10 indicated. Restoration modified from those of Ingersoll et al.

(2014), Bennett et al. (2016), and Haxel et al. (2018).

75

anticlinorium is widely accepted as being offset by, and thus documenting the cumulative slip on, strands of the San Andreas fault system (Hill and Dibblee, 1953; Crowell, 1962, 1975), active since the middle Miocene (e.g., Nicholson et al., 1994; Ingersoll and Rumelhart, 1999;

Oskin et al., 2001; Crowell, 2003; Oskin and Stock, 2003). The original architecture of the anticlinorium, however, is debated (e.g., Haxel et al., 2002, 2018; Bennett et al., 2016; Grove and Jacobson, 2016). If segments of the anticlinorium were originally oriented differently than at present, then the orientations of the structures listed above must be corrected for these changes before their tectonic implications can be properly inferred. Furthermore, if the anticlinorium has bent during its separation along strands of the San Andreas fault system, then this bending must be included with the amounts of separation along these strands to determine the full extent of

Pacific-North America transform motion recorded by the anticlinorium. In this paper, I present new lineation measurements from offset segments of the anticlinorium, then compare orientations of lineations within the anticlinorium with paleomagnetic declinations along its flanks. I consider how apparent disagreement of these datasets can best be reconciled, and the resulting implications for the original architecture of the anticlinorium and evolution of the San

Andreas fault system.

GEOLOGIC BACKGROUND

The Chocolate Mountains Anticlinorium

The Pelona, Orocopia, Rand, and related schists (herein, “the schist”) are correlated units generally interpreted as supra-crustal material of the Farallon plate underplated to the base of

North America during flat-slab subduction in the Late Cretaceous and early Paleogene (e.g.,

76

Yeats, 1968; Burchfiel and Davis, 1981; Jacobson, 1983a; Jacobson et al., 2007, 2011; Haxel et al., 2015; Chapman, 2017; but see Haxel and Dillon, 1978; Barth and Schneiderman, 1996;

Haxel et al., 2002; Chapman, 2017). The schist is exposed at the surface in various parts of southern California and southwestern Arizona, including in the cores of elongate structural domes that define a belt referred to as the Chocolate Mountains anticlinorium (e.g., Haxel and

Dillon, 1978; Fig. 2-1). Exposures of the schist at Mount Pinos, Sierra Pelona, and Blue Ridge are thought to be segments of the anticlinorium displaced along the San Andreas, Punchbowl, and San Gabriel faults; some reconstructions restore these segments directly adjacent to the rest of the anticlinorium (e.g., Hill and Dibblee, 1953; Crowell, 1962, 1975; Dibblee, 1967, 1968;

Ehlig, 1968; this study; Fig. 2-1B, C), whereas others restore them approximately 60-80 km farther northwest (e.g., Smith, 1977; Powell, 1981, 1993; Matti and Morton, 1993).

The present trend of the anticlinorium is substantially curved (Fig. 2-1A). This could represent either the original shape of the anticlinorium (e.g., Haxel et al., 2018; Fig. 2-1B), or deformation of an originally linear trend (e.g., Ingersoll et al., 2014; Bennett et al., 2016; Grove and Jacobson, 2016; Fig. 2-1C). In the latter case, vertical-axis clockwise rotations of variable magnitudes would be required to explain present orientations of the anticlinorium fold axis and the belts of sedimentary and volcanic strata along the flanks of the anticlinorium (e.g., Ingersoll et al., 2014; Bennett et al., 2016; c.f. geologic maps in Dibblee, 1997a, b; Haxel et al., 2002;

Ricketts et al., 2011; Ingersoll et al., 2014).

Penetrative Lineations

The schist contains penetrative lineations, some of which have been interpreted as associated with prograde metamorphism (e.g., Jacobson, 1983a; Jacobson and Dawson, 1995;

Chapman et al., 2010). Prograde metamorphism is generally thought to have occurred during

77 subduction underthrusting and underplating of the schist (e.g., Jacobson and Dawson, 1995;

Haxel et al., 2018) during the late Cretaceous and early Paleogene (Grove et al., 2003). In many exposures of the schist, including along the anticlinorium, penetrative lineations are presently oriented approximately NE/SW (Harvill, 1969; Jacobson and Dawson, 1995; Oyarzabal et al.,

1997; Jacobson et al, 2002). This is compatible with lineation development via simple shear caused by subduction underthrusting, as the direction of Farallon-plate subduction relative to

North America was approximately northeastward during this time (Engebretson et al., 1985;

Stock and Molnar, 1988; Doubrovine and Tarduno, 2008). If this was the developmental mechanism of the penetrative lineations, and their orientation was not modified as subduction continued, then substantial vertical-axis rotations of segments of the anticlinorium are precluded, as restoration of these rotations would restore these penetrative lineations to orientations no longer parallel with the subduction direction (Haxel et al., 2018). The present, bent shape of the anticlinorium would thus represent its original architecture (Fig. 2-1B).

Paleomagnetic Declinations

Paleomagnetic declinations determined in ca. 25-21 Ma volcanic rocks along the flanks of the anticlinorium at Sierra Pelona, the Orocopia Mountains, and the Gavilan Hills (Fig. 2-1) imply variable clockwise vertical-axis rotations (Terres, 1984; Costello, 1985; Terres and

Luyendyk, 1985), correction for which would align these segments in a consistent, approximately NE/SW orientation. This suggests that the present, bent shape of the anticlinorium represents deformation of an originally linear architecture (Fig. 2-1C). Thus, penetrative lineations, if assumed to have formed and remained parallel to the direction of subduction, and paleomagnetic data appear contradictory.

78

The significance of some of the paleomagnetic data reported from along the anticlinorium can be argued, because, in most locations, too few flows were sampled to adequately average secular variation (e.g., Butterworth, 1984; Calderone and Butler, 1984; Costello, 1985; Terres and Luyendyk, 1985; Viseth, 1985; Calderone et al., 1990). The relevance of some of the paleomagnetic data to the anticlinorium is also debatable: in the Orocopia Mountains, for example, the volcanic flows sampled are separated from the anticlinorium by the Clemens Well fault, which could potentially have accommodated differential vertical-axis rotation on either side, especially if it is a major (80-110 km of dextral slip) Miocene strike-slip fault, as some models have argued (Powell, 1981, 1993; but see Goodmacher et al., 1989; Robinson and Frost,

1996; Ebert, 2004; Jacobson et al., 2007; Ingersoll et al., 2014; Coffey et al., 2019).

METHODS

I searched for penetrative lineations and crenulation lineations within outcrops of the schist at Mount Pinos, Sierra Pelona, and the Crafton Hills (Fig. 2-1), mostly along roadcuts.

Penetrative lineations were defined, as in other segments of the anticlinorium (e.g., Jacobson,

1983a; Jacobson and Dawson, 1995; Oyarzabal et al., 1997), by ductile mineral stretching of mica, feldspar, and quartz (common; Fig. 2-2A, B), preferred alignment of elongate minerals, and intersection lineation of compositional layering with metamorphic foliation (rare).

Crenulation lineations were defined, as in other segments of the anticlinorium (e.g., Jacobson and Dawson, 1995; Oyarzabal et al., 1997), by weakly to moderately developed crenulations, manifested in the study areas as mm-scale crinkling of the foliation surface (Fig. 2-2C).

Slickenlines were locally present, but ignored. Where penetrative and/or crenulation lineations were discernable, and outcrop appeared in-place and free of structural complexity on the outcrop

79

Figure 2-2: Field photos showing lineations on foliation surfaces exposed at Sierra Pelona: A: penetrative lineations within mica schist of Pelona Schist. B: penetrative lineations within quartzite

of Pelona Schist. C: crenulation lineations within mica schist of Pelona Schist. D: penetrative lineations within mylonitic gneiss structurally overlying the Pelona Schist along the Vincent fault.

Mechanical pencil for scale.

scale (approximately 1-10 m), I recorded the type of lineation, and measured the location (with a handheld GPS), and orientation of foliation and rake of lineation (with a Brunton pocket transit).

In the Crafton Hills and the few places at Sierra Pelona and Mount Pinos where I encountered mylonitic gneiss structurally above the schist, I took the same measurements of its penetrative lineations (Fig. 2-2D). I conducted this fieldwork in 2018, January to October.

Arching of the anticlinorium postdated development of the penetrative lineations, and either post-dated or was coeval with development of the crenulation lineations (e.g., Jacobson and Dawson, 1995; Jacobson et al., 2002; Ricketts et al., 2011). Thus, when these lineations

80 formed, the foliation of the schist was planar. Foliation within the schist is subparallel to both originally low-angle faults and to bedding of Oligocene-Miocene strata in the hanging walls of these faults (e.g., Ehlig, 1981; Dibblee, 1997a, b; Jacobson et al., 2002, 2007; Ricketts et al.,

2011; Coffey et al., 2019), implying that it was originally subhorizontal. Accordingly, I corrected lineation orientations for tilting: for each measurement location, I untilted orientations of the foliation and lineation(s) to horizontal (i.e., rotated both about an axis parallel to foliation strike by an angle equal to foliation dip). I performed these calculations using Stereonet 9

(Allmendinger et al., 2012; Cardozo and Allmendinger, 2013). Because lineations are bidirectional, I calculated the overall trend of tilt-corrected lineations for each region with

Bingham axial distributions, using Stereonet 9.

RESULTS

Mount Pinos

Foliation in the schist near Mount Pinos generally dips southward (Figs. 2-3A, 2-4A), away from the San Andreas fault, as documented by Kellogg and Miggins (2002); no clear anticlinorium fold axis is defined. Crenulation lineations are not well developed in the schist near

Mount Pinos (Figs. 2-3B, 2-4C).

Tilt-corrected penetrative lineations are dominantly WNW/ESE through most of the schist exposure (Figs. 2-3B, 2-4E), but at the western and eastern ends, the dominant orientation is NW/SE to NNW/SSE (Figs. 2-3B, 2-4D, F). This coincides with the low-angle segments of the Sawmill Mountain fault (as mapped by Kellogg and Miggins, 2002). The several lineation measurements in the schist near Mount Pinos reported by Kellogg and Miggins (2002) have the same range of orientations, and are generally compatible with this spatial variability. The two

81

Figure 2-3: Map of Mount Pinos area, showing measurements of A: foliation within schist (black)

and mylonitic gneiss (red), and B: tilt-corrected penetrative (black) and crenulation (blue) lineations within schist, and penetrative lineations (red) within mylonitic gneiss. Schist is divided into three regions (W, M, and E) by gray lines, and primary axis of Bingham axial distribution of penetrative lineations within each region is shown as a heavy black line (lineations are plotted by region in Fig. 2-4). Area of figure shown in Fig. 2-1A. Geology from Kellogg and Miggins (2002).

82

Figure 2-4: Foliation and lineation data from Mount Pinos area. A-C: lower-hemisphere, equal-

area projections. A: poles to foliation (i.e., schistosity). B: present (not tilt-corrected) penetrative lineations. C: present (not tilt-corrected) crenulation lineations. D-F: rose diagrams of penetrative-

lineation data by region (see Fig. 2-3), after tilt correction. Primary axis of Bingham axial distribution shown on edge of each diagram in black; that of crenulation-lineation data of the same

region shown in blue for comparison. D: region W. E: region M. F: region E.

lineation measurements that I made in mylonitic gneiss along the Sawmill Mountain fault have tilt-corrected NNW/SSE orientations, subparallel to the nearest penetrative lineations measured within the schist (Fig. 2-3B). Original foliation and lineation measurements for the schist and mylonitic gneiss near Mount Pinos are in Table S2-1.

Sierra Pelona

Foliation in the schist of Sierra Pelona generally dips NNW and SSE (Figs 2-5A, 2-6A), defining the approximately ENE/WSW-trending fold axis of the anticlinorium, as documented by Jahns and Muehlberger (1954), Muehlberger and Hill (1958), and Dibblee (1997a, b). Present orientations of both penetrative (Fig. 2-6B) and crenulation lineations (Fig. 2-6C) are diverse; tilt

83

84

Figure 2-5 (previous page): Map of Sierra Pelona, showing measurements of A: foliation within schist (black) and mylonitic gneiss (red), and B: tilt-corrected penetrative (black) and crenulation

(blue) lineations within schist, and penetrative lineations (red) within mylonitic gneiss. Schist is divided into five regions (NW, SW, NE, SE, and MSZ, the last being a mylonitic shear zone) by gray lines, and primary axis of Bingham axial distribution of penetrative lineations within each region is

shown as a heavy black line (lineations are plotted by region in Fig. 2-6). Area of figure shown in

Fig. 2-1A. Geology from Dibblee (1997a, b).

Figure 2-6: Foliation and lineation data from schist of Sierra Pelona. A-C: lower-hemisphere,

equal-area projections. A: poles to foliation (i.e., schistosity), with overall trend and plunge of

anticlinorium fold axis calculated from these data plotted as large square. B: present (not tilt-

corrected) penetrative lineations. C: present (not tilt-corrected) crenulation lineations. D-G: rose diagrams of penetrative-lineation data by region (see Fig. 2-5), after tilt correction. Primary axis of

Bingham axial distribution shown on edge of each diagram in black; that of crenulation-lineation data of the same region shown in blue for comparison. D: region SW. E: region NW. F: region SE.

G: region NE. 85 correction of lineation orientations improves clustering of these data, but substantial variability remains.

Tilt-corrected penetrative lineations are dominantly NNE/SSW through much of the schist exposure (Figs. 2-5B, 2-6D), but along its northern edge, they are deflected clockwise, to an ENE/WSW to E/W dominant orientation, with increased scatter (Figs. 2-5B, 2-6E).

Orientations also deviate from NNE/SSW within a mylonitic shear zone (region “MSZ” in Fig.

2-5B). Tilt-corrected penetrative lineations in the eastern half of Sierra Pelona, both away from

(Fig. 2-6F) and along (Fig. 2-6G) its northern edge, are generally about 20°-30° clockwise of those in equivalent positions in the western half (Figs. 2-6D and 2-6E, respectively). These measurements corroborate and expand upon those of Harvill (1969), which were collected in the western part of Sierra Pelona, and mostly concentrated within several of the quartzite bodies found within the schist there.

Tilt-corrected crenulation lineations, which are more abundant along the northern edge of the schist exposure, are generally subparallel to tilt-corrected penetrative lineations, showing similar clockwise deflections along the northern edge of the schist exposure, and in the eastern vs. western half (Figs. 2-5B, 2-6D-G). This is consistent with the few previously reported orientations of lineations defined by “wrinkling in the plane of foliation” (Muehlberger and Hill,

1958, p. 641). The few lineations that I measured in mylonitic gneiss along the Vincent fault are subparallel to nearby penetrative lineations within the schist (Fig. 2-5B). Original foliation and lineation measurements for the schist and mylonitic gneiss of Sierra Pelona are in Table S2-2.

Crafton Hills

Foliation in the schist and mylonitic gneiss of the Crafton Hills generally dips southward, but with substantial local variation (Figs. 2-7A, 2-8A), as documented by Matti et al. (2003). Due

86

Figure 2-7: Map of northeastern Crafton Hills, showing measurements of A: foliation within schist

(black) and mylonitic gneiss (red), and B: tilt-corrected penetrative (black) and crenulation (blue) lineations within schist, and penetrative lineations (red) within mylonitic gneiss. Area is divided into

three regions, one of schist (NE) and two of mylonitic gneiss (M, E; boundary between shown as

white line), and primary axis of Bingham axial distribution of penetrative lineations within each

region is shown as a heavy red (mylonitic gneiss) or black (schist) line (lineations are plotted by

region in Fig. 2-8). Area of figure shown in Fig. 2-1A. Geology from Matti et al. (2003).

to limited access and exposure, most of my measurements are of penetrative lineations in the mylonitic gneiss, with only three measurements each of penetrative and crenulation lineations in the schist.

87

Tilt-corrected penetrative lineations are NE/SW through most of the mylonitic gneiss exposure (Figs. 2-7B, 2-8D), transitioning to E/W near the eastern edge (Figs. 2-7B, 2-8E). The few penetrative and crenulation lineations measured in the schist, near the eastern edge of its exposure, are NE/SW (Figs. 2-7B, 2-8F). The several lineation measurements in the schist and mylonitic gneiss reported by Matti et al. (2003) closely match these patterns. Original foliation and lineation measurements for the mylonitic gneiss and schist at the Crafton Hills are in Table

S2-3.

Figure 2-8: Foliation and lineation data from northeastern Crafton Hills. A-C: lower-hemisphere,

equal-area projections. A: poles to foliation (both schistosity of schist and foliation of mylonitic gneiss). B: present (not tilt-corrected) penetrative lineations within mylonitic gneiss. C: present (not

tilt-corrected) penetrative and crenulation lineations within schist. D-F: rose diagrams of

penetrative-lineation data by region (see Fig. 2-7), after tilt correction. Primary axis of Bingham axial distribution shown on edge of each diagram in black; that of crenulation-lineation data of the

same region shown in blue for comparison in F. D: region M (mylonitic gneiss). E: region E

(mylonitic gneiss). F: region NE (schist).

88

DISCUSSION

My data from Sierra Pelona support the longstanding interpretation that Sierra Pelona is an offset segment of the anticlinorium. The schist of Sierra Pelona contains the same lineation types, and exhibits along its northern margin the same clockwise deflection of lineations, greater scatter in orientations, and increased abundance of crenulation lineations as the schist of the anticlinorium in the Orocopia Mountains and Gavilan Hills, where they have been attributed to shear adjacent to post-subduction faults (Jacobson and Dawson, 1995; Oyarzabal et al., 1997;

Jacobson et al., 2002). Clockwise deflection of lineation orientations in the eastern half of Sierra

Pelona relative to the western half is likely due to the bend of comparable angle (approximately

20°) in this body of schist (Fig. 2-5; the anticlinorium axial trace is also bent, but to a lesser extent). Some of the localized anomalous lineation orientations not explained by the spatial trends discussed above likely represent localized structural complexities and misidentification of slide blocks as true outcrop. Both were more difficult to avoid in the eastern half of Sierra

Pelona, where outcrops are typically fewer and smaller than in the western half, potentially explaining the greater variability of these data (Fig. 2-6F, 2-G vs. Fig. 2-6D, E).

Present (tilt-corrected) orientations of penetrative lineations away from the northern margin of schist exposure are comparable with the approximately northeastward estimated subduction direction of the Farallon plate during schist underplating (Engebetson et al., 1985;

Stock and Molnar, 1988; Doubrovine and Tarduno, 2008; this observation was previously made by Haxel et al., 2018, based on data of Harvill, 1969). Therefore, if penetrative lineations are interpreted to have formed parallel to the direction of subduction, and not undergone reorientation during continuing subduction, then the schist of Sierra Pelona cannot have experienced substantial vertical-axis rotations since its emplacement in the early Paleogene.

89

However, paleomagnetic data from volcanic rocks approximately 5-10 km south of the schist of

Sierra Pelona indicate 37° ± 12° of net clockwise vertical-axis rotation since ca. 25 Ma (Terres and Luyendyk, 1985; Frizzell and Weigand, 1993; Coffey, 2015; Fig. 2- 9). Slip magnitudes on surrounding, post-ca. 25 Ma sinistral faults are compatible with this amount of clockwise rotation (Luyendyk et al., 1980; Terres and Luyendyk, 1985), supporting its validity. Therefore, as along other segments of the anticlinorium, if penetrative lineations are interpreted to have formed parallel to the direction of subduction, they appear to contradict nearby paleomagnetic data. Reconciliation of these datasets is necessary.

Decoupling-Surface Model

Haxel et al. (2018) proposed that vertical-axis rotations of late Oligocene-early Miocene volcanic rocks implied by paleomagnetic data were not experienced by the schist, which contains the penetrative lineations. Haxel et al. (2018) suggested that the originally low-angle faults along which the schist is juxtaposed beneath North America basement acted as the decoupling surface, an expansion of an idea proposed by Terres and Luyendyk (1985) for the Vincent fault, which separates the schist from North America basement in the San Gabriel Mountains (e.g., Ehlig,

1958; Xia and Platt, 2018). This model predicts an originally curved architecture for the anticlinorium, with penetrative lineations oriented NE/SW across all schist exposures (Fig. 2-

1B), and provides a simple solution to the apparent disagreement of structural and paleomagnetic data. The relative timing of disruption of the decoupling surface and rotation of the volcanic rocks, however, may be problematic.

Through much of the anticlinorium, low-angle faults between the schist and upper

Oligocene-lower Miocene volcanic rocks are interpreted as detachment faults, with movement constrained to the mid-Cenozoic (e.g., Jacobson and Dawson, 1995; Oyarzabal et al., 1997;

90

Haxel et al., 2002; Jacobson et al., 2002, 2007; Ricketts et al., 2011). In their initial geometry, such faults could readily have acted as decoupling surfaces. However, at least at Mount Pinos,

Sierra Pelona, Blue Ridge, and the Orocopia Mountains, sedimentologic data indicate that the topographic highs along which the schist of the anticlinorium is presently exposed were established and actively uplifting ca. 25-23 Ma (e.g., Hendrix and Ingersoll, 1987; Cole and

Stanley, 1995; Hendrix et al., 2010; Ingersoll et al., 2014; Coffey et al., 2019), when the nearby volcanic rocks were erupted (Frizzell and Weigand, 1993; Coffey, 2015). The schist was likely being uplifted in the subsurface of these highs at this time, and possibly arching to form the anticlinorium (e.g., Hendrix, 1993; Ingersoll et al., 2014). This interpretation is supported by thermochronometric data: K-feldspar 40Ar/39Ar dating of the schist in all of these regions, as well as in the Gavilan Hills, indicates cooling ca. 28-22 Ma (Jacobson et al., 2002, 2007; Grove and

Jacobson, 2016), and apatite (U-Th)/He dating of granitic rock structurally overlying the schist at

Mount Pinos similarly implies that rapid cooling was occurring by ca. 22 Ma (Spotila et al.,

2007; Niemi et al., 2013). The schist first appeared as clasts in the local sedimentary record ca.

17.2 Ma (Ehlert, 2000; Prothero et al., 2008; Hoyt et al., 2018). Exhumation and/or arching of low-angle faults above the schist should have inactivated them, and their ability to serve as a decoupling surface, during or shortly after eruption of the volcanic rocks ca. 25-23 Ma, and certainly before 17.2 Ma. This leaves little to no time for the volcanic rocks to have undergone vertical-axis rotation independent of the schist. Timing is particularly problematic at Sierra

Pelona, where the clockwise vertical-axis rotation implied by paleomagnetism (Terres and

Luyendyk, 1985) was probably accommodated by slip on a system of sinistral faults (Luyendyk et al., 1980; Terres and Luyendyk, 1985), which offset strata containing detritus derived from the schist of Sierra Pelona (e.g., Dibblee, 1996; Ehlert, 2000; Hoyt et al., 2018). This implies that the

91 schist was exposed and eroding during the time of vertical-axis rotation, rather than at depth and decoupled from overlying rocks.

Origin of Penetrative Lineations Within the Schist

The presently NE/SW orientation of penetrative lineations across most exposures of the schist (e.g., Haxel et al., 2018) precludes vertical-axis rotations only if it is assumed that these lineations formed, and remained, parallel to subduction (e.g., Ehlig, 1958; Haxel and Dillon,

1978; Jacobson and Dawson, 1995; Haxel et al., 2018, but see discussion in Jacobson et al.,

1988), which was directed approximately northeastward at the time of schist underplating

(Engebretson et al., 1985; Stock and Molnar, 1988; Doubrovine and Tarduno, 2008). There are two potential challenges to this assumption. First, lineations may not be oriented parallel to the regional transport direction, particularly in zones of oblique convergence (e.g., Ridley, 1986;

Tikoff and Teyssier, 1994; Tikoff and Greene, 1997; Miller et al., 2006). Second, recent studies suggest that penetrative lineations within the schist formed subsequent to their emplacement via underthrusting, in which case they need not be subduction-parallel. In northern schist exposures

(, Rand Mountains, and Sierra de Salinas; Fig. 2-1A), even those penetrative lineations which developed during prograde metamorphism, which was a result of subduction beneath hot lower crust of the North America plate (e.g., Peacock, 1987), are associated with top-to-NNE shear (Chapman et al., 2010). This is opposite of that which would have resulted from subduction underthrusting, implying that the schist of these exposures underwent additional deformation, such as passive-roof thrusting (Yin, 2002), return flow

(Oyarzabal et al., 1997; Saleeby, 2003), or channelized extrusion (Chapman et al., 2010), during or shortly after emplacement via subduction underthrusting, while still at prograde metamorphic conditions (Chapman et al., 2010). In the southern San Gabriel Mountains, the Vincent fault

92

(formerly, “Vincent thrust”), has been considered a rare example of the original subduction thrust responsible for emplacement of the schist, preserved without significant reactivation

(Yeats, 1968; Jacobson, 1983a, b, 1997; Jacobson and Dawson, 1995). Subsequent studies, however, have documented multiple, pervasive retrograde deformational events within the schist in its footwall (e.g., Behr et al., 2008; Xia and Platt, 2017, 2018). Penetrative lineations in the schist and structurally overlying mylonite have been attributed to these retrograde events, rather than to earlier, prograde metamorphism presumably caused by subduction (Xia and Platt, 2017,

2018). The earliest phase of deformation documented in the schist of the Orocopia Mountains has been interpreted as subduction-related, based on comparison with deformation along the

Vincent fault (Jacobson and Dawson, 1995). Thus, penetrative lineations associated with this early phase of deformation in the Orocopia Mountains presumably also post-date emplacement via subduction underthrusting. In light of these recent findings, development of penetrative lineations within the schist and structurally overlying mylonitic gneiss cannot conclusively be attributed to subduction underthrusting, and may instead correspond to a subsequent deformational event.

NNW-Lineations Model

If the penetrative lineations within the schist did not form parallel to the direction of subduction underthrusting (i.e., approximately NE/SW), either as a consequence of oblique convergence, or because they record a subsequent deformational event, then apparent contradictions with paleomagnetic data are eliminated. I hypothesize that penetrative lineations in the schist in and along the western part of the anticlinorium formed or were reoriented approximately NNW/SSE, within a few million years of emplacement of the schist via subduction underthrusting (Coffey, 2018). Subsequent, variable vertical-axis rotations of bodies

93 of the schist, together with overlying volcanic rocks, would have resulted in the presently observed lineation orientations. This model predicts an approximately linear original architecture of the anticlinorium (Fig. 2-1C), and eliminates the need for a decoupling surface.

Comparison of the Two Models

Because only 19±14°of clockwise vertical-axis rotation is implied by paleomagnetic data from volcanics near Mount Pinos (Prothero et al., 2008), I had hoped that measuring penetrative lineations in the schist there would be an effective test of these two hypotheses: the decoupling- surface model would predict an approximately NE/SW orientation, whereas my NNW-lineations model would predict an approximately N/S orientation (i.e., an approximately 19° clockwise vertical-axis rotation of an originally NNW/SSE orientation). Unfortunately, the dominantly

WNW/ESE orientation of (tilt-corrected) penetrative lineations (Figs. 2-3B, 2-4E) is far from either prediction. The schist is substantially closer to the San Andreas fault than the volcanic rocks where paleomagnetic data were collected, and there are intervening faults (e.g., Kellogg and Miggins, 2002), so it is conceivable that the schist underwent a coherent block rotation independent of these volcanic rocks. However, I consider a more likely explanation to be modification of the orientations of penetrative lineations via transpression associated with the adjacent San Andreas fault. The (tilt-corrected) penetrative lineations are generally subparallel to the trace of the fault, with variation in lineation orientation from west to east even partly mirrored by variation in the fault trace (Fig. 2-3B). Across the San Andreas fault, at present less than 2 km from the schist of the Mount Pinos area, exposures of the schist in the San Emigdio

Mountains (Fig. 2-1A) have a range of penetrative-lineation orientations similar to that of Mount

Pinos, but anomalous compared to those of the Rand Mountains and Sierra de Salinas, which would have been adjacent prior to slip on the San Andreas fault system and Garlock fault

94

(Postlethwaite and Jacobson, 1987; Chapman et al., 2010; Fig. 2-1B). Chapman et al. (2010) attributed this to transpression along the San Andreas fault in the Pliocene-Quaternary.

To compare the decoupling-surface and NNW-lineations models, I compiled reported vertical-axis rotations implied by paleomagnetic data collected near exposures of the schist

(Table 2-1; secular variation not adequately averaged for most values), and representative orientations of penetrative lineations within each exposure of the schist, both at present, and if corrected for these rotations (Table 2-2; details of determination of representative orientations in

Supplemental Table S4). These data are plotted on reconstructions of the anticlinorium in

Figures 2-9 and 2-10. In Fig. 2-9, no vertical-axis rotations are restored, except for approximately 16° counterclockwise rotation of the San Gabriel block, which occurred after ca.

12-11 Ma (Table 2-1), by which time the schist was exhumed and any decoupling surface was deactivated (see discussion above), and approximately 60° clockwise rotation of the eastern

Castle Dome Mountains, implied by the dominant trend of Miocene dikes within this exposure of the schist, which is approximately 60° clockwise of that in the schist of the western Castle Dome

Mountains and Neversweat Ridge (Haxel et al., 2002). This is the approximate architecture of the anticlinorium implied by the decoupling-surface model (Haxel et al., 2018). In Fig. 2-10, each exposure of the schist, together with surrounding pre-Pliocene rocks, are corrected for the amount of vertical-axis rotation implied by the nearest paleomagnetic data (the eastern Castle

Dome Mountains are again corrected for the approximately 60° of clockwise rotation implied by dike orientations, rather than for the approximately 35° implied by nearby paleomagnetic data; both corrections are shown in the circular plot; the uncertainty of the paleomagnetic estimate and the inadequate averaging of secular variation are probably sufficient to account for this difference). This is the approximate architecture of the anticlinorium implied by my NNW-

95

Table 2-1. Summary of Paleomagnetic Data from Rocks in the Vicinity of Schist Exposures Schist exposure (and abbreviation) Vertical-axis rotation implieda Approximate age of (degrees clockwise) measured rocks

WESTERN EXPOSURES Mount Pinos (MP) 19°b, c 27-23 Mad Sierra Pelona (SP) 37°e, f 25 Mad, h -16°f, g 12-11 Maf Blue Ridge (BR) 37°e, i 25 Mai -28°j 8.5 Maj Orocopia Mountains (OM) 90°k 24-21 Mad northern Chocolate Mountains (NCM)l no data - southern Chocolate Mountains (SCM)l 25°m Oligocene-Miocenen Peter Kane Mountain (PKM) 50°m Oligocene-Miocenen Gavilan Hills (GH) 50°m Oligocene-Miocenen southern San Gabriel Mountains (SSGMN/SSGMS) 37°o 25 Mao -16°g, o 12-11 Mao Crafton Hills (CH) no data -

EASTERN EXPOSURES Picacho District/ (PD/TM) 28°m Oligocene-Miocenep Middle Mountains (MM) no data - , western exposure (CDMW) 0°q, r Oligocene-Miocenes Castle Dome Mountains, eastern exposure (CDME) 35°q, r, t Oligocene-Miocenes Neversweat Ridge (NSR) 10°r, u 18 Mar Plomosa Mountains (PM) -20°r, u, w 17-16 Max Cemetery Ridge (CR) 10°q, r, y, z 18 Max

NORTHERN EXPOSURES Tehachapi Mountains (TeM) 59°†, § 87-77 Ma†, # Rand Mountains, , Sierra de Salinas, 59°* 87-77 Ma* and Portal Ridge (RM/SEMW/SEME/SdS/PR) aSecular variation typically not adequately averaged. bdata of Terres (1984) imply -10°, but thought by Prothero (2006) to represent remagnetization; other measurements in Eocene-Miocene rocks in the same fault block imply 31°-45° (Prothero, 2006; Prothero et al., 2008). cProthero (2006). dFrizzell and Weigand (1993). eNet amount of vertical-axis rotation since ca. 25 Ma. fTerres and Luyendyk (1985). gAmount of vertical-axis rotation since ca. 12-11 Ma. hCoffey (2015). iExtrapolated from data of Sierra Pelona, with which Blue Ridge was contiguous prior to initiation of the Punchbowl fault (e.g., Coffey et al., 2019). jW. Liu (1990). lMost of the Chocolate Mountains are inaccessible because they are within the Chocolate Mountain Aerial Gunnery Range. mCostello (1985). nCrowe et al. (1979). oExtrapolated from data of Sierra Pelona. pNeedy et al. (2007); Biggs (2008). qButterworth (1984). rCalderone et al. (1990). sGrubensky et al. (1993). tOrientations of Miocene dikes in the Castle Dome Mountains and at Neversweat Ridge suggest 60° (Haxel et al., 2002). uNo data from immediate vicinity; data from surrounding mountain ranges suggest 0°-20°; I assume 10° as a compromise. vNo data from immediate vicinity; data from approximately 35-40 km north and south suggest -10° and -30°, respectively; I assume -20° as a compromise. wCalderone and Butler (1984). xShafiqullah et al. (1980). yVeseth (1985). zNo data from immediate vicinity; data from the , to the west, indicate 10°. †McWilliams and Li (1985). §Decreasing to the northeast to 45° (Kanter and McWilliams, 1982; McWilliams and Li, 1985). #Paleomagnetic data from ca. 23-22 Ma volcanic rocks suggest that much or all of this rotation may have occurred after ca. 23-22 Ma; see text for discussion and references. *Extrapolated from data of the Tehachapi Mountains. See text for discussion.

96

Table 2-2. Summary of Structural Data from Schist Exposures Schist exposure Orientation of Approximate orientation of anticlinorium axial trace a (and abbreviation) penetrative lineations (azimuth degrees) WESTERN EXPOSURES Present Correctedb Present Correctedb Mount Pinos (MP) 100°c, d 81°d not definede - Sierra Pelona (SP) 44°f, g, h 171°h E/Wg, i NE/SW 28° f, g, h (ENE/WSW)g, i Blue Ridge (BR) inconsistentj WNW/ESE-NNW/SSEg, k, l NE/SW-E/Wl (E/W-NW/SE)g, k, l Orocopia Mountains (OM) 62°h, m 152°h NW/SE m NE/SW northern Chocolate Mountains (NCM)n no data - no data - southern Chocolate Mountains (SCM)n 173°o 148° E/W-NW/SEo, p ENE/WSW-WNW/ESE Peter Kane Mountain (PKM) 4° o 134° E/Wo, p NE/SW Gavilan Hills (GH) 30°h, q 160°h E/Wr NE/SW southern San Gabriel Mountains, northern part 148°g, s 148° not part of CMA - (SSGMN) 132°g, s southern San Gabriel Mountains, southern part 16°g, t 16° not part of CMA - (SSGMS) 0°g, t Crafton Hills (CH) 56° no data Not part of CMA - EASTERN EXPOSURES Picacho District (PD) 28°o 0° E/Wo ENE/WSW Trigo Mountains (TM) 30°u 2° ENE/WNWu NE/SW Middle Mountains (MM) inconsistentu, v not definedu - Castle Dome Mtns., western exposure (CDMW) 43°u 43° not definedw - Castle Dome Mtns., eastern 47°u, x 47°y not definedw - u, x y Exposure (CDME) 107° 72° Neversweat Ridge (NSR) 67°u 57° not definedz - Plomosa Mountains (PM) 46°†, § 66°§ not part of CMA - Cemetery Ridge (CR) 42°# 32° not part of CMA - NORTHERN EXPOSURES* San Emigdio Mtns., western exposure (SEMW) 98°d, †† 39°d not part of CMA - San Emigdio Mtns., eastern exposure (SEME) 117°d, †† 58°d not part of CMA - Tehachapi Mountains (TeM) no data - not part of CMA - Rand Mountains (RM) 19°§§ 140° not part of CMA - Sierra de Salinas (SdS) 30°†† 151° not part of CMA - Portal Ridge (PR) no data - not part of CMA - Note: Mtns. = Mountains; CMA = Chocolate Mountains anticlinorium. aDetails of determination of representative orientations in Appendix A. bCorrected for vertical-axis rotations implied by paleomagnetic data (summarized in Table 2-1). cData of this study (region M). dLikely altered by Pliocene-Quaternary deformation related to the San Andreas fault system; see Chapman et al. (2010) and discussion in text. eKellogg and Miggins (2002). fData of this study (region SW). gFirst value is corrected for 16° of post-12 Ma, counterclockwise vertical-axis rotation of the San Gabriel block (see Table 2-1 and discussion in text), but not for any earlier rotation; second value (italicized) is the uncorrected, modern orientation. hLineations are deflected clockwise of this orientation along the northern margin of schist exposure. iJahns and Muehlberger (1954); Muehlberger and Hill (1958). jJacobson et al. (1988). kDibblee (2002a, b, c). lThe more clockwise of this range of orientations is found adjacent to the dextral Punchbowl fault, and may be the result of drag folding. mJacobson and Dawson (1995). nMost of the Chocolate Mountains are inaccessible because they are within the Chocolate Mountain Aerial Gunnery Range. oDillon et al. (1990). pDillon (1976). qJacobson et al. (2002). rOyarzabal et al. (1997). sJacobson (1983a). tMay and Walker (1989). uHaxel et al. (2002). vPresumably due to low sample size (n = 6). wGrubensky et al. (1993). xFirst value is corrected for 60° of post-Miocene rotation implied by relative orientations of Miocene dikes in the western and eastern Castle Dome Mountains and at Neversweat Ridge (Haxel et al., 2002); second value (italicized) is the uncorrected, modern orientation. yFirst value is corrected for 60° of post-Miocene rotation implied by orientations of Miocene dikes (see previous footnote); second value (italicized) is correction of the modern orientation using paleomagnetic data, and ignoring the Miocene-dike argument of Haxel et al. (2002). zGrubensky et al. (1995). †Strickland et al. (2018). §These lineations are related to Miocene detachment faulting, not deformation during emplacement of the schist (Strickland et al., 2018). #Haxel et al. (2015). *Correction based on paleomagnetism may be inappropriate for these exposures, even if appropriate for all others; see text for discussion. ††Chapman et al. (2010). §§Postlethwaite and Jacobson (1987). 97

Figure 2-9: Paleotectonic reconstruction of exposures of Pelona, Orocopia, Rand, and related

schists, including those along the Chocolate Mountains anticlinorium, ca. 25 Ma, according to the decoupling-surface model. Shown are representative orientations of penetrative lineations, senses- 98

of-shear related to early Cenozoic deformation, anticlinorium axial traces, and paleomagnetic

declinations near schist exposures (references in Tables 2-1 and 2-2, and in text). Slip on faults of the San Andreas fault system and the Garlock fault has been restored; San Gabriel block, including all schist exposures denoted by “*” and adjacent paleomagnetic declinations, has been corrected for

16° of post-ca. 12 Ma, counterclockwise vertical-axis rotation; eastern Castle Dome Mountains,

including schist exposure denoted by “†” and adjacent paleomagnetic declination, have been corrected for 60° of clockwise vertical-axis rotation (gray and black lines in circular plots indicate

uncorrected and corrected orientations, respectively; see discussion in text); no other corrections

for vertical-axis rotations have been performed. Note generally NE/SW orientation of penetrative lineations, and several exceptions in western exposures. S. = Southern. Areas of figure shown in Fig.

2-1B. Determination of representative orientations is discussed in Appendix A. Restoration

modified from Haxel et al. (2018).

Figure 2-10 (following page): Paleotectonic reconstruction of exposures of Pelona, Orocopia, Rand,

and related schists, including those along the Chocolate Mountains anticlinorium, ca. 25 Ma, according to the NNW-lineations model (but note that correction for rotation of schist in northern exposures is only a possibility, and not central to the model; see discussion in text). Details same as

in Fig. 2-9, but with schist exposures corrected for all vertical-axis rotations implied by

paleomagnetic data, and schematic restoration of hypothesized related strike-slip faulting and

extension caused by Basin and Range normal faulting east of the Colorado River. Map and black

lines in circular plot marked by “§” assume that schist of southern San Gabriel Mountains

experienced 37° of net clockwise vertical-axis rotation documented in northern San Gabriel

Mountains; gray lines marked by “§” are corrected only for 16° of counterclockwise rotation since

12 Ma; see discussion in text. Map and black line in circular plot marked by “#” are corrected for

60° of clockwise vertical-axis rotation of eastern Castle Dome Mountains implied by dike

99

orientations; gray line marked by “#” is corrected for 35° implied by nearby paleomagnetic data; see discussion in text. Areas of figure shown in Fig. 2-1C. Restoration modified from Ingersoll et al.

(2014), Bennett et al. (2016), and Haxel et al. (2018).

100 lineations model.

The present orientations of penetrative lineations across exposures of the schist are dominantly close to NE/SW, and this consistency across hundreds of kilometers has been interpreted as supporting a lack of vertical-axis rotations, and thus the decoupling-surface model

(Haxel et al., 2018). However, dividing the schist into western, eastern, and northern exposures reveals that the northern and eastern exposures are largely responsible for this trend, with orientations in the western exposures varying substantially (Fig. 2-9). In the decoupling-surface model, this requires ad-hoc postulation of counterclockwise vertical-axis rotations of approximately 33° in the southern Chocolate Mountains and 28° in the Peter Kane Mountain block, and up to 70-90° of vertical-axis rotation in the southern San Gabriel Mountains (Haxel et al., 2018). If orientations are corrected for vertical-axis rotations implied by paleomagnetism, those in the western exposures become notably more consistent, clustering around an approximately NNW/SSE orientation (Fig. 2-10), with only two outliers. One, at Mount Pinos, is anomalous in both models, and likely modified by subsequent deformation, as discussed above.

The other, in the northern part of the southern San Gabriel Mountains, deviates 48° from that of the nearby southern part; without independent paleomagnetic measurements adjacent to both parts, they must be treated as a single, coherent block, and thus, the orientation for at least one will deviate from any predicted orientation. If, alternatively, the southern San Gabriel Mountains experienced only the 16° ± 30° of counterclockwise vertical-axis rotation that occurred since ca.

12-11 Ma (Terres and Luyendyk, 1985), which was likely experienced by the entire San Gabriel block as a result of counterclockwise rotation of the Mojave segment of the San Andreas fault

(Terres and Luyendyk, 1985; Ingersoll and Coffey, 2017), and not the net 37° of clockwise vertical-axis rotation since ca. 25 Ma (Table 2-1), which is based on data from the opposite end

101 of the San Gabriel block, then lineations in this northern part of the schist exposure would restore to the predicted NNW/SSE orientation, but those in the southern part would become anomalous

(gray lines in Fig. 2-10). The schist of Crafton Hills is part of a 17 km-by-4 km fault-bounded block between the dextral San Andreas and San Jacinto faults (e.g., Matti et al., 2003; Fig. 2-9), which may have undergone substantial rotation independent of other exposures, particularly if slip has been transferred between these faults via the normal faults bounding the Crafton Hills

(Matti et al., 1992). With no paleomagnetic data reported from this block, the orientations of penetrative lineations there cannot be restored, and consequently cannot be evaluated in the

NNW-lineations model (and are not plotted in Fig. 2-10). Correction for vertical-axis rotations implied by paleomagnetism also aligns the axial traces of the segments of the anticlinorium to a generally consistent, approximately NE/SW orientation (Fig. 2-10).

Orientations of penetrative lineations in the eastern exposures of the schist generally remain approximately NE/SW after correction for vertical-axis rotations implied by paleomagnetism (Figs. 2-9, 2-10). The deviation at Neversweat Ridge is slightly reduced, whereas that at Cemetery Ridge is slightly increased. Increased deviation at the Plomosa

Mountains is not relevant, as these penetrative lineations are related to Miocene detachment faulting (Strickland et al., 2018). The NE/SW orientation of corrected lineations in the eastern exposures requires that the hypothesized formation or reorientation of penetrative lineations in a

NNW/SSE orientation in my NNW-lineations model did extend this far east. Correction of the orientations of penetrative lineations at the Picacho District and the Trigo Mountains, located between the western exposures and the other eastern exposures, results in approximately N/S orientations (Fig. 2-10). This may represent a transition from lineations formerly oriented

NNW/SSE farther west to lineations originally oriented NE/SW farther east.

102

The schist of the northern exposures is substantially older than (ca. 95-75 Ma vs. 75-60

Ma; Grove et al., 2003), and originated nearly 300 km north of, the schist of the western exposures (Fig. 2-1B). Accordingly, its history may differ substantially. Igneous rocks in the nearby Tehachapi Mountains indicate clockwise vertical-axis rotation between ca. 87 and 16 Ma

(Table 2-1; Fig. 2-9). Chapman et al. (2010) suggested that rotation occurred ca. 88-86 Ma, decoupled from the underlying, recently underplated schist. If so, then penetrative lineations within the northern schist exposures would remain NE/SW to NNE/SSW even after reversing this rotation (Fig. 2-9). Dickinson (1996) suggested that rotation occurred ca. 23-16 Ma, based on paleomagnetic data from ca. 23-22 Ma volcanic rocks (Turner, 1970; written communication of J. Plescia, 1987, in Goodman et al., 1989; written communication of J. Plescia, 1986, in

Goodman and Malin, 1992) implying between 27° and 44° of clockwise rotation (Plescia and

Calderone, 1986; Graham et al., 1990; Plescia et al., 1994), comparable to the rotation implied by ca. 87-80 plutonic rocks (Table 2-1). If rotation did occur during this later interval, then it is more likely that the schist and the igneous rocks of the Tehachapi Mountains rotated together.

Correction for this vertical-axis rotation would restore penetrative lineations to an approximately

NNW/SSE orientation, consistent with corrected orientations in the western exposures (Fig. 2-

10). Extrapolation of the rotation implied by paleomagnetic data from the Tehachapi Mountains to the other northern schist exposures is reasonable because it occurred prior to 20-16 Ma

(McWilliams and Li, 1985; unpublished data of B. Turrin in McWilliams and Li, 1985), when the major faults that currently separate the northern schist exposures (i.e., San Andreas fault system; Garlock fault; ) had not formed (e.g., Stein and Thatcher, 1981;

Burbank and Whistler, 1987; Nicholson et al., 1994). This was recognized by Dickinson (1996), who suggested that this rotation may have extended across the future San Andreas fault system

103 to the area of Sierra de Salinas. As discussed above, orientations of penetrative lineations in the

San Emigdio Mountains exposures, anomalous in either model (Figs. 2-9, 2-10), have likely been altered by subsequent deformation (Chapman et al., 2010).

Formation of NNW/SSE-Oriented Penetrative Lineations

The formerly NNW/SSE orientation of penetrative lineations required by the NNW- lineations model is not problematic. As mentioned above, lineations in a zone of oblique convergence may not be parallel to the direction of subduction (e.g., Ridley, 1986; Tikoff and

Teyssier, 1994; Tikoff and Greene, 1997; Miller et al., 2006). Even if this was the case, a subsequent deformational event capable of producing NNW/SSE lineations is feasible. In the

Orocopia Mountains, asymmetric folds within the schist record top-to-NE to top-to-E shear

(Jacobson and Dawson, 1995; Fig. 2-9); muscovite recrystallized within the hinges of these folds during their development yields the same ca. 44 Ma 40Ar/39Ar ages as muscovite in the surrounding schist, indicating that this shear occurred prior to ca. 44 Ma (Jacobson et al., 2007).

Top-to-N shear has been documented along the Vincent thrust in the southern part of the schist exposure in the southern San Gabriel Mountains (May and Walker, 1989; Fig. 2-9). In both cases, correcting for vertical-axis rotations implied by paleomagnetism restores these senses of shear to top-to-NNW (Fig. 2-10). The latter case is probably equivalent to the top-to-NW shear documented along the Vincent thrust in the northern part of this same exposure (Fig. 2-9), which occurred within a few million years of emplacement by subduction, and prior to all other post- subduction deformation (Xia and Platt, 2017). Approximately top-to-NE shear has been determined in the Gavilan Hills (Simpson, 1986) and from combined data of the southeastern

Chocolate Mountains, Gavilan Hills, and Picacho District (Dillon et al., 1990; Fig. 2-9).

Correcting for vertical-axis rotations implied by paleomagnetism restores these senses of shear to

104 top-to-N to top-to-NNW (Fig. 2-10). Collectively, this evidence suggests widespread, top-to-

NNW shear within the schist of the western exposures shortly after emplacement by subduction.

Similar, earlier top-to-NNW shear may have occurred in the older schist (Grove et al.,

2003) of the northern exposures (i.e., Tehachapi, Rand, and San Emigdio Mountains, Sierra de

Salinas, and Portal Ridge). The earliest deformation to follow underthrusting via subduction in the schist of these northern exposures was top-to-NE shear (Nourse, 1989; Chapman et al., 2010;

Fig. 2-9), which would restore to top-to-NNW after correcting for vertical-axis rotation implied by paleomagnetic data from the Tehachapi Mountains (Fig. 2-10). However, as discussed above, the schist may not have experienced these vertical-axis rotations (Chapman et al., 2010), and this early post-subduction deformation occurred prior to that in the schist of the western and eastern exposures.

Top-to-NNW shear could have been caused by the geometry of the subducting conjugate

Hess rise: if the long dimension of this rise was slightly counterclockwise of the subduction direction, then progressively farther southeastward parts of the western edge of the North

America plate would have been underlain by this rise through time (see Jacobson et al., 2011, who proposed this mechanism as a possible cause of sinistral slip on the Nacimiento fault, based on modification of a proposal of Barth and Schneiderman, 1996). A second possible cause is detachment faulting following isostatic uplift of the schist (e.g., Jacobson et al., 1996, 2007; but see Yin, 2002, Chapman, 2017). A third possible cause for NNW/SSE-oriented lineations could be strain partitioning of oblique subduction into trench-normal and trench-parallel components, with the latter forming these lineations. However, the trench-parallel component of subduction was probably dextral (Engebretson et al., 1985; Stock and Molnar, 1988; Doubrovine and

Tarduno, 2008; but see Singleton and Cloos, 2013), in other words, SSE motion of the North

105

America plate relative to the Farallon plate. This would cause top-to-SSE shear, rather than top- to-NNW.

Possible Relationship to Nacimiento Fault

I suspect there may be a connection between the top-to-NNW shear discussed above and the Nacimiento fault, although I admit this is rather speculative. Pervasive sinistral shear has been documented on approximately NW-striking, high-angle faults in mélange of the Franciscan

Complex near San Simeon (Singleton and Cloos, 2013), which formed in the same subduction zone as the schist (e.g., Hamilton, 1969; Chapman, 2017). This shear is constrained to post-85

Ma and pre-20 Ma, and thought to have occurred in the latest Cretaceous to early Cenozoic

(Singleton and Cloos, 2013). This is the same time interval during which the schist of the western exposures was deformed by the likely top-to-NNW shear discussed above (Grove et al.,

2003; Xia and Platt, 2017), and these shearing events may have been synchronous and related.

This sinistral shear within the Franciscan Complex has been attributed to sinistral motion on the nearby Nacimiento fault (Singleton and Cloos, 2013), which was probably active between ca. 75-

70 Ma and ca. 60 Ma (e.g., Dickinson et al., 2005; Jacobson et al., 2011; Hall and Saleeby,

2013). However, it is actively debated whether the Nacimiento fault is sinistral (see Ingersoll,

2019, and references therein) or a subduction megathrust, potentially reactivated as a normal fault with top-to-SW motion (see Hall, 1991; Hall and Saleeby, 2013, and references therein). I suspect that the Nacimiento-Vincent fault may be a single, primarily low-angle fault that reactivated, or formed subparallel to, the former subduction megathrust (as in model of Hall and

Saleeby, 2013, and references therein), but that motion was top-to-NNW (close to the top-to-W to top-to-NW thrusting direction proposed by Hall, 1991, and Hall and Saleeby, 2013), and that

106 the surface trace of the Nacimiento fault is a lateral ramp, along which slip was sinistral (as in the model of Dickinson, 1983, and Ingersoll, 2019, and references therein; Fig. 2-11).

The Sierra de Salinas shear zone / Rand fault has been interpreted as a top-to-NE detachment fault (e.g., Chapman et al., 2010). If vertical-axis rotation implied by paleomagnetic data adjacent to the northern exposures of the schist was decoupled from the schist itself (Chapman et al., 2010; see discussion above), then this interpretation would remain unchanged. However, if such rotation was experienced by the schist, then shearing would have been top-to-NNW, and this fault could be an older, structurally higher predecessor to the Nacimiento-Vincent fault (Fig.

2-11). The Late Cretaceous ages of the schist in the northern exposures (Grove et al., 2003) and of the Sierra de Salinas shear zone / Rand fault (Chapman et al., 2010) versus the latest

Cretaceous-earliest Paleogene ages of the schist of the western exposures (Grove et al., 2003;

Jacobson et al., 2018) and of the Nacimiento-Vincent fault (see discussion in Jacobson et al.,

2011; Xia and Platt, 2017) may be due to subduction of the conjugate Shatsky rise beneath the area of the northern schist exposures during the Late Cretaceous versus subduction of the conjugate Hess rise beneath the area of the western schist exposures during the latest Cretaceous- earliest Paleogene (e.g., Saleeby, 2003; Liu et al., 2010).

Origin of Neogene-Quaternary Clockwise Vertical-Axis Rotations

The clockwise vertical-axis rotations of varying magnitudes since the latest Oligocene- early Miocene, implied by paleomagnetic data for western segments of the anticlinorium, and necessary in the NNW-lineations model, also have potential explanations. Much of the western part of the anticlinorium seems to have undergone approximately 40°-50° of clockwise vertical- axis rotation, with subsequent counterclockwise rotation of Sierra Pelona and the rest of the San

Gabriel block, and additional clockwise rotation of the Orocopia Mountains (for a total of

107

Figure 2-11: Schematic block diagram illustrating hypothesized geometry and kinematics of the

Nacimiento-Vincent fault, ca. 70 Ma, along which top-to-NNW shear may have occurred in the

NNW-lineations model. Note that top of block diagram (map view) is essentially identical to

sinistral-strike-slip models for the Nacimiento fault, whereas side (cross-section view) matches

architecture (though not kinematics) of thrust-reactivated-as-normal-fault models for the

Nacimiento/Sur/Vincent faults. If sense-of-shear indicators in northern schist exposures restore to top-to-NNW, then the older, inactive, structurally higher Rand fault / Sierra de Salinas shear zone may have experienced similar top-to-NNW shear. The Nacimiento-Vincent fault likely reactivated a

former position of the subduction megathrust; un-reactivated part of this former megathrust is

shown in gray between the active subduction megathrust and the Rand fault / Sierra de Salinas

shear zone. Teeth indicate hanging wall of a megathrust. No vertical exaggeration. See discussion

and references in text. Figure adapted from those of Jacobson et al. (2011).

108 approximately 90°; Table 2-1). One possible explanation for the approximately 40°-50° of rotation is that it is related to clockwise vertical-axis rotation of the western Transverse Ranges

(Hornafius et al., 1986; Nicholson et al., 1994; Ingersoll and Coffey, 2017). The later rotations of smaller segments can be readily related to deformation related to the geometry of the San

Andreas fault system (Terres and Luyendyk, 1985; Ingersoll and Coffey, 2017).

Implications of the NNW-Lineations Model

The NNW-lineations model predicts an originally linear, NE/SW- to ENE/WSW-trending anticlinorium (Fig. 2-10). For several segments of the anticlinorium, this implies original orientations of structural and sedimentologic features drastically different from present ones, requiring alternate interpretations. A good example is the Orocopia Mountains fault: there has been ongoing debate over whether this fault is an extensional detachment (e.g., Jacobson and

Dawson, 1995; Jacobson et al., 1996, 2007; Robinson and Frost, 1996), a passive-roof thrust

(Yin, 2002), or the upper bounding surface of return flow or channelized extrusion in a subduction channel (e.g., Chapman, 2017). Critical to arguments in this debate is the orientation of the Orocopia Mountains fault and its sense-of-shear indicators relative to the direction of

Farallon-plate subduction beneath the North America plate. However, restoring the approximately 90° of clockwise vertical-axis rotation that occurred according to the NNW- lineations model would completely change these relationships. More broadly, an originally linear anticlinorium could have formed in ways that an originally curved anticlinorium could not have

(e.g., parallel to the , as suggested by Grove and Jacobson, 2016).

Bending of the anticlinorium from an originally linear trend would also imply on the order of 100 km of Pacific-North America transform motion (Fig. 2-1B vs. Fig. 2-1C), in addition to the approximately 310 km previously inferred from displacements of segments of the

109 anticlinorium along the San Andreas fault system (e.g., Hill and Dibblee, 1953; Crowell, 1962,

1975, 2003; Carman, 1964; Ingersoll et al., 2014). As mentioned above, this could be related to clockwise vertical-axis rotation of the western Transverse Ranges, which accommodated hundreds of kilometers of transform motion that must be added to slip on the southern San

Andreas fault system (together with transform motion accommodated by Basin and Range extension) to account for total relative motion of the North America and Pacific plates since the mid-Miocene (Dickinson, 1996; Dickinson and Wernicke, 1997).

CONCLUSIONS

Penetrative lineations within the schist at Sierra Pelona are generally NNE/SSW, but are deflected clockwise along the northern edge of its exposure, where crenulation lineations are more common. This matches patterns observed in the Orocopia Mountains and the Gavilan Hills of the Chocolate Mountains anticlinorium, lending further support to the generally accepted idea that exposures of the schist at Mount Pinos, Sierra Pelona, and Blue Ridge are segments of this anticlinorium that have been displaced along strands of the San Andreas fault system. At Sierra

Pelona, as elsewhere along the anticlinorium, if penetrative lineations within the schist are assumed to have developed and remained parallel to the former direction of subduction, then their present orientation implies little to no vertical-axis rotation, at odds with the substantial clockwise rotation implied by paleomagnetic data from nearby, ca. 25 Ma volcanic rocks. This apparent discrepancy must be reconciled in order to understand the original architecture of the

Chocolate Mountains anticlinorium, and the evolution of the transform boundary in southern

California that has deformed it.

110

The decoupling-surface model (Haxel et al., 2018) reconciles lineation orientations with paleomagnetic data by postulating the existence of a decoupling surface between the schist and overlying volcanic rocks, which would have allowed the penetrative lineations within the schist to avoid the rotations of the volcanic rocks implied by paleomagnetic data. However, timing of rotations are potentially problematic. This model predicts an originally curved shape of the

Chocolate Mountains anticlinorium.

My NNW-lineations model is a fundamentally different way to reconcile these datasets.

It suggests that penetrative lineations in the schist along the western part of the anticlinorium developed in, or were reoriented to, a NNW/SSE orientation, and that these bodies of schist then rotated in conjunction with overlying volcanic rocks to achieve their present orientations. This implies an originally linear shape of the anticlinorium, trending approximately NE/SW to

ENE/WSW. An originally linear trend of the Chocolate Mountains anticlinorium would have important implications for its mechanism of formation, the significance of the orientation of sedimentologic and structural features along it, and the magnitude of transform motion implied by its present architecture. The NNW-lineations model is a viable alternative to the decoupling- surface model of Haxel et al. (2018); both models should be considered in future studies of the

Chocolate Mountains anticlinorium and reconstructions of the southern San Andreas fault system.

Additional work could test the competing, decoupling-surface and NNW-lineations hypotheses. As discussed above, a potential problem with the decoupling-surface model is the relative timing of exhumation of the schist, and thus deactivation of any decoupling surface above it, and of the rotations implied by paleomagnetism. Along the flanks of much of the anticlinorium, including at Sierra Pelona, the sedimentary record spans from before eruption of

111 the volcanic deposits that have yielded paleomagnetic data, to after exposure and erosion of the schist (e.g., Jahns and Muehlberger, 1954; Terres and Luyendyk, 1985; Hendrix and Ingersoll,

1987; Ehlert, 2000; Ricketts et al., 2011; Hoyt et al., 2018; Coffey et al., 2019). Collection of paleomagnetic data throughout this stratigraphic section could better constrain the timing of rotation of the volcanic rocks structurally above the schist, establishing whether the underlying schist having been decoupled from this rotation is feasible.

112

REFERENCES CITED

Allmendinger, R.W., Cardozo, N., and Fisher, D., 2012, Structural geology algorithms: Vectors

and tensors in structural geology: Cambridge, Cambridge University Press, 289 p.

Barth, A.P., and Schneiderman, J.S., 1996, A comparison of structures in the Andean orogen of

northern Chile and exhumed midcrustal structures in southern California, USA: An

analogy in tectonic style?: International Geology Review, v. 38, p. 1075-1085,

https://doi.org/10.1080/00206819709465383

Behr, W.M., Alpert, L.A., Economos, R., Johanesen, K., Platt, J.P., and Zhang, T., 2008,

Microstructural analysis of deformation and metamorphism along the Vincent thrust in

the eastern San Gabriel Mountains, California: Geological Society of America Abstracts

with Programs, v. 40, n. 1, p. 47.

Bennett, S.E.K., Darin, M.H., Dorsey, R.J., Skinner, L.A., Umhoefer, P.J., and Oskin, M.E.,

2016, Animated tectonic reconstruction of the Lower Colorado River region: implications

for late Miocene to present deformation: in Reynolds, R.E., ed., Going LOCO:

investigations along the Lower Colorado River: 2016 Desert Symposium, Desert Studies

Center, California, p. 73-86.

Biggs, M.A., 2008, Miocene volcanic rocks and conglomerates, SE California: evidence for

Neogene reactivation of the Chocolate Mountains anticlinorium [M.S. thesis]: San Diego,

San Diego State University, 30 p.

Burbank, D.W., and Whistler, D.P., 1987, Temporally constrained tectonic rotations derived

from magnetostratigraphic data: Implications for the initiation of the Garlock fault,

California: Geology, v. 15, p. 1172-1175, https://doi.org/10.1130/0091-

7613(1987)15<1172:TCTRDF>2.0.CO;2

113

Burchfiel, B.C., and Davis, G.A., 1981, Mojave Desert and environs, in Ernst, W.G., ed., The

Geotectonic Development of California (Rubey Volume I): Englewood Cliffs, New

Jersey, Prentice-Hall, p. 217-252.

Butterworth, J.E., 1984, Paleomagnetic investigation of mid-Tertiary volcanic rocks in the Castle

Dome and Southern Kofa Mountains, Yuma County, Southwestern Arizona [M.S. thesis]:

San Diego, San Diego State University, 195 p.

Calderone, G., and Butler, R.F., 1984, Paleomagnetism of Miocene volcanic rocks from

southwestern Arizona: Tectonic implications: Geology, v. 12, p. 627-630,

https://doi.org/10.1130/0091-7613(1984)12<627:POMVRF>2.0.CO;2

Calderone, G.J., Butler, R.F., and Acton, G.D., 1990, Paleomagnetism of Middle Miocene

volcanic rocks in the Mojave-Sonora Desert region of western Arizona and southeastern

California: Journal of Geophysical Research, v. 95, p. 625-647,

https://doi.org/10.1029/JB095iB01p00625

Cardozo, N., and Allmendinger, R.W., 2013, Spherical projections with OSXStereonet:

Computers & Geosciences, v. 51, p. 193 – 205,

https://doi.org/10.1016/j.cageo.2012.07.021

Carman, M.F., Jr., 1964, Geology of the Lockwood Valley area, Kern and Ventura Counties,

California: California Division of Mines and Geology Special Report 81, 62 p.

Chapman, A.D., 2017, The Pelona-Orocopia-Rand and related schists of southern California: a

review of the best-known archive of shallow subduction on the planet: International

Geology Review, v. 59, p. 664-701, https://doi.org/10.1080/00206814.2016.1230836

Chapman, A.D., Kidder, S., Saleeby, J.B., and Ducea, M.N., 2010, Role of extrusion of the Rand

and Sierra de Salinas schists in Late Cretaceous extension and rotation of the southern

114

Sierra Nevada and vicinity: Tectonics, v. 29, TC5006,

https://doi.org/10.1029/2009TC002597

Coffey, K.T., 2015, Oligocene-Miocene sedimentary and volcanic strata of the Vincent Gap

region, eastern San Gabriel Mountains, southern California, USA, and their tectonic

significance [M.S. thesis]: Los Angeles, University of California, 174 p.

Coffey, K.T., 2018, Can structural and paleomagnetic data be reconciled to determine the

original geometry of the Chocolate Mountains anticlinorium of southern California and

southwestern Arizona?: Geological Society of America Abstracts with Programs, v. 50,

n. 5, https://doi.org/10.1130/abs/2018RM-313789

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019: Stratigraphy, provenance and tectonic

significance of the Punchbowl block, San Gabriel Mountains, California, U.S.A.:

Geosphere, v. 15, https://doi.org/10.1130/GES02025.1

Cole, R.B., and Stanley, R.G., 1995, Middle Tertiary extension recorded by lacustrine fan-delta

deposits, Plush Ranch basin, Western Transverse Ranges, California: Journal of

Sedimentary Research, v. 65B, p. 455-468, https://doi.org/10.1306/D4268284-2B26-

11D7-8648000102C1865D

Costello, S.C., 1985, A paleomagnetic investigation of mid-Tertiary volcanic rocks in the lower

Colorado River area, Arizona and California [M.S. thesis]: San Diego, San Diego State

University, 109 p.

Crowe, B.M., Crowell, J.C., and Krummenacher, D., 1979, Regional stratigraphy, K-Ar ages,

and tectonic implications of Cenozoic volcanic rocks, southeastern California: American

Journal of Science, v. 279, p. 186-216, https://doi.org/10.2475/ajs.279.2.186

115

Crowell, J.C., 1962, Displacement along the San Andreas fault, California: Geological Society of

America Special Paper 71, 61 p., https://doi.org/10.1130/SPE71

Crowell, J.C., 1975, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Crowell, J.C., 2003, Introduction to geology of Ridge Basin, southern California, in Crowell,

J.C., ed., Evolution of Ridge Basin, southern California: An interplay of sedimentation

and tectonics: Geological Society of America Special Paper 367, p. 1-15,

https://doi.org/10.1130/0-8137-2367-1.1

Dibblee, T.W., Jr., 1967, Areal geology of the western Mojave Desert, California: U.S.

Geological Survey Professional Paper, v. 522, 153 p., 2 plates, 1:125000.

Dibblee, T.W., Jr., 1968, Displacements on the San Andreas fault system in the San Gabriel, San

Bernardino, and San Jacinto Mountains, Southern California in Dickinson, W.R., and

Grantz, A., eds., Proceedings of conference on geological problems of San Andreas fault

system: Stanford University Publications in Geological Sciences, v. 11, p. 260-278.

Dibblee, T.W., Jr., 1996, Geologic Map of the Agua Dulce quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-58, 1:24,000.

Dibblee, T.W., Jr., 1997a, Geologic Map of the Green Valley quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-65, 1:24,000.

Dibblee, T.W., Jr., 1997b, Geologic Map of the Sleepy Valley & Ritter Ridge quadrangles, Los

Angeles County, California: Dibblee Geological Foundation, Map DF-66, 1:24,000.

Dibblee, T.W., Jr., 2002a, Geologic Map of the Valyermo quadrangle, Los Angeles County,

California: Dibblee Geological Foundation, Map DF-80, 1:24,000.

116

Dibblee, T.W., Jr., 2002b, Geologic Map of the Mescal Creek quadrangle, Los Angeles and San

Bernardino Counties, California: Dibblee Geological Foundation, Map DF-81, 1:24,000.

Dibblee, T.W., Jr., 2002c, Geologic Map of the Mount San Antonio quadrangle, Los Angeles

and San Bernardino Counties, California: Dibblee Geological Foundation, Map DF-81,

1:24,000.

Dickinson, W.R., 1983, Cretaceous sinistral strike slip along Nacimiento fault in coastal

California: AAPG Bulletin, v. 67, p. 624-645.

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,

https://doi.org/10.1130/SPE305

Dickinson, W.R., and Wernicke, B.P., 1997, Reconciliation of San Andreas slip discrepancy by a

combination of interior Basin and Range extension and transrotation near the coast:

Geology, v. 25, p. 663-665, https://doi.org/10.1130/0091-

7613(1997)025<0663:ROSASD>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., and Brabb, E.E., 2005, Net dextral slip, Neogene San

Gregorio- zone, coastal California: Geologic evidence and tectonic

implications: Geological Society of America Special Paper 391, 43 p.

https://doi.org/10.1130/0-8137-2391-4.1

Dillon, J.T., 1976, Geology of the Chocolate and Cargo Muchacho mountains, southeasternmost

California [Ph.D. thesis]: Santa Barbara, University of California, 575 p.

117

Dillon, J.T., Haxel, G.B., and Tosdal, R.M., 1990, Structural evidence for northeastward

movement on the Chocolate Mountains Thrust, southeasternmost California: Journal of

Geophysical Research, v. 95, p. 19953-19971, https://doi.org/10.1029/JB095iB12p19953

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, https://doi.org/10.1029/2008JB005585

Ebert, K.A., 2004, Exhumational history of the Orocopia Schist and development of hanging-

wall structures of the Orocopia fault, southeastern California [M.S. thesis]: Los Angeles,

University of California, 124 p.

Ehlert, K.W., 2000, Offset of the Miocene Mint Canyon and Caliente Formations along the San

Andreas and San Gabriel Faults, southern California [M.S. thesis]: Los Angeles,

California State University, 115p.

Ehlig, P.L., 1958, The geology of the region of the San Gabriel Mountains,

California [Ph.D. thesis]: Los Angeles, University of California, 192 p.

Ehlig, P.L., 1968, Causes of distribution of Pelona, Rand, and Orocopia Schists along the San

Andreas and Garlock faults, in Dickinson, W.R., and Grantz, A., eds., Proceedings of

conference on geological problems of San Andreas fault system: Stanford University

Publications in Geological Sciences, v. 11, p. 294-305.

Ehlig, P.L., 1981, Origin and tectonic history of the basement terrane of the San Gabriel

Mountains, central Transverse Ranges, in Ernst, W.G., ed., The Geotectonic

Development of California (Rubey Volume I): Englewood Cliffs, New Jersey, Prentice-

Hall, p. 253-283.

118

Engebretson, D.C., Cox, A., and Gordon, R.G., 1985, Relative motions between oceanic and

continental plates in the Pacific basin: Geological Society of America Special Paper 206,

59 p., https://doi.org/10.1130/SPE206-p1

Frizzell, V.A., Jr., and Weigand, P.W., 1993, Whole-rock K-Ar ages and geochemical data from

middle Cenozoic volcanic rocks, southern California: A test of correlations across the

San Andreas fault, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San

Andreas fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 273-288,

https://doi.org/10.1130/MEM178-p273

Goodmacher, J., Barnett, L., Buckner, G., Ouachrif, L., Vidigal, A., and Frost, E., 1989, The

Clemens Well fault in the Orocopia Mountains of southern California: a strike-slip or

normal fault structure?: Geological Society of America Abstracts with Programs, v. 21,

no. 5, p. 85.

Goodman, E.D., and Malin, P.E., 1992, Evolution of the southern San Joaquin basin and mid-

Tertiary “transitional” tectonics, central California: Tectonics, v. 11, p. 478-498,

https://doi.org/10.1029/91TC02871

Goodman, E.D., Malin, P.E., Ambos, E.L., and Crowell, J.C., 1989, The southern San Joaquin

Valley as an example of Cenozoic basin evolution in California, in Price, R.A., ed.,

Origin and evolution of sedimentary basins and their energy and mineral resources:

American Geophysical Union Monograph 48, p. 87-107,

https://doi.org/10.1029/GM048p0087

Graham, S.A., DeCelles, P.G., Carroll, A.R., and Goodman, E.D., 1990, Middle Tertiary

contractile deformation, uplift, extension, and rotation in the San Emigdio Range,

119

southern California (abstract): American Association of Petroleum Geologists Bulletin, v.

74, p. 665.

Grove, M., and Jacobson, C.E., 2016, Fossil imprint of the Mendocino Fracture Zone across the

San Andreas fault: Geological Society of America Abstracts with Programs, v. 48, n. 7,

https://doi.org/10.1130/abs/2016AM-287480

Grove, M., Jacobson, C.E., Barth, A.P., and Vućić, A., 2003, Temporal and spatial trends of Late

Cretaceous–early Tertiary underplating of Pelona and related schist beneath southern

California and southwestern Arizona, in Johnson, S.E., Patterson, S.R., Fletcher, J.M.,

Girty, G.H., Kimbrough, D.L., and Martin-Barajas, A., eds., Tectonic Evolution of

Northwestern Mexico and the Southwestern USA: Geological Society of America Special

Paper 374, p. 381–406, https://doi.org/10.1130/0-8137-2374-4.381

Grubensky, M.J., Haxel, G.B., and Koch, R.D., 1993, Geologic map of the Castle Dome

Mountains, southwestern Arizona: U.S. Geological Survey Miscellaneous Investigations

Series Map I-2138, 1:62500.

Grubensky, M.J., Haxel, G.B., and Demsey, K.A., 1995, Geologic map of the southeastern Kofa

Mountains and western , southwestern Arizona: U.S. Geological Survey

Miscellaneous Investigations Series Map I-2454, 1:62500.

Hall, C.A., Jr., 1991, Geology of the Point Sur–Lopez Point Region, Coast Ranges, California; A

Part of the Southern California Allochthon: Geological Society of America Special Paper

266, 40 p., https://doi.org/10.1130/SPE266

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

120

California: International Geology Review, v. 55, p. 1575-1615,

https://doi.org/10.1080/00206814.2013.825141

Hamilton, W., 1969, Mesozoic California and the Underflow of Pacific Mantle: Geological

Society of America Bulletin, v. 80, p. 2409-2430, https://doi.org/10.1130/0016-

7606(1969)80[2409:MCATUO]2.0.CO;2

Harvill, L.L., 1969, Deformational history of the Pelona Schist, northwestern Los Angeles

County, California [Ph.D. thesis]: Los Angeles, University of California, 117 p.

Haxel, G.B., and Dillon, J.T., 1978, The Pelona-Orocopia Schist and Vincent–Chocolate

Mountain thrust system, southern California, in Howell, D.G., and McDougall, K.A.,

eds., Mesozoic Paleogeography of the Western United States: Los Angeles, Pacific

Section, SEPM (Society for Sedimentary Geology), Pacific Coast Paleogeography

Symposium 2, p. 453-469.

Haxel, G.B., Jacobson, C.E., Richard, S.M., Tosdal, R.M., and Grubensky, M.J., 2002, The

Orocopia Schist in southwest Arizona: Early Tertiary oceanic rocks trapped or

transported far inland, in Barth, A., ed., Contributions to Crustal Evolution of the

Southwestern United States: Geological Society of America Special Paper 365, p. 99-

128, https://doi.org/10.1130/0-8137-2365-5.99

Haxel, G.B., Jacobson, C.E., and Wittke, J.H., 2015, Mantle peridotite in newly discovered far-

inland subduction complex, southwest Arizona: initial report: International Geology

Review, v. 57, p. 871-892, https://doi.org/10.1080/00206814.2014.928916

Haxel, G.B., Beard, L.S., and Jacobson, C.E., 2018, Significance of northeast-southwest

orientation of prograde lineation in the Pelona-Orocopia-Rand-Schist low-angle

subduction complex, southern California and southwest Arizona: Geological Society of

121

America Abstracts with Programs. v. 50, no. 5, https://doi.org/10.1130/abs/2018RM-

313363

Hendrix, E.D., 1993, Soledad Basin, Central Transverse Ranges, Calif., in Sherrod, D.R., and

Nielson, J.E., eds., Tertiary stratigraphy of highly extended terranes, California, Arizona,

and Nevada: U.S. Geological Survey Bulletin 2053, p. 243-350.

Hendrix, E.D., and Ingersoll, R.V., 1987, Tectonics and alluvial sedimentation of the upper

Oligocene/lower Miocene Vasquez Formation, Soledad Basin, southern California:

Geological Society of America Bulletin, v. 98, p. 647-663, https://doi.org/10.1130/0016-

7606(1987)98<647:TAASOT>2.0.CO;2

Hendrix, E.D., Cole, R.B., and Ingersoll, R.V., 2010, Soledad and Plush Ranch basins: mid-

Tertiary extensional terrane dismembered by the San Andreas fault system, in Clifton,

H.E., and Ingersoll, R.V., eds., Geologic excursions in California and Nevada: tectonics,

stratigraphy and hydrogeology: Pacific Section, SEPM (Society for Sedimentary Geology)

Book 108, p. 103-171.

Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults, California: a

study of the character, history, and tectonic significance of their displacements:

Geological Society of America Buletin v. 64, p. 443-458, https://doi.org/10.1130/0016-

7606(1953)64[443:SAGABP]2.0.CO;2

Hornafius, J.S., Luyendyk, B.P., Terres, R.R., and Kamerling, M.J., 1986, Timing and extent of

Neogene tectonic rotation in the western Transverse Ranges, California: Geological

Society of America Bulletin, v. 97, p. 1476-1487, https://doi.org/10.1130/0016-

7606(1986)97<1476:TAEONT>2.0.CO;2

122

Hoyt, J.F., Coffey, K.T., Ingersoll, R.V., and Jacobson, C.E., 2018, Paleogeographic and

paleotectonic setting of the middle Miocene Mint Canyon and Caliente formations,

southern California: An integrated provenance study, in Ingersoll, R.V., Lawton, T.F, and

Graham, S.A., eds., Tectonics, Sedimentary Basins, and Provenance: A Celebration of

the Career of William R. Dickinson: Geological Society of America Special Paper 540, p.

463-480, https://doi.org/10.1130/2018.2540(21)

Ingersoll, R.V., 2019, Forearc strike-slip displacement as an alternative to subduction erosion,

with examples from Mexico and Calfornia (sinistral Nacimiento fault): Canadian Journal

of Earth Sciences, v. 56 (in press).

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Ingersoll, R.V., and Rumelhart, P.E., 1999, Three-stage evolution of the Los Angeles basin,

southern California: Geology, v. 27, p. 593-596, https://doi.org/10.1130/0091-

7613(1999)027<0593:TSEOTL>2.3.CO;2

Ingersoll, R.V., Pratt, M.J., Davis, P.M., Caracciolo, L., Day, P.P., Hayne, P.O., Petrizzo, D.A.,

Gingrich, D.A., Cavazza, W., Critelli, S., Diamond, D.S., Coffey, K.T., Stang, D.M.,

Hoyt, J.F., Reith, R.C., and Hendrix, E.D., 2014, Paleotectonics of a complex Miocene

half graben formed above a detachment fault: the Diligencia basin, Orocopia Mountains,

southern California: Lithosphere, v. 6, p. 157-176, https://doi.org/10.1130/L334.1

Jacobson, C.E., 1983a, Structural geology of the Pelona Schist and Vincent thrust, San Gabriel

Mountains, California: Geological Society of America Bulletin, v. 94, p. 753–767,

https://doi.org/10.1130/0016-7606(1983)94<753:SGOTPS>2.0.CO;2

123

Jacobson, C.E., 1983b, Relationship of deformation and metamorphism of the Pelona Schist to

movement on the Vincent thrust, San Gabriel Mountains, Southern California: American

Journal of Science, v. 283, p. 587-604, https://doi.org/10.2475/ajs.283.6.587

Jacobson, C.E., 1997, Metamorphic convergence of the upper and lower plates of the Vincent

thrust, San Gabriel Mountains, southern California, USA: Journal of Metamorphic

Geology, v. 15, p. 155-165, https://doi.org/10.1111/j.1525-1314.1997.00009.x

Jacobson, C.E., and Dawson, M.R., 1995, Structural and metamorphic evolution of the Orocopia

Schist and related rocks, southern California: Evidence for late movement on the

Orocopia fault: Tectonics, v. 14, p. 933-944, https://doi.org/10.1029/95TC01446

Jacobson, C.E., Dawson, M.R., and Postlethwaite, C.E., 1988, Structure, metamorphism, and

tectonic significance of the Pelona, Orocopia, and Rand Schists, southern California, in

Ernst, W.G., ed., Metamorphism and crustal evolution of the western United States:

Rubey Vol. VII, Prentice-Hall, Englewood Cliffs, New Jersey, p. 976-997.

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,

https://doi.org/10.1130/0091-7613(1996)024<0547:SAEOTP>2.3.CO;2

Jacobson, C.E., Grove, M., Stamp, M.M., Vućić, A., Oyarzabal, F.R., Haxel, G.B., Tosdal, R.M.,

and Sherrod, D.R., 2002, Exhumation history of the Orocopia Schist and related rocks in

the Gavilan Hills area of southeasternmost California, in Barth, A., ed., Contributions to

Crustal Evolution of the Southwestern United States: Geological Society of America,

Special Paper 365, p. 129–154, https://doi.org/10.1130/0-8137-2365-5.129

Jacobson, C.E., Grove, M., Vućić, A., Pedrick, J.N., and Ebert, K.A., 2007, Exhumation of the

Orocopia Schist and associated rocks of southeastern California: Relative rates of

124

erosion, synsubduction tectonic denudation, and middle Cenozoic extension, 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. 1-37, https://doi.org/10.1130/SPE419

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, https://doi.org/10.1130/B30238.1

Jacobson, C.E., Sharman, G.R., Hourigan, J.K., Grove, M., Johnston, S.M., Barth, A., Haxel,

G.B., Ingersoll, R.V., Chapman, A.D., Howard, J.L., and Wooden, J.L., 2018, Late

Cretaceous—Paleogene tectonic and paleogeographic evolution of southern California

based on detrital zircon from the forearc basin, Pelona-Orocopia-Rand Schists, and

Franciscan Complex: Geological Society of America Abstracts with Programs, v. 50, n.

5, https://doi.org/10.1130/abs/2018RM-313919

Jahns, R.H., and Muehlberger, W.P., 1954, Geology of the Soledad Basin, Los Angeles County:

California Division of Mines Bulletin 170, v. 2, Map Sheet 6, 1:85000.

Kanter, L.R., and McWilliams, M.O., 1982, Rotation of the southernmost Sierra Nevada,

California: Journal of Geophysical Research, v. 87, p. 3819-3830,

https://doi.org/10.1029/JB087iB05p03819

Kellogg, K.S., and Miggins, D.P., 2002, Geologic map of the Sawmill Mountain quadrangle,

Kern and Ventura Counties, California: U.S. Geological Survey Open-File Report 02-

406, 1:24000.

125

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. 353-

357, https://doi.org/10.1038/NGEO829

Liu, W., 1990, Paleomagnetism of Miocene sedimentary rocks in the Transverse Ranges: the

implications for tectonic history [Ph.D. thesis]: Pasadena, California Institute of

Technology, 233 p.

Luyendyk, B.P., Kamerling, M.J., and Terres, R., 1980, Geometric model for Neogene crustal

rotations in southern California: Geological Society of America Bulletin, v. 91, p. 211-

217, https://doi.org/10.1130/0016-7606(1980)91<211:GMFNCR>2.0.CO;2

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

Matti, J.C., Morton, D.M., and Cox, B.F., 1992, The San Andreas fault system in the vicinity of

the central Transverse Ranges province, Southern California: U.S. Geological Survey

Open-File Report 92-354, 49 p.

Matti, J.C., Morton, D.M., Cox, B.F., Carson, S.E., and Yetter, T.J., 2003, Geologic map and

digital database of the Yucaipa 7.5’ quadrangle, San Bernardino and Riverside Counties,

California: U.S. Geological Survey Open-File Report 03-301, 1:24000.

May, D.J., and Walker, N.W., 1989, Late Cretaceous juxtaposition of metamorphic terranes in

the southeastern San Gabriel Mountains, California: Geological Society of America

126

Bulletin, v. 101, p. 1246-1267, https://doi.org/10.1130/0016-

7606(1989)101<1246:LCJOMT>2.3.CO;2

McWilliams, M., and Li, Y., 1985, Oroclinal bending of the southern :

Science, v. 230, p. 172-175, https://doi.org/10.1126/science.230.4722.172

Miller, R.B., Paterson, S.R., Lebit, H., Alsleben, H., and Lüneburg, C., 2006, Siginificance of

composite lineations in the mid- to deep crust: a case study from the North Cascades,

Washington: Journal of Structural Geology, v. 28, p. 302-322,

https://doi.org/10.1016/j.jsg.2005.11.003

Muehlberger, W.R., and Hill, H.S., 1958, Geology of the central Sierra Pelona, Los Angeles

County, California: American Journal of Science, v. 256, p. 630-643,

https://doi.org/10.2475/ajs.256.9.630

Needy, S.K., Wooden, J.L., Barth, A.P., and Jacobson, C.E., 2007, Geochronology of igneous

rocks in the Chocolate Mountains region as a means to interpret tectonic evolution of

southeastern California: Geological Society of America Abstracts with Programs, v. 39,

n. 6, p. 407.

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. 491-495,

https://doi.org/10.1130/0091-7613(1994)022<0491:MCROTW>2.3.CO;2

Niemi, N.A., Buscher, J.T., Spotila, J.A., House, M.A., and Kelley, S.A., 2013, Insights from

low-temperature thermochronometry into transpressional deformation and crustal

exhumation along the San Andreas fault in the western Transverse Ranges, California:

Tectonics, v. 32, p. 1602-1622, https://doi.org/10.1002/2013TC003377

127

Nourse, J.A., 1989, Geological evolution of two crustal scale shear zones; Part I, The Rand thrust

complex, northwestern Mojave Desert, California; Part II, The Magdalena metamorphic

core complex, north central Sonora, Mexico [Ph.D. thesis]: Pasadena, California Institute

of Technology, 394 p.

Oskin, M., and Stock, J., 2003, Marine incursion synchronous with plate-boundary localization in

the Gulf of California: Geology, v. 31, p. 23-26, https://doi.org/10.1130/0091-

7613(2003)031<0023:MISWPB>2.0.CO;2

Oskin, M., Stock, J., and Martin-Barajas, A., 2001, Rapid localization of Pacific-North America

plate motion in the Gulf of California: Geology, v. 29, p. 459-462,

https://doi.org/10.1130/0091-7613(2001)029<0459:RLOPNA>2.0.CO;2

Oyarzabal, F.R., Jacobson, C.E., and Haxel, G.B., 1997, Extensional reactivation of the

Chocolate Mountains subduction thrust in the Gavilan Hills of southeastern California:

Tectonics, v. 16, p. 650-661, https://doi.org/10.1029/97TC01415

Peacock, S.M., 1987, Creation and preservation of subduction-related inverted metamorphic

gradients: Journal of Geophysical Research, v. 92, p. 12763-12781,

https://doi.org/10.1029/JB092iB12p12763

Plescia, J.B., and Calderone, G.J., 1986, Paleomagnetic constraints on the timing and extent of

rotation of the Tehachapi Mountains: Geological Society of America Abstracts with

Programs, v. 18, p. 171.

Plescia, J.B., Calderone, G.J., and Snee, L.W., 1994, Paleomagnetic analysis of Miocene basalt

flows in the Tehachapi Mountains, California: U.S. Geological Survey Bulletin 2100, 11

p.

128

Postlethwaite, C.E., and Jacobson, C.E., 1987, Early history and reactivation of the Rand thrust,

southern California: Journal of Structural Geology, v. 9, p. 195-205,

https://doi.org/10.1016/0191-8141(87)90025-3

Powell, R.E., 1981, Geology of the crystalline basement complex, eastern Transverse Ranges,

southern California: constraints on regional tectonic interpretation [Ph.D. thesis]:

Pasadena, California Institute of Technology, 441 p.

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

Prothero, D.R., 2006, Paleomagnetism and tectonic rotation of the southern Coast Ranges,

California, in Girty, G.H., and Cooper, J.D., eds., Using stratigraphy, sedimentology, and

geochemistry to unravel the geologic history of the southwestern Cordillera: a volume in

honor of Patrick L. Abbott: Society for Sedimentary Geology (SEPM), Pacific Section,

Book 101.

Prothero, D.R., Kelly, T.S., Mccardel, K.J., and Wilson, E.L., 2008, Magnetostratigraphy,

biostratigraphy, and tectonic rotation of the Miocene Caliente Formation, Ventura

County, California: New Mexico Museum of Natural History and Science Bulletin, v. 44,

p. 225-272.

129

Richard, S.M., 1993, Palinspastic reconstruction of southeastern California and southwestern

Arizona for the middle Miocene: Tectonics, v. 12, p. 830-854,

https://doi.org/10.1029/92TC02951

Ricketts, J.W., Girty, G.H., Sainsbury, J.S., Muela, K.K., Sutton, L.A., Biggs, M.A., and Voyles,

E.M., 2011, Episodic growth of the Chocolate Mountains anticlinorium recorded by the

Neogene Bear Canyon conglomerate, southeastern California, U.S.A.: Journal of

Sedimentary Research, v. 81, p. 859-873, https://doi.org/10.2110/jsr.2011.66

Ridley, J., 1986, Parallel stretching lineations and fold axes oblique to a shear displacement

direction – a model and observations: Journal of Structural Geology, v. 8, p. 647-653,

https://doi.org/10.1016/0191-8141(86)90070-2

Robinson, K.L., and Frost, E.G., 1996, Orocopia Mountains detachment system: progressive

development of a tilted crustal slab and half-graben sedimentary basin during regional

extension, in Abbott, P.L., and Cooper, J.D., eds., Field Conference Guide 1996: Pacific

Section, AAPG GS 73 and Pacific Section, SEPM (Society for Sedimentary Geology)

Book 80, p. 277-284.

Saleeby, J., 2003, Segmentation of the Laramide Slab – evidence from the southern Sierra

Nevada region: Geological Society of America Bulletin, v. 115, p. 655-668,

https://doi.org/10.1130/0016-7606(2003)115<0655:SOTLSF>2.0.CO;2

Shafiqullah, M., Damon, P.E., Lynch, D.J., Reynolds, S.J., Rehrig, W.A., and Raymond, R.H.,

1980, K-Ar geochronology and geologic history of southwestern Arizona and adjacent

areas: Arizona Geological Society Digest, v. 12, p. 201-260.

130

Sherrod, D.R., and Tosdal, R.M., 1991, Geologic setting and Tertiary structural evolution of

southwestern Arizona and southeastern California: Journal of Geophysical Research, v.

96, p. 12407-12423, https://doi.org/10.1029/90JB02688

Simpson, C., 1986, Microstructural evidence for northeastward movement on the Vincent-

Orocopia- Chocolate Mountain thrust system: Geological Society of America Abstracts

with Programs, v. 18, n. 2, p. 185.

Singleton, J.S., and Cloos, M., 2013, Kinematic analysis of mélange fabrics in the Franciscan

Complex near San Simeon, California: Evidence for sinistral slip on the Nacimiento fault

zone?: Lithosphere, v. 5, p. 179-188, https://doi.org/10.1130/L259.1

Smith, D.P., 1977, San Juan-St Francis fault – Hypothesized major middle tertiary right-lateral

fault in central and southern California: California Division of Mines and Geology

Special Report 129, p. 41-50.

Spotila, J.A., House, M.A., Niemi, N.A., Brady, R.C., Oskin, M., and Buscher, J.T., 2007,

Patterns of bedrock uplift along the San Andreas fault and implications for mechanisms

of transpression, 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. 15-33, https://doi.org/10.1130/SPE434

Stein, R.S., and Thatcher, W., 1981, Seismic and aseismic deformation associated with the 1952

Kern County, California, earthquake and relationship to the Quaternary history of the

White Wolf Fault: Journal of Geophysical Research, v. 86, p. 4913-4928,

https://doi.org/10.1029/JB086iB06p04913

131

Stock, J., and Molnar, P., 1988, Uncertainties and implications of the Late Cretaceous and

Tertiary position of North America relative to the Farallon, Kula, and Pacific plates:

Tectonics, v. 7, p. 1339-1384, https://doi.org/10.1029/TC007i006p01339

Strickland, E.D., Singleton, J.S., and Haxel, G.B., 2018, Orocopia Schist in the northern Plomosa

Mountains, west-central Arizona: A Laramide subduction complex exhumed in a

Miocene metamorphic core complex: Lithosphere, v. 10, p. 723-742,

https://doi.org/10.1130/L742.1

Terres, R.R., 1984, Paleomagnetism and tectonics of the central and eastern Transverse Ranges,

southern California [Ph.D. thesis]: Santa Barbara, University of California, 323 p.

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Tikoff, B., and Greene, D., 1997, Stretching lineations in transpressional shear zones: an

example from the Sierra Nevada Batholith, California: Journal of Structural Geology, v.

19, p. 29-39, https://doi.org/10.1016/S0191-8141(96)00056-9

Tikoff, B., and Teyssier, C., 1994, Strain modeling of displacement-field partitioning in

transpressional orogens: Journal of Structural Geology, v. 16, p. 1575-1588,

https://doi.org/10.1016/0191-8141(94)90034-5

Turner, D.L., 1970, Potassium-argon dating of Pacific Coast Miocene foraminiferal stages, in

Bandy, O.L., ed., Radiometric dating and paleontologic zonation: Geological Society of

America Special Paper 124, p. 91-130, https://doi.org/10.1130/SPE124-p91

132

Veseth, M.K., 1985, Paleomagnetics of Mid-Tertiary and younger volcanic rocks of the Kofa

Mountains, southeastern Arizona [M.S. thesis]: San Diego, San Diego State University,

168 p.

Xia, H., and Platt, J.P., 2017, Structural and rheological evolution of the Laramide subduction

channel in southern California: Solid Earth, v. 8, p. 379-403, https://doi.org/10.5194/se-8-

379-2017

Xia, H., and Platt, J.P., 2018, Is the Vincent fault in southern California the Laramide subduction

zone megathrust?: Geological Society of America Bulletin, v. 131, p. 120-136,

https://doi.org/10.1130/B31870.1

Yeats, 1968, Southern California structure, sea-floor spreading, and history of the Pacific Basin:

Geological Society of America Bulletin, v. 79, p. 1693-1702,

https://doi.org/10.1130/0016-7606(1968)79[1693:SCSSSA]2.0.CO;2

Yin, A., 2002, Passive-roof thrust model for the emplacement of the Pelona-Orocopia Schist in

southern California, United States: Geology, v. 30, p. 183-186,

https://doi.org/10.1130/0091-7613(2002)030<0183:PRTMFT>2.0.CO;2

133

CHAPTER 3: STRAIN PARTITIONING EXPLAINS

DISPARATE CROSS-FAULT CORRELATIONS ALONG

THE SAN ANDREAS FAULT, SOUTHERN CALIFORNIA,

U.S.A.

ABSTRACT

When strain partitioning is taken into account, disparate cross-fault correlations along the southern San Andreas fault become compatible. Comparison of these correlations has resulted in decades of debate, predicated on the assumption that cumulative slip on the Mojave segment of the San Andreas fault should be the same as that on adjacent segments. However, along the anomalously oriented Mojave segment of the San Andreas fault, dextral slip experienced by adjacent segments is accommodated by a combination of dextral slip on the fault itself, and a zone of north-south shortening beside it. Therefore, it is the vector sum of the dextral slip along this segment and the magnitude of north-south shortening beside it that should equal the dextral slip on adjacent segments of the San Andreas fault. This relationship allows reconciliation of the approximately 215 km of cumulative dextral slip implied for the San Andreas fault system

(including slip on the San Jacinto fault, but less slip transferred off the San Andreas fault in the eastern Transverse Ranges) south of the Mojave segment with the approximately 160 km of cumulative dextral slip implied along the Mojave segment.

134

INTRODUCTION

The San Andreas fault is important both because of its seismic hazard and because it is a well studied example of a continental transform plate boundary. For most of its length, including along the Coachella Valley (the Coachella Valley segment; Matti and Morton, 1993), the San

Andreas fault is subparallel to the direction of relative motion of the Pacific and North America plates (e.g., DeMets et al., 1990; Sauber et al., 1994); northeast of Los Angeles, however, the approximately 255 km Mojave segment (Matti and Morton, 1993; Sauber et al., 1994) is oriented about 25° counterclockwise to this typical orientation (Fig. 3-1). This geometry, termed a restraining double bend by McClay and Bonora, 2001, results in strike-slip motion being accompanied by significant shortening (transpression).

A long-standing cross-fault correlation along the San Andreas fault in southern California is between conglomerate clasts in the Mint Canyon Formation (SW side) interpreted to have been derived from rapakivi-textured quartz-latite porphyry in the northern Chocolate Mountains

(NE side), implying approximately 240 km of dextral slip (Ehlig and Ehlert, 1972; Ehlig et al.,

1975; Joseph and Davis, 1977; Ehlert, 2000, 2003; Fig. 3-1). This correlation builds upon, and is supported by, correlation of basement rocks, Cenozoic strata, and structural features within the same areas (e.g., Hill and Dibblee, 1953; Crowell, 1962, 1975; Carman, 1964). This correlation, however, has been challenged based on correlation of Triassic megaporphyritic monzogranite outcrops on either side of the Mojave segment of the San Andreas fault, which implies 160 ± 10 km of dextral slip (Matti et al., 1985; Frizzell et al., 1986; Matti and Morton, 1993; Fig. 3-1). All correlated units are older than the southern San Andreas fault (e.g., Terres and Luyendyk, 1985;

Matti and Morton, 1993), which initiated ca. 8-5 Ma (e.g., Nicholson et al., 1994; Ingersoll and

135

Rumelhart, 1999; Oskin et al., 2001; Crowell, 2003; Oskin and Stock, 2003, Dorsey et al., 2011), and thus represent estimates of cumulative slip.

The above correlations are apparently contradictory regarding the magnitude of cumulative fault offset, which has resulted in decades of debate. Most attempts to reconcile them

Figure 3-1: Map of southern California, showing Mojave (SAf-M) and Coachella Valley (SAf-CV)

segments of the San Andreas fault, and related strike-slip faults, as well as approximate extent of zone of north-south shortening, and cross-fault correlations discussed in text (orientation of tie-line connecting MCF to San Andreas fault based on inferred orientation of MCF paleodrainage; Ehlert,

2003). Approximate amounts of cumulative strike slip on dextral faults other than San Andreas are

given in parentheses. approx. = approximate; ETR = eastern Transverse Ranges; Gf = Garlock

fault; LA = Los Angeles; MCF = Mint Canyon Formation; PRb = batholith;

SBM = San Bernardino Mountains; SEM = San Emigdio Mountains; SGM = San Gabriel

Mountains; SHf = Sheep Hole fault; SJf = San Jacinto fault; SNb = Sierra Nevada batholith; TM =

Tehachapi Mountains. Figure modified from Darin and Dorsey (2013).

136 have involved discarding one of them as spurious (e.g., Meisling and Weldon, 1986; Matti and

Morton, 1993; Powell, 1993). Darin and Dorsey (2013) proposed a reconciliation in which both are valid. They suggested that extension associated with rotation of the eastern Transverse

Ranges (ETR; Fig. 3-1), too far south to have affected separation of the megaporphyritic monzogranite, was responsible for 20-25 km of separation of the Mint Canyon Formation from the quartz-latite porphyry of the northern Chocolate Mountains. They further suggested, using a geometric argument based on eastward extrapolation of paleocurrent data from the Mint Canyon

Formation, that the Mint Canyon Formation / northern Chocolate Mountains correlation implies only 200 ± 14 km of slip, rather than the previously agreed upon approximately 240 km. These proposals reduce the discrepancy in slip implied by the two correlations to approximately 20 km or less, which Darin and Dorsey (2013) suggested may have been transferred off the San

Andreas fault south of the Mojave segment, into the eastern California shear zone and/or into distributed shortening southwest of the Mojave segment.

I agree with Darin and Dorsey’s (2013) assertion that extension associated with rotation of the ETR added approximately 25 km of dextral slip to the Coachella Valley segment of the

San Andreas fault south of the ETR. This extension likely represents transfer of approximately

25 km of dextral slip from the Sheep Hole fault: approximately 30 km of dextral slip has been proposed for the Sheep Hole fault to the north (McQuarrie and Wernicke, 2005), but only 5 km or less is suggested along the ETR (Jagiello et al., 1992; Langenheim and Powell, 2009). Darin and Dorsey’s (2013) argument for reducing the approximately 240 km cumulative-slip estimate to 200 ± 14 km is reasonable, using paleocurrent data from the Mint Canyon Formation to reinterpret the orientation of the drainage system that transported clasts into this formation from quartz-latite porphyry in the northern Chocolate Mountains. However, because this argument

137 relies on extrapolation of these paleocurrent data eastward across the present San Gabriel

Mountains and San Andreas fault, the original estimate also remains viable (i.e., these paleocurrent data may not be representative of the entire length of the drainage). Below, I argue that, after compensating for the approximately 25 km of slip transferred off of the San Andreas fault by the eastern Transverse Ranges, accounting for strain partitioning along the Mojave segment of the San Andreas fault brings the two estimates into closer, potentially complete agreement.

STRAIN-PARTITIONING MODEL

Because of its anomalous orientation within a restraining double bend, if the Mojave segment of the San Andreas fault accommodated the same differential motion as the Coachella

Valley segment, its slip would be oblique reverse-dextral. Instead, only a dextral component of this differential motion is accommodated by the Mojave segment, which is almost purely strike- slip (e.g., Sieh, 1978; Savage, 1983; Argus et al., 2005). The remaining differential motion is accommodated within a zone of shortening beside it, including the San Gabriel, San Bernardino,

San Emigdio, and Tehachapi Mountains (e.g., Stein and Thatcher, 1981; Meisling and Weldon,

1989; Argus et al., 2005; Niemi et al., 2013; Fig. 3-1), which may extend offshore (Weldon and

Humphreys, 1986); an approximately north-south present orientation of shortening is implied by north/south-directed reverse faults in the San Bernardino (e.g., Meisling and Weldon, 1989;

Spotila and Sieh, 2000), San Gabriel (e.g., Crowell, 1973), and San Emigdio (e.g., Keller et al.,

2000) Mountains, as well as by geodetic data (e.g., Argus et al., 2005). This is a straightforward example of strain partitioning in an oblique reverse-strike-slip fault system, a phenomenon well described by Jones and Tanner (1995). Thus, the assumption that the Mojave segment should have experienced the same magnitude of strike slip as the Coachella Valley segment (including 138 slip on the San Jacinto fault) is fundamentally wrong. Rather, it is the vector sum of the strike slip experienced by the Mojave segment and the shortening that should equal the strike slip experienced by the Coachella Valley segment (Fig. 3-2A). It is widely recognized that shortening

Figure 3-2: Diagrams showing vector-sum relationship of differential motion within the restraining

double bend of San Andreas fault. Dextral slip rate along the Mojave segment and north-south shortening rate within a zone beside it sum to total differential rate of motion, which is equivalent

to dextral slip rate on the Coachella Valley segment. θ is angle between Coachella Valley and

Mojave segments. γ is angle between Coachella Valley segment and north-south. Faults and zone of shortening shown in Figure 3-1. Mathematical relationships between vectors given in Appendix B.

A: General proposed relationship. B: Modern slip rates. Total rate of differential motion implied by geodetic data (Shen et al., 1996) is decomposed into dextral-slip and north-south-shortening rates as in part A. C: Cumulative slip, geometry of restraining double bend fixed. Approximately 215 km of

total differential motion implied by separation of Mint Canyon Formation from northern

Chocolate Mountains is decomposed into cumulative dextral slip and north-south shortening. D:

Cumulative slip, with geometry of restraining double bend varying with time. Same as part C,

except θ increases from 14.4° to 25° as slip accumulates, according to expression derived in

Appendix C (γ remains fixed). See Appendix D for details. Note agreement of resulting 160 km

(part C) and 167 km (part D) estimates for dextral slip on the Mojave segment with 160 ± 10 km implied by separation of megaporphyritic monzogranite. See Fig. 3-3 for modification of part C to

incorporate slip on Garlock fault. 139

along the restraining double bend contributes to Pacific-North America relative plate motion

(e.g., Bohannon and Howell, 1982; Bird and Rosenstock, 1984; Weldon and Humphreys, 1986;

Spotila and Sieh, 2000; Argus et al., 2005; but see Weldon and Humphreys, 1986), and some studies have suggested this shortening as a way to accommodate some of the cumulative dextral slip of the Coachella-Valley-segment San Andreas fault or the San Jacinto fault (e.g., Anderson,

1971; Matti and Morton, 1993; Darin and Dorsey, 2003; McGill et al., 2013). However, the vector-sum relationship of cumulative slip along different segments of the San Andreas fault has not, to my knowledge, been explicitly suggested.

Modern slip rates are compatible with this vector-sum relationship, supporting its validity. Geodetic data collected throughout the region surrounding the Mojave segment of the

San Andreas fault indicate approximately 32 mm/yr of relative motion, approximately parallel to the Coachella Valley segment (Shen et al., 1996; Fig. 3-2B). This is compatible with an estimated 35 mm/yr of relative motion surrounding the Coachella Valley segment

(approximately 20 mm/yr on the Coachella Valley segment itself and approximately 15 mm/yr on the San Jacinto fault: Field et al., 2013, and references therein; see also discussion in McGill et al., 2013). It is also compatible with the 34 mm/yr slip rate measured along the San Andreas fault north of the Mojave segment (Sieh and Jahns, 1984). Decomposing this 32 mm/yr into

Mojave-segment strike slip and north-south shortening (see Appendix B for trigonometric equations) yields a slip rate for the Mojave segment of 24 mm/yr (Fig. 3-2B). Slip-rate estimates for the Mojave segment range from 14 to 36 mm/yr, with those based on geodetic data generally lower than those based on geologic offsets (see discussion and references in Bird, 2009; Chuang and Johnson, 2011; O’Banion, 2012). However, 24 mm/yr is close to some of the more recent

140 estimates, based both on geologic offsets, i.e., 21.9 ± 3.85 mm/yr (calculated from all available previous estimates by Bird, 2009) and 25-40 mm/yr (best estimate of the Uniform California

Earthquake Rupture Forecast, Version 3; Field et al., 2013; see also references therein), and geodetic data, i.e., 25.1 ± 0.3 mm/yr (McCaffrey, 2005), 26.0 ± 1.5 mm/yr (Chuang and Johnson,

2011), and 25.5 ± 0.5 mm/yr (Daout et al., 2016). The vector-sum relationship yields a rate of north-south shortening of 14 mm/yr, identical to the 14 mm/yr of shortening predicted beside the southeastern part of the Mojave segment by the model of Bird and Rosenstock (1984).

Modeling the relationship between cumulative displacements is more complicated, as the directions of relative motion may have varied with time. A simple, first-order approximation is to assume that the geometry of the restraining double bend has remained fixed (i.e., constant θ and

γ; Fig. 3-2), such that the vector-sum relationship between cumulative displacements is identical to that between modern velocities. The Mint Canyon Formation / northern Chocolate Mountains correlation spans much of both the Mojave and Coachella Valley segments. However, because the Mint Canyon Formation is well away from the Mojave segment, at the edge of the zone of shortening (Fig. 3-1), nearly all of this shortening has contributed to its separation from the northern Chocolate Mountains. Thus, after subtracting slip transferred to the Sheep Hole fault, this correlation gives an estimate of 215 km of total differential motion accumulated within the restraining double bend. The fixed-geometry, vector-sum relationship predicts approximately

160 km of cumulative dextral slip on the Mojave segment (Fig. 3-2C), in very close agreement with the 160 ± 10 km implied by separation of the megaporphyritic monzogranite.

As alluring as the above reconciliation is, it may not be entirely realistic. Two assumptions are a fixed geometry for the restraining double bend (i.e., constant θ and γ; Fig. 3-

2), which has likely tightened through time (Garfunkel, 1974; Bohannon and Howell, 1982; but

141 see Bird and Rosenstock, 1984), and that potential effects of the Garlock fault are insignifcant.

The ends of the restraining double bend wrap around the ends of major batholiths (the Sierra

Nevada and Peninsular Ranges batholiths; Fig. 3-1). Differential motion of these batholiths due to slip on the San Andreas fault has likely been responsible for tightening the restraining double bend (e.g., Bohannon and Howell, 1982; Powell, 2016; Ingersoll and Coffey, 2017). Thus, using published estimated ages of initiation of the southern San Andreas and San Jacinto faults, and assuming a constant combined San Andreas/San Jacinto slip rate, cumulative displacements on these faults can be used to model the evolution of the restraining double bend through time (see

Appendix C for details and references). Such a geometric model yields a tightening (Δθ) of

10.6°. This is well within the uncertainty of the 16° ± 30° of clockwise rotation of the Mint

Canyon Formation implied by paleomagnetic data (Terres and Luyendyk, 1985), which was likely caused by tightening of the restraining double bend (Terres and Luyendyk, 1985; Ingersoll and Coffey, 2017), although rotation during entry into the restraining double bend is an alternative possibility (Ingersoll and Coffey, 2017). This model also yields an expression for the tightening of the bend through time (θ(t); see Appendix C). This can be used to calculate

Mojave-segment-strike-slip and adjacent shortening in a restraining double bend of changing geometry (see Appendix D for details). Again assuming approximately 215 km of total differential motion, and a constant γ, predicted cumulative dextral slip on the Mojave segment is approximately 167 km (Fig. 3-2D), within uncertainty of the 160 ± 10 km implied by separation of the megaporphyritic monzogranite.

Different tectonic models have been proposed for the sinistral Garlock fault (see discussion in McGill et al., 2009). One interpretation is that the Garlock fault is a conjugate shear to the San Andreas fault, accommodating eastward extrusion of the intervening Mojave block

142

(Hill and Dibblee, 1953). If so, and if the northern Chocolate Mountains are outside of the

Mojave block, then this motion must be included in the vector-sum relationship (Fig. 3-3A).

Modifying the fixed-geometry model to include either 12 km, 32 km, or 64 km of San-Andreas- related sinistral slip on the western Garlock fault predicts Mojave-segment dextral slip of approximately 171 km, 188 km, or 216 km, respectively (Fig. 3-3B-D; see Appendix E for calculations and references for Garlock fault slip). The first of these estimates overlaps in uncertainty with the 160 ± 10 km implied by separation of the megaporphyritic monzogranite; the other two are too large, and thus would require additional reconciliation. Such reconciliation is possible via the revised orientation of the Mint Canyon drainage system proposed by Darin and Dorsey (2013), which reduces the predicted total differential motion within the restraining double bend of the San Andreas fault from approximately 215 km to approximately 175 km.

With this revised value, 32 km of San Andreas-related Garlock-fault slip predicts 159 km of cumulative dextral slip on the Mojave segment (Fig. 3-3F), in good agreement with the the 160 ±

10 km implied by separation of the megaporphyritic monzogranite. For 64 km of estimated San

Andreas-related Garlock-fault slip, 187 km of cumulative dextral slip is predicted for the Mojave segment (Fig. 3-3G). Predicted amounts of cumulative north-south shortening, which may be unrealistically large in the first models presented (Fig. 3-2C, D), are progressively reduced with larger estimates of cumulative, San Andreas-related Garlock-fault slip (Fig. 3-3).

All of the above models are simplifications. A closer approximation of strain partitioning along the Mojave segment of the San Andreas fault could likely be achieved by allowing for changes in orientations through time of not only the Mojave segment itself (i.e., changing θ), but also of the direction of shortening (i.e., changing γ) and, if Garlock-fault slip is relevant, of the

Garlock fault (i.e., changing α). However, the models presented here demonstrate that even

143

Figure 3-3: Diagrams showing vector-sum relationships of differential motion within the

restraining double bend of San Andreas fault, as in Fig. 3-2C, except with addition of slip on

Garlock fault related to eastward extrusion of Mojave block added as an additional component.

Mathematical relationships between vectors, and discussion and references of cumulative slip on

Garlock fault and its possible relationship to San Andreas fault given in Appendix E.

simplified, first-order approximations of strain partitioning along the Mojave segment of the San

Andreas fault can bring long-debated and otherwise-conflicting cumulative-slip estimates for the

San Andreas fault into agreement.

CONCLUSIONS

When strain partitioning within the restraining double bend of the San Andreas fault is taken into account, cross-fault correlations that have been regarded as contradictory become more compatible, and potentially corroborative. In a simple model in which the geometry of the restraining double double bend is fixed, the approximately 215 km of combined, cumulative dextral slip on the San Andreas and San Jacinto faults south of the restraining double bend, 144 implied by separation of the Mint Canyon Formation from the northern Chocolate Mountains, is partitioned within this bend into approximately 160 km of dextral slip on the San Andreas fault, consistent with the 160 ± 10 km separation of a Triassic megaporphyritic monzogranite, and approximately 100 km of north-south shortening beside it. Modification of this model to account for either tightening of the restraining double bend through time, or eastward extrusion of the

Mojave block via sinistral slip on the Garlock fault, results in higher estimates of cumulative dextral slip on the San Andreas fault within the bend, and lower estimates of cumulative north- south shortening. Extrapolation of estimates of either slip rate or cumulative slip from the segment of the San Andreas fault within the restraining double bend to those outside it, or vice versa, ignores strain partitioning, and is thus unsound.

145

REFERENCES CITED

Allen, C.R., 1957, San Andreas fault zone in San Gorgonio Pass, southern California: Geological

Society of America Bulletin, v. 68, p. 315-350, https://doi.org/10.1130/0016-

7606(1957)68[315:SAFZIS]2.0.CO;2

Anderson, 1971, The San Andreas fault: Scientific American, v. 225, p. 52-71.

Argus, D.F., Heflin, M.B., Peltzer, G., Crampé, F., and Webb, F.H., 2005, Interseismic strain

accumulation and anthropogenic motion in metropolitan Los Angeles: Journal of

Geophysical Research, v. 110, B04401, https://doi.org/10.1029/2003JB002934

Bird, P., 2009, Long‐term fault slip rates, distributed deformation rates, and forecast of

seismicity in the western United States from joint fitting of community geologic,

geodetic, and stress direction data sets: Journal of Geophysical Research, v. 114, B11403,

https://doi.org/10.1029/2009JB006317

Bird, P., and Rosenstock, R.W., 1984, Kinematics of present crust and mantle flow in southern

California: Geological Society of America Bulletin, v. 95, p. 946-957,

https://doi.org/10.1130/0016-7606(1984)95<946:KOPCAM>2.0.CO;2

Bohannon, R.G., and Howell, D.G., 1982, Kinematic evolution of the junction of the San

Andreas, Garlock, and Big Pine faults, California: Geology, v. 10, p. 358-363,

https://doi.org/10.1130/0091-7613(1982)10<358:KEOTJO>2.0.CO;2

Carman, M.F., Jr., 1964, Geology of the Lockwood Valley area, Kern and Ventura Counties,

California: California Division of Mines and Geology Special Report 81, 62 p.

Chuang, R.Y., and Johnson, K.M., 2011, Reconciling geologic and geodetic model fault slip-rate

discrepancies in southern California: Consideration of nonsteady mantle flow and lower

crustal fault creep: Geology, v. 39, p. 627-630, https://doi.org/10.1130/G32120.1

146

Crowell, J.C., 1962, Displacement along the San Andreas fault, California: Geological Society of

America Special Paper 71, 61 p., https://doi.org/10.1130/SPE71

Crowell, J.C., 1973, Problems concerning the San Andreas fault system in southern California, in

Proceedings of the conference on tectonic problems of the San Andreas fault system:

Stanford University Publication, Geological Sciences, v. 13: Stanford University,

California.

Crowell, J.C., 1975, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Crowell, J.C., 2003, Introduction to geology of Ridge Basin, southern California, in Crowell,

J.C., ed., Evolution of Ridge Basin, southern California: An interplay of sedimentation

and tectonics: Geological Society of America Special Paper 367, p. 1-15,

https://doi.org/10.1130/0-8137-2367-1.1

Daout, S., Barbot, S., Peltzer, G., Doin, M.-P., Liu, Z., and Jolivet, R., 2016, Constraining the

kinematics of metropolitan Los Angeles faults with a slip-partitioning model:

Geophysical Research Letters, v. 43, p. 11192-11201,

https://doi.org/10.1002/2016GL071061

Darin, M.H., and Dorsey, R.J., 2013, Reconciling disparate estimates of total offset on the

southern San Andreas fault: Geology, v. 41, p. 975-978,

https://doi.org/10.1130/G34276.1

DeMets, C., Gordon, R.G., Argus, D.F., and Stein, S., 1990, Current plate motions: Geophysical

Journal International, v. 101, p. 425-478, https://doi.org/10.1111/j.1365-

246X.1990.tb06579.x

147

Dorsey, R.J., Housen, B.A., Janecke, S.U., Fanning, C.M., and Spears, A.L.F., 2011,

Stratigraphic record of basin development within the San Andreas fault system: Late

Cenozoic Fish Creek-Vallecito basin, southern California: Geological Society of America

Bulletin, v. 123, p. 771-793, https://doi.org/10.1130/B30168.1

Ehlert, K.W., 2000, Offset of the Miocene Mint Canyon and Caliente Formations along the San

Andreas and San Gabriel Faults, southern California [M.S. Thesis]: Los Angeles,

California State University, 115 p.

Ehlert, K.W., 2003, Tectonic significance of the middle Miocene Mint Canyon and Caliente

Formations, southern California, in Crowell, J.C., ed., Evolution of Ridge Basin, southern

California: An interplay of sedimentation and tectonics: Geological Society of America

Special Paper 367, p. 113-130, https://doi.org/10.1130/0-8137-2367-1.113

Ehlig, P.L., and Ehlert, K.W., 1972, Offset of Miocene Mint Canyon Formation from volcanic

source along San Andreas fault, Southern California: Geological Society of America

Abstracts with Programs, v. 4, p. 154.

Ehlig, P.L., Ehlert, K.W., and Crowe, , B.M., 1975, Offset of the upper Miocene Caliente and

Mint Canyon formations along the San Gabriel and San Andreas faults: California

Division of Mines and Geology Special Report 118 p. 83-92.

Field, E.H., Biasi, G.P., Bird, P., Dawson, T.E., Felzer, K.R., Jackson, D.D., Johnson, K.M.,

Jordan, T.H., Madden, C., Michael, A.J., Milner, K.R., Page, M.T., Parsons, T., Powers,

P.M., Shaw, B.E., Thatcher, W.R., Weldon, R.J., II, Zeng, Y., 2013, Uniform California

Earthquake Rupture Forecast, version 3 (UCERF3): The time-independent model: U.S.

Geological Survey Open-File Report 2013–1165 / California Geological Survey Special

148

Report 228 / Southern California Earthquake Center Publication 1792, 97 p.,

http://pubs.usgs.gov/of/2013/1165/

Frizzell, V.A., Jr., Mattinson, J.M., and Matti, J.C., 1986, Distinctive Triassic megaporphyritic

monzogranite: evidence for only 160 km offset along the San Andreas fault, southern

California: Journal of Geophysical Research, v. 91, p. 14080-14088,

https://doi.org/10.1029/JB091iB14p14080

Garfunkel, Z., 1974, Model for the Late Cenozoic tectonic history of the Mojave Desert,

California, and for its relation to adjacent regions: Geological Society of America

Bulletin, v. 85, p. 1931-1944, https://doi.org/10.1130/0016-

7606(1974)85<1931:MFTLCT>2.0.CO;2

Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults, California: a

study of the character, history, and tectonic significance of their displacements:

Geological Society of America Bulletin v. 64, p. 443-458, https://doi.org/10.1130/0016-

7606(1953)64[443:SAGABP]2.0.CO;2

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Ingersoll, R.V., and Rumelhart, P.E., 1999, Three-stage evolution of the Los Angeles basin,

southern California: Geology, v. 27, p. 593-596, https://doi.org/10.1130/0091-

7613(1999)027<0593:TSEOTL>2.3.CO;2

Jagiello, K., Christie, J.M., and Blom, R.G., 1992, Horizontal separation of major late Cenozoic

strike-slip faults in the Twentynine Palms region, Mojave Desert, California, in Richard,

S.M., ed., Deformation associated with the Neogene Eastern California shear zone,

149

southwestern Arizona and southeastern California: San Bernardino County Museum

Special Publication 92–1, p. 48–53.

Jones, R.R., and Tanner, P.W.G., 1995, Strain partitioning in transpression zones: Journal of

Structural Geology, v. 17, p. 793-802, https://doi.org/10.1016/0191-8141(94)00102-6

Joseph, S.E., and Davis, T.E., 1977, 87Sr/86Sr correlation of rapakivi-textured porphyry to

measure offset on the San Andreas fault: Geological Society of America Abstracts with

Programs, v. 9, p. 443.

Keller, E.A., Seaver, D.B., Laduzinsky, D.L., Johnson, D.L., and Ku, T.L., 2000, Tectonic

geomorphology of active folding over buried reverse faults: front,

southern , California: Geological Society of America Bulletin, v. 112,

p. 86-97, https://doi.org/10.1130/0016-7606(2000)112<86:TGOAFO>2.0.CO;2

Langenheim, V.E., and Powell, R.E., 2009, Basin geometry and cumulative offsets in the Eastern

Transverse Ranges, southern California: Implications for transrotational deformation

along the San Andreas fault system: Geosphere v. 5, p. 1–22,

https://doi.org/10.1130/GES00177.1

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

Matti, J.C., Morton, D.M., and Cox, B.F., 1985, Distribution and geologic relations of fault

systems in the vicinity of the central Transverse Ranges, Southern California: U.S.

Geological Survey Open-File Report 85-365, https://doi.org/10.3133/ofr85365

150

Matti, J.C., Morton, D.M., and Cox, B.F., 1992, The San Andreas fault system in the vicinity of

the central Transverse Ranges province, Southern California: U.S. Geological Survey

Open-File Report 92-354, 49 p., https://doi.org/10.3133/ofr92354

McCaffrey, R., 2005, Block kinematics of the Pacific-North America plate boundary in the

southwestern United States from inversion of GPS, seismologoical, and geologic data:

Journal of Geophysical Research, v. 110, B07401, https://doi.org/10.1029/2004JB003307

McClay, K., and Bonora, M., 2001, Analog models of restraining stepovers in strike-slip fault

systems: American Association of Petroleum Geologists Bulletin, v. 85, p. 233-260,

https://doi.org/10.1306/8626C7AD-173B-11D7-8645000102C1865D

McGill, S.F., Wells, S.G., Fortner, S.K., Kuzma, H.A., and McGill, J.D., 2009, Slip rate of the

western Garlock fault, at Clark Wash, near Lone Tree Canyon, Mojave Desert,

California: Geological Society of America Bulletin, v. 121, p. 536-554,

https://doi.org/10.1130/B26123.1

McGill, S.F., Owen, L.A., Weldon, R.J., II, and Kendrick, K.J., 2013, Latest Pleistocene and

Holocene slip rate for the San Bernardino strand of the San Andreas fault, Plunge Creek,

Southern California: Implications for strain partitioning within the southern San Andreas

fault system for the last ~35 k.y.: Geological Society of America Bulletin, v. 125, p. 48-

72, https://doi.org/10.1130/B30647.1

McQuarrie, N., and Wernicke, B.P., 2005, An animated tectonic reconstruction of southwestern

North America since 36 Ma: Geosphere, v. 1, p. 147–172,

https://doi.org/10.1130/GES00016.1

151

Meisling, K.E., and Weldon, R.J., II, 1986, Cenozoic uplift of the San Bernardino Mountains:

Possible thrusting across the San Andreas fault, Geological Society of America Abstracts

with Programs, v. 18, n. 2, p. 157.

Meisling, K.E., and Weldon, R.J., II, 1989, Late Cenozoic tectonics of the northwestern San

Bernardino Mountains, southern California: Geological Society of America Bulletin, v.

101, p. 106-128, https://doi.org/10.1130/0016-7606(1989)101<0106:LCTOTN>2.3.CO;2

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. 491-495,

https://doi.org/10.1130/0091-7613(1994)022<0491:MCROTW>2.3.CO;2

Niemi, N.A., Buscher, J.T., Spotila, J.A., House, M.A., and Kelley, S.A., 2013, Insights from

low-temperature thermochronometry into transpressional deformation and crustal

exhumation along the San Andreas fault in the western Transverse Ranges, California:

Tectonics, v. 32, p. 1602-1622, https://doi.org/10.1002/2013TC003377

O’Banion, T.C., 2012, Exploration of the Holocene record of fault slip for the Mojave section of

the San Andreas fault [M.S. thesis]: Davis, University of California, 140 p.

Oskin, M., and Stock, J., 2003, Marine incursion synchronous with plate-boundary localization

in the Gulf of California: Geology, v. 31, p. 23-26, https://doi.org/10.1130/0091-

7613(2003)031<0023:MISWPB>2.0.CO;2

Oskin, M., Stock, J., and Martin-Barajas, A., 2001, Rapid localization of Pacific-North America

plate motion in the Gulf of California: Geology, v. 29, p. 459-462,

https://doi.org/10.1130/0091-7613(2001)029<0459:RLOPNA>2.0.CO;2

152

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

Powell, R.E., 2016, Relations between crustal blocks and evolution of the San Andreas fault

system: Geological Society of America Abstracts with Programs, v. 48, n. 4,

https://doi.org/10.1130/abs/2016CD-274564

Sauber, J., Thatcher, W., Solomon, S.C., and Lisowski, M., 1994, Geodetic slip rate for the

eastern California shear zone and the recurrence time of Mojave desert earthquakes:

Nature, v. 367, p. 264-266, https://doi.org/10.1038/367264a0

Savage, J.C., 1983, Strain accumulation in western United States: Annual Review of Earth and

Planetary Sciences, v. 11, p. 11-41.

Sharp, R.V., 1967, in the Peninsular Ranges of southern California:

Geological Society of America Bulletin, v. 78, p. 705-730, https://doi.org/10.1130/0016-

7606(1967)78[705:SJFZIT]2.0.CO;2

Shen, Z., Jackson, D.D., and Ge, B.X., 1996, Crustal deformation across and beyond the Los

Angeles basin from geodetic measurements: Journal of Geophysical Research, v. 101, p.

27957-27980, https://doi.org/10.1029/96JB02544

Sieh, K.E., 1978, Slip along the San Andreas fault associated with the great 1857 earthquake:

Bulletin of the Seismological Society of America, v. 68, p. 1421-1448.

153

Sieh, K.E., and Jahns, R.H., 1984, Holocene activity of the San Andreas fault at Wallace Creek,

California: Geological Society of America Bulletin, v. 95, p. 883-896,

https://doi.org/10.1130/0016-7606(1984)95<883:HAOTSA>2.0.CO;2

Spotila, J.A., and Sieh, K., 2000, Architecture of transpressional thrust faulting in the San

Bernardino Mountains, southern California, from deformation of a deeply weathered

surface: Tectonics, v. 19, p. 589-615, https://doi.org/10.1029/1999TC001150

Stein, R.S., and Thatcher, W., 1981, Seismic and aseismic deformation associated with the 1952

Kern County, California, earthquake and relationship to the Quaternary history of the

White Wolf fault: Journal of Geophysical Research, v. 86, p. 4913-4928,

https://doi.org/10.1029/JB086iB06p04913

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Weldon, R., and Humphreys, E., 1986, A kinematic model of southern California: Tectonics, v.

5, p. 33-48, https://doi.org/10.1029/TC005i001p00033

154

CHAPTER 4: CONSTRAINING THE GEOMETRY AND

TIMING OF NORMAL FAULTING AND ASSOCIATED

VOLCANISM IN THE SAN GABRIEL MOUNTAINS AND

ITS CROSS-FAULT CORRELATIVES, SOUTHERN

CALIFORNIA, U.S.A.

ABSTRACT

The Sierra Pelona anticlinorium of the San Gabriel Mountains has been correlated with the Chocolate Mountains anticlinorium in the Orocopia Mountains. These anticlinoria are associated with Oligocene-Miocene normal faults, but the relationship between the anticlinoria and these faults, and the timing of their development, is unclear. Previous workers have suggested that the Sierra Pelona anticlinorium is the exhumed footwall of a detachment fault, exposed as the Pelona fault. However, kinematic data from the Pelona fault, and its geometric relationship with other faults, indicates that it is actually an antithetic normal fault within the hanging wall of a detachment. Multiple subsequent deformational events have modified the normal-fault system, but sequential palinspastic reconstructions suggest that the anticlinorium is part of a hanging-wall fault block uplifted and warped by isostasy. U-Pb dating of zircon from volcanic units indicates that extension and volcanism initiated concurrently in the San Gabriel and Orocopia Mountains, and on both sides of the anticlinorium. Overlapping chemical compositions of these volcanics underscore their close genetic relationship and original proximity. These improved palinspastic reconstructions help constrain the architecture and history of crustal blocks that have been offset by the San Andreas fault system. 155

INTRODUCTION

Within the San Gabriel Mountains of southern California, an anticlinorium of schist is exposed at Sierra Pelona (Fig. 4-1). This anticlinorium, which has been correlated with the western end of the Chocolate Mountains anticlinorium in the Orocopia Mountains (Crowell,

1962, 1975a; Fig. 4-1), provides an important constraint on the cumulative slip on the southern

San Andreas fault. However, it is debated whether the two regions are truly correlative, and whether the differing orientations of these anticlinoria were originally consistent.

Correlation of the anticlinoria at Sierra Pelona and the Orocopia Mountains, and of surrounding lithologies and structures, has been supported and refined by numerous studies (e.g.,

Crowell, 1962, 1975a; Carman, 1964; Ehlig and Ehlert, 1972; Bohannon, 1975; Ehlert, 1982,

2003; Weigand, 1982; Frizzell and Weigand, 1993; Ingersoll et al., 2014), but challenged by others (e.g., Smith, 1977; Powell, 1981, 1993; Matti and Morton, 1993; Weldon et al., 1993).

Most of these alternate reconstructions are based on cross-fault correlation of a Triassic megaporphyritic monzogranite (Matti et al., 1985; Frizzell et al., 1986; Matti and Morton, 1993), which seems to imply substantially less cumulative slip on the southern San Andreas fault. In contradiction to these alternate reconstructions, Chapter 1 (Coffey et al., 2019) has added support to the correlation of Sierra Pelona and the Orocopia Mountains, and recent models (e.g., Darin and Dorsey, 2013; Chapter 3) have potentially reconciled this correlation with that of the monzogranite.

An originally linear correlated anticlinorium of Sierra Pelona and the Orocopia

Mountains, as suggested by various studies (e.g., Ingersoll et al., 2014; Bennett et al., 2016;

Grove and Jacobson, 2016), is compatible with paleomagnetic data from these regions (Terres,

1984; Terres and Luyendyk, 1985), but has been challenged with arguments based on the

156

Figure 4-1: Regional map, showing southern San Andreas fault system, insets of three correlated regions separated by it, and extent of Fig. 4-5A. Shown in insets are faults (red: ca. 12 Ma –present

San Andreas fault system; blue: ca. 18-12 Ma sinistral-fault system; black: ca. 26-18 normal-fault system), and locations of fault-kinematic measurements (red-filled circles), volcanic and hypabyssal

samples dated by U-Pb methods and analyzed for geochemistry (Oligocene: white-filled circles;

Miocene: blue-filled circle), and volcanic samples only analyzed for geochemistry (solid-black

circles). Cf = Clearwater fault; CWf = Clemens Well fault; Gf = Garlock fault; Ltf = Lonetree

fault; OM = Orocopia Mountains; Pf = Pelona fault; Puf = Punchbowl fault; SAf = San Andreas

fault; SCf = Salton Creek fault; Sdf = Soledad fault; SFf = San Francisquito fault; SGf = San

Gabriel fault; SGM = San Gabriel Mountains; SJf = San Jacinto fault; SM/Cuf = Sierra

Madre/Cucamonga fault; VCf = Vasquez Canyon fault. Shaded-relief basemap generated using

GeoMapApp (http://www.geomapapp.org; topography from Ryan et al., 2009).

157 orientations of penetrative lineations within the schist of these anticlinoria (e.g., Haxel et al.,

2018). Chapter 2 presents a model which reconciles these lineation data with an originally linear

Sierra Pelona/Orocopia Mountains anticlinorium, and Ingersoll and Coffey (2017) demonstrated how these anticlinoria may have undergone subsequent rotations to achieve their present, disparate orientations.

Although the debate is far from settled, the recent contributions discussed above provide substantial support for original continuity and linearity of the Sierra Pelona and Orocopia

Mountains anticlinoria. However, the relationship of these anticlinoria to Oligocene-Miocene normal faults that bound them remains unclear, as does relative timing of normal faulting and associated volcanism. I present, herein, kinematic data, sequential palinspastic reconstructions, and volcanic-rock age and chemistry data to address these questions.

SIERRA PELONA AND ASSOCIATED NORMAL-FAULT SYSTEM

Southeast of the anticlinorium at Sierra Pelona are the Vasquez Canyon and Soledad faults (e.g., Muehlberger, 1958) (Fig. 4-1). These are normal faults which were active during deposition of the Vasquez Formation (Sharp, 1935) on their northern (hanging-wall) sides

(Muehlberger, 1958; Hendrix and Ingersoll, 1987). The Lonetree fault (Fig. 4-1), which has a similar strike and is cut by the same middle Miocene system of sinistral faults (e.g., Dibblee,

1996a, b), is likely part of this normal-fault system (Hendrix, 1993). The San Francisquito fault, which is also probably part of this system (e.g., Konigsberg, 1967; Hendrix and Ingersoll, 1987;

Bunker and Bishop, 2001; Yan et al., 2005), has alternatively been interpreted primarily as a dextral fault related to the San Andreas fault system (Smith, 1977; Powell, 1981, 1993). Much of the evidence for dextral slip, however, comes from northeast of its junction with the younger,

158 dextral Clearwater fault, and thus is likely indicative of motion on this younger fault, and not the

San Francisquito fault proper. The Pelona fault is constrained as the same age as these other normal faults, and is considered part of the same system; it originally dipped south (e.g., Dibblee,

1997a), however, in constrast to the other faults.

Using the criteria of Friedmann and Burbank (1995), the normal-fault system at Sierra

Pelona more closely resembles a supradetachment system than a “traditional” rift system: duration of sedimentation was only a few million years (e.g., Woodburne, 1975; Hendrix and

Ingersoll, 1987); sediment was derived largely from the hanging walls of normal faults and deposited as alluvial fans (Hendrix and Ingersoll, 1987; Hendrix, 1993); associated volcanism was subalkaline to calc-alkaline (Frizzell and Weigand, 1993); and the crust of the San Gabriel

Mountains had experienced significant tectonism in the preceding 150 Myr (e.g., Ehlig, 1981).

Top-to-north detachment faulting is implied (or top-to-northwest after correcting for vertical-axis rotations implied by paleomagnetism; Terres and Luyendyk, 1985; Hendrix and Ingersoll, 1987;

Hendrix, 1993).

It has been suggested that the anticlinorium at Sierra Pelona is the footwall of the detachment system, uplifted and arched due to isostatic uplift associated with detachment faulting (Bishop, 1990; Bishop and Ehlig, 1990; Hendrix, 1993). In this model, the originally south-dipping Pelona fault bounding the southern flank of the anticlinorium, and the north- dipping San Francisquito fault bounding the northern flank, are both exposures of the detachment surface, folded about the anticlinorium (Fig. 4-2). Bishop (1990) interpreted both strands of the

Pelona fault as detachment surfaces, whereas Hendrix (1993) interpreted the northern strand as the detachment and the southern strand as an antithetic normal fault.

159

Figure 4-2: Schematic cross section across Sierra Pelona, illustrating hypothesis that Pelona and

San Francisquito faults are an exhumed detachment fault exposed on either limb of the Sierra

Pelona anticlinorium (Bishop, 1990; Bishop and Ehlig, 1990; Hendrix, 1993). In model of Hendrix

(1993), southern strand of Pelona fault, which is not depicted separately in this figure, exhibits

south-side-down motion. Adapted from Hendrix (1993).

Kinematics of the Pelona Fault

I measured kinematic indicators along the Pelona fault to test the Pelona-detachment- fault hypothesis. Within the Vasquez Formation conglomerate, in the damage zone approximately 3-15 m from the fault core, I measured orientations of slickenlines within fault planes (Fig. 4-3A), and of fault planes along which sense-of-shear was indicated by offset conglomerate clasts (Fig. 4-3B-D). Two measurement sites were along the southern strand of the

Pelona fault, and a third was east of where the two strands merge (Fig. 4-1). At all three sites, slickenline lineations indicate dip-slip (Fig. 4-4; phase of strike-slip suggested in Fig. 4-4C likely occurred subsequently on nearby sinistral fault; e.g., Dibblee, 1997a). Offset conglomerate clasts indicate south-side-down motion along fault planes with dips within approximately 30° of that of the main fault trace (blue planes in Fig. 4-4), and top-to-north motion along fault planes with dips approximately 60°-80° from that of the main fault trace (red planes in Fig. 4-4); I interpret

160 these fault planes as Riedel shears (R) and conjugate Riedel shears (R’), respectively (Fig. 4-4D-

G). All these observations are consistent with south-side-down slip on the Pelona fault (or southeast side down after correcting for vertical-axis rotations implied by paleomagnetism;

Terres and Luyendyk, 1985).

Figure 4-3: Field photos from Vasquez Formation within damage zone of Pelona fault. A: a fault

surface with slickenline lineations indicating dip slip; rock hammer for scale. B-D: Conglomerate

clasts offset by faults: offset clasts are outlined in black; offsetting faults are traced in red.

Mechanical pencil for scale. A is at Fall Canyon (middle site in Fig. 4-1A; Fig. 4-4B); B is at Rowher

Canyon (easternmost site in Fig. 4-1A; Fig. 4-4C); C and D are at Bouquet Canyon (westernmost

site in Fig. 4-1A; Fig. 4-4A).

161

Figure 4-4: Kinematic data measured along Pelona fault and their interpretation. A-C: Kinematic

data (equal-area, lower-hemisphere projections). Note that trends of slickenline lineations are approximately perpendicular to the strike direction of the fault, indicating dip-slip. Orientation of

Pelona fault varies between locations, with bedding/foliation on either side of fault varying in same way; presumably result of subsequent deformation. A: Southern strand of Pelona fault at Bouquet

Canyon (western exposure in Fig. 4-1A). B: Southern strand of Pelona fault at Fall Canyon (middle

exposure in Fig. 4-1A). C: Pelona fault, east of convergence of northern and southern strands, at

162

Rowher Canyon (eastern exposure in Fig. 4-1A). D-G: Interpretation of data of A-C, shown as

projections and schematic cross sections. Coloration as in A-C; blue and red faults interpreted as

Riedel (R) shears and conjugate Riedel (R’) shears, respectively.

South-side-down slip on the Pelona fault is opposite the sense of motion predicted by the

Pelona-detachment hypothesis (Bishop, 1990; Bishop and Ehlig, 1990). In the model of Hendrix

(1993), the southern strand of the Pelona fault is a high-angle, antithetic fault, and only the northern strand is a detachment surface; however, even this model has the antithetic fault

(southern Pelona fault) terminating against the detachment surface (northern Pelona fault), and would thus predict south-side-up slip east of where these two strands converge (Fig. 4-4C). It is possible that the above kinematic data are the result of subsequent reactivation of a folded detachment fault in a south-side-down sense. However, both strands of the Pelona fault exhibit brittle deformation, without the mylonitic zone that would be expected in the footwall of an exhumed detachment fault (e.g., Coney, 1980); mylonitic gneiss in the hanging wall of the

Vincent fault (formerly: “Vincent thrust”) is locally juxtaposed against the Pelona fault (e.g.,

Dibblee, 1997a, b), but is genetically unrelated. Furthermore, whereas supradetachment normal faults should sole into the detachment, the Pelona fault appears to terminate against the Vasquez

Canyon fault (based on compilation of mapping by Muehlberger, 1958; Bishop, 1990; Dibblee,

1997a, b). All of these observations are readily explained by both strands of the Pelona fault being parts of a supradetachment normal fault, antithetic to, and terminating against, the Vasquez

Canyon fault. The Pelona-detachment-fault hypothesis (Bishop, 1990; Bishop and Ehlig, 1990;

Hendrix, 1993), though not disproven, involves postulation of an earlier phase of deformation for which no evidence is apparent.

163

SEQUENTIAL RECONSTRUCTION OF THE SAN GABRIEL BLOCK

If the Pelona fault is an antithetic, supradetachment normal fault, and the geometry proposed in the detachment-fault model (Fig. 4-2) is incorrect, then what is the actual geometry of this normal fault system? Whereas the San Gabriel Mountains have been extensively mapped, and the surface traces and approximate orientations of these normal faults have been established, the normal faults have been extensively modified by multiple younger deformational events. It is necessary to sequentially reverse younger to older deformational events in order to ascertain the original geometry of this normal-fault system. I performed such sequential restorations via a series of geologic maps and cross sections of the San Gabriel block (bounded on the northeast by the San Andreas fault, the west by the San Gabriel fault, and the south by the Sierra

Madre/Cucamonga fault; Fig 4-1). For each time step, details of the type and amount of slip restored on each fault, the amount of vertical-axis rotation restored, the geologic units removed, the evidence upon which these restorations are based, constraints on timing, and references are given in Appendix F.

Ca. 5 Ma Reconstruction

Beginning with the present (Fig. 4-5A), I first restored deformation in the southern San

Gabriel block back to ca. 5 Ma (Fig. 4-5B): dextral slip on strands of the San Andreas and San

Jacinto faults, oblique reverse sinistral slip on a system of faults conjugate to these dextral faults, and associated uplift and tilting. This restoration was closely based on that of Nourse (2002), with the exception that I assumed conjugate dextral and sinistral faults in the southern San

Gabriel block to be coeval, and restored them simultaneously, whereas Nourse (2002) interpreted the dextral faults as younger. In my interpretation, fault blocks moved northeast via sinistral slip, occupying space vacated by southeastward extrusion of fault-bound slivers via dextral slip (Fig.

164

4-5A, B). This restoration drastically reduces the space problem that is created if dextral slip is restored prior to restoration of sinistral slip (as in Nourse, 2002). I also corrected the San Gabriel block for counterclockwise rotation implied by paleomagnetic data (Terres and Luyendyk,

1985).

Ca. 14 Ma Reconstruction

My second step was to restore deformation associated with oblique reverse dextral slip in the northern San Gabriel block, along the Clearwater, Liebre, North Liebre, and Bald Mountain faults, resulting in a ca. 14 Ma reconstruction (Fig. 4-5C). Slip on most of these faults was dominantly dextral slip at their northwestern ends, but largely reverse slip at their southeastern ends, where they intersect the cross-section line, due to changes in their orientations. As documented by the middle Miocene Mint Canyon Formation (Kew, 1924), this transpression was accompanied by folding, which I also restored: the anticlinorium at Sierra Pelona tightened during this time, and a syncline developed south of Sierra Pelona (Fig. 4-5B, C).

Ca. 18 Ma Reconstruction

My third step was to restore deformation on a system of sinistral faults south of Sierra

Pelona, as well as clockwise vertical-axis rotation of the San Gabriel block, which was likely related (Luyendyk et al., 1980; Terres and Luyendyk, 1985), producing a ca. 18 Ma reconstruction (Fig. 4-5D). In this reconstruction, the last of the breaks in the cross-section line present in previous steps have been eliminated. This reconstruction shows the San Gabriel block after the Oligocene-Miocene normal-fault system had ceased activity, but prior to subsequent deformation. It demonstrates how mapped relationships are consistent with a supradetachment system of normal faults (e.g., Gibbs, 1984), in which the near-vertical Lonetree fault is a listric

165

Figure 4-5 (continued on

subsequent pages): Sequential

palinspastic reconstructions of the

San Gabriel block, as paired maps

and cross sections, from present

(A) to ca. 26 Ma (E). In maps, only

lateral movement and vertical-

axis rotations are restored; units

and contacts are not adjusted to

account for erosion or burial (i.e.,

units shown in maps would not

necessarily be exposed at surface

at that time). Extent of Fig. 4-5A

is shown in Fig. 4-1. Geologic map

of Fig. 4-5A modified from

sources listed in Table S4-1.

Details of assumptions used to

generate these reconstructions,

and their bases, are given in Table

S4-1. For full-size figure, see Fig.

S4-1 (A-E) in Supplementary

Materials.

166

167

168

169

170 breakaway, the Soledad, Vasquez Canyon, and San Francisquito faults are approximately planar normal faults accommodating extension via rotation of intervening fault blocks, and the Pelona fault is an antithetic normal fault terminating against the Vasquez Canyon fault (Fig. 4-5D). In this model, the Sierra Pelona anticlinorium, broader prior to transpressional tightening ca. 14-5

Ma, is the lower part of one of the rotated fault blocks, isostatically uplifted and warped, together with the underlying detachment fault (Fig. 4-5D; but see model for early Paleogene development of the Chocolate Mountains anticlinorium in Yin, 2002). Predicted cumulative normal slip on the

Soledad fault is 4.5 km, matching the upper limit of the range of cumulative slip considered likely by Muehlberger (1958). Slip of 5 km is predicted for the listric Lonetree fault, and 6 km for the San Francisquito fault, on which motion could have continued after isostatic uplift had inactivated the other supradetachment faults in the system (Fig. 4-5D). Late-stage coarse-grained alluvial-fan sedimentation in the related Charlie Canyon basin north of the San Francisquito fault is consistent with this interpretation (e.g., Hendrix and Ingersoll, 1987).

Ca. 26 Ma Reconstruction

My final step was to undo slip on the system of normal faults, and associated isostatic uplift and warping, producing a ca. 26 Ma reconstruction (Fig. 4-5E). This reconstruction shows the San Gabriel block prior to initiation of normal faulting, and associated volcanism and sedimentation. The Vincent fault is restored to a continuous, subhorizontal fault, and the original geometry of the fault exposed in the Mill Canyon window is shown, together with the location of its predicted continuation north of the Soledad fault (as the “Tierra Subida fault”; Fig. 4-5E).

This continuation follows a contact between different crystalline units, previously mapped as intrusive (e.g., Dibblee, 1997b), but which I verified is a fault zone. This reconstruction predicts that the supradetachment normal-fault system accommodated approximately 30% extension,

171 which matches the expected value (Wernicke and Burchfiel, 1982), given my inference of normal faults of initial dip of approximately 60° rotating to dips of approximately 40° (ignoring tilting due to isostatic uplift and subsequent deformation).

TIMING OF NORMAL FAULTING AND ANTICLINORIUM DEVELOPMENT

Normal faulting surrounding Sierra Pelona in the San Gabriel block was coeval with deposition of the Oligocene-Miocene Vasquez Formation (e.g., Muehlberger, 1958; Hendrix and

Ingersoll, 1987), and normal faulting in the Orocopia Mountains was coeval with deposition of the correlated Diligencia Formation (e.g., Spittler and Arthur, 1982; Law et al., 2001; Ingersoll et al., 2014). In the stratigraphically lower parts of these formations are interbedded volcanic flows

(Sharp, 1935; Muehlberger, 1958; Crowell, 1975b; Spittler and Arthur, 1982; Hendrix and

Ingersoll, 1987; Frizzell and Weigand, 1993), whose ages constrain the timing of normal faulting, and thus likely of development of the anticlinorium (but see Yin, 2002).

K-Ar Dating of Volcanic Flows (Previous Work)

Whole-rock K-Ar dating has suggested an age of the Vasquez Formation volcanic flows between 25.6 and 23.1 Ma (Frizzell and Weigand, 1993), but with individual ages ranging from

37.5 ± 1.0 Ma to 14.7 ± 0.4 Ma (Frizzell and Weigand, 1993), and plagioclase K-Ar ages ranging from 25.6 ± 2.2 Ma to 20.7 ± 0.8 Ma (Crowell, 1973; Spittler, 1974; Woodburne, 1975; recalculated after Dalrymple, 1979). Whole-rock K-Ar ages for the Diligencia Formation volcanic flows range from 23.6 ± 0.5 Ma to 21.3 ± 0.6 Ma (Frizzell and Weigand, 1993), and plagioclase K-Ar ages range from 23.0 ± 3.0 Ma to 19.1 ± 2.0 Ma (Crowell, 1973; Spittler, 1974; recalculated after Dalrymple, 1979). The seemingly younger ages of the Diligencia Formation volcanic flows would suggest that volcanism, and probably associated extension, migrated across

172 this extensional system through time (e.g., Hendrix et al., 2010). However, such interpretations are hampered by the large uncertainties in these ages. Furthermore, some of these ages may be partially reset by subsequent thermal disturbance; in the southern San Gabriel Mountains, for example, the Telegraph Peak pluton, dated by U-Pb analysis of zircon as 25.6 ± 1 Ma (May and

Walker, 1989) to 27.5 ± 0.7 Ma (Dykstra et al., 2018), has yielded biotite K-Ar ages ranging from 27 ± 3 Ma to 14.4 ± 0.4 Ma (Hsu et al., 1963; Miller and Morton, 1977; recalculated after

Dalrymple, 1979), interpreted by Nourse (2002) as the result of widespread emplacement of dikes in the middle Miocene. I performed U-Pb dating of zircon from the volcanic flows of the

Vasquez and Diligencia formations, and from rhyolite domes along the Salton Creek fault, southeast of the Orocopia Mountains (Fig. 4-1), to more accurately and precisely determine the timing of magmatism in these areas.

Zircon U-Pb Dating of Volcanic Flows and Domes

Methods

I collected eight approximately 1-2 kg samples of volcanic and hypabyssal rock for zircon analysis: three from the volcanic flows of the Vasquez Formation, one from a hypabyssal intrusion approximately 1 km south of these flows, one from the volcanic flows of the Diligencia

Formation, and three from rhyolite domes along Salton Creek fault (Fig. 4-1). I crushed and sieved the samples, then separated them by density and magnetism using: (1) a Mineral

Technologies MD Gemini shaking table at Pomona College, (2) a neodymium hand magnet and a model L-1 Frantz isodynamic magnetic separator, and (3) methylene iodide (ρ = 3.32 g/cm3).

At the University of California, Santa Cruz, approximately 100 zircon crystals from each sample was mounted, together with grains of zircon primary- and secondary-standard reference materials (Temora 2 as primary: 416.78 ± 0.33 Ma 2σ; Black et al., 2004), on epoxy plugs, which

173 were polished to a depth of ~30 μm to expose crystal interiors. These zircon crystals were then dated by U-Pb methods using laser-ablation–inductively coupled plasma–mass spectrometry

(LA-ICP-MS) at the University of California, Santa Cruz, in two sessions during November and

December of 2018, using protocols described by Sharman et al. (2013). To minimize the effects of instrument drift during the analytical session, unknowns and secondary standards were analyzed in a “round robin” pattern. Analyses of Temora2 standards bracketed the beginning and ending of the session, and were interspersed every five analyses throughout the routine. Material was ablated from pits 28 μm in diameter using a Photon Machines Analyte 193 excimer laser, then carried in helium into the plasma source of an Element XR ICP-MS. Intensities of atomic masses 202Hg, 204Pb, 206Pb, 207Pb, 208Pb, 232Th, 235U, and 238U were measured in single-collector mode. Raw data were reduced using the Iolite add-on for IgorPro to correct for down-hole fractionation and instrument background and to propagate associated random errors (Paton et al.,

2010, 2011). Corrected ratios and propagated uncertainties of 206Pb/238U, 207Pb/235U, and

207Pb/206Pb (uncorrected for common Pb) from individual analyses were then exported, uncorrected for common Pb.

I used IsoplotR (Vermeesch, 2018) to calculate 207Pb-corrected 206Pb/238U ages for individual analyses (Table S4-1), and 207Pb-corrected 206Pb/238U Terra-Wasserburg (inverse- concordia) and weighted-mean ages. I used a 207Pb correction-based approach to correct for nonradiogenic Pb in measured isotopic ratios. I assumed that common Pb was incorporated at the time of zircon crystallization, with a 207Pb/206Pb ratio of 0.837 (Stacey and Kramers, 1975). I used this 207Pb-based correction, rather than a 204Pb-based correction, because 204Pb count rates rarely exceed the limits of detection (as discussed in Chang et al., 2006). Agreement of the determined ages for the secondary standards (Appendix G; Table S4-1) with their published ages

174

(Fish Canyon Tuff: 28.48 ± 0.03 Ma; Schmitz and Bowring, 2001; Dromedary: 99.12 ± 0.03 Ma;

Schoene et al., 2006; Plešovice: 337.13 ± 0.37 Ma; Sláma et al., 2008; FC1: 1099.0 ± 0.6; Paces and Miller, 1993; all errors 2σ) demonstrates the effective accuracy of the Temora2 primary standard. Many of the samples contain individual analyses that are notably discordant, defining linear arrays extending toward 207Pb/206Pb ratios of approximately 0.8 (Stacey and Kramers,

1975; Appendix G). Younger zircon is particularly susceptible to this effect because it contains lower levels of radiogenic Pb. After correction for common Pb, these analyses yield consistent ages. Most samples yield weighted-mean 207Pb-corrected 206Pb/238U ages with MSWD (Wendt and Carl, 1991) values which indicate a coherent age population and a sound error-propagation model (Appendix G). The 207Pb-corrected 206Pb/238U inverse-concordia age results from each sample are reported in Fig. 4-6 and are shown alongside weighted mean plots in Appendix G.

Ages

In contrast to previously determined K-Ar ages, U-Pb dating of zircon yields consistent, ca. 25 Ma ages for the volcanic flows of both the Vasquez and Diligencia formations, as well as for the rhyolite domes along Salton Creek fault (Fig. 4-6). This indicates that volcanism and normal faulting initiated ca. 26-25 Ma, both along Sierra Pelona and the Orocopia Mountains, and on both sides of the anticlinorium.

The hypabyssal intrusion approximately 1 km south of the Vasquez Formation volcanic flows yielded an age of ca. 18 Ma (Fig. 4-6), indicating it is related to widespread Miocene magmatism in southern California (e.g., Eaton, 1958; Stanley et al., 2000), which had not previously been documented in the San Gabriel block, except south of the north branch of the

San Gabriel fault as the Glendora volcanics (e.g., Shelton, 1955; Nourse et al., 1998).

175

Geochemistry of Volcanic Flows and Domes

To further examine relationships between the volcanic rocks beside Sierra Pelona and those of the Orocopia Mountains, I determined their major- and trace-element compositions, as well as those of volcanic rocks from the intervening Punchbowl block (Fig. 4-6), found as interbedded flows and conglomerate clasts within strata equivalent to the Vasquez Formation

(Chapter 1: Coffey et al., 2019).

Figure 4-6: Palinspastic reconstruction of Sierra Pelona, Blue Ridge, and Orocopia Mountains anticlinoria and surrounding areas (Fig. 4-1A, B, and C, respectively), ca. 18 Ma, showing ages (and

2σ errors) of volcanic units determined by U-Pb dating of zircon. As in Fig. 4-5, units and contacts

176

are not adjusted to account for erosion or burial. White space indicates either areas covered by

Neogene-Quaternary deposits or gaps generated by the reconstruction. The two ages with anomalously high errors are determined from anomalously few analyses (n = 2 and 3). BRf = Blue

Ridge fault; CWf = Clemens Well fault; Ff = Fenner fault; fMCw = fault of Mill Canyon window;

Ltf = Lonetree fault; OMf = Orocopia Mountains fault; Pf = Pelona fault; Sdf = Soledad fault; SFf

= San Francisquito fault; TSf = Tierra Subida fault; VCf = Vasquez Canyon fault; Vf = Vincent

fault. Punchbowl-block age is recalculated from data of Chapter 1 (Coffey et al., 2019); all other data are of this study. Sierra Pelona geology from Fig. 4-5D; Punchbowl block geology after Hoyt et al. (2018) and Chapter 1 (Coffey et al., 2019); Orocopia Mountains geology after Law et al. (2001),

Ingersoll et al. (2014), and Hoyt et al. (2018). Restoration of vertical-axis rotations based on paleomagnetic data of Terres (1984) and Terres and Luyendyk (1985), and models of Ingersoll and

Coffey (2017) and Chapter 2.

Methods

The eight volcanic-rock samples for which I obtained U-Pb ages of zircon, as well as two additional samples from the Diligencia Formation volcanics, and four samples from the volcanics of the Punchbowl block, were prepared and analyzed via X-ray fluorescence and inductively coupled plasma mass spectrometry at the Peter Hooper GeoAnalytical Lab at

Washington State University. Samples were heated to determine volatile content via loss on ignition (LOI), then ground into a powder, mixed with dilithium tetraborate (Li2B4O7) flux, and fused in graphite crucibles at 1000°C. The resulting glass beads were reground. For analysis by

X-ray-fluorescence, the resulting powder was refused, polished with 600-grit silicon carbide, cleaned, and mounted, together with one bead each of internal standards BCR-P and GSP-1. For

177 analysis by inductively coupled plasma mass spectrometry, the resulting powder was dissolved, using HNO3, HF, HClO4, H2O2, deionized water, and heat.

Concentrations of 10 major elements and 19 trace elements were determined using a

Thermo ARL ADVANT’XP+ sequential X-ray-fluorescence spectrometer. X-ray intensities were measured for each element using a rhodium target run at 50 kV and 50 mA at full vacuum and a 25 mm mask. Elemental concentrations were then determined by comparing these measurements with the measured intensities from two beads each of nine USGS standard samples (PCC-1, BCR-1, BIR-1, DNC-1, W-2, AGV-1, GSP-1, G-2, and STM-1), using the values recommended by Govindaraju (1994), as well as two beads of pure vein quartz to record background levels for all elements except Si. Internal standards BCR-P and GSP-1 were analyzed every 28 samples as a check on instrumental performance and precision. The intensities for all elements were automatically corrected for line interference and absorption effects due to other elements using the fundamental-parameter method. A thorough discussion of the methods used, including a discussion of the precision and accuracy of the spectrometer and an element- by-element comparison with results from other laboratories, can be found in Johnson et al.

(1999).

Concentrations of 27 trace elements were determined using an Agilent 7700 ICP quadrupole mass spectrometer. The plasma is typically operated at 1250 W, and typical argon gas flow conditions are 0.25 L/s for the plasma gas, 0.017 L/s for the auxiliary gas, 0.013 L/s for the nebulization gas and 0.005 L/s for the make-up gas. Tuning parameters are software controlled. Each sample was diluted an additional 10x at the time of analysis using the Agilent’s

Integrated Sample Introduction System, yielding a final dilution factor of 1:4800. The electron multiplier detects particles by pulse counting when the beam intensity is <3.0 MHz and by

178 current integration when the beam intensity is >3.0 MHz. Instrument drift was corrected using

Ru, In and Re as internal standards. Internal standardization for rare-earth elements (REEs) was performed using a linear interpolation between In and Re after Doherty (1989) to compensate for mass-dependent differences in instrumental drift. Tuning was optimized to maintain a CeO/Ce ratio of less than 0.5% to minimize interference of light REE oxides with the signals of intermediate and heavy REEs. The remaining interferences were rectified using correction factors estimated using two mixed-element solutions, one containing Ba, Pr and Nd, the other containing Tb, Sm, Eu and Gd. Concentrations were standardized by comparison with analyses

Figure 4-7: Classification of volcanic samples according to Le Bas et al. (1986) classification. Note

range of compositions of most sample sets. Raw data in Table S4-2.

179 of duplicates of three in-house rock standards, which were interspersed within each set of 18 samples.

Results

For the XRF data, the sum of determined major- and trace-element abundances as oxides and the mass lost on ignition (LOI) is between 99.24% and 100.08% for all samples, and trace- element abundances determined by XRF and by ICPMS are in close agreement (Tables S4-2, S4-

3), collectively suggesting that the results are reliable.

Using the total-alkali-silica (TAS) classification scheme of Le Bas et al. (1986), the samples range in composition from basaltic andesite to rhyolite (Fig. 4-7). All samples from the rhyolite domes along Salton Creek fault are rhyolite; most other sample sets, however, show

Figure 4-8: Plots of A: Th, B: U, and C: Hf vs. Ta for samples shown in Figs. 4-1 and 4-7, and from

Frizzell and Weigand (1993), and Cole and Basu (1995). The ca. 18 Ma hypabyssal sample plots

outside area of these plots. Note strong effect of silica content in A and B. Raw data are given in

Table S4-3.

180

Figure 4-9: Chondrite-normalized A: rare-earth-element and B: trace-element abundances for

samples shown in Figs. 4-1 and 4-6. Note strong effect of silica content in A, and distinct

composition of ca. 18 Ma hyabyssal sample. Raw data are given in Table S4-3.

substantial compositional variability. In plots of Th, U and Hf vs. Ta (Fig. 4-8), used to compare flows of the Vasquez and Diligencia formations by Frizzell and Weigand (1993), there is much overlap between the different sample sets, with clustering of data due more to silica content than to sample location. The ca. 18 Ma hypabyssal sample, in contrast, plots outside the area of these plots (Fig. 4-8), despite its overlapping silica content and intrusion through similar basement rocks (Fig. 4-7). In plots of rare-earth-element and other trace-element compositions (Fig. 4-9), the various sample sets again overlap substantially, with silica content again more responsible for clustering than sample location. In particular, higher silica contents correspond to greater negative Eu anomalies (Fig. 4-9). Again, the ca. 18 Ma hypabyssal sample, with similar silica content and surrounding country rock, is distinct in composition, particularly in rare-earth

181 elements (Fig. 4-9A). Collectively, these data support the close genetic relationship between these volcanic units suggested by their ages, structural positions, and hypothesized palinspastic correlation.

CONCLUSIONS

In the San Gabriel Mountains, the area around Sierra Pelona experienced supradetachment normal faulting and associated volcanism beginning ca. 26-25 Ma. This faulting and volcanism extended across the future San Andreas fault to the Orocopia Mountains, where volcanic units have the same age and chemistry as those of the San Gabriel Mountains.

Kinematic data and map relationships indicate that the Pelona fault, which bounds the Sierra

Pelona anticlinorium on its southern side, is not the exhumed detachment fault as some studies have suggested. Rather, it is an antithetic normal fault in the supradetachment system, terminating against the coeval Vasquez Canyon fault. The normal-fault system of the San

Gabriel Mountains has been modified by subsequent deformational events. Sequential palinspastic restoration of these deformational events, working back from the present, is necessary to reveal the original architecture of this normal-fault system. These results further constrain offsets across the San Andreas fault system and facilitate palinspastic reconstruction of southern California.

182

REFERENCES CITED

Bennett, S.E.K., Darin, M.H., Dorsey, R.J., Skinner, L.A., Umhoefer, P.J., and Oskin, M.E.,

2016, Animated tectonic reconstruction of the Lower Colorado River region: implications

for late Miocene to present deformation, in Reynolds, R.E., ed., Going LOCO:

investigations along the Lower Colorado River: 2016 Desert Symposium, Desert Studies

Center, California, p. 73-86.

Bishop, K.M., 1990, Tectonic model for the evolution of the Texas Canyon basin and Pelona

fault, Soledad basin, southern California, in Buckley, C.I., and Larson, R.A., eds.,

Geology and engineering geology of the western Soledad basin, Los Angeles County,

California: Association of Engineering Geologists, Southern California Section, Annual

Field Trip Guidebook, p. 35-56.

Bishop, K.M., and Ehlig, P.L., 1990, The Pelona fault, central Transverse Ranges, southern

California: an extensional detachment fault?: Geological Society of America Abstracts

with Programs, v. 22, n. 3, p. 8.

Black, L.P., Kamo, S.L., Allen, C.M., Davis, D.W., Aleinikoff, J.N., Valley, J.W., Mundil, R.,

Campbell, I.H., Korsch, R.J., Williams, I.S., and Foudoulis, C., 2004, Improved

206Pb/238U microprobe geochronology by the monitoring of a trace-element-related matrix

effect; SHRIMP, ID-TIMS, ELA-ICP-MS and oxygen isotope documentation for a series

of zircon standards: Chemical Geology, v. 205, p. 115-140,

https://doi.org/10.1016/j.chemgeo.2004.01.003

Bohannon, R.G., 1975, Mid-Tertiary conglomerates and their bearing on Transverse Range

tectonics, southern California, in Crowell, J.C., ed., The San Andreas fault in southern

183

California: A guide to the San Andreas fault from Mexico to Carrizo Plain: California

Division of Mines and Geology Special Report, v. 118, p. 75-82.

Bunker, J.V., and Bishop, K.M., 2001, The San Francisquito fault, southern California: A normal

fault related to uplift of the Sierra Pelona anticline?: Geological Society of America

Abstracts with Programs, v. 33, n. 3, p. 73.

Carman, M.F., Jr., 1964, Geology of the Lockwood Valley area, Kern and Ventura Counties,

California: California Division of Mines and Geology Special Report 81, 62 p.

Chang, Z., Vervoort, J.D., McClelland, W.C., and Knaack, C., 2006, U-Pb dating of zircon by

LA-ICP-MS: Geochemistry, Geophysics, Geosystems, v. 7, Q05009,

https://doi.org/10.1029/2005GC001100

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019: Stratigraphy, provenance and tectonic

significance of the Punchbowl block, San Gabriel Mountains, California, U.S.A.:

Geosphere, v. 15, p. 479-501, https://doi.org/10.1130/GES02025.1

Cole, R.B., and Basu, A.R., 1995, Nd-Sr isotopic geochemistry and tectonics of ridge subduction

and middle Cenozoic volcanism in western California: Geological Society of America

Bulletin, v. 107, p. 167-179, https://doi.org/10.1130/0016-

7606(1995)107<0167:NSIGAT>2.3.CO;2

Coney, P.J., 1980, Cordilleran metamorphic core complexes: An overview, in Crittenden, M.D.,

Jr., Coney, P.J., and Davis, G.H., eds., Cordilleran metamorphic core complexes:

Geological Society of America Memoir 153, p. 7-34.

Crowell, J.C., 1962, Displacement along the San Andreas fault, California: Geological Society of

America Special Paper 71, 61 p., https://doi.org/10.1130/SPE71-p1

184

Crowell, J.C., 1973, Problems concerning the San Andreas fault system in southern California:

Stanford University Publications in Geological Science, v. 13, p. 125-135.

Crowell, J.C., 1975a, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Crowell, J.C., 1975b, Geologic sketch of the Orocopia Mountains, southeastern California, in

Crowell, J.C., ed., San Andreas fault in southern California: A guide to the San Andreas

fault from Mexico to Carrizo Plain: California Division of Mines and Geology Special

Report, v. 118, p. 99–110.

Dalrymple, G.G., 1979, Critical tables for conversion of K-Ar ages from old to new constants:

Geology, v. 7, p. 558-560, https://doi.org/10.1130/0091-

7613(1979)7<558:CTFCOK>2.0.CO;2

Darin, M.H., and Dorsey, R.J., 2013, Reconciling disparate estimates of total offset on the

southern San Andreas fault: Geology, v. 41, p. 975-978,

https://doi.org/10.1130/G34276.1

Dibblee, T.W., Jr., 1996a, Geologic Map of the Agua Dulce Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-58, scale 1:24,000.

Dibblee, T.W., Jr., 1996b, Geologic Map of the Acton Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-59, scale 1:24,000.

Dibblee, T.W., Jr., 1997a, Geologic Map of the Green Valley Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-65, scale 1:24,000.

185

Dibblee, T.W., Jr., 1997b, Geologic Map of the Sleepy Valley and Ritter Ridge Quadrangles,

Los Angeles County, California: Dibblee Geological Foundation Map DF-66, scale

1:24,000.

Doherty, W., 1989, An internal standardization procedure for the determination of yttrium and

the rare earth elements in geological materials by inductively coupled plasma-mass

spectrometry: Spectrochimica Acta, v. 44B, p. 263-280, https://doi.org/10.1016/0584-

8547(89)80031-X

Dykstra, M.R., Van Buer, N.J., and Nourse, J.A., 2018, Whole-rock geochemistry, petrography

and preliminary U-Pb zircon analysis of Oligocene intrusives in the eastern San Gabriel

Mountains: Geological Society of America Abstracts with Programs, v. 50, n. 5,

https://doi.org/10.1130/abs/2018RM-314342

Eaton, G.P., 1958, Miocene volcanic activity in the Los Angeles basin, in Higgins, J.W., ed., A

guide to the geology and oil fields of the Los Angeles and Ventura regions: Los Angeles,

California, American Association of Petroleum Geologists, Pacific Section, p. 55-58.

Ehlert, K.W., 1982, Basin analysis of the Miocene Mint Canyon Formation, southern California,

in Ingersoll, R.V., and Woodburne, M.O., eds., Cenozoic nonmarine deposits of

California and Arizona: Los Angeles, Pacific Section, Society of Economic

Paleontologists and Mineralogists (SEPM), p. 51-64.

Ehlert, K.W., 2003, Tectonic significance of the middle Miocene Mint Canyon and Caliente

Formations, southern California, in Crowell, J.C., ed., Evolution of Ridge Basin, southern

California: An interplay of sedimentation and tectonics: Geological Society of America

Special Paper 367, p. 113-130, https://10.1130/SPE367-p113

186

Ehlig, P.L., 1981, Origin and tectonic history of the basement terrane of the San Gabriel

Mountains, central Transverse Ranges, in Ernst, W.G., ed., The Geotectonic

Development of California (Rubey Volume I): Englewood Cliffs, New Jersey, Prentice-

Hall, p. 253-283.

Ehlig, P.L., and Ehlert, K.W., 1972, Offset of Miocene Mint Canyon Formation from volcanic

source along San Andreas fault, Southern California: Geological Society of America

Abstracts with Programs, v. 4, p. 154.

Frizzell, V.A., Jr., and Weigand, P.W., 1993, Whole-rock K-Ar ages and geochemical data from

middle Cenozoic volcanic rocks, southern California: A test of correlations across the

San Andreas fault, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San

Andreas fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 273-288,

https://doi.org/10.1130/MEM178-p273

Frizzell, V.A., Jr., Mattison, J.M., and Matti, J.C., 1986, Distinctive Triassic megaporphyritic

monzogranite: Evidence for only 160 km offset along the San Andreas fault, southern

California: Journal of Geophysical Research, v. 91, p. 14,080-14,088,

https://10.1029/JB091iB14p14080

Gibbs, A.D., 1984, Structural evolution of extensional basin margins: Journal of the Geological

Society, v. 141, p. 609-620, https://doi.org/10.1144/gsjgs.141.4.0609

Govindaraju, K., 1994, 1994 compilation of working values and sample description for 383

geostandards: Geostandards Newsletter, v. 18, p. 1-158, https://doi.org/10.1046/j.1365-

2494.1998.53202081.x-i1

187

Grove, M., and Jacobson, C.E., 2016, Fossil imprint of the Mendocino Fracture Zone across the

San Andreas fault: Geological Society of America Abstracts with Programs, v. 48, n. 7,

https://doi.org/10.1130/abs/2016AM-287480

Haxel, G.B., Beard, L.S., and Jacobson, C.E., 2018, Significance of northeast-southwest

orientation of prograde lineation in the Pelona-Orocopia-Rand-Schist low-angle

subduction complex, southern California and southwest Arizona: Geological Society of

America Abstracts with Programs. v. 50, no. 5, https://doi.org/10.1130/abs/2018RM-

313363

Hendrix, E.D., 1993, Soledad Basin, Central Transverse Ranges, California: U.S. Geological

Survey Bulletin 2053, p. 243-350.

Hendrix, E.D., and Ingersoll, R.V., 1987, Tectonics and alluvial sedimentation of the upper

Oligocene/lower Miocene Vasquez Formation, Soledad Basin, southern California:

Geological Society of America Bulletin, v. 98, p. 647-663, https://doi.org/10.1130/0016-

7606(1987)98<647:TAASOT>2.0.CO;2

Hendrix, E.D., Cole, R.B., and Ingersoll, R.V., 2010, Soledad and Plush Ranch basins: mid-

Tertiary extensional terrane dismembered by the San Andreas fault system, in Clifton,

H.E., and Ingersoll, R.V., eds., Geologic excursions in California and Nevada: tectonics,

stratigraphy and hydrogeology: Pacific Section, SEPM (Society for Sedimentary Geology)

Book 108, p. 103-171.

Hoyt, J.F., Coffey, K.T., Ingersoll, R.V., and Jacobson, C.E., 2018, Paleogeographic and

paleotectonic setting of the middle Miocene Mint Canyon and Caliente formations,

southern California: An integrated provenance study, in Ingersoll, R.V., Lawton, T.F, and

Graham, S.A., eds., Tectonics, Sedimentary Basins, and Provenance: A Celebration of

188

the Career of William R. Dickinson: Geological Society of America Special Paper 540, p.

463-480, https://doi.org/10.1130/2018.2540(21)

Hsu, K.J., Edwards, G., and McLaughlin, W.A., 1963, Age of the intrusive rocks of the

southeastern San Gabriel Mountains, California: Geological Society of America Bulletin,

v. 74, p. 507-512, https://doi.org/10.1130/0016-7606(1963)74[507:AOTIRO]2.0.CO;2

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Ingersoll, R.V., Pratt, M.J., Davis, P.M., Caracciolo, L., Day, P.P., Hayne, P.O., Petrizzo, D.A.,

Gingrich, D.A., Cavazza, W., Critelli, S., Diamond, D.S., Coffey, K.T., Stang, D.M.,

Hoyt, J.F., Reith, R.C., and Hendrix, E.D., 2014, Paleotectonics of a complex Miocene

half graben formed above a detachment fault: the Diligencia basin, Orocopia Mountains,

southern California: Lithosphere, v. 6, p. 157-176, https://doi.org/10.1130/L334.1

Johnson, D.M., Hooper, P.R., and Conrey, R.M., 1999, XRF analysis of rocks and minerals for

major and trace elements on a single low dilution Li-tetraborate fused bead: Advances in

X-ray Analysis, v. 41, p. 843-867.

Kew, W.S.W., 1924, Geology and oil resources of a part of Los Angeles and Ventura Counties,

California: U.S. Geological Survey Bulletin 753, 202 p.

Konigsberg, W., 1967, Geology along the San Francisquito fault [M.S. thesis]: Los Angeles,

University of Calfornia, 84 p.

Law, R.D., Eriksson, K., and Davisson, C., 2001, Formation, evolution, and inversion of the

middle Tertiary Diligencia basin, Orocopia Mountains, southern California: Geological

189

Society of America Bulletin, v. 113, p. 196-221, https://doi.org/10.1130/0016-

7606(2001)113<0196:FEAIOT>2.0.CO;2

Le Bas, M.J., Lemaitre, R.W., Streckeisen, A., and Zanettin, B., 1986, A chemical classification

of volcanic rocks based on the total alkali silica diagram: Journal of Petrology, v. 27, p.

745-750, https://doi.org/10.1093/petrology/27.3.745

Luyendyk, B.P., Kamerling, M.J., and Terres, R., 1980, Geometric model for Neogene crustal

rotations in southern California: Geological Society of America Bulletin, v. 91, p. 211-

217, https://doi.org/10.1130/0016-7606(1980)91<211:GMFNCR>2.0.CO;2

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

Matti, J.C., Morton, D.M., and Cox, B.F., 1985, Distribution and geologic relations of fault

systems in the vicinity of the central Transverse Ranges, Southern California: U.S.

Geological Survey Open-File Report 85-365, https://doi.org/10.3133/ofr85365

May, D.J., and Walker, N.W., 1989, Late Cretaceous juxtaposition of metamorphic terranes in

the southeastern San Gabriel Mountains, California: Geological Society of America

Bulletin, v. 101, p. 1246-1267, https://doi.org/10.1130/0016-

7606(1989)101<1246:LCJOMT>2.3.CO;2

Miller, F.K., and Morton, D.M., 1977, Comparison of granitic intrusions in the Pelona and

Orocopia schists, southern California: Journal of Research of the U.S. Geological Survey,

v. 5, p. 643-649.

190

Muehlberger, W.R., 1958, Geology of northern Soledad basin, Los Angeles County, California:

American Association of Petroleum Geologists Bulletin, v. 42, p. 1812-1844.

Nourse, J.A., 2002, Middle Miocene reconstruction of the central and eastern San Gabriel

Mountains, southern California, with implications for evolution of the San Gabriel fault

and Los Angeles basin in Barth, A., ed., Contributions to Crustal Evolution of the

Southwestern United States: Geological Society of America Special Paper 365, p. 161-

185, https://doi.org/10.1130/0-8137-2365-5.161

Nourse, J.A., Weigand, P.W., and Hazelton, G.B., 1998, Igneous and tectonic response of the

eastern San Gabriel Mountains to Neogene extension and rotation of the Transverse

Ranges block in Bell, R.J., ed., Field Trip Program, 94th Annual Meeting of the

Cordilleran Section, Geological Society of America, April 3-10, 1998: Long Beach,

California, California State University Long Beach, p. 10-1 to 10-15.

Paces, J.B., and Miller, J.D., 1993, Precise U-Pb ages of Duluth Complex and related mafic

intrusions, northeastern Minnesota: geochronological insights to physical, petrogenetic,

paleomagnetic, and tectonomagmatic processes associated with the 1.1 Ga Midcontinent

Rift System: Journal of Geophysical Research, v. 98, p. 13997-14013,

https://doi.org/10.1029/93JB01159

Paton, C., Woodhead, J., Hellstrom, J., Hergt, J.M., Greig, A., and Maas, R., 2010, Improved

laser ablation U-Pb zircon geochronology through robust downhole fractionation

correction: Geochemistry Geophysics Geosystems, v. 11,

https://doi.org/10.1029/2009GC002618

191

Paton, C., Hellstrom, J., Paul, B., Woodhead, J., and Hergt, J., 2011, Iolite: Freeware for the

visualization and processing of mass spectrometric data: Journal of Analytical Atomic

Spectrometry, v. 26, p. 2508-2518, https://doi.org/10.1039/c1ja10172b

Powell, R.E., 1981, Geology of the crystalline basement complex, eastern Transverse Ranges,

southern California: constraints on regional tectonic interpretation [Ph.D. dissertation]:

Pasadena, California Institute of Technology, 441 p.

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://10.1130/MEM178-p1

Ryan, W.B.F., S.M. Carbotte, J.O. Coplan, S. O'Hara, A. Melkonian, R. Arko, R.A. Weissel, V.

Ferrini, A. Goodwillie, F. Nitsche, J. Bonczkowski, and R. Zemsky, 2009, Global Multi-

Resolution Topography synthesis: Geochemistry, Geophysics, Geosystems, v. 10, p.

Q03014, https://doi.org/10.1029/2008GC002332

Schmitz, M.D., and Bowring, S.A., 2001, U-Pb zircon and titanite systematics of the Fish

Canyon Tuff: an assessment of high-precision U-Pb geochronology and its application to

young volcanic rocks: Geochimica et Cosmochimica Acta, v. 65, p. 2571-2587,

https://doi.org/10.1016/S0016-7037(01)00616-0

Schoene, B., Crowley, J.L., Condon, D.J., Schmitz, M.D., and Bowring, S.A., 2006, Reassessing

the uranium decay constants for geochronology using ID-TIMS U-Pb data: Geochmicia

et Cosmochimica Acta, v. 70, p. 426-445, https://doi.org/10.1016/j.gca.2005.09.007

192

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, https://doi.org/10.1130/G34214.1

Sharp, R.P., 1935, Geology of the Ravenna quadrangle, Los Angeles County, California [M.S.

thesis]: Pasadena, California Institute of Technology, 81 p.

Shelton, J.S., 1955, Glendora volcanic rocks, Los Angeles basin, California: Geological Society

of America Bulletin, v. 66, p. 45-90, https://doi.org/10.1130/0016-

7606(1955)66[45:GVRLAB]2.0.CO;2

Sláma, J., Košler, 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, Plešovice zircon – A new natural reference material

for U-Pb and Hf isotopic microanalysis: Chemical Geology, v. 249, p. 1-35,

https://doi.org/10.1016/j.chemgeo.2007.11.005

Smith, D.P., 1977, San Juan-St Francis fault – Hypothesized major middle tertiary right-lateral

fault in central and southern California: California Division of Mines and Geology

Special Report 129, p. 41-50.

Spittler, T.E., 1974, Volcanic petrology and stratigraphy of non-marine strata, Orocopia

Mountains: their bearing on Neogene slip on the San Andreas fault, southern California

[M.S. thesis]: Riverside, University of California, 115 p.

Spittler, T.E., and Arthur, M.A., 1982, The lower Miocene Diligencia Formation of the Orocopia

Mountains, southern California: Stratigraphy, petrology, sedimentology, and structure, in

Ingersoll, R.V., and Woodburne, M.O., eds., Cenozoic nonmarine deposits of California

193

and Arizona: Los Angeles, Pacific Section, Society of Economic Paleontologists and

Mineralogists (SEPM), p. 83-89.

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. 207-211,

https://doi.org/10.1016/0012-821X(75)90088-6

Stanley, R.G., Wilson, D.S., and McCrory, P.A., 2000, Locations and ages of middle Tertiary

volcanic centers in coastal California: USGS Open-File Report 00-154, 27 p.

Terres, R.R., 1984, Paleomagnetism and tectonics of the central and eastern Transverse Ranges,

southern California [Ph.D. thesis]: Santa Barbara, University of California, 323 p.

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Vermeesch, P., 2018, IsoplotR: a free and open toolbox for geochronology: Geoscience

Frontiers, v. 9, p. 1479-1493, https://doi.org/10.1016/j.gsf.2018.04.001

Weigand, P.W., 1982, Middle Cenozoic volcanism of the Western Transverse Ranges, in Fife,

D.L., and Minch, J.A., eds., Geology and mineral wealth of the California Transverse

Ranges: Santa Ana, South Coast Geological Society Annual Symposium and Guidebook

10, p. 170-188.

Weldon, R.J., II, Meisling, K.E., and Alexander, J., 1993, A speculative history of the San

Andreas fault in the central Transverse Ranges, California, in Powell, R.E., Weldon, R.J.,

II, and Matti, J.C., eds., The San Andreas fault system: displacement, palinspastic

reconstruction, and geologic evolution: Geological Society of America Memoir 178, p.

161-198, https://doi.org/10.1130/MEM178-p161

194

Wendt, I., and Carl, C., 1991, The statistical distribution of the mean squared weighted

deviation: Chemical Geology: Isotope Geoscience section, v. 86, p. 275-285,

https://doi.org/10.1016/0168-9622(91)90010-T

Wernicke, B., and Burchfiel, B.C., 1982, Modes of extensional tectonics: Journal of Structural

Geology, v. 4, p. 105-115, https://doi.org/10.1016/0191-8141(82)90021-9

Woodburne, M.O., 1975, Cenozoic stratigraphy of the Transverse Ranges and adjacent areas,

southern California: Geological Society of America Special Paper 162, 91 p.,

https://doi.org/10.1130/SPE162-p91

Yan, Z., Calyton, R.W., and Saleeby, J., 2005, Seismic refraction evidence for steep faults

cutting highly attenuated continental basement in the central Transverse Ranges,

California: Geophysical Journal International, v. 160, p. 651-666,

https://doi.org/0.1111/j.1365-246X.2005.02506.x

Yin, A., 2002, Passive-roof thrust model for the emplacement of the Pelona-Orocopia Schist in

southern California, United States: Geology, v. 30, p. 183-186,

https://doi.org/10.1130/0091-7613(2002)030<0183:PRTMFT>2.0.CO;2

195

AFTERWORD

The geology of the San Gabriel Mountains, including anticlinoria of schist offset from the

Chocolate Mountains anticlinorium, provides valuable constraints on the evolution of the southern San Andreas fault system. The four chapters of this dissertation combine new data with previous work to provide new insights on actively debated aspects of the San Andreas fault system and Chocolate Mountains anticlinorium. Chapter 1 (Coffey et al., 2019) demonstrates how the previously understudied geology of the Punchbowl block supports original continuity of the anticlinoria of schist in the San Gabriel Mountains with the Chocolate Mountains anticlinorium (e.g., Crowell, 1975), and suggests that 80-110 km of dextral slip on the San

Francisquito-Fenner-Clemens Well fault (e.g., Powell, 1993) is unlikely. Chapter 2 demonstrates how the apparent contradiction of orientations of penetrative lineations within schist of the

Chocolate Mountains anticlinorium with nearby paleomagnetic data (e.g., Haxel et al., 2018) can be resolved if lineations along the western part of the anticlinorium did not form parallel to the former subduction direction, but rather in a NNW/SSE orientation. In contrast to an alternative model for reconciling these datasets (Haxel et al., 2018), my model implies an originally linear, approximately NE/SW to ENE/WSW architecture of the anticlinorium. Chapter 3 proposes a strain-partitioning model which, together with an insight by Darin and Dorsey (2013), can explain the disparate estimates of cumulative slip on the southern San Andreas fault implied by two robust and long-standing cross-fault correlations (e.g., Crowell, 1975; Frizzell et al., 1986;

Matti and Morton, 1993; Ehlert, 2003). This eliminates the apparent incompatability of the smaller of these cumulative-slip estimates with the reconstructions of Chapters 1 and 2. Chapter

4 combines fault-kinematic data with previous mapping of the San Gabriel Mountains to model how the Chocolate Mountains anticlinorium, as reconstructed in the other chapters, developed 196 within a detachment-fault system, and uses geochronologic data to constrain when this development and faulting occurred.

The San Gabriel Mountains expose a vast and varied array of lithologies and structures, providing crucial evidence for the evolution of southern California through geologic time. Its geology has been studied for a century, and much information has been published on the subject.

This information is invaluable; it would take a prohibitive effort to replicate, and in numerous locations, would be impossible due to a lack of access and extensive land development. On the other hand, much of this work was done prior to critical advances in the field of geology, such as the theory of plate tectonics and the methods of radiometric dating, as well as important regional and local findings. Accordingly, existing data are often incomplete or biased, and their original interpretations outdated. This dissertation highlights how thoughtful collection of new data can complement, correct, and leverage existing findings, revealing new implications and allowing for more robust interpretations. Additionally, the series of deformational events experienced by the

San Gabriel Mountains, which make them such a valuable geologic record, complicates understanding the details of older events. As highlighted by Chapter 4 in particular, in order to rigorously examine older events, younger deformation must be sequentially reversed. I hope that geologic research will continue in the San Gabriel Mountains and throughout southern

California, that it will use our ever-improving understanding of younger deformational events to better establish the original geometries of older structures and units, and that it will take full advantage of the body of existing work while reevaluating the often entrenched interpretations that accompany it.

197

REFERENCES CITED

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019, Stratigraphy, provenance and tectonic

significance of the Punchbowl block, San Gabriel Mountains, California, USA:

Geosphere, v. 15, p. 479-501, https://doi.org/10.1130/GES02025.1

Crowell, J.C., 1975, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Darin, M.H., and Dorsey, R.J., 2013, Reconciling disparate estimates of total offset on the

southern San Andreas fault: Geology, v. 41, p. 975-978,

https://doi.org/10.1130/G34276.1

Ehlert, K.W., 2003, Tectonic significance of the middle Miocene Mint Canyon and Caliente

Formations, southern California, in Crowell, J.C., ed., Evolution of Ridge Basin, southern

California: An interplay of sedimentation and tectonics: Geological Society of America

Special Paper 367, p. 113-130, https://doi.org/10.1130/0-8137-2367-1.113

Frizzell, V.A., Jr., Mattinson, J.M., and Matti, J.C., 1986, Distinctive Triassic megaporphyritic

monzogranite: evidence for only 160 km offset along the San Andreas fault, southern

California: Journal of Geophysical Research, v. 91, p. 14080-14088,

https://doi.org/10.1029/JB091iB14p14080

Haxel, G.B., Beard, L.S., and Jacobson, C.E., 2018, Significance of northeast-southwest

orientation of prograde lineation in the Pelona-Orocopia-Rand-Schist low-angle

subduction complex, southern California and southwest Arizona: Geological Society of

America Abstracts with Programs. v. 50, no. 5, https://doi.org/10.1130/abs/2018RM-

313363

198

Matti, J.C., and Morton, D.M., 1993, Paleogeographic evolution of the San Andreas fault in

southern California: A reconstruction based on a new cross-fault correlation, in Powell,

R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas fault system: displacement,

palinspastic reconstruction, and geologic evolution: Geological Society of America

Memoir 178, p. 107-159, https://doi.org/10.1130/MEM178-p107

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

199

APPENDIX A: CHAPTER 2 – BASIS OF REPRESENTATIVE

ORIENTATIONS OF PENETRATIVE LINEATIONS USED

IN FIGURE 2-9 AND TABLE 2-2

For lineation data of this study, representative orientations are mean values of tilt- corrected orientations of penetrative lineations (from region M of the Mount Pinos area, region

SW of Sierra Pelona, and region NE of the Crafton Hills; see Figs. 2-3 through 2-7). For previously published lineation data, such mean values generally cannot be calculated, and so representative orientations were approximated visually, as in Haxel et al. (2018). For the sake of consistency and greater objectivity, I typically (but with some exceptions, explained below) used the representative orientations determined in this same manner by Haxel et al. (2018).

Representative orientations, the data from which they were determined, and a comparison to those of Haxel et al. (2018) are given below.

Schist exposure Represent- Data from which representative Representative orientation of penetrative ative orientation was selected lineations determined by Haxel et al. (and abbreviation) orientation (2018; azimuth degrees) (azimuth degrees) eastern Castle Dome 47.0°* Haxel et al. (2002) Grouped with CDMW, NSR, and TM (TM Mtns. (CDME) (107.0°)* dominates): group representative orientation 34.5°. western Castle Dome 42.5° Haxel et al. (2002) Grouped with CDME, NSR, and TM (TM Mtns. (CDMW) dominates): group representative orientation 34.5°. Crafton Hills (CH) 56.0° This study (region NE) None determined. Cemetery Ridge (CR) 42.0° Haxel et al. (2015) Identical. Gavilan Hills (GH) 30.0° Jacobson et al. (2002) (Fig. 3B, 40.5°, because lineations deflected away from northern and clockwise along the northern and eastern edges of schist eastern edges of schist exposure (see exposure) discussion in text) were included. Mount Pinos (MP) 99.5°† This study (region M) None determined.

200

Neversweat Ridge 67.0° Haxel et al. (2002) Grouped with CDME, CDMW, and TM (NSR) (TM dominates): group representative orientation 34.5°. Orocopia Mountains 61.5° Jacobson and Dawson (1995) 38.5°, because a different dataset was (OM) (Fig. 3, & Fig. 4 away from used (i.e., Jacobson et al., 1988). northern edge of schist exposure) Picacho District (PD) 28.0° Dillon et al. (1990) Identical. Peter Kane Mountain 4.0° Dillon et al. (1990) Identical. (PKM) Plomosa Mtns. (PM) 46.0° Strickland et al. (2018) None determined. Rand Mtns. (RM) 18.5° Postlethwaite and Jacobson Identical. (1987) southern Chocolate 172.5° Dillon et al. (1990) Identical. Mountains (SCM) Sierra de Salinas 29.5° Chapman et al. (2010) Identical. (SdS) eastern San Emigdio 116.5°† Chapman et al. (2010) None determined. Mountains (SEME) western San Emigdio 97.5°† Chapman et al. (2010) None determined. Mountains (SEMW) Sierra Pelona (SP) 43.5°§ This study (region SW) 48° (32°)§, because a different dataset (27.5°)§ was used (i.e., Harvill, 1969). northern part of 148.0°§ Jacobson (1983a) Identical. southern San Gabriel (132.0°)§ Mtns. (SSGMN) southern part of 16.0°§ May and Walker (1989) None determined. southern San Gabriel (0.0°)§ Mtns. (SSGMS) Trigo Mountains (TM) 30.0° Haxel et al. (2002) Grouped with CDME, CDMW, and NSR (TM dominates): group representative orientation 34.5°.

Note: Mtns. = Mountains. *First value is corrected for 60° of clockwise vertical-axis rotation of the eastern Castle Dome Mountains (see discussion in text); second value (italicized, in parentheses) is the uncorrected, modern orientation. †These orientations have likely been altered by Pliocene-Quaternary deformation related to the San Andreas fault system: see Chapman et al. (2010) and discussion in text. §First value is corrected for 16° of post-ca. 12 Ma, counterclockwise vertical-axis rotation of the San Gabriel block (see discussion in text); second value (italicized, in parentheses) is the uncorrected, modern orientation.

201

APPENDIX B: CHAPTER 3 – RELATIONSHIP BETWEEN

RATE OF TOTAL DIFFERENTIAL MOTION, DEXTRAL

SLIP RATE, AND DIFFUSE, NORTH-SOUTH

SHORTENING RATE IN THE RESTRAINING DOUBLE

BEND OF THE SAN ANDREAS FAULT

Along the restraining double bend of the San Andreas fault, the rate of total differential movement, Ṫ, is equal in orientation and magnitude to the dextral slip rate on the San Andreas fault outside of this bend. The segment of the San Andreas fault within the restraining double bend (the Mojave segment) is oriented counterclockwise to the rest of the San Andreas fault

(e.g., the Coachella Valley segment) by an angle θ, and has a dextral slip rate of Ḋ. The rate of north-south shortening, Ṡ, is oriented at an angle γ to Ṫ. The vector sum of Ḋand Ṡ is Ṫ. This defines a triangle, which can be divided into two right triangles by a line segment, L, as shown below:

202

Note that:

(1) 퐿 = Ṡ sin(훾) = Ḋ sin(휃)

Also note that:

Ṫ = Ṡ cos(훾) + Ḋ cos(휃)

Ṫ − Ḋ cos(휃) (2) Ṡ = cos (훾)

Substituting (2) into (1):

Ṫ − Ḋ cos(휃) sin(훾) = Ḋ sin(휃) cos(훾)

Ḋ sin(휃) cos(훾) = Ṫ sin(훾) − Ḋ sin(훾)cos(휃)

Ḋ sin(휃) cos(훾) + Ḋ sin(훾)cos(휃) = Ṫ sin(훾)

Ḋ [sin(휃) cos(훾) + sin (훾)cos(휃)] = Ṫ sin(훾)

퐬퐢퐧(휸) (ퟑ)Ḋ = Ṫ 퐬퐢퐧(휽) 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬(휽)

This equation can be used to calculate the dextral slip rate expected on the Mojave segment of the San Andreas fault for a given value of total differential movement rate.

203

Solving (1) for Ṡ:

Ṡ sin(훾) = Ḋ sin(휃)

sin(휃) (4) Ṡ = Ḋ sin(훾) or, substituting (3) into (4):

sin(휃) sin(훾) Ṡ = [ ] Ṫ sin(훾) sin(휃) cos(훾) + sin(훾)cos(휃)

퐬퐢퐧(휽) (ퟓ) Ṡ = Ṫ 퐬퐢퐧(휽) 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬(휽)

This equation can be used to calculate the diffuse, north-south shortening rate expected beside the Mojave segment of the San Andreas fault for a given value of total differential movement rate.

Assuming a fixed geometry for the restraining double bend (i.e., a constant θ), then all trigonometric terms in (3) and (5) are constants. Thus, these equations can easily be integrated with respect to time to give the cumulative dextral slip on the Mojave segment, D, and cumulative north-south shortening, S, as functions of the cumulative total differential motion, T:

sin(훾) 퐷 = 푇 + 퐶 sin(휃) cos(훾) + sin(훾)cos(휃) 1

sin(휃) 푆 = 푇 + 퐶 sin(휃) cos(훾) + sin(훾)cos(휃) 2

204

But, since this model assumes that Mojave-segment slip and diffuse shortening initiate at the same time as the rest of the southern San Andreas fault, then for T = 0, D = 0 and S = 0, so:

퐬퐢퐧(휸) 푫 = 푻 퐬퐢퐧(휽) 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬(휽)

퐬퐢퐧(휽) 푺 = 푻 퐬퐢퐧(휽) 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬(휽)

These equations are the same simple trigonometric relationships as in (3) and (5), and can be used to calculate the cumulative dextral slip on the Mojave segment, D, and the cumulative north-south shortening, S, for a given value of total differential movement, T.

Assuming a total differential movement of T = 215 km (see discussion in text), θ = 25°, and γ = 44°, then:

D = 160 km

S = 97 km

205

APPENDIX C: CHAPTER 3 – GEOMETRIC MODEL FOR

EVOLUTION OF THE RESTRAINING DOUBLE BEND OF

THE SAN ANDREAS FAULT

The geometry of the restraining double bend of the San Andreas fault, shown in Figure 3-

1, can be modeled as the simplified geometry shown below:

In this model, the Sierra Nevada and Peninsular Ranges batholiths move relative to one-another due to motion on the San Andreas fault (including the Banning fault). Points A and B, which define the ends of the restraining double bend, move with these batholiths. Thus, the geometry of the restraining double bend changes with time.

The restraining double bend has a length, m (at present, 255 km). This can be decomposed into a length, y, parallel to the San Andreas fault outside the restraining double bend, and a length, x, perpendicular to it (at present, yp = 231 km and xp = 108 km). The segment of the San Andreas fault within the restraining double bend (the Mojave segment) is oriented 206 counterclockwise to y and to the rest of the San Andreas fault (e.g., the Coachella Valley segment) by an angle θ (at present, θp = 25°).

Using estimates of the timing of initiation of the faults in this model, and their cumulative slip outside the restraining double bend, and assuming that the combined slip rate on the San

Andreas and San Jacinto faults has been constant, average slip rates for these faults can be calculated:

fault: cumulative dextral slip (km) time of initiation (Ma) slip rate (km/Ma) 5 (e.g., Nicholson et al., 1994; Ingersoll and Rumelhart, 1999; Oskin et al., 2001; combined approximately 215 Crowell, 2003; Oskin and Stock, 2003; San Andreas 43 (see discussion in Chapter 3) some slip likely started occurring & San Jacinto accumulating at a lower rate ca. 8 Ma (Dorsey et al., 2011; Bennett et al., 2016; Darin et al., 2016)

approximately 190 1.5-0 Ma: 26.3 San Andreas see above (see discussion in Chapter 3) 5-1.5 Ma: 43

1.5 25 (Morton and Matti, 1993; Albright, San Jacinto 16.7 (Sharp, 1967) 1999; Dorsey, 2001; but see discussion in Dorsey and Roering, 2006)

Slip on the San Andreas fault reduces the length of y. Slip on the San Jacinto fault does not cause any relative movement of points A and B, and thus does not alter the geometry of the restraining double bend. Thus, y and x can be written as functions of time, t (in millions of years before present, Ma):

(1) 푦(푡) = 푦p + 푟SA푡, and

(2) 푥(푡) = 푥p, where rSA is the average slip rate on the San Andreas fault outside of the restraining double bend

(in km/Ma), and yp and xp are the present lengths of y and x, respectively (in km).

207

푥 휃 = tan−1 ( ) 푦

Substituting in equations (1) and (2) yields an expression for θ as a function of time, t:

푥p 휃(푡) = tan−1 [ ] 푦p + 푟SA푡

Substituting in the values listed above gives the following expressions for θ as a function of time:

ퟏퟎퟖ 퐤퐦 휽(풕) = 퐭퐚퐧−ퟏ [ ] , ퟎ 퐌퐚 ≤ 풕 ≤ ퟏ. ퟓ 퐌퐚 퐤퐦 ퟐퟑퟏ 퐤퐦 + ퟐퟔ. ퟑ 풕 퐌퐚

ퟏퟎퟖ 퐤퐦 휽(풕) = 퐭퐚퐧−ퟏ [ ] , 풕 > ퟏ. ퟓ 퐌퐚 퐤퐦 퐤퐦 ퟐퟑퟏ 퐤퐦 + ퟐퟔ. ퟑ (ퟏ. ퟓ 퐌퐚) + ퟒퟑ (풕 − ퟏ. ퟓ 퐌퐚) 퐌퐚 퐌퐚 Note that there is one expression for times between the present and 1.5 Ma, and a second for times before 1.5 Ma. This is because of the change in slip rate of the San Andreas fault, rSA, 1.5

Ma, when the San Jacinto fault initiated.

For t = 5 Ma, when the southern San Andreas fault initiated (see references in table above), the second of the above expressions for θ reduces to:

휽풕=ퟓ 퐌퐚 = ퟏퟒ. ퟒ°

And, the angle, Δθ, by which the restraining double bend has tightened is thus:

휟휽 = 휽풕=ퟎ 퐌퐚 − 휽풕=ퟓ 퐌퐚

휟휽 = ퟐퟓ° − ퟏퟒ. ퟒ° = ퟏퟎ. ퟔ°

208

26 24

22 20 18

16

(t) (degrees)(t)

14 θ 12 10 5 4 3 2 1 0 time (Ma)

REFERENCES CITED:

Albright, L.B., III, 1999, Magnetostratigraphy and biochronology of the San Timoteo Badlands,

southern California, with implications for local Pliocene–Pleistocene tectonic and

depositional patterns: Geological Society of America Bulletin, v. 111, p. 1265-1293,

https://doi.org/10.1130/0016-7606(1999)111<1265:MABOTS>2.3.CO;2

Bennett, S.E.K., Oskin, M.E., Iriondo, A., and Kunk, M.J., 2016, Slip history of the La Cruz

fault: Development of a late Miocene transform in response to increased rift obliquity in

the northern Gulf of California: Tectonophysics, v. 693, p. 409-435,

https://doi.org/10.1016/j.tecto.2016.06.013

Crowell, J.C., 2003, Introduction to geology of Ridge Basin, southern California, in Crowell,

J.C., ed., Evolution of Ridge Basin, southern California: An interplay of sedimentation

and tectonics: Geological Society of America Special Paper 367, p. 1-15,

https://doi.org/10.1130/0-8137-2367-1.1

Darin, M.H., Bennett, S.E.K., Dorsey, R.J., Oskin, M.E., and Iriondo, A., 2016, Late Miocene

extension in coastal Sonora, México: Implications for the evolution of dextral shear in the

209

proto-Gulf of California oblique rift: Tectonophysics, v. 693, p. 378-408,

https://doi.org/10.1016/j.tecto.2016.04.038

Dorsey, R.J., 2001, Two-stage evolution of the San Jacinto fault zone: Crustal response to

Pleistocene oblique collision along the San Andreas fault: EOS Transactions of the AGU

Fall Meeting Supplement, v. 82, p. 933.

Dorsey, R.J., and Roering, J.J., 2006, Quaternary landscape evolution in the San Jacinto fault

zone, Peninsular Ranges of Southern California: Transient response to strike-slip fault

initiation: Geomorphology, v. 73, p. 16-32,

https://doi.org/10.1016/j.geomorph.2005.06.013

Dorsey, R.J., Housen, B.A., Janecke, S.U., Fanning, C.M., and Spears, A.L.F., 2011,

Stratigraphic record of basin development within the San Andreas fault system: Late

Cenozoic Fish Creek-Vallecito basin, southern California: Geological Society of America

Bulletin, v. 123, p. 771-793, https://doi.org/10.1130/B30168.1

Ingersoll, R.V., and Rumelhart, P.E., 1999, Three-stage evolution of the Los Angeles basin,

southern California: Geology, v. 27, p. 593-596, https://doi.org/10.1130/0091-

7613(1999)027<0593:TSEOTL>2.3.CO;2

Morton, D.M., and Matti, J.C., 1993, Extension and contraction within an evolving divergent

strike-slip fault complex: The San Andreas and San Jacinto fault zones at their

convergence in southern California, in Powell, R.E., Weldon, R.J., II, and Matti, J.C.,

eds., The San Andreas fault system: displacement, palinspastic reconstruction, and

geologic evolution: Geological Society of America Memoir 178, p. 217-230,

https://doi.org/10.1130/MEM178-p217

210

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. 491-495,

https://doi.org/10.1130/0091-7613(1994)022<0491:MCROTW>2.3.CO;2

Oskin, M., and Stock, J., 2003, Marine incursion synchronous with plate-boundary localization in

the Gulf of California: Geology, v. 31, p. 23-26, https://doi.org/10.1130/0091-

7613(2003)031<0023:MISWPB>2.0.CO;2

Oskin, M., Stock, J., and Martin-Barajas, A., 2001, Rapid localization of Pacific-North America

plate motion in the Gulf of California: Geology, v. 29, p. 459-462,

https://doi.org/10.1130/0091-7613(2001)029<0459:RLOPNA>2.0.CO;2

Sharp, R.V., 1967, San Jacinto fault zone in the Peninsular Ranges of southern California:

Geological Society of America Bulletin, v. 78, p. 705-730, https://doi.org/10.1130/0016-

7606(1967)78[705:SJFZIT]2.0.CO;2

211

APPENDIX D: CHAPTER 3 – CALCULATION OF

COMPONENTS OF DEXTRAL SLIP AND NORTH-SOUTH

SHORTENING IN A RESTRAINING DOUBLE BEND OF

TIME-VARYING GEOMETRY

As explained in Chapter 3 and Appendix B, within the restraining double bend of the San

Andreas fault, where the Mojave segment is oriented counterclockwise to the rest of the San

Andreas fault by an angle θ, the total rate of differential movement, Ṫ, is expressed as a dextral slip rate, Ḋ, on the Mojave segment of the San Andreas fault, coupled with a north-south shortening rate, Ṡ, beside it. These rates are related by the following equations:

sin(훾) Ḋ = Ṫ sin(휃) cos(훾) + sin(훾)cos(휃)

sin(휃) Ṡ = Ṫ sin(휃) cos(훾) + sin(훾)cos(휃) where γ is the angle between north-south shortening and the typical orientation of the San

Andreas fault.

In Appendix B, these expressions were integrated with respect to time, assuming a fixed geometry for the restraining double bend (i.e., constant θ). However, as explained in Chapter 3 and Appendix C, the restraining double bend has likely tightened through time (i.e., θ is a function of time). In this case:

퐬퐢퐧(휸) Ḋ = Ṫ 퐬퐢퐧[휽(풕)] 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬[휽(풕)]

212

퐬퐢퐧[휽(풕)] Ṡ = Ṫ 퐬퐢퐧[휽(풕)] 퐜퐨퐬(휸) + 퐬퐢퐧(휸)퐜퐨퐬[휽(풕)] where θ(t) is θ as a function of time.

The expressions for θ(t) derived in Appendix C make integration of the above expressions to give analytical solutions for cumulative dextral slip, D, and cumulative north- south shortening, S, a daunting task. However, numerical solutions can readily be determined using Riemann sums.

Using the expressions for θ(t) derived in Appendix C (one for times between the present and 1.5 Ma, when the San Jacinto fault initiated, and one for times between 1.5 and 5 Ma, when the southern San Andreas fault initiated; see references in Appendix C), and assuming, as in previous calculations, γ = 44° and T = 215 km (see discussion in Chapter 3), then:

푫 ≈ ퟏퟔퟒ 퐤퐦

푺 ≈ ퟕퟕ 퐤퐦

250

200

150

T

100 D

S cumulative movement (km) 50

0 5 4 3 2 1 0 time (Myr ago)

213

APPENDIX E: CHAPTER 3 – CALCULATION OF

COMPONENTS OF DEXTRAL SLIP AND NORTH-SOUTH

SHORTENING INCORPORATING EASTWARD

EXTRUSION OF THE MOJAVE BLOCK ALONG THE

GARLOCK FAULT

In Appendix B, I proposed the following vector-sum relationship, assuming a fixed geometry for the restraining double bend:

where T is the total cumulative slip along the San Andreas fault south of the restraining double bend, D is the cumulative dextral slip on the Mojave segment of the San Andreas fault, within the restraining double bend, and S is the cumulative north-south shortening beside the San

Andreas fault within the restraining double bend.

The Mojave block may experience eastward extrusion via sinistral slip on the Garlock fault (Hill and Dibblee, 1953).

214

The northern Chocolate Mountains, part of the cross-fault correlation indicating approximately

215 km (240 km - 25 km of transferred slip) of cumulative dextral slip on the San Andreas fault south of the restraining double bend (see Chapter 3), may lie outside of the Mojave block. If both of these possibilities are true, then the above relationship must be modified to include slip on the

Garlock fault, G, oriented an angle α clockwise of a line perpendicular to T:

Note that:

푁2 = [푇 + 퐺sin(훼)]2 + [퐺cos(훼)]2

(1) 푁 = √[푇 + 퐺sin(훼)]2 + [퐺cos(훼)]2

215

Also note that:

퐺cos(α) (2) 훽 = tan−1 [ ] 푇 + 퐺sin(α) and:

훿 + 훽 = 휃

(3) 훿 = 휃 − 훽

Substituting (2) into (3):

퐺cos(α) (4) 훿 = 휃 − tan−1 [ ] 푇 + 퐺sin(α)

Also note that:

(180° − 휀) + 훾 + 훽 = 180°

(5) 휀 = 훾 + 훽

Substituting (2) into (5):

퐺cos(α) (6) 휀 = 훾 + tan−1 [ ] 푇 + 퐺sin(α)

The triangle with sides D, S, and N is analogous to the triangle with sides D, S, and T in the first figure and in supplementary file S1. Using the same math as in supplementary file S1, D and S can be expressed in terms of N, δ, and ε:

216

푁sin(휀) (7) 퐷 = sin(훿) cos(휀) + sin(휀)cos(훿)

푁sin(훿) (8) 푆 = sin(훿) cos(휀) + sin(휀)cos(훿)

Substituting (1), (4), and (6) into (7) and (8):

푮퐜퐨퐬(훂) √[푻 + 푮퐬퐢퐧(휶)]ퟐ + [푮퐜퐨퐬(휶)]ퟐ퐬퐢퐧 {휸 + 퐭퐚퐧−ퟏ [ ]} 푻 + 푮퐬퐢퐧(훂) = 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 퐬퐢퐧 {휽 − 퐭퐚퐧−ퟏ [ ]} 퐜퐨퐬 {휸 + 퐭퐚퐧−ퟏ [ ]} + 퐬퐢퐧 {휸 + 퐭퐚퐧−ퟏ [ ]} 퐜퐨퐬 {휽 − 퐭퐚퐧−ퟏ [ ]} 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂)

푮퐜퐨퐬(훂) √[푻 + 푮퐬퐢퐧(휶)]ퟐ + [푮퐜퐨퐬(휶)]ퟐ퐬퐢퐧 {휽 − 퐭퐚퐧−ퟏ [ ]} 푻 + 푮퐬퐢퐧(훂) = 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 푮퐜퐨퐬(훂) 퐬퐢퐧 {휽 − 퐭퐚퐧−ퟏ [ ]} 퐜퐨퐬 {휸 + 퐭퐚퐧−ퟏ [ ]} + 퐬퐢퐧 {휸 + 퐭퐚퐧−ퟏ [ ]} 퐜퐨퐬 {휽 − 퐭퐚퐧−ퟏ [ ]} 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂) 푻 + 푮퐬퐢퐧(훂)

These equations can be used to calculate the cumulative dextral slip on the Mojave segment, D, and the cumulative north-south shortening, S, for given values of total differential movement, T, and associated slip on the Garlock fault, G.

The appropriate value for G in this model is debatable. Cumulative slip on the Garlock fault is approximately 64 km (Smith, 1962; Smith et al., 1968; Smith and Ketner, 1970;

Monastero et al., 1997), and initiated ca. 10 Ma (Burbank and Whistler, 1987). This would imply an average slip rate of 6.4 mm/yr, which is within error of the 7.6 mm/yr Quaternary slip rate determined for the western Garlock fault by McGill et al. (2009). All of this slip may be related to the San Andreas fault, especially given recent, ca. 8 Ma estimates for initiation of the southern

San Andreas fault (references in Chapter 3 and Appendix C). Alternatively, if one assumes that the Garlock slip rate has been approximately constant since initiation, but that substantial slip did 217 not occur on the southern San Andreas until ca. 5 Ma (references in Chapter 3 and Appendix C), then approximately half of this slip, i.e., 32 km, would have been coeval with, and thus potentially related to, slip on the southern San Andreas fault.

Slip on the Garlock fault related to slip on the San Andreas fault and to eastward extrusion of the Mojave block should extend to the western end of the Garlock fault. Dibblee

(1967) estimated 72 km of cumulative slip at the western end of the Garlock fault, which would be broadly consistent with the above estimate. Powell (1993), however, estimated no more than

12 km of cumulative slip at the western end of the Garlock fault, in which case additional slip farther east presumably had a different cause, and 12 km or less should be used when calculating strain partitioning within the restraining double bend.

Assuming, as in previous calculations, a total differential movement of T = 215 km (see discussion in Chapter 3), θ = 25°, and γ = 44°, and assuming α = 10°; then:

If G = 64 km, then: D = 216 km, S = 41 km

If G = 32 km, then: D = 188 km, S = 69 km

If G = 12 km, then: D = 171 km, S = 87 km

REFERENCES CITED:

Burbank, D.W., and Whistler, D.P., 1987, Temporally constrained tectonic rotations derived

from magnetostratigraphic data: Implications for the initiation of the Garlock fault,

California: Geology, v. 15, p. 1172-1175, https://doi.org/10.1130/0091-

7613(1987)15<1172:TCTRDF>2.0.CO;2

Dibblee, T.W., Jr., 1967, Areal geology of the western Mojave Desert, California: U.S.

Geological Survey Professional Paper 522, 153 p.

218

McGill, S.F., Wells, S.G., Fortner, S.K., Kuzma, H.A., and McGill, J.D., 2009, Slip rate of the

western Garlock fault, at Clark Wash, near Lone Tree Canyon, Mojave Desert,

California: Geological Society of America Bulletin, v. 121, p. 536-554,

https://doi.org/10.1130/B26123.1

Monastero, F.C., Sabin, A.E., and Walker, J.D., 1997, Evidence for post-early Miocene initiation

of movement on the Garlock fault from offset of the Cudahy Camp Formation, east-

central California: Geology, v. 25, p. 247-250, https://doi.org/10.1130/0091-

7613(1997)025<0247:EFPEMI>2.3.CO;2

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

Smith, G.I., 1962, Large lateral displacement on Garlock fault, California, as measured from

offset dike swarm: American Association of Petroleum Geologists Bulletin, v. 46, p. 85–

104.

Smith, G.I., and Ketner, K.B., 1970, Lateral displacement on the Garlock fault, southeastern

California, suggested by offset sections of similar metasedimentary rocks: U.S.

Geological Survey Professional Paper 700-D, p. D1-D9.

Smith, G.I., Troxel, B.W., Gray, C. H., Jr., and von Huene, R. E., 1968, Geologic reconnaissance

of the Slate Range, San Bernardino and Inyo counties, California: California Division of

Mines and Geology Special Report 96, 33 p.

219

APPENDIX F: CHAPTER 4 – DETAILS OF PALINSPASTIC

RECONSTRUCTIONS (FIG. 4-5)

Geologic map in A modified from Ehlig (1981), Dibblee (1989, 1991a, b, 1993, 1996a, b, c, d, e,

1997a, b, c, d, 1998, 2001, 2002a, b, c, d, e, f, g, h, i, j, k, l, m, 2003a, b, c, d, 2004, 2006a, b,

2008), Nourse (2002), and Chapter 1 (Coffey et al., 2019).

Step 1 - From A (present) to B (ca. 5 Ma):

Geologic units removed: Pliocene-Pleistocene Saugus Formation (this unit had not yet formed at the time of B; Kew, 1924).

Vertical-axis rotation restorations: 16° of counterclockwise rotation is restored based on 16° ±

30° of counterclockwise rotation implied by paleomagnetism of tuff beds within the upper part of the Mint Canyon Formation, dated ca. 11 or 10 Ma (Terres and Luyendyk, 1985). This rotation has been interpreted as related to the transpressional restraining double bend in the southern San Andreas fault (Terres and Luyendyk, 1985; Ingersoll and Coffey, 2017), which initiated ca. 8-5 Ma (references below), and thus this rotation is restored in B (ca. 5 Ma), rather than C or D. This rotation is broadly supported by paleomagnetic evidence from the Punchbowl block (Liu, 1990). Paleomagnetic data from Ridge Basin Group deposits suggest that no major rotations occurred in the northwestern San Gabriel block after 8.5 Ma (Ensley and Verosub,

1982), however, approximately 16° of counterclockwise rotation since ca. 5 Ma is within error of the subset of these measurements from deposits younger than ca. 5 Ma.

220

Fault restorations: (listed east to west)

Fault Abbrev. Slip restored Evidence Timing constraints References

Southern none Kilometers of San Andreas is a dextral The southern San e.g.: Hill and San Andreas dextral slip, transform plate boundary, Andreas initiated ca. Dibblee, 1953; likely either as indicated by offset 8-5 Ma, and is Crowell, 1975; approximately features, geodesy, currently active. Any Powell, 1993; 240 km or 160- earthquake focal pre-5 Ma slip was Nicholson et 185 km, mechanisms, trenching, etc. likely at a lower rate al., 1994; including slip on than post-5 Ma slip. Oskin et al., the Nf and Pf 2001; Dorsey (see below). et al., 2011; (precise Bennett et al., restoration of 2016; Darin et San Andreas al., 2016 fault slip is not attempted here)

Glen Helen GHf 5.0 km of dextral Part of the dextral San Must have initiated Cramer and slip, merging Jacinto fault system. Right after the southern Harrington, with ESJf to the separation of Telegraph San Andreas (ca. 8- 1987; Morton northwest for a Peak pluton/Pelona Schist 5 Ma), likely much and Matti, combined 12.0 intrusive contact and of later (≤1.5 Ma); 1993; Norum, km of dextral internal stratigraphic microseismicity 1997; Nourse, slip. 1.0 km of markers and foliation indicates San 2002 northwest-side- domains within the Pelona Jacinto fault system up reverse slip. Schist indicate 5.0 km of is currently active. offset.

San Jacinto SJf 10.5 km of Part of the dextral San Same as above Same as dextral slip Jacinto fault system. above Combination of 7.0 and 3.5 km of dextral slip on the ESJf and Western WSJ/Sf, which merge to the southwest (see below).

Eastern San ESJf 7.0 km of dextral Part of the dextral San Same as above Same as Jacinto slip, merging Jacinto fault system. Dextral above with GHf to the separation of Telegraph northwest for a Peak pluton/Pelona Schist combined 12.0 intrusive contact and of km of dextral internal stratigraphic slip; merging markers and foliation with WSJ/Sf to domains within the Pelona the southeast to Schist indicate 7.0 km of form the San offset. Jacinto fault (SJf) for a combined 10.5 km of dextral

221

slip

Western San WSJ/Sf 3.5 km of dextral Part of the dextral San Same as above Same as Jacinto / slip Jacinto fault system. Right above Scotland separation of Telegraph (merging with Peak pluton/Pelona Schist WSJ/Sf to the intrusive contact and of southeast to internal stratigraphic form the San markers and foliation Jacinto fault domains within the Pelona (SJf) for a Schist indicate 5.0 km of combined 10.5 offset. km of dextral slip

Nadeau and Nf, Pf Kilometers of An inactive strand of the Initiated ca. 8-5 Ma e.g., Dibblee, Punchbowl dextral slip, southern San Andreas fault, with the southern 1968; Ehlig, likely and thus dextral. Correlation San Andreas fault, 1968 approximately of offset equivalents was inactive by ca. 40-50 km suggests approximately 40- 3 Ma. (precise 50 km of dextral slip. restoration of Nf- Pf slip is not attempted here)

Lytle Creek LCf 0 km of dextral Presumably part of the Same as above Morton and slip dextral San Jacinto fault Matti, 1993; system, and thus dextral. No Dibblee, known clear offset features. 2003b, c

San Gabriel Cuf, Df, Top-to-south This is a major zone of Active, as e.g., Crook et block frontal SMf reverse slip reverse faulting, as evidenced by al., 1987; faults: (precise indicated by ongoing uplift of Quaternary Pechmann, Cucamonga, restoration of the San Gabriel Mountains, displacements; 1987 Duarte and frontal-fault slip offset alluvial deposits, initiation post-dated Sierra Madre is not attempted earthquake focal inception of here) mechanisms, trenching, etc. southern San Andreas fault ca. 8- 5 Ma.

Grapevine Gf 0 km of dextral Presumably dextral, like the Fault postdates Dibblee, slip similarly oriented LCf and Cretaceous 2003b, c DuCf immediately to the plutonics; given northeast and southwest, proximity to and respectively. matching kinematics with San Jacinto fault zone, likely similar timing (see above). Fault appears to terminate against the West Lytle Creek fault (WLCf).

222

Duncan DuCf 1.5 km of dextral Has been interpreted as and Same as above Morton and Canyon slip is almost certainly dextral, Matti, 1993; given orientation, straight Dibblee, trace, proximity to San 2003b, c Jacinto fault system and right-separation of contacts. Right-separation of Cretaceous plutonics/”black- belt” mylonite and “black- belt” mylonite/Cucamonga granulite contacts suggests approximately 1.5 km of dextral slip.

Morse MC/DCf 1.0 km of dextral Has been interpreted as and Same as above Morton and Canyon / slip is almost certainly dextral, Matti, 1993; Day Canyon given orientation, straight Dibblee, 2003c trace, proximity to San Jacinto fault system and right-separation of contacts. Right-separation of Cretaceous plutonics/”black- belt” mylonite and “black- belt” mylonite/Cucamonga granulite contacts suggests approximately 1.0 km of dextral slip.

Demens DeCf 0 km of dextral Has been interpreted as and Fault postdates Same as Canyon / slip is likely dextral, given Cretaceous above Demon’s orientation, straight trace Cucamonga Canyon and proximity to the nearby Granulite; timing Gf, DuCf and MC/DCf (see likely similar to the above). However, no known nearby Gf, DuCf offset features. and MC/DCf, and thus, by extension, to the San Jacinto fault system (see above).

West Lytle WLCf Variable Geometric requirement if Termination of DuCf Dibblee, Creek amounts of top- DuCf and MC/DCf terminate and MC/DCf against 2002k, 2003c to-southeast against WLCf as mapped. these faults reverse slip, as Consistent with the suggests coeval necessary to consistently northwest dip of activity. This would accommodate the WLCf and SCf. likely be during San dextral slip on Jacinto fault zone DuCf and activity (see above). MC/DCf (see above).

Stoddard SCf 2.0 km of Consistent, 1.5-2.0 km left Displaces, and thus Nourse, 2002; Canyon sinistral slip; separation of Icehouse post-dates, ICf from Dibblee, 2003c additionally, Canyon fault (ICf) from MFLCf, thought to 223

southwest of correlated Middle Fork Lytle have been active junction with Creek fault (MFLCf) and of until ca. 5 Ma (see WLCf, also steeply dipping below); terminates variable Paleoproterozoic against the active amounts of top- metasedimentary WSJ/Sf (see to-southeast rock/Cretaceous plutonics above). reverse slip, as contact. Part of a set of necessary to northeast-southwest- accommodate striking, high-angle sinistral dextral slip on faults (see below). DuCf and MC/DCf (see above).

San Antonio SACf, 3.5 km of Consistent left separation of Displaces, and thus Nourse, 2002 Canyon and uf1 sinistral slip north branch San Gabriel post-dates, nbSGf unnamed (0.25 km of fault (nbSGf) from ICf, of from ICf, thought to fault which was Vincent thrust, and of folds have been restored on within the Pelona Schist and connected and unnamed fault Cretaceous mylonite along active until ca. 5 Ma splay uf1), Vincent thrust, suggesting (see below); merging with 3.0-3.5 km of sinistral slip. terminates against SCf to the Offset of ICf by the active WSJ/Sf southwest for a approximately 0.25 km by (see above). combined 5.5 unnamed fault splay uf1 km of sinistral suggests that 0.25 km of slip, and this sinistral slip was merging with the accommodated by this Sunset Ridge splay. fault (SRf) to the northeast for a combined 4.0 km of sinistral slip. Variable dip slip, increasing to the northeast, as required by curved trace of fault: along cross-section line, 1.0 km of top-to-southeast reverse slip.

Sunset Ridge SRf 0.5 km of Left separation of nbSGf Displaces, and thus Nourse, 2002 sinistral slip and of body of post-dates, middle Paleoproterozoic augen Miocene dikes and gneiss within Jurassic nbSGf, thought to porphyritic granodiorite, both have been active suggesting 0.3-0.8 km of until ca. 5 Ma (see sinistral slip. below); merges with SACf, which terminates against

224

the active WSJ/Sf (see above).

San Dimas SDCf, 0.5 km of Approximately 0.5 km left Displaces, and thus Dibblee, 2002i; Canyon and CwWf sinistral slip, 1.0 separation of high-angle post-dates, nbSGf, Nourse, 2002 Coldwater / km of top-to- nbSGf. Greater left thought to have Weber southeast separation of crystalline been active until ca. reverse slip units and of Vincent thrust 5 Ma (see below); along presumably the result of a terminates against northeastern top-to-southeast reverse-slip the active WSJ/Sf end, decreasing component (foliation in (see above). to the southwest crystalline units and Vincent due to curvature thrust are both shallowly of fault. southwest-dipping in this area).

Pine PMf 1.0 km of Fault shows subhorizontal Displaces, and thus Dibblee, Mountain sinistral slip, 0.5 slickenlines and post-dates, nbSGf, 2002b; Nourse, km of top-to- approximately 1.0 km left thought to have 2002 southeast separation of high-angle been active until ca. reverse slip. nbSGf. 5 Ma (see below); terminates against the Punchbowl fault (Pf), which likely initiated ca. 8-5 Ma, and was inactive by ca. 3 Ma.

Unnamed uf2 0 km of sinistral Separation along this fault is Is a strand of, and Nourse, 2002 strand of PMf slip minimal and inconsistent, thus coeval with, suggesting <0.5 km of PMf (see above) sinistral slip.

Crystal Lake CLf 0 km of sinistral Left separation varies along Same as above Dibblee, 2002g slip length from 0-2.0 km. Because of inconsistent separation, and lack of offset of nbSGf, no slip was restored in this reconstruction.

Small, none 0 km of slip Based on their orientations Likely coeval with Nourse, 2002 unlabeled and, in two cases, small left sinistral faults (SCf, faults separations of nbSGf, these SACF, SRf, SDCf, faults are likely synthetic to CWf, PMf). Some the larger sinistral faults displace, and thus discussed above. Left post-date, nbSGf. separations suggest ≤0.5 One displaces, and km of sinistral slip in all thus post-dates, the cases. Because of this, and Upper Oligocene their small mapped lengths, Telegraph Peak no slip was restored in this pluton. reconstruction.

225

Step 2 - From B (ca. 5 Ma) to C (ca. 14 Ma):

Geologic units removed: Upper Miocene-Pliocene Ridge Basin Group, Upper Miocene Castaic

Formation and Violin Breccia (these units had not yet formed at the time of C; Ensley and

Verosub, 1982). Miocene Mint Canyon and Punchbowl formations (the majority of these units had not yet formed at the time of D; only the Tick Canyon unit and other basal strata are likely older than ca. 14 Ma; Kew, 1924; Stirton, 1933; Woodburne, 1975; Terres and Luyendyk, 1985;

Liu, 1990; personal communications of Allen and Whistler in Liu, 1990).

Vertical-axis rotation restorations: none (net rotation since ca. 11 or 10 Ma implied by Mint

Canyon Formation already restored in B (ca. 5 Ma; see above), except inferred 8° clockwise rotation of block between dextral Lf and Cf, as a way to accommodate some of the shortening that would otherwise occur on the Lf along the Cf.

Fault restorations: (listed north to south)

Fault Abbrev. Slip restored Evidence Timing constraints References

Bald BMf 1 km of heave of Mapped as reverse fault Stratigraphy indicates Faggioli, 1952; Mountain top-to-north of variable dip angle, more recent activity Dibblee, reverse faulting. causing juxtaposition of than the Liebre fault 2002m, 2008; distinct geologic units (see below); fault Crowell, 2003 and considered presumably became regionally significant. No inactive with initiation of offset features to southern San Andreas indicate amount of slip; 1 fault; thus, activity km of heave restored as constrained to ca. 7-5 a conservative amount Ma. sufficient to separate mismatched geologic units.

Unlabeled none 0 km of slip Likely related to BMf, Likely related to, and Dibblee, faults NLf, Lf and Cf, with thus coeval with, BMF, 2002d, m between BMf oblique reverse dextral NLf, Lf and Cf, in which and Lf slip. Because of the lack case, activity was likely of clear offset features between ca. 7 Ma and and the minimal ca. 5 Ma (see above 226

mismatch of geologic and below). Fault units across these faults, activity progressively no slip was restored. Slip younger to the north was likely much less (see above and below). than on BMf, NLf, Lf and Cf.

North Liebre NLf 2 km of dextral Right separation of Magnetostratigraphy Faggioli, 1952; slip (merging with granitic-bearing constrains activity to Ensley and Lf to southeast conglomerate suggests 6.0-5.0 Ma. Verosub, 1982; for combined 6.5 ≥1.8 km of dextral slip; Dibblee, km of dextral Dip-separation of 2002m; slip) features suggests 0.2- Crowell, 2003; 0.6 km of top-to- Yan et al., southwest reverse slip. 2005

Liebre Lf 4.5 km of dextral Mapped by multiple Magnetostratigraphy Faggioli, 1952; slip (merging with studies as an oblique constrains activity to Ensley and NLf to southeast reverse dextral fault, 7.3-6.1 Ma. Verosub, 1982; for combined 6.5 steeply dipping along Dibblee, km of dextral much of its length, but 2002m; slip); variable shallowly dipping in Crowell, 2003; top-to-south places. Dextral slip Yan et al., reverse slip, thought to be 2005 increasing to the approximately 6.5-8 km, southeast, as which roughly aligns geometrically Cretaceous plutonics / necessary, Paleoproterozoic gneiss increasing to 4.5 contacts and gneiss- km of reverse bearing conglomerate of slip along cross- Upper Miocene-Pliocene section line. Ridge Basin Group with Paleoproterozoic gneiss; reverse slip thought to be substantial.

Clearwater Cf, SFf, 2 km of dextral Considered oblique Magnetostratigraphy Stanley, 1966; and eastern uf3 slip (including 1 reverse dextral slip by constrains activity to Ensley and part of San km along fault numerous studies. 8.1-7.8 Ma, with an Verosub, 1982; Francisquito branch uf3); Coeval with, and thought earlier phase of motion Crowell, 2003; variable top-to- to be related to, the likely. Yan et al., the-south reverse dextral San Gabriel fault. 2005 slip as Steeply north-dipping to geometrically near vertical, with both necessary, right separation and increasing to 4 down-to-south km of reverse separation. Separation slip along cross- of a facies change within section line. the Paleogene San Francisquito Formation and of a near-vertical Cretaceous plutonics / Paleoproterozoic gneiss

227

contact, and offset of correlated surfaces within a syncline in the Paleogene San Francisquito Formation, indicate approximately 1.6 km of dextral slip and ≥0.6 km of down-to-the- south dip slip; other workers give comparable estimates of ≥1.8 km of dextral slip and likely 0.5-1.0 km of reverse slip. No reverse slip is restored in map view except as necessary due to geometry of fault because, due to near- vertical orientation, reverse slip would have been manifested dominantly as throw, not heave.

San Gabriel, none Kilometers of Strata and their source Activity began ca. 11 e.g., Crowell, south branch dextral slip, likely rocks exhibit kilometers Ma, and ceased ca. 5 1952, 2003; San Gabriel approximately of right separation. Ma, as constrained by Ensley and 20-45 km, with Correlation of offset ages of pre-, syn- and Verosub, 1982; additional slip on equivalents suggests post-growth strata. Powell, 1993 the related approximately 20-45 km Canton fault of dextral slip, or (precise approximately 60-70 km restoration of when combined with slip San Gabriel fault on the related Canton slip is not fault. attempted here)

Maple MCf 0 km of slip Based on proximity to Presumably coeval with Dibblee, 2002e Canyon and matching orientation nbSGf (see below). with nbSGf, presumably related. Because of the lack of clear offset features and the minimal mismatch of geologic units across these faults, no slip was restored.

North branch nbSGf, Kilometers of Major mismatch of Presumably coeval with e.g., Crowell, San Gabriel, ICf, dextral slip geologic units across the the San Gabriel fault 1952; Powell, Icehouse (some of the slip nbSGf-ICf-MFLCf, which (see above), with some 1993; Nourse, Canyon and MFLCf of the San was a strand of the San subsequent reactivation 2002 Middle Fork Gabriel fault to Gabriel fault (see possible. Lytle Creek the northwest, above), a dextral

228

see above); transform boundary. possible substantial dip- slip (precise restoration of nbSGf-ICf-MFLCf slip is not attempted here)

South sbSGf/ Kilometers of Major mismatch of Presumably coeval with e.g., Crowell, Branch San dextral slip (the geologic units across the the San Gabriel fault 1952; Nourse, Gabriel fault VCkf dextral slip of the sbSGf, which was a (see above). 2002; Dibblee, / Vasquez San Gabriel fault strand of the San Gabriel 2002e Creek fault to the northwest, fault (see above), a less slip dextral transform accommodated boundary. by the nbSGf; see above; precise restoration of sbSGf slip is not attempted here).

Step 3 - From C (ca. 14 Ma) to D (ca. 18 Ma):

Geologic units removed: None (Middle Miocene Glendora volcanics already removed from figure as part of nbSGf-ICf-MFLCf restoration).

Vertical-axis rotation restorations: 53° of clockwise rotation is restored, based on 37° ± 12° of net clockwise rotation implied by paleomagnetism of volcanic flows within the lower part of the

Vasquez Formation (Terres and Luyendyk, 1985), dated ca. 26-23 Ma (Frizzell and Weigand,

1993), coupled with 16° ± 30° of earlier counterclockwise rotation (see step #1 above). This rotation was probably accommodated by slip on a system of sinistral faults (Luyendyk et al.,

1980; Terres and Luyendyk, 1985; details of these faults below), which were active until ca. 14

Ma (see below). This rotation is assumed to have applied to the entire San Gabriel block (as suggested by model of Terres and Luyendyk, 1985).

229

Fault restorations: (listed north to south; all orientations listed are post-53° of clockwise vertical-axis rotation, i.e., the orientations shown in C, not D).

Fault Abbrev. Slip restored Evidence Timing References constraints

Mint Canyon MCyf 1.0 km of sinistral Mapped as sinistral; Offsets only the Sharp, 1935; slip (inferred to northeast-southwest basal layers of Muehlberger, merge with TCf to orientation, steep dip and the Mint Canyon 1954, 1958; the northeast for a timing constraints Formation, Dibblee, 1996b, combined 1.5 km of consistent with system of indicating activity c sinistral slip). 0.2 km sinistral faults in area. ceased early in of top-to-southeast Approximately 1.5 km of Mint Canyon reverse slip inferred left separation of basal Formation (0.2 km per 1.0 km of Mint Canyon Formation deposition, ca. sinistral slip; suggests 1.5 km of 14 Ma. Offsets, assumed for all sinistral slip, but and thus post- sinistral faults in this structures within the dates, the Upper system). Pelona Schist to the Oligocene-Lower northeast do not allow Miocene this much slip, so 1.0 km Vasquez of sinistral slip is Formation, restored. Remaining 0.5 deposited until km of separation of the ca. 20 Ma. basal Mint Canyon Activity thus Formation may either constrained to represent increasing slip between ca. 20 to the southwest, or a and ca. 14 Ma. dip-slip component (the basal Mint Canyon Formation is shallowly dipping).

Tick Canyon TCf 0.5 km of sinistral Mapped as sinistral; Same as above Same as above slip (distributed northeast-southwest across multiple fault orientation, straight trace strands; inferred to and timing constraints merge with MCyf to consistent with system of the northeast for a sinistral faults in area. combined 1.5 km of Approximately 1.0 km of sinistral slip). 0.1 km left separation of Upper of top-to-southeast Oligocene-Lower reverse slip inferred Miocene Vasquez (0.2 km per 1.0 km of Formation / Upper sinistral slip; Oligocene Vasquez assumed for all Formation volcanics sinistral faults in this contacts suggests system). approximately 1.0 km of sinistral slip, but structures within the Pelona Schist to the 230

northeast do not allow this much slip, so 0.5 km of sinistral slip is restored. Remaining 0.5 km of separation may represent increasing slip to the southwest.

Green Ranch GRf, Ef 1.5 km of sinistral Mapped as sinistral; Same as above Same as above and Elkhorn slip (split equally northeast-southwest between two inferred orientation, straight trace strands to the and timing constraints northeast). 0.3 km of consistent with system of top-to-southeast sinistral faults in area. reverse slip inferred 1.5-2 km of left (0.2 km per 1.0 km of separation of Middle sinistral slip; Miocene Mint Canyon assumed for all Formation / Upper sinistral faults in this Oligocene-Lower system). Miocene Vasquez Formation contact and 1.0-1.5 km of left separation of Upper Oligocene-Lower Miocene Vasquez Formation / Mesoproterozoic anorthosite and gneiss contact suggests approximately 1.5 km of sinistral slip.

Agua Dulce ADf 0.5 km of sinistral Mapped as sinistral; Same as above personal slip (less along northeast-southwest communication inferred continuation orientation, straight trace of R.B. to the northeast due and timing constraints Campbell in to accommodation by consistent with system of Muehlberger, reverse slip along sinistral faults in area. 1954; KVf). 0.1 km of top- Approximately 0.5 km of Muehlberger, to-southeast reverse left separation of Upper 1954, 1958; slip inferred (0.2 km Oligocene-Lower Dibblee, 1996b, per 1.0 km of sinistral Miocene Vasquez c slip; assumed for all Formation / Upper sinistral faults in this Oligocene Vasquez system). Formation volcanics contacts suggests approximately 0.5 km of sinistral slip. An earlier phase of motion has also been postulated.

Kashmere KVf variable top-to-east Juxtaposes Triassic Lowe Offsets, and thus Dibblee, 1996d Valley reverse slip. pluton and Upper post-dates,

231

Oligocene Vasquez Upper Oligocene Formation volcanics; Vasquez reverse slip a geometric Formation necessity if KVf is a right volcanics. double-bend in a system Interpretation as of sinistral slip faults, as part of system of here interpreted. sinistral slip faults (see above and below) suggests activity within the interval ca. 20-14 Ma.

Pole Canyon PCf 0.5 km of sinistral Near-vertical orientation Offsets only the Muehlberger, slip. 0.1 km of top-to- and subhorizontal basal layers of 1954, 1958; southeast reverse slickenlines indicate the Mint Canyon Dibblee, 1996b, slip inferred (0.2 km strike-slip; left separation Formation, c; this study per 1.0 km of sinistral indicates sinistral slip, indicating activity slip; assumed for all and orientation and ceased early in sinistral faults in this timing consistent with Mint Canyon system). system of sinistral faults Formation in area. Left separation of deposition, ca. the Soledad fault, which 14 Ma. Offsets, juxtaposes Upper and thus post- Oligocene-Lower dates, the Upper Miocene Vasquez Oligocene-Lower Formation and Miocene Mesoproterozoic Vasquez anorthosite and gneiss, Formation, suggests ≥2 km of deposited until sinistral slip, however, if ca. 20 Ma. the Pole Canyon fault Activity thus reactivated a segment of constrained to the Soledad fault, then between ca. 20 sinistral slip would be and ca. 14 Ma. less. Contacts to the northeast exhibit ≤0.5 km of left separation, and the fault likely dies out to the northeast. Near-vertical orientation of the fault restores to steeply north- dipping after undoing tilting documented by the Mint Canyon Formation; this is compatible with a component of top-to- southeast reverse slip on this system of sinistral faults. inferred fault1 if1 0.5 km of sinistral Apparent left separation Offsets, and thus Dibblee, 1997d slip. 0.1 km of top-to- of Mesoproterozoic post-dates,

232

southeast reverse syenite / Triassic diorite Upper Oligocene slip inferred (0.2 km and gabbro contact Vasquez per 1.0 km of sinistral suggests ≤0.5 km of Formation slip; assumed for all sinistral slip; juxtaposition volcanics. sinistral faults in this of Cretaceous plutonics Extrapolation system). and Upper Oligocene from nearby Vasquez Formation sinistral faults volcanics along (see above and northwest-dipping fault below) suggests suggests top-to- activity within the southeast reverse slip. interval ca. 20-14 Ma.

Magic MMf 1.0 km of sinistral Mapped as sinistral; Offsets Upper Muehlberger, Mountain slip. 0.2 km of top- northeast-southwest Oligocene-Lower 1958; Dibblee, to-southeast reverse orientation, steep dip and Miocene 1996b, c, d slip inferred (0.2 km timing constraints Vasquez per 1.0 km of sinistral consistent with system of Formation, but is slip; assumed for all sinistral faults in area. not mapped as sinistral faults in this Approximately 1.0 km of continuing into system), increasing left separation of the Mint Canyon to the northeast due Mesoproterozoic syenite / Formation, to change in fault Mesoproterozoic suggesting orientation. anorthosite and gabbro activity within the contacts suggests interval ca. 20-14 approximately 1.0 km of Ma. sinistral slip. Along inferred continuation to the northeast, left separation of contacts are ≤0.5 km. 0.5 km of sinistral slip was restored along this continuation, and 0.5 km of sinistral slip was restored along an inferred splay in , for a total of 1.0 km inferred fault if2 1.0 km of sinistral Allows accommodation of Inferred as Dibblee, 1996d 2 slip. 0.2 km of top-to- slip experienced by ADf, offsetting, and southeast reverse PCf, MMf transferred via thus post-dating, slip inferred (0.2 km KVf, Upper Oligocene per 1.0 km of sinistral Vasquez slip; assumed for all Formation sinistral faults in this volcanics, ca. 25 system). Ma. Extrapolation from nearby sinistral faults (see above) suggests activity within the interval

233

ca. 20-14 Ma. unnamed uf4 1.0 km of sinistral Allows accommodation of Offsets, and thus Dibblee, 1996d fault 4 slip, merging with TLf slip experienced by ADf, post-dates, to northeast for a PCf, MMf transferred via Cretaceous combined 2.5 km of KVf, plutonics. sinistral slip. 0.2 km Presumably of top-to-southeast same age as TLf reverse slip inferred (see below), (0.2 km per 1.0 km of which it is sinistral slip; interpreted as a assumed for all branch of, and sinistral faults in this thus likely active system). ca. 20-14 Ma.

Transmission TLf 1.5 km of sinistral Mapped as sinistral; Offsets, and thus Muehlberger, Line slip, merging with uf4 northeast-southwest post-dates, Lf, 1958; Dibblee, to northeast for a orientation and straight which itself 1991a, 1996b, c combined 2.5 km of trace consistent with offsets Triassic sinistral slip. 0.3 km system of sinistral faults Lowe pluton. Is of top-to-southeast in area. At southwestern overlain by, and reverse slip inferred end, branches into a thus pre-dates, (0.2 km per 1.0 km of horsetail structure quaternary sinistral slip; characteristic of strike- alluvium. assumed for all slip faults, with left Extrapolation sinistral faults in this separation of features from nearby system). within crystalline sinistral faults basement along all (see above) splays. Approximately 1.5 suggests activity km left separation of within the interval gneiss/anorthosite ca. 20-14 Ma. complex contact at western end of fault suggests approximately 1.5 km of sinistral slip. Restoration of 1.5 km of sinistral slip also aligns offset segments of Ltf, and, along inferred continuation to northeast, Paleoproterozoic gneiss / Triassic diorite and gabbro contact.

Mill Creek MCkf 1.0 km of sinistral Northeast-southwest Offsets, and thus Dibblee, 2002e, slip. 0.2 km of top-to- orientation and near- post-dates, f southeast reverse vertical dip consistent Cretaceous slip inferred (0.2 km with system of sinistral plutonics. Is per 1.0 km of sinistral faults in area. overlain by, and slip; assumed for all Approximately 1.0 km left thus pre-dates, sinistral faults in this separation of several quaternary system). basement contacts alluvium. suggests approximately Extrapolation

234

1.0 km of sinistral slip. from nearby sinistral faults (see above) suggests activity within the interval ca. 20-14 Ma.

Small, none 0 km of slip Northeast-southwest Likely coeval with Muehlberger, unlabeled orientations and straight nearby sinistral 1954, 1958; faults traces consistent with faults (MCyf, TCf, Bishop, 1990; system of sinistral faults GRf/Ef, ADf, PCf, Dibblee, 1996c in area; some exhibit MMf, TLf, MCkf). small (<0.5 km) amounts of left separation. Because of minimal left separations and lack of mapped continuations to the northeast or southwest, no slip was restored on these faults.

Step 4 - From D (ca. 18 Ma) to E (ca. 26 Ma):

Geologic units removed: Upper Oligocene-Lower Miocene Vasquez Formation and Upper

Oligocene Vasquez Formation volcanics. The volcanics were erupted ca. 25 Ma (this study), and are near the base of, and thus largely predate, the Vasquez Formation (e.g., Hendrix and

Ingersoll, 1987).

Vertical-axis rotation restorations: None (no paleomagnetic data have been reported from units older than the Upper Oligocene Vasquez Formation volcanics).

Fault restorations: (listed west to east; all orientations are those corrected for subsequent vertical-axis rotation implied by paleomagnetic data; see above)

Fault Abbrev. Slip restored Evidence Timing References constraints

San SFf 6.0 km of down- Interpreted as a major Offsets, and thus Konigsberg, Francisquito to-west normal north-dipping normal postdates, the 1967; Hendrix

235 fault slip. fault active during Paleogene San and Ingersoll, deposition of Vasquez Francisquito 1987; Smith, Formation by Formation. 1977; Powell, numerous studies. Offsets the Upper 1981, 1993; The San Francisquito Oligocene-Lower Dibblee, fault has alternatively Miocene Vasquez 1997b; been interpreted Formation; Bunker and primarily as a dextral sedimentologic Bishop, 2001; fault related to the evidence Yan et al., San Andreas fault suggests that 2005 system; much of the fault motion was evidence for this, coeval with however, comes from deposition; fault north of the junction offsets only the with the younger, basal part of the dextral Clearwater Middle Miocene fault, and thus is likely Mint Canyon indicative of motion on Formation, and this younger fault, and thus is mostly not the San older than this Francisquito fault unit. Thus, activity proper. North-dipping constrained to ca. orientation compatible 60-14 Ma, and with normal motion, probably ca. 26- and, following 14 Ma. An earlier restoration of tilting phase of motion implied by the Mint is inferred in this Canyon Formation, reconstruction to similar to orientation create the basin of normal faults to the in which the San east (VCf, Sf). Francisquito Formation was deposited.

Pelona fault Pef 0.5 km of down- Kinematic Offsets Upper Hendrix and to-east normal measurements along Oligocene-Lower Ingersoll, slip (antithetic to the Pelona fault (this Miocene Vasquez 1987; Bishop, main normal study) indicate down- Formation; 1990; faults) to-east normal slip. overlain by, and Dibblee, thus predates, the 1996b, 1997c basal Mint Canyon Formation. This indicates that fault motion was coeval with deposition of the Vasquez 236

Formation.

Vasquez VCf 1.5 km of down- Mapped as down-to- Offsets Upper Muehlberger, Canyon fault to-west normal west normal fault Oligocene-Lower 1958; Hendrix slip (potentially active during Miocene Vasquez and Ingersoll, originally deposition of Upper Formation; 1987; Bishop, greater, but Oligocene-Lower overlain by, and 1990 reduced by Miocene Vasquez thus predates, the reverse motion Formation. basal Mint during isostatic Canyon uplift of Sierra Formation. This Pelona indicates that anticline). fault motion was coeval with deposition of the Vasquez Formation.

Soledad Sf 4.5 km of down- Mapped as down-to- Offsets Upper Muehlberger, fault to-west normal west normal fault Oligocene-Lower 1958; Hendrix slip. active during Miocene Vasquez and Ingersoll, deposition of Upper Formation; 1987; Oligocene-Lower overlain by, and Dibblee, Miocene Vasquez thus predates, the 1996c Formation, with up to basal Mint 4.5 km of slip Canyon considered Formation. This reasonable. indicates that fault motion was coeval with deposition of the Vasquez Formation.

Lonetree Ltf 5.0 km of down- Suggested to be part Offsets, and is Hendrix, fault to-west normal of the nearby normal- thus younger 1993; slip. fault system (SFf, Pef, than, the Triassic Dibblee, VCf, Sf). Steeply east- Lowe pluton. 1996c, d dipping orientation Appears to be restores to steeply offset by, and west-dipping after thus predates, the correcting for tilting TLf, which is part implied by the Mint of a system of Canyon Formation; sinistral faults this is compatible with active ca. 18-14 down-to-west normal Ma (see above). slip, particularly if this Interpreted as fault represents a part of the nearby

237

listric breakaway fault, system of normal in contrast to more faults (SFf, Pef, planar faults to the VCf, Sf), in which west. Strike matches case, slip would that of nearby normal have been coeval faults, and is with deposition of anomalous compared the Vasquez to nearby, younger Formation. sinistral faults (see above). Furthermore, mismatch of crystalline units on either side of fault cannot be reconciled by restoring a few km of sinistral slip, and the Ltf appears to be offset by the sinistral TLf.

REFERENCES CITED:

Bennett, S.E.K., Oskin, M.E., Iriondo, A., and Kunk, M.J., 2016, Slip history of the La Cruz

fault: Development of a late Miocene transform in response to increased rift obliquity in

the northern Gulf of California: Tectonophysics, v. 693, p. 409-435,

https://doi.org/10.1016/j.tecto.2016.06.013

Bishop, K.M., 1990, Tectonic model for the evolution of the Texas Canyon basin and Pelona

fault, Soledad basin, southern California, in Buckley, C.I., and Larson, R.A., eds.,

Geology and engineering geology of the western Soledad basin, Los Angeles County,

California: Association of Engineering Geologists, Southern California Section, Annual

Field Trip Guidebook, p. 35-56.

Bunker, J.V., and Bishop, K.M., 2001, The San Francisquito fault, southern California: A normal

fault related to uplift of the Sierra Pelona anticline?: Geological Society of America

Abstracts with Programs, v. 33, n. 3, p. 73. 238

Coffey, K.T., Ingersoll, R.V., and Schmitt, A.K., 2019: Stratigraphy, provenance and tectonic

significance of the Punchbowl block, San Gabriel Mountains, California, U.S.A.:

Geosphere, v. 15, p. 479-501, https://doi.org/10.1130/GES02025.1

Cramer, C.H., and Harrington, J.M., 1987, Seismicity and tectonics of the Cucamonga fault and

the eastern San Gabriel Mountains, in Holocene reverse faulting of the Transverse

Ranges, California: U.S. Geological Survey Professional Paper 1339, p. 7–26.

Crook, Richard, Jr., Allen, C.R., Kamb, B., Payne, C.M., and Proctor, R.J., 1987, Quaternary

geology and seismic hazard of the Sierra Madre and associated faults, western San

Gabriel Mountains, in Morton, D.M., and Yerkes, R.F., eds., Recent reverse faulting in

the Transverse Ranges, California: U.S. Geological Survey Professional Paper 1339, U.S.

Geological Survey , Washington, D.C., p. 27-63.

Crowell, J.C., 1952, Probable large lateral displacement on San Gabriel fault, southern

California: Amiercan Association of Petroleum Geologists Bulletin, v. 36, p. 2026-2035.

Crowell, J.C., 1975, The San Andreas fault in southern California, in Crowell, J.C., ed., San

Andreas fault in southern California: A guide to the San Andreas fault from Mexico to

Carrizo Plain: California Division of Mines and Geology Special Report, v. 118, p. 7-27.

Crowell, J.C., 2003, Tectonics of Ridge Basin region, southern California, Introduction to

geology of Ridge Basin, southern California, in Crowell, J.C., ed., Evolution of Ridge

Basin, southern California: An interplay of sedimentation and tectonics: Geological

Society of America Special Paper 367, p. 157-203, https://doi.org/10.1130/0-8137-2367-

1.1

Darin, M.H., Bennett, S.E.K., Dorsey, R.J., Oskin, M.E., and Iriondo, A., 2016, Late Miocene

extension in coastal Sonora, México: Implications for the evolution of dextral shear in the

239

proto-Gulf of California oblique rift: Tectonophysics, v. 693, p. 378-408,

https://doi.org/10.1016/j.tecto.2016.04.038

Dibblee, T.W., Jr., 1968, Displacements on the San Andreas fault system in the San Gabriel, San

Bernardino, and San Jacinto Mountains, Southern California in Dickinson, W.R., and

Grantz, A., eds., Proceedings of conference on geological problems of San Andreas fault

system: Stanford University Publications in Geological Sciences, v. 11, p. 260-278.

Dibblee, T.W., Jr., 1989, Geologic Map of the Pasadena Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-23, scale 1:24,000.

Dibblee, T.W., Jr., 1991a, Geologic Map of the Sunland and Burbank (north ½) Quadrangles,

Los Angeles County, California: Dibblee Geological Foundation Map DF-32, scale

1:24,000.

Dibblee, T.W., Jr., 1991b, Geologic Map of the San Fernando and (north ½)

Quadrangles, Los Angeles County, California: Dibblee Geological Foundation Map DF-

33, scale 1:24,000.

Dibblee, T.W., Jr., 1993, Geologic Map of the Val Verde Quadrangle, Los Angeles and Ventura

Counties, California: Dibblee Geological Foundation Map DF-50, scale 1:24,000.

Dibblee, T.W., Jr., 1996a, Geologic Map of the Newhall Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-56, scale 1:24,000.

Dibblee, T.W., Jr., 1996b, Geologic Map of the Mint Canyon Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-57, scale 1:24,000.

Dibblee, T.W., Jr., 1996c, Geologic Map of the Agua Dulce Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-58, scale 1:24,000.

240

Dibblee, T.W., Jr., 1996d, Geologic Map of the Acton Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-59, scale 1:24,000.

Dibblee, T.W., Jr., 1996e, Geologic Map of the Cobblestone Mountain Quadrangle, Ventura and

Los Angeles Counties, California: Dibblee Geological Foundation Map DF-62, scale

1:24,000.

Dibblee, T.W., Jr., 1997a, Geologic Map of the Whitaker Peak Quadrangle, Los Angeles and

Ventura Counties, California: Dibblee Geological Foundation Map DF-63, scale

1:24,000.

Dibblee, T.W., Jr., 1997b, Geologic Map of the Warm Springs Mountain Quadrangle, Los

Angeles County, California: Dibblee Geological Foundation Map DF-64, scale 1:24,000.

Dibblee, T.W., Jr., 1997c, Geologic Map of the Green Valley Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-65, scale 1:24,000.

Dibblee, T.W., Jr., 1997d, Geologic Map of the Sleepy Valley and Ritter Ridge Quadrangles,

Los Angeles County, California: Dibblee Geological Foundation Map DF-66, scale

1:24,000.

Dibblee, T.W., Jr., 1998, Geologic Map of the Mt. Wilson and Azusa Quadrangles, Los Angeles

County, California: Dibblee Geological Foundation Map DF-67, scale 1:24,000.

Dibblee, T.W., Jr., 2001, Geologic Map of the Pacifico Mountain and Palmdale (south half)

Quadrangles, Los Angeles County, California: Dibblee Geological Foundation Map DF-

76, scale 1:24,000.

Dibblee, T.W., Jr., 2002a, Geologic Map of the Juniper Hills Quadrangle (and southern

Littlerock Quadrangle), Los Angeles County, California: Dibblee Geological Foundation

Map DF-79, scale 1:24,000.

241

Dibblee, T.W., Jr., 2002b, Geologic Map of the Valyermo Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-80, scale 1:24,000.

Dibblee, T.W., Jr., 2002c, Geologic Map of the Lake Hughes and Del Sur Quadrangles, Los

Angeles County, California: Dibblee Geological Foundation Map DF-82, scale 1:24,000.

Dibblee, T.W., Jr., 2002d, Geologic Map of the Burnt Peak Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-83, scale 1:24,000.

Dibblee, T.W., Jr., 2002e, Geologic Map of the Condor Peak Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-84, scale 1:24,000.

Dibblee, T.W., Jr., 2002f, Geologic Map of the Chilao Flat Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-85, scale 1:24,000.

Dibblee, T.W., Jr., 2002g, Geologic Map of the Waterman Mountain Quadrangle, Los Angeles

County, California: Dibblee Geological Foundation Map DF-86, scale 1:24,000.

Dibblee, T.W., Jr., 2002h, Geologic Map of the Crystal Lake Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-87, scale 1:24,000.

Dibblee, T.W., Jr., 2002i, Geologic Map of the Mount San Antonio Quadrangle, Los Angeles

and San Bernardino Counties, California: Dibblee Geological Foundation Map DF-88,

scale 1:24,000.

Dibblee, T.W., Jr., 2002j, Geologic Map of the Glendora Quadrangle, Los Angeles County,

California: Dibblee Geological Foundation Map DF-89, scale 1:24,000.

Dibblee, T.W., Jr., 2002k, Geologic Map of the Mount Baldy Quadrangle, Los Angeles and San

Bernardino Counties, California: Dibblee Geological Foundation Map DF-90, scale

1:24,000.

242

Dibblee, T.W., Jr., 2002l, Geologic Map of the Black Mountain Quadrangle, Los Angeles and

Ventura Counties, California: Dibblee Geological Foundation Map DF-92, scale

1:24,000.

Dibblee, T.W., Jr., 2002m, Geologic Map of the Liebre Mountain Quadrangle, Los Angeles

County, California: Dibblee Geological Foundation Map DF-93, scale 1:24,000.

Dibblee, T.W., Jr., 2003a, Geologic Map of the Cajon Quadrangle, San Bernardino County,

California: Dibblee Geological Foundation Map DF-104, scale 1:24,000.

Dibblee, T.W., Jr., 2003b, Geologic Map of the Devore Quadrangle, San Bernardino County,

California: Dibblee Geological Foundation Map DF-105, scale 1:24,000.

Dibblee, T.W., Jr., 2003c, Geologic Map of the Cucamonga Peak Quadrangle, San Bernardino

County, California: Dibblee Geological Foundation Map DF-106, scale 1:24,000.

Dibblee, T.W., Jr., 2003d, Geologic Map of the Telegraph Peak and Phelan Quadrangles, San

Bernardino County, California: Dibblee Geological Foundation Map DF-107, scale

1:24,000.

Dibblee, T.W., Jr., 2004, Geologic Map of the San Bernardino North/north ½ of San Bernardino

South Quadrangles, San Bernardino County, California: Dibblee Geological Foundation

Map DF-127, scale 1:24,000.

Dibblee, T.W., Jr., 2006a, Geologic Map of the Frazier Mountain and Lebec Quadrangles, Los

Angeles, Ventura, and Kern Counties, California: Dibblee Geological Foundation Map

DF-198, scale 1:24,000.

Dibblee, T.W., Jr., 2006b, Geologic Map of the Alamo Mountain Quadrangle, Ventura and Los

Angeles Counties, California: Dibblee Geological Foundation Map DF-254, scale

1:24,000.

243

Dibblee, T.W., Jr., 2008, Geologic Map of the Neenach and Willow Springs 15’ Quadrangles,

Kern and Los Angeles Counties, California: Dibblee Geological Foundation Map DF-

383, scale 1:62,500.

Dorsey, R.J., Housen, B.A., Janecke, S.U., Fanning, C.M., and Spears, A.L.F., 2011,

Stratigraphic record of basin development within the San Andreas fault system: Late

Cenozoic Fish Creek-Vallecito basin, southern California: Geological Society of America

Bulletin, v. 123, p. 771-793, https://doi.org/10.1130/B30168.1

Ehlig, P.L., 1968, Causes of distribution of Pelona, Rand, and Orocopia Schists along the San

Andreas and Garlock faults, in Dickinson, W.R., and Grantz, A., eds., Proceedings of

conference on geological problems of San Andreas fault system: Stanford University

Publications in Geological Sciences, v. 11, p. 294-305.

Ehlig, P.L., 1981, Origin and tectonic history of the basement terrane of the San Gabriel

Mountains, central Transverse Ranges, in Ernst, W.G., ed., The Geotectonic

Development of California (Rubey Volume I): Englewood Cliffs, New Jersey, Prentice-

Hall, p. 253-283.

Ensley, R.A., and Verosub, K.L., 1982, A magnetostratigraphic study of the sediments of the

Ridge Basin, southern California and its tectonic and sedimentologic implications: Earth

and Planetary Science Letters, v. 59, p. 192-207, https://doi.org/10.1016/0012-

821X(82)90125-X

Faggioli, R.E., 1952, The geology of the Liebre fault [M.A. thesis]: Los Angeles, University of

California, 56 p.

Frizzell, V.A., Jr., and Weigand, P.W., 1993, Whole-rock K-Ar ages and geochemical data from

middle Cenozoic volcanic rocks, southern California: A test of correlations across the

244

San Andreas fault, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San

Andreas fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 273-288,

https://doi.org/10.1130/MEM178-p273

Hendrix, E.D., 1993, Soledad Basin, Central Transverse Ranges, Calif., in Sherrod, D.R., and

Nielson, J.E., eds., Tertiary stratigraphy of highly extended terranes, California, Arizona,

and Nevada: U.S. Geological Survey Bulletin 2053, p. 243-350.

Hendrix, E.D., and Ingersoll, R.V., 1987, Tectonics and alluvial sedimentation of the upper

Oligocene/lower Miocene Vasquez Formation, Soledad Basin, southern California:

Geological Society of America Bulletin, v. 98, p. 647-663, https://doi.org/10.1130/0016-

7606(1987)98<647:TAASOT>2.0.CO;2

Hill, M.L., and Dibblee, T.W., Jr., 1953, San Andreas, Garlock, and Big Pine faults, California: a

study of the character, history, and tectonic significance of their displacements:

Geological Society of America Bulletin v. 64, p. 443-458, https://doi.org/10.1130/0016-

7606(1953)64[443:SAGABP]2.0.CO;2

Ingersoll, R.V., and Coffey, K.T., 2017, Transrotation induced by crustal blocks moving through

restraining double bends, with southern California examples: The Journal of Geology, v.

125, p. 551-559, https://doi.org/10.1086/692654

Kew, W.S.W., 1924, Geology and oil resources of a part of Los Angeles and Ventura Counties,

California: U.S. Geological Survey Bulletin 753, 202 p.

Konigsberg, W., 1967, Geology along the San Francisquito fault [M.S. thesis]: Los Angeles,

University of Calfornia, 84 p.

245

Liu, W., 1990, Paleomagnetism of Miocene sedimentary rocks in the Transverse Ranges: the

implications for tectonic history [Ph.D. thesis]: Pasadena, California Institute of

Technology, 233 p.

Luyendyk, B.P., Kamerling, M.J., and Terres, R., 1980, Geometric model for Neogene crustal

rotations in southern California: Geological Society of America Bulletin, v. 91, p. 211-

217, https://doi.org/10.1130/0016-7606(1980)91<211:GMFNCR>2.0.CO;2

Morton, D.M., and Matti, J.C., 1993, Extension and contraction within an evolving divergent

strike-slip fault complex: The San Andreas and San Jacinto fault zones at their

convergence in southern California, in Powell, R.E., Weldon, R.J., II, and Matti, J.C.,

eds., The San Andreas fault system: displacement, palinspastic reconstruction, and

geologic evolution: Geological Society of America Memoir 178, p. 217-230,

https://doi.org/10.1130/MEM178-p217

Muehlberger, W.R., 1954, Deposition and deformation in the northern Soledad basin, Los

Angeles County, California [Ph.D. thesis]: Pasadena, California Institute of Technology,

87 p.

Muehlberger, W.R., 1958, Geology of northern Soledad basin, Los Angeles County, California:

American Association of Petroleum Geologists Bulletin, v. 42, p. 1812-1844.

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. 491-495,

https://doi.org/10.1130/0091-7613(1994)022<0491:MCROTW>2.3.CO;2

Norum, B.V., 1997, Stratigraphy and structure of the Pelona Schist in the north fork Lytle Creek

drainage of the San Gabriel Mountains: Implications for palinspastic reconstruction

246

[Senior thesis]: Pomona, California, California State Polytechnic University, 40 p., plus

one plate.

Nourse, J.A., 2002, Middle Miocene reconstruction of the central and eastern San Gabriel

Mountains, southern California, with implications for evolution of the San Gabriel fault

and Los Angeles basin in Barth, A., ed., Contributions to Crustal Evolution of the

Southwestern United States: Geological Society of America Special Paper 365, p. 161-

185, https://doi.org/10.1130/0-8137-2365-5.161

Oskin, M., Stock, J., and Martin-Barajas, A., 2001, Rapid localization of Pacific-North America

plate motion in the Gulf of California: Geology, v. 29, p. 459-462,

https://doi.org/10.1130/0091-7613(2001)029<0459:RLOPNA>2.0.CO;2

Pechmann, J.C., 1987, Tectonic implications of small earthquakes in the central Transverse

Ranges, in Morton, D.M., and Yerkes, R.F., eds., Recent reverse faulting in the

Transverse Ranges, California: U.S. Geological Survey Professional Paper 1339, U.S.

Geological Survey , Washington, D.C., p. 97-111.

Powell, R.E., 1981, Geology of the crystalline basement complex, eastern Transverse Ranges,

southern California: constraints on regional tectonic interpretation [Ph.D. thesis]:

Pasadena, California Institute of Technology, 441 p.

Powell, R.E., 1993, Balanced palinspastic reconstruction of pre-late Cenozoic paleogeology,

southern California: geologic and kinematic constraints on evolution of the San Andreas

fault system, in Powell, R.E., Weldon, R.J., II, and Matti, J.C., eds., The San Andreas

fault system: displacement, palinspastic reconstruction, and geologic evolution:

Geological Society of America Memoir 178, p. 1-106, https://doi.org/10.1130/MEM178-

p1

247

Sharp, R.P., 1935, Geology of the Ravenna Quadrangle, Los Angeles County, California [M.S.

thesis]: Pasadena, California, California Institute of Technology, 81 p.,

https://www.doi.org/10.7907/4364-VE20

Smith, D.P., 1977, San Juan-St Francis fault – Hypothesized major middle tertiary right-lateral

fault in central and southern California: California Division of Mines and Geology

Special Report 129, p. 41-50.

Stanley, K.O., 1966, The structural history of the Clearwater fault, northwest Los Angeles

County, California [M.A. thesis]: Los Angeles, University of California, 73 p.

Stirton, R.A., 1933, Critical review of the Mint Canyon mammalian fauna and its correlative

significance: American Journal of Science, v. 26, p. 569.

Terres, R.R., and Luyendyk, B.P., 1985, Neogene tectonic rotation of the San Gabriel region,

California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,

p. 12467-12484, https://doi.org/10.1029/JB090iB14p12467

Woodburne, M.O., 1975, Cenozoic stratigraphy of the Transverse Ranges and adjacent areas,

southern California: Geological Society of America Special Paper 162, 91 p.,

https://doi.org/10.1130/SPE162

Yan, Z., Calyton, R.W., and Saleeby, J., 2005, Seismic refraction evidence for steep faults

cutting highly attenuated continental basement in the central Transverse Ranges,

California: Geophysical Journal International, v. 160, p. 651-666,

https://doi.org/0.1111/j.1365-246X.2005.02506.x

248

APPENDIX G: CHAPTER 4 – INVERSE-CONCORDIA AND

WEIGHTED-MEAN AGE RESULTS AND PLOTS

For each sample, shown left to right are: 1. A Weatherill concordia (206Pb/238U vs. 207Pb/235U) plot of all ages, A Terra-Wasserburg inverse-concordia (207Pb/206Pb vs. 207Pb/235U) plot of all

Oligocene-Miocene ages, and 3. A weighted-mean plot of all Oligocene-Miocene ages. Input for all plots is ratios reported in Table S4-1. Details of age calculations are below. Plots generated using IsoplotR (Vermeesch, 2018).

Terra-Wasserburg inverse-concordia (207Pb/206Pb vs. 207Pb/235U) plot of all Oligocene-Miocene ages:

 Age is calculated by fitting a discordia line through the data using the maximum

likelihood algorithm of Ludwig (1998), which assumes that the scatter of the data is

solely due to the analytical uncertainties. The upper intercept is the common-Pb

207 206 composition, ( Pb/ Pb)O, assumed to be 0.837 (Stacey and Kramers, 1975). The

lower intercept yields the age. For secondary standards older than Oligocene-

Miocene, age was calculated as a concordia age.

 Uncertainty is shown as ± x | y (| z ), where x is the analytical uncertainty (2σ) for the

age; y is the studentised 95% confidence interval for the age using the appropriate

number of degrees of freedom; z, only reported if MSWD (Wendt and Carl, 1991) >

1, is the approximate 95% confidence interval for the age with overdispersion,

calculated as z = y√MSWD. Ages reported in Fig. 4-6 show z, if calculated, or y

otherwise.

 n is the number of analyses used in the calculation 249

 MSWD is the Mean Square of the Weighted Deviates (MSWD) for the isochron fit

 p(χ2) is the Chi-squared p-value for the isochron fit

Weighted-mean plot of all Oligocene-Miocene ages:

 Age is calculated by first calculating a common-Pb-corrected (Stacey and Kramers,

1975) 206Pb/238U age and standard error for each individual analysis, then calculating

a weighted-mean, assuming that any dispersion between the different analyses is due

solely to the analytical uncertainty. Details in Vermeesch (2018).

 Outliers (shown in blue) rejected using a modified version of Chauvenet's criterion.

Details in Vermeesch (2018).

 Uncertainty is shown as ± x | y (| z ), where x is twice the standard error (2σ) for the

mean age; y is the width of the 95% confidence interval for the mean age, shown as a

grey band on the plot; z is the approximate 95% confidence interval for t with

overdispersion, calculated as z = y√MSWD.

 n is the number of analyses used in the calculation (outliers excluded)

 MSWD is the Mean Square of the Weighted Deviates (MSWD)

 p(χ2) is the Chi-squared p-value

Note: Latitude/Longitude in WGS84 reference coordinate system

250

251

252

253

254

255

Secondary standards:

Ages from Sessions 1 & 2, calculated as discordia-line ages for Oligocene-Miocene standard

(as with unknowns), and as concordia ages for standards older than Oligocene-Miocene (n

= number of analyses):

Published Session 1 Session 2 Standard: Reference n n age (Ma) age (Ma) age (Ma)

Dromedary 99.12 ± 0.1 Schoene et al. (2006) 98.4 ± 2.6 8 99.6 ± 1.4 22

FC1 1099.0 ± 0.6 Paces and Miller (1993) 1091 ± 14 15 1094 ± 6 39

Fish Canyon Schmitz and Bowring 28.48 ± 0.03 27.8 ± 0.6 10 27.8 ± 0.4 22 Tuff (2001)

Plešovice 337.13 ± 0.37 Sláma et al. (2008) 337 ± 7 Ma 6 342 ± 5 16

Ages from Sessions 1 & 2, calculated as common-Pb-corrected weighted-mean 206Pb/238U ages (as with unknowns) (n = number of analyses):

Published Session 1 Session 2 Standard: Reference n n age (Ma) age (Ma) age (Ma)

Dromedary 99.12 ± 0.1 Schoene et al. (2006) 98.8 ± 2.3 7 98.5 ± 1.0 22

FC1 1099.0 ± 0.6 Paces and Miller (1993) 1104 ± 16 15 1100 ± 7 39

Fish Canyon Schmitz and Bowring 28.48 ± 0.03 27.8 ± 0.6 10 27.7 ± 0.3 21 Tuff (2001)

Plešovice 337.13 ± 0.37 Sláma et al. (2008) 336 ± 9 6 338 ± 3 16

Weatherill concordia, Terra-Wasserburg inverse-concordia, and weighted-mean plots for secondary standards are below:

256

257

258

259

260

REFERENCES CITED

Ludwig, K.R., 1998, On the treatment of concordant uranium-lead ages: Geochimica et

Cosmochimica Acta, v. 62, p. 665-676, https://doi.org/10.1016/S0016-7037(98)00059-3

Paces, J.B., and Miller, J.D., 1993, Precise U-Pb ages of Duluth Complex and related mafic

intrusions, northeastern Minnesota: geochronological insights to physical, petrogenetic,

paleomagnetic, and tectonomagmatic processes associated with the 1.1 Ga Midcontinent

Rift System: Journal of Geophysical Research, v. 98, p. 13997-14013,

https://doi.org/10.1029/93JB01159

Schmitz, M.D., and Bowring, S.A., 2001, U-Pb zircon and titanite systematics of the Fish

Canyon Tuff: an assessment of high-precision U-Pb geochronology and its application to

young volcanic rocks: Geochimica et Cosmochimica Acta, v. 65, p. 2571-2587,

https://doi.org/10.1016/S0016-7037(01)00616-0

Schoene, B., Crowley, J.L., Condon, D.J., Schmitz, M.D., and Bowring, S.A., 2006, Reassessing

the uranium decay constants for geochronology using ID-TIMS U-Pb data: Geochmicia

et Cosmochimica Acta, v. 70, p. 426-445, https://doi.org/10.1016/j.gca.2005.09.007

Sláma, J., Košler, 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, Plešovice zircon – A new natural reference material

for U-Pb and Hf isotopic microanalysis: Chemical Geology, v. 249, p. 1-35,

https://doi.org/10.1016/j.chemgeo.2007.11.005

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. 207-211,

https://doi.org/10.1016/0012-821X(75)90088-6

261

Vermeesch, P., 2018, IsoplotR: a free and open toolbox for geochronology: Geoscience

Frontiers, v. 9, p. 1479-1493, https://doi.org/10.1016/j.gsf.2018.04.001

Wendt, I., and Carl, C., 1991, The statistical distribution of the mean squared weighted

deviation: Chemical Geology: Isotope Geoscience section, v. 86, p. 275-285,

https://doi.org/10.1016/0168-9622(91)90010-T

262