"Science, Faculty of"@en . "Earth, Ocean and Atmospheric Sciences, Department of"@en . "DSpace"@en . "UBCV"@en . "Caruthers, Andrew Harry"@en . "2013-04-17T09:12:32Z"@en . "2013"@en . "Doctor of Philosophy - PhD"@en . "University of British Columbia"@en . "The Pliensbachian\u00E2\u0080\u0093Toarcian marine extinction is observable at the species and generic levels. Ammonite diversity data from Europe and parts of the Arctic suggest a multi-phased event with diversity declining over six separate intervals. The main-phase of decline begins at the Pliensbachian\u00E2\u0080\u0093Toarcian boundary and extends into the Early Toarcian to a level correlative with the Tenuicostatum / Serpentinum Zone boundary. To date only this main-phase has been demonstrated as being global in extent, affecting multiple taxonomic groups. The entire Pliensbachian-Toarcian extinction has been attributed to regional and global controlling mechanisms associated with the Volcanic Greenhouse Scenario, an hypothesis linking eruption of the Karoo\u00E2\u0080\u0093Ferrar large igneous province (LIP) to global warming and mass extinction, specifically involving the release of methane hydrate from shelf reservoirs and a global marine anoxic event in the Early Toarcian (the T\u00E2\u0080\u0093OAE).\nThe study presented herein uses paleontology and isotope geochemistry to investigate the duration and potential controlling mechanisms of this protracted extinction. A primary objective is to compare new data from western North America with previously established records in Europe, testing: 1) the multi-phased nature of this extinction, 2) its magnitude within two taxonomic groups (ammonoids and foraminifera) in western North America and 3) its controlling mechanisms, relating to methane hydrate release and geographic extent of the T-OAE.\nResults show that all six phases of species decline are recognizable in western North America, even the oldest episode which was previously thought to be an event restricted to the Tethys Ocean area of Europe. This research strongly supports a correlation between the timing of the entire multi-phased extinction and formation of the Karoo igneous province. The study also provides one of the first records of the Early Toarcian \u00E2\u0080\u0098negative carbon-isotope excursion\u00E2\u0080\u0099 outside the Tethys Ocean area (concomitant with the main-phase of extinction) which implicates global methane hydrate release. Lastly, geochemical results do not support the presence of an anoxic water mass in the northeast paleo Pacific Ocean at the time of the so-called global Toarcian event (T-OAE)."@en . "https://circle.library.ubc.ca/rest/handle/2429/44233?expand=metadata"@en . "Pliensbachian\u00E2\u0080\u0093Toarcian (Early Jurassic) Extinction in Western North America by Andrew Harry Caruthers M.S., The University of Montana, 2005 B.S., The University of Puget Sound, 2002 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in THE FACULTY OF GRADUATE STUDIES (Geological Sciences) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) April 2013 \u00C2\u00A9 Andrew Harry Caruthers, 2013 ii Abstract The Pliensbachian\u00E2\u0080\u0093Toarcian marine extinction is observable at the species and generic levels. Ammonite diversity data from Europe and parts of the Arctic suggest a multi-phased event with diversity declining over six separate intervals. The main-phase of decline begins at the Pliensbachian\u00E2\u0080\u0093Toarcian boundary and extends into the Early Toarcian to a level correlative with the Tenuicostatum / Serpentinum Zone boundary. To date only this main-phase has been demonstrated as being global in extent, affecting multiple taxonomic groups. The entire Pliensbachian-Toarcian extinction has been attributed to regional and global controlling mechanisms associated with the Volcanic Greenhouse Scenario, an hypothesis linking eruption of the Karoo\u00E2\u0080\u0093Ferrar large igneous province (LIP) to global warming and mass extinction, specifically involving the release of methane hydrate from shelf reservoirs and a global marine anoxic event in the Early Toarcian (the T\u00E2\u0080\u0093OAE). The study presented herein uses paleontology and isotope geochemistry to investigate the duration and potential controlling mechanisms of this protracted extinction. A primary objective is to compare new data from western North America with previously established records in Europe, testing: 1) the multi- phased nature of this extinction, 2) its magnitude within two taxonomic groups (ammonoids and foraminifera) in western North America and 3) its controlling mechanisms, relating to methane hydrate release and geographic extent of the T-OAE. Results show that all six phases of species decline are recognizable in western North America, even the oldest episode which was previously thought to be an event restricted to the Tethys Ocean area of Europe. This research strongly supports a correlation between the timing of the entire multi-phased extinction and formation of the Karoo igneous province. The study also provides one of the first records of the Early Toarcian \u00E2\u0080\u0098negative carbon-isotope excursion\u00E2\u0080\u0099 outside the Tethys Ocean area (concomitant with the main-phase of extinction) which implicates global methane hydrate release. Lastly, geochemical results do not support the presence of an anoxic water mass in the northeast paleo Pacific Ocean at the time of the so-called global Toarcian event (T-OAE). iii Preface Data presented in this study is the result of collaboration between UBC\u00E2\u0080\u0099s paleontology laboratory and the Stable-Isotope Biogeochemistry Laboratory at Durham University (UK), led by Dr. Paul Smith, Dr. Darren Gr\u00C3\u00B6cke, and Andrew Caruthers. I was responsible for all of the geochemical sample collection / preparation of all of the samples, and for analyzing the Yakoun River 2010, Whiteaves Bay 2010 and South Barrow #3 datasets (a total of 1243 samples). I was also responsible for compiling the stratigraphic range charts, performing the extinction analysis, and was the primary writer in all published material. Dr. Smith supervised the sample collection, compilation of stratigraphic range charts, extinction analysis, and the writing process. Dr. Gr\u00C3\u00B6cke supervised the sample collection and isotope analysis (in Durham). Dr. Gr\u00C3\u00B6cke was also responsible for analyzing the Yakoun River 2008 and Whiteaves Bay 2008 datasets (a total of 353 samples). This thesis is composed of published and unpublished material. Published material is spread-out throughout all chapters of this work and does not constitute the entirety of any one chapter. In particular, modified versions of Figures 1.9, 1.11, 2.1, 3.5, 4.1, 4.4\u00E2\u0080\u00934.6, 5.7, 5.8, 6.3, Plate 1 and the related text within each chapter have been published in the following contributions: Caruthers, Andrew H., Gr\u00C3\u00B6cke, D.R., and Smith, P.L., 2011. The significance of an Early Jurassic (Toarcian) carbon-isotope excursion in Haida Gwaii (Queen Charlotte Islands), British Columbia, Canada. Earth and Planetary Science Letters, 307, 19\u00E2\u0080\u009326. I collected and prepared the samples for isotope geochemistry analysis and wrote the manuscript. Dr. Gr\u00C3\u00B6cke supervised sample collection and performed the isotopic analysis of these samples from the Yakoun River section. Dr. Smith supervised the sample collection and the writing process. Caruthers, Andrew H., and Smith, P.L., 2012. Pliensbachian ammonoids from the Talkeetna Mountains (Peninsular Terrane) of Southern Alaska. Review de Pal\u00C3\u00A9obiologie Volume sp\u00C3\u00A9cial 11, 365\u00E2\u0080\u0093378. I collected the ammonites and wrote the manuscript. Dr. Smith supervised the fossil identification and writing process. iv Table of Contents Abstract .................................................................................................................. ii Preface .................................................................................................................. iii Table of Contents .................................................................................................. iv List of Figures ........................................................................................................ vi List of Plates .......................................................................................................... ix Acknowledgements ............................................................................................... x Dedication ............................................................................................................ xii Chapter 1 Introduction ........................................................................................... 1 1.1 Mass extinctions ........................................................................................ 1 1.2 Mass extinction and the eruption of large igneous provinces .................... 9 1.3 The Pliensbachian\u00E2\u0080\u0093Toarcian extinction event ......................................... 14 1.3.1 Geochemical data from northwest Europe and the Mediterranean Tethys .......................................................................................................... 20 1.3.2 Controlling mechanisms .................................................................... 26 1.4 Study objectives ...................................................................................... 30 Chapter 2 The Pliensbachian\u00E2\u0080\u0093Toarcian time scale ............................................. 32 2.1 Pliensbachian zonal scheme .................................................................. 35 2.2 Toarcian zonal scheme .......................................................................... 40 2.3 Calibration with U\u00E2\u0080\u0093Pb ages .................................................................... 44 2.4 Zonal standards for the high-Arctic ......................................................... 45 2.4.1 Foraminiferal zone scheme for the high Arctic .................................. 47 Chapter 3 Methods: extinction & geochemical analyses ..................................... 51 3.1 Extinction analysis .................................................................................. 51 3.1.1 Diversity measurements ................................................................... 61 3.2 Geochemical analysis ............................................................................. 63 Chapter 4 Regional geology ................................................................................ 67 4.1 Haida Gwaii ............................................................................................ 69 4.1.1 Lithostratigraphy ............................................................................... 71 4.2 Northern Alaska ...................................................................................... 75 4.2.1 Lithostratigraphy ............................................................................... 78 4.3 Southern Alaska ..................................................................................... 80 4.3.1 Lithostratigraphy ............................................................................... 81 4.3.1.1 Biostratigraphy of the Hicks and Camp Creek sections ............... 83 4.3.1.1.1 Camp Creek section ............................................................... 85 4.3.1.1.2 Hicks Creek section ............................................................... 89 Chapter 5 Results ............................................................................................... 91 5.1 Extinction data and patterns of diversity ................................................. 91 5.1.1 Extinction and Origination patterns ................................................... 98 v 5.2 Geochemical data ................................................................................. 101 5.2.1 Whiteaves Bay section ................................................................... 101 5.2.2 Yakoun River section ...................................................................... 107 5.2.2.2 Higher resolution sampling ........................................................ 115 5.2.3 South Barrow #3 core ..................................................................... 117 3.2.3.1 Temporal constraints ................................................................. 120 Chapter 6 Discussion ........................................................................................ 123 6.1 A multi-phased Pliensbachian\u00E2\u0080\u0093Toarcian mass extinction .................... 123 6.1.1 Correlation with the Karoo\u00E2\u0080\u0093Ferrar magmatism ............................... 128 6.2 The negative CIE interval & the long-term carbon-isotope record ........ 133 6.2.1 The long-term carbon-isotope record ............................................. 137 6.3 Global vs. regional marine anoxia ........................................................ 147 6.3.1 Total organic carbon ....................................................................... 149 6.3.2 Nitrogen-isotope data ..................................................................... 153 6.3.2.1 Regional denitrification during the Early Pliensbachian ............. 157 Chapter 7 Systematic paleontology .................................................................. 162 Chapter 8 Conclusions ...................................................................................... 207 References ........................................................................................................ 215 Appendices ....................................................................................................... 245 Appendix A: Diversity and Rate Metrics ........................................................ 245 Appendix B: Geochemical Data ..................................................................... 249 \t\r \u00C2\u00A0 vi List of Figures Figure 1.1 Extinction rate and potential causes of major extinction events throughout the Phanerozoic .......................................................... 3 Figure 1.2 The Volcanic Greenhouse Scenario ............................................. 5 Figure 1.3 Continental and oceanic large igneous provinces ...................... 10 Figure 1.4 A comparison in timing of mass extinction events and eruption ages of large igneous provinces ................................................. 12 Figure 1.5 Discrepancy in timing of events related to the Pliensbachian\u00E2\u0080\u0093 Toarcian extinction ...................................................................... 14 Figure 1.6 Family-level extinction data for the Pliensbachian\u00E2\u0080\u0093Toarcian extinction event ........................................................................... 16 Figure 1.7 Species-level extinction data for the Pliensbachian\u00E2\u0080\u0093Toarcian extinction event from northwest Europe ...................................... 18 Figure 1.8 Multi-phased extinction data for the Pliensbachian\u00E2\u0080\u0093Toarcian extinction event ........................................................................... 19 Figure 1.9 Negative carbon-isotope excursion data from northwest Europe and parts of the Mediterranean ....................................... 22 Figure 1.10 Correlative chemostratigraphy of northwest Europe and parts of the Mediterranean .................................................................... 25 Figure 1.11 Global paleogeographic map for the Early Toarcian showing locations for localities discussed ................................................ 27 Figure 1.12 Schematic diagram of the restricted basin hypothesis ................ 29 Figure 2.1 Correlative Pliensbachian\u00E2\u0080\u0093Toarcian time scale .......................... 34 Figure 2.2 Stratigraphic ranges of ammonite species in western North America ....................................................................................... 37 Figure 2.3 Stratigraphic ranges of ammonite species in western North America ....................................................................................... 39 Figure 2.4 Stratigraphic ranges of ammonite species in western North America ....................................................................................... 42 Figure 3.1 Stratigraphic ranges of foraminiferal species in western North America ............................................................................. 52 Figure 3.2 Stratigraphic ranges of foraminiferal species in western North America ............................................................................. 54 Figure 3.3 Stratigraphic ranges of foraminiferal species in western North America ............................................................................. 55 Figure 3.4 Stratigraphic ranges of foraminiferal species in western North America ............................................................................. 56 Figure 3.5 Map of western North America showing localities in this Study .......................................................................................... 57 Figure 3.6 Correlation chart of previously published stratigraphic vii sections ...................................................................................... 58 Figure 3.7 Parameters and metrics used to analyze extinction ................... 63 Figure 4.1 Map showing major tectonic elements of the North American Cordillera .......................................................... 68 Figure 4.2 Pliensbachian\u00E2\u0080\u0093Toarcian stratigraphy of Haida Gwaii ................. 71 Figure 4.3 Summary of the depositional sequence of the Kingak Shale, Northern Alaska ............................................................... 78 Figure 4.4 Litho- and biostratigraphy of the Hicks Creek and Camp Creek sections, Talkeetna Mountains Alaska ........... 84 Figure 4.5 Distribution of Pliensbachian\u00E2\u0080\u0093Toarcian ammonite taxa identified in this study ......................................................... 86, 163 Figure 4.6 U\u00E2\u0080\u0093Pb concordia diagram for Ash Sample CCA-1, Camp Creek section .............................................................................. 88 Figure 5.1 Ammonite and foraminiferal species diversity data .................... 92 Figure 5.2 Foraminiferal species diversity data, combined and without long-ranging taxa ........................................................... 96 Figure 5.3 Extinction rate data for ammonites and foraminifera without longer stratigraphic ranges ........................................... 100 Figure 5.4 Litho- and biostratigraphy of the Whiteaves Bay section, Haida Gwaii ................................................................. 102 Figure 5.5 Geochemistry of the Whiteaves Bay section, Haida Gwaii .............................................................................. 105 Figure 5.6 Litho- and biostratigraphy of the Yakoun River section, Haida Gwaii ................................................................. 108 Figure 5.7 Geochemistry of the Yakoun River section, Haida Gwaii .............................................................................. 111 Figure 5.8 Seawater 87Sr/86Sr for the Late Pliensbachian\u00E2\u0080\u0093 Early Toarcian of Yorkshire (UK) .............................................. 114 Figure 5.9 Higher-resolution carbon-isotope data for the Yakoun River section, Haida Gwaii .......................................... 117 Figure 5.10 Geochemical data for the South Barrow #3 core, northern Alaska ........................................................................ 119 Figure 5.11 Litho- and biostratigraphy for the South Barrow #3 core, northern Alaska ............................................................... 122 Figure 6.1 Correlative multi-phased species level extinction of western North America, Europe and the Arctic ........................ 124 Figure 6.2 Comparison of timing between the eruptions of the Karoo\u00E2\u0080\u0093Ferrar province and the multi-phased extinction ........... 131 Figure 6.3 Correlative global negative carbon-isotope excursion data .......................................................................... 135 Figure 6.4 Late Triassic\u00E2\u0080\u0093Early Jurassic carbon-isotope data viii from Haida Gwaii ...................................................................... 139 Figure 6.5 Combined Pliensbachian\u00E2\u0080\u0093Toarcian carbon-isotope data from Haida Gwaii and northern Alaska ............................. 142 Figure 6.6 A comparison of Late Pliensbachian carbon-isotope data between western North America and Portugal ................. 143 Figure 6.7 A composite Late Permian\u00E2\u0080\u0093Middle Triassic carbon- isotope record from southern China ......................................... 146 Figure 6.8 Combined Pliensbachian\u00E2\u0080\u0093Toarcian organic carbon data from Haida Gwaii and northern Alaska ............................. 152 Figure 6.9 Combined Early Pliensbachian and Toarcian nitrogen- isotope data from Haida Gwaii and northern Alaska ................ 156 Figure 6.10 Adapted flow chart for the Volcanic Greenhouse Scenario showing the potential effects of global warming in restricted vs. unrestricted basins .............................................................. 157 Figure 6.11 Combined Early Pliensbachian and Toarcian total nitrogen data from Haida Gwaii and northern Alaska ............... 160 ix List of Plates Plate 7.1 Late Pliensbachian ammonoids from the Talkeetna Mountains ................................................................................. 189 Plate 7.2 Toarcian ammonoids from Haida Gwaii .................................... 192 Plate 7.3 Specimens of the Toarcian genus Cleviceras from the Yakoun River section ................................................................ 195 Plate 7.4 Toarcian ammonoids from Haida Gwaii .................................... 197 Plate 7.5 A specimen from the genus Phymatoceras from the Yakoun River section ................................................................ 199 Plate 7.6 Toarcian ammonoids from Haida Gwaii .................................... 201 Plate 7.7 A specimen from the genus Rarenodia, Yakoun River section ...................................................................................... 203 Plate 7.8 Toarcian ammonoids from the Yakoun River section ............... 205 x Acknowledgements This research could not have been accomplished without support and guidance from countless individuals. My supervisors, Paul Smith and Darren Gr\u00C3\u00B6cke, were instrumental in this undertaking! I would like to express my sincerest gratitude for their unwavering intellectual support, generous funding, and continuous enthusiasm for this research topic. I would also like to thank my committee, Stuart Sutherland, Jim Mortensen and Mike Orchard, for their assistance and support. Stuart and Mike, for the many insightful tips and conversations during my committee meetings and Jim (as well as Rich Friedman at the Pacific Centre for Isotopic and Geochemical Research, UBC), for conducting the U\u00E2\u0080\u0093Pb TIMS isotope analyses. Funding for this work was generously provided by NSERC research grants to Paul Smith (#8493) and Darren Gr\u00C3\u00B6cke (#288321), a NERC grant to Darren Gr\u00C3\u00B6cke (NE/H021868/1), financial support to the Stable-Isotope Biogeochemistry Laboratory (SIBL) at Durham University was provided by TOTAL UK, a grant from the Alaska Geological Society (to AHC), and tuition support from UBC. I would like to express my sincerest gratitude to those who helped me complete the various portions of this research. In Alaska: Robert B. Blodgett for pointing me in the direction of the Talkeetna Mountains back in 2004, for access to previously collected material from the Camp Creek section, and for his guidance in potential contacts for the South Barrow #3 Core; Curvin Metzler was of critical assistance while in the field, and provided much logistical support as well as access to previously collected material from both sections at Hicks Creek and Camp Creek; Ken Papp, Kurt Johnson and Jean Riordan at the Geologic Materials Center in Eagle River Alaska for allowing me to collect samples from the South Barrow #3 Core, and for their patience!; Ken Bird (USGS) and John Reeder (DNR) provided much assistance in selecting the South Barrow #3 Core, and allowed access to published (and unpublished) material; and Boris Nikitenko generously provided an unpublished version of the foraminiferal zone scheme for the South Barrow #3 core, which greatly aided the level of biostratigraphic constraint for geochemical results. J\u00C3\u00B3zsef P\u00C3\u00A1lfy and Giselle Jakobs helped locate the starting point of their sections along the Yakoun River, which ultimately led to the understanding of new exposures in the lower part of the section. Joanne Peterkin at the Stable-Isotope Biogeochemistry Laboratory (SIBL) at Durham University provided analytical assistance during sample analysis. Marc Bustin is thanked for access to his centrifuge and Wayne Vogl (at the Life Sciences Centre, UBC) for \u00E2\u0080\u0098ultrapure\u00E2\u0080\u0099 Milli-Q water during sample preparation and George Stanley Jr. at the University of Montana Paleontology Center for allowing the curation of ammonites from the Talkeetna Mountains. This research was also greatly aided by the discussions and suggestions of Jim Haggart, Guillaume Dera, Steve Calvert and Mike Foote who provided much insight toward various parts of the extinction and geochemical analyses. I would also like to thank many people who helped me along the way, either in the field by collecting ammonites (the fun part!) or by collecting samples xi for geochemistry (the not-so-fun part!); or in lab by keeping my spirits up through their constant enthusiasm and humor. A very special thanks to Luke Beranek, Mike Valentine, Allison Perrigo, Erika Kercher, and Malcolm Brown for assistance while in the field; and to my lab-mates Melissa Grey, Louise Longridge, Emily Hopkin, Farshad Shirohammad, Martyn Golding, Pengfei Hou, and Sarah Porter. Lastly, this work could not have been accomplished without the support, encouragement and love of my girlfriend (Ellie Young), parents (Jennie and Marvin Caruthers), aunt (Sally Smoly \u00E2\u0080\u0098the great one\u00E2\u0080\u0099), baba (Mary Smoly), as well as the rest of my family (Jon Caruthers, Alexander Caruthers, Fiona Caruthers- Jucker, and Amy Nafziger). Thanks for being there for me through the tough times! And finally, how could I forget my furry companions, AnnaBelle, Monkey and cousin Cheryl, who always kept my spirits up by providing a constant source of entertainment and companionship. xii Dedication To my parents Jennie and Marvin Caruthers, who were always insistent that I \u00E2\u0080\u009CDo my homework\u00E2\u0080\u009D and constantly urged me to see the bigger picture of life and the potential opportunities ahead. You are my biggest inspiration! 1 Chapter 1 Introduction 1.1 Mass extinctions Mass extinctions involve the global disruption of a broad suite of environments causing a significant proportion of the world\u00E2\u0080\u0099s biota to disappear over a narrow interval of geological time (Hallam and Wignall, 1997). They force the biosphere to restructure and, along with other background extinctions and recoveries, help form the basis of biostratigraphy (Raup, 1994). Recently, mass extinctions have become an important topic of interest because several studies have indicated that the world is currently experiencing a major decline in biodiversity that may be human-induced (Wilson, 1993; Lawton and May, 1995; Leakey and Lewin, 1995; Pimm et al., 1995; Thomas et al., 2004a,b). Understanding the dynamics, causes and extent of extinctions is a central question for science and society. Throughout the Phanerozoic, five first order and up to 27 second order mass extinction events have been identified (Sepkoski, 1986; Raup, 1994; Hallam and Wignall, 1997). First order events show rapid declines in biodiversity that are measurable at the family, genus and species levels (Raup and Sepkoski, 1982; Sepkoski, 1993; Benton and Storrs, 1994; Jablonski, 1995; Benton, 1995; Hallam and Wignall, 1997). During the five well-known first order events (known as the \u00E2\u0080\u0098Big Five\u00E2\u0080\u0099), an estimated 75\u00E2\u0080\u009395% of existing species are thought to have become extinct (Raup, 1994). These events occur in the Ordovician, Devonian, Permian, Triassic, and Cretaceous Periods (Figure 1.1A) and show extinction 2 intensities between 70-95% among families and 76-95% among genera (Jablonski, 1995; Hallam and Wignall, 1997; Bambach, 2006). Second order events, on the other hand, are less well understood. They are often restricted to a specific environment or biogeographic realm and are characterized by much lower extinction rates (estimated to be 8%, 26%, and 33\u00E2\u0080\u009353% at the family, genus, and species levels respectively using data for two events discussed by Harries and Little, 1999). Due to the lower extinction intensity at the family level, second order events are often best identifiable at the genus and species levels. 3 Figure 1.1 - A) Extinction rate throughout the Phanerozoic for Paleozoic evolutionary fauna (adapted from Peters, 2008). Paleozoic evolutionary fauna refers to the low-biomass, epibenthic suspension feeding organisms that were dominant during the Paleozoic time period (Peters, 2008). Error bars show 95% confidence. Major \u00E2\u0080\u0098Big Five\u00E2\u0080\u0099 mass extinction events are indicated, sequentially, in red. B) Potential cause and calculated extinction intensity for major \u00E2\u0080\u0098Big Five\u00E2\u0080\u0099 mass extinctions (adapted from Hallam and Wignall, 1997). O = Ordovician, S = Silurian, D = Devonian, C = Carboniferous, P = Permian, Tr = Triassic, J = Jurassic, K = Cretaceous, Pg = Paleogene, Ng = Neogene. 4 Currently there are a wide variety of causal factors and mechanisms invoked to explain first and second order mass extinction events (Figure 1.1B). Nearly all of these factors are Earth-bound and associated with periods of prolonged environmental change relating to events such as eustatic sea-level change, global warming and/or global cooling (Figure 1.1B). Recently, one mechanism-chain in particular has emerged that links global warming to both first and second order extinction events throughout the Late Paleozoic and Mesozoic (Figure 1.2). It is referred to as the \u00E2\u0080\u0098Volcanic Greenhouse Scenario\u00E2\u0080\u0099 by Wignall (2005) and it proposes that the eruptions forming large igneous provinces (LIPs) are responsible for initiating a cascade of events which include global warming, marine anoxia and mass extinction (P\u00C3\u00A1lfy and Smith, 2000; Wignall, 2001; 2005). It is this model that is explored in the current study. 5 Figure 1.2 - The \u00E2\u0080\u0098Volcanic Greenhouse Scenario\u00E2\u0080\u0099 a model showing the potential chain of events leading to environmental change and mass extinction (modified from Wignall, 2001; 2005). Note, figure is composite from many individual events, not every isotopic system is necessarily affected. Green bubbles are specific events that are addressed in my study. The volcanic greenhouse scenario model invokes outgassing of volcanogenic CO2 and other greenhouse gasses during LIP eruption (Figure 1.2) caused an escalation of temperatures and prolonged global warming (P\u00C3\u00A1lfy and Smith, 2000; Wignall, 2001). Warmer water temperature is thought to have initiated a series of events that resulted in large-scale environmental change (P\u00C3\u00A1lfy and Smith, 2000; Hesselbo et al., 2000; Wignall, 2001). These events are summarized in Figure 1.2 and include: 1) the catastrophic release of methane 6 hydrate along the continental shelf sediment reservoir, 2) disruption of the thermohaline circulation of the ocean, 3) increased weathering and erosion on the continents and 4) ocean stagnation and marine anoxia. The resulting geological evidence of these large-scale events is a geochemical record that contains sizable perturbations in many isotopic systems (Figure 1.2); these include carbon- (Jenkyns, 1988; 2003; 2010; Holser et al., 1989; Magaritz et al., 1992; Gr\u00C3\u00B6cke et al., 1999; Hesselbo et al., 2000; 2002; 2004; Wignall, 2001; P\u00C3\u00A1lfy et al., 2001; Ward et al., 2001; Payne et al., 2004; Williford et al., 2007; Whiteside et al., 2007; 2010; Suan et al., 2008; 2010;), nitrogen- (Jenkyns et al., 2001), molybdenum- (Pearce et al., 2008; McArthur et al., 2008), sulfur- (Gill et al., 2011), strontium- (McArthur et al., 2000) and osmium- (Cohen and Coe, 2002; 2007; Cohen et al., 2004; Kuroda et al., 2010) isotopes. It should be noted that Figure 1.2 is a composite diagram showing the total number of isotopic systems potentially affected by LIP eruption and mass extinction, although not every isotopic system listed in Figure 1.2 is necessarily affected simultaneously in each mass extinction event. The development of regionally (or globally?) extensive marine anoxia may result from four factors that are attributed to warmer water temperatures: 1) warm water holds less dissolved oxygen than cold water and therefore the greenhouse effect caused by increasing CO2, from LIP eruption alone, can contribute directly to anoxic marine conditions (Wignall, 2005), 2) warmer surface and bottom water on continental shelves causes a destabilization in the methane hydrate reservoir, this releases a large amount of methane into the water column and the 7 atmosphere which oxidizes quickly into CO2 and thereby reinforces greenhouse conditions and reduces the concentration of dissolved oxygen in the water column (Hesselbo et al., 2000; Wignall, 2005), 3) warmer water disrupts water current dynamics and oceanic circulation patterns by decreasing the amount of cold and dense surface water in polar regions, this decreases the amount of oxygen that can descend to deep water environments and contributes to marine anoxia (Wignall, 2005), 4) global warming causes an increase in continental weathering and runoff, through increased precipitation, which increases the supply of nutrients into the ocean, this increases productivity and lowers the concentration of dissolved oxygen in the water column through the process of biomass decay (Wignall, 2005). Under normal marine redox conditions degradation (oxidation) of organic matter by bacteria is accomplished through the consumption of oxygen, which produces an abundance of carbon dioxide, nitrate, and phosphate (Froelich et al., 1979). However, as oxygen becomes less abundant or absent from the environment, this process begins to consume nitrate. Nitrate is commonly known as the second-best (to oxygen) electron acceptor within this system (Froelich et al., 1979) and under extreme conditions, after the available nitrates are depleted, sulfate-reducing bacteria continue this process. These bacteria reduce a variety of other compounds and minor elements which include manganese, iron hydroxide, sulfate, and methane fermentation. This produces free hydrogen sulfide in the water column and results in a geochemical record with perturbations in the affected isotopic systems (Figure 1.2). 8 Recently, work has also linked anoxic marine conditions with perturbations (in the geological record) in molybdenum and molybdenum-isotopes (McArthur et al., 2008; Pearce et al., 2008). It is believed that in the presence of free hydrogen sulfide, molybdenum is removed from the anoxic (or euxinic) water column by reduction to thiomolybdate complexes (MoOxS2\u00E2\u0080\u00934\u00E2\u0080\u0093x) and is sequestered in organic matter and/or pyrite, which is also formed in the water column (Helz et al., 1996; Erickson and Helz, 2000; Vorlicek and Helz, 2002; Tribovillard et al., 2004a,b, 2006, 2008; McArthur et al., 2008). This process of anaerobic decay of biomatter altogether is less efficient than that mediated by aerobic microbial processes, so that more of the settling particulate flux survives transport through the water column, leading to organic enrichment of the resulting sediment (Figure 1.2). These periods of increased organic carbon (wt% TOC) in the sedimentary record have been used to identify periods of suspected ocean anoxia in the geological past (Jenkyns, 1988; Hallam, 1995; Jenkyns and Clayton, 1997; Hallam and Wignall, 1997; P\u00C3\u00A1lfy, 2003; Jenkyns, 2003; 2010). The study presented herein is a critical evaluation of this model. It utilizes aspects of marine invertebrate paleontology as well as carbon- and nitrogen- isotope geochemistry to investigate the second order Pliensbachian\u00E2\u0080\u0093Toarcian mass extinction and its correlation with the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP. It specifically addresses the timing and geographic extent of this correlation with respect to: 1) mass extinction, 2) methane release, and 3) marine anoxia (green bubbles in Figure 1.2). 9 1.2 Mass extinction and the eruption of large igneous provinces Large igneous provinces are massive, large-scale, eruptions of predominantly mafic extrusive and associated intrusive igneous rocks (Figure 1.3). They include continental and oceanic flood basalts, volcanic passive margins, oceanic plateaus, submarine ridges and seamount groups (Coffin and Eldholm, 1994). They consist mainly of subaerial sheet flows that can range in volume from several hundred to several thousand cubic kilometers (Baksi, 1990; Arndt et al. 1993; Coffin and Eldholm, 1993; 1994; Courtillot et al., 1999; Wignall, 2001; Bryan et al., 2010) and occur globally on continents as well as oceanic basins (Figure 1.3A). With the exception of basalt eruptions associated with spreading centers, LIPs are the largest accumulations of mafic material on the Earth\u00E2\u0080\u0099s surface (Coffin and Eldholm, 1994). The formation of LIPs, which are largely thought to be the result of mantle plumes or hotspots (Wilson, 1963; Morgan, 1971; 1972; 1981; Coffin and Eldholm, 1994) produce considerable quantities of CO2 and other greenhouse gasses. As previously discussed, the long-term outgassing during LIP eruptions is thought to cause a significant global climate change (P\u00C3\u00A1lfy and Smith, 2000; Wignall, 2001; 2005). 10 Figure 1.3 - A) Map of the world showing areal extent of various continental and oceanic Large Igneous Provinces (after Wignall, 2001). Karoo Traps include main lavas, dike swarms, and sills (areal extent after Jourdan et al., 2008). Position and areal extent of Ferrar Group after Morrison and Reay (1995). B) List of Large Igneous Provinces with their radiometric age, estimated volume, and corresponding extinction event (compiled from data in Wignall, 2001; Duncan, 2002; Courtillot and Renne, 2003; Jourdan et al., 2008; Wignall et al., 2009). 11 Extensive research over the last two decades has revealed that many mass extinction events and LIP eruptions are approximately coeval (Courtillot, 1999; P\u00C3\u00A1lfy and Smith, 2000; Wignall, 2001; 2005 Courtillot and Renne, 2003; Suan et al., 2008; Caswell et al., 2009). Studies have shown that of the 12 major LIP episodes, 10 coincide with known extinction events (Figure 1.3B, 1.4). Four of these occur consecutively throughout the mid-Phanerozoic and include, in stratigraphic order: 1) the Emeishan flood basalts (~263 Ma; in Sun et al., 2010) and the Middle Permian extinction; 2) the Siberian Traps (~250 Ma) and the end Permian extinction (Wignall, 2001); 3) the Central Atlantic Magmatic Province (CAMP) flood basalt (~200 Ma) and the end Triassic extinction (Wignall, 2001); and 4) the Karoo\u00E2\u0080\u0093Ferrar flood basalt (~183 Ma) and the Pliensbachian\u00E2\u0080\u0093Toarcian extinction (orange circle in Figure 1.4; P\u00C3\u00A1lfy and Smith, 2000). In the case of the Deccan Traps (~65 Ma) and the end Cretaceous extinction, this correlation has been validated (Wignall, 2001; Keller, 2008); however there is strong evidence that this mass extinction event was also heavily influenced by a major bolide impact (Wignall, 2001 and references therein). Other mass extinctions occur at times of low sea level (Figure 1.1B) when loss of shelf space would certainly affect biodiversity. Regression may reflect the eustatic influence of glaciations or tectonics (changes in mid-oceanic ridges) or the more regional, epeirogenic influence of continental doming (followed by fragmentation and the formation of LIPS). However, the interrelationships between regression and other agents of environmental change are beyond the scope of my study. 12 Figure 1.4 - A diagram comparing the timing of mass extinction versus the eruption ages of Large Igneous Provinces over the past 300 million years (after Courtillot, 1999). Note that ages of the Karoo\u00E2\u0080\u0093Ferrar Traps are approximated from data in Jourdan et al. (2008) and that the Pliensbachian\u00E2\u0080\u0093Toarcian extinction event, the subject of this study, is denoted in orange. One of the most intriguing aspects of this correlation is the affect that voluminous outgassing of CO2 during LIP eruption has on the global carbon reservoir. Several studies have shown that during times of LIP eruption, there are large perturbations in the carbon-isotope (\u00CE\u00B413C) record, which form the basis for hypotheses concerning environmental change (P\u00C3\u00A1lfy and Smith, 2000; Hesselbo et al., 2000; Wignall, 2001; Jenkyns, 2010). Recently, however, discrepancies have been noted for some of these events concerning the relative timing of extinction, flood basalt eruption, perturbation in the carbon-isotope record, and 13 marine anoxia (Wignall et al., 2005; Wignall and Bond, 2009; Wignall et al., 2009). In some instances it is argued that the main extinction interval precedes either flood basalt eruption or the initial perturbation in carbon isotopes, which causes concern for the validity of a LIP/mass extinction correlation. The Pliensbachian\u00E2\u0080\u0093Toarcian extinction is one such event in which this timing has been questioned (Wignall et al., 2005). In the study by Wignall et al. (2005), a composite dataset was generated from a well-known Early Toarcian succession in northwest Europe showing a clear discrepancy in timing between the extinction interval (with co-occurring anoxic marine facies) and the large negative perturbation in carbon-isotope values (Figure 1.5). Their conclusions have therefore cast some doubt on the timing within the Volcanic Greenhouse Scenario and further motivate this study. 14 Figure 1.5 - Summary of Early Toarcian sedimentation, ichnofacies, benthic macrofauna, and carbon-isotope gechemistry along the Yorkshire coast at Kettleness (UK), modified from Wignall et al. (2005). Figure shows: 1) a good correlation between the extinction horizon of benthic organisms at Kettleness and the appearance of organic-rich shale (indicating anoxic conditions), and 2) a discrepancy in timing between these events and the well-known negative CIE. Dashed red line denotes main extinction of benthic macrofauna (in Wignall et al., 2005) and approximate position of extinction horizon (ii) in Caswell et al. (2009). Diagonal lines indicate the approximate lower limit of anoxic facies, as indicated in Wignall et al. (2005). Serp. = Serpentinum, Exar. = Exaratum, Jet = Jet Rock, HCS = Hummocky cross-stratification. 1.3 The Pliensbachian\u00E2\u0080\u0093Toarcian extinction event The Pliensbachian\u00E2\u0080\u0093Toarcian extinction is best recognized in marine organisms at the family, genus and species levels of northwest Europe 15 (Sepkoski, 1982; 1992; 1996; Hallam, 1986; 1987; Little, 1996; Little and Benton, 1995; Harries and Little, 1999; Caswell et al., 2009). A calculation by Harries and Little (1999, p.45) suggests that this event significantly impacted taxa at the species level in northwest Europe, with a species extinction intensity of 90% among ammonites, belemnites, bivalves (deposit-feeding infaunal, suspension- feeding infaunal, and semi-infaunal), gastropods, and brachiopods. However, a more recent calculation of ammonite species extinction in the northwest Tethyan and Arctic domains by Dera et al. (2010) suggests this group sustained losses that are measured between 40\u00E2\u0080\u009365% at the subzone level and 70\u00E2\u0080\u009390% at the zone level. In Europe, this extinction event also affected a variety of other benthic and pelagic taxonomic groups including: radiolarians, foraminifera, ostracods, dinocysts, crinoids, asteroids, crustaceans, coleoids, marine reptiles, and fish (Hallam, 1986; 1987; Little and Benton, 1995; Hallam and Wignall, 1997; Harries and Little, 1999; V\u00C3\u00B6r\u00C3\u00B6s, 2002; Cecca and Macchioni, 2004; Zakharov et al., 2006; Caswell et al., 2009; Garc\u00C3\u00ADa Joral et al., 2011). 16 Figure 1.6 - Number of extinct marine families throughout the Early Jurassic (from Little and Benton, 1995). Data show heightened global family-level extinction in five ammonite biozones spanning the Late Pliensbachian to Middle Toarcian. NW = northwest, Eur. = Europe, Teth. = Tethys, Aust. = Austria, Bor. = Boreal, G. = Grammoceras, D. = Dactylioceras, A. = Arnioceras. Family level data show a broad decline in marine biodiversity that spans five ammonite biozones from the Late Pliensbachian to Early Toarcian with extinction peaks in the Late Pliensbachian Margaritatus and Early Toarcian Serpentinum (Falciferum) Zones of northwest Europe (Figure 1.6). Throughout this interval, 33 of 49 global and 18 of 33 northwest European families went extinct (Little and Benton, 1995). 17 At the species and genus level, there is a major decline in biodiversity that occurred within the Early Toarcian at a level correlative with the Tenuicostatum / Serpentinum (Falciferum) Zone boundary (Benton, 1993; Little and Benton, 1995; Harries and Little, 1999; Zakharov et al., 2006; Caswell et al., 2009; Bilotta et al., 2010). This boundary is considered to be the main extinction interval of this event and is thought to have affected many benthic and pelagic marine groups (Figure 1.7). It has been observed in many Tethyan and Boreal regions and has been argued to be global in geographic extent (Zakharov et al., 2006; Caswell et al., 2009; Dera et al., 2010). It is this correlative boundary interval that some argue, coincides with the deposition of organic rich black shale in northwest Europe (Hallam, 1987; Jenkyns, 1988; Little, 1995; Little and Benton, 1995; Jenkyns and Clayton, 1997; Harries and Little, 1999). However the exact timing of extinction at the subzonal level with respect to black shale deposition within northwest Europe (Figure 1.5), and in other outcrops, has been questioned (Wignall et al., 2005; Gomez et al., 2008; McArthur et al., 2008; Garcia Joral et al., 2011). 18 Figure 1.7 - A summary of composite taxonomic ranges of Late Pliensbachian to Middle Toarcian marine species in four localities of northwest Europe (originally compiled by Harries and Little, 1999). Ranges show heightened extinction among marine groups at two main intervals, one at the Pliensbachian / Toarcian boundary and the other within the Early Toarcian at the Tenuicostatum / Serpentinum Zone boundary (denoted as dashed-grey boxes). Intervals of heightened extinction were originally marked by Harries and Little (1999) as four separate steps (dashed line at upper and lower boundary) and were subsequently shown to also occur in the Tethyan and Arctic Realms, steps (i) and (iii) in Caswell et al. (2009). Tenui. = Tenuicostatum, Sp. = Spinatum. Specific identification of marine taxa can be found in Harries and Little (1999). Ammonite zonal scheme after Page (2003). A study by Dera et al. (2010) of Pliensbachian\u00E2\u0080\u0093Toarcian ammonite biodiversity and morphology in the NW Tethyan and Arctic domains has greatly expanded the known temporal extent of this event. In this study a multi-phased event is shown where species and genus level biodiversity declined over six specific intervals of time throughout the Pliensbachian\u00E2\u0080\u0093Toarcian Stages (Figure 1.8). These drops in biodiversity include the: 1) Ibex\u00E2\u0080\u0093Davoei Zone, 2) the 19 Gibbosus subzone, 3) the Pliensbachian / Toarcian boundary, 4) Semicelatum subzone, 5) Bifrons\u00E2\u0080\u0093Variabilis Zones and 6) Dispansum Zone, with the major main-phase extinction event at a level that is correlative with the Tenuicostatum / Serpentinum Zone boundary. In Dera et al.\u00E2\u0080\u0099s work, events 2\u00E2\u0080\u00936 are interpreted as contributing to the Pliensbachian\u00E2\u0080\u0093Toarcian extinction. Event number 1, occurring in the Early Pliensbachian at the Ibex\u00E2\u0080\u0093Davoei Zone boundary, was originally identified and interpreted by Dommergues et al. (2009) as a regional decline in ammonite species diversity across the Mediterranean and northwest European parts of the Tethys Ocean. 20 Figure 1.8 - Species and generic ammonite diversity in the Pliensbachian and Toarcian of Europe (modified from Dera et al., 2010). Six major declines in biodiversity are recognized with the largest being in the Early Toarcian at the Tenuicostatum / Serpentinum Zone boundary. Major declines in biodiversity are numbered and in red. \u00E2\u0080\u0098Weighted richness\u00E2\u0080\u0099 refers to the total number of taxa per measured time interval; \u00E2\u0080\u0098geographic singletons\u00E2\u0080\u0099 refer to taxa that are confined to a specific geographic area; \u00E2\u0080\u0098single-interval taxa\u00E2\u0080\u0099 are confined to a specific subzone; and \u00E2\u0080\u0098boundary crossers\u00E2\u0080\u0099 refer to taxa that occur in two separate time intervals. Marg. = Margaritatus, Sp. = Spinatum, Ten. = Tenuicostatum, Ser. = Serpentinum, Bif. = Bifrons, Variab. = Variabilis, Thouar. = Thouarsense, Dis. = Dispansum, Ps. = Pseudoradiosa, Aal. = Aalensis. To date, this multi-phased extinction has not been established in other taxonomic groups, except for the two phases at the Pliensbachian / Toarcian boundary and the Tenuicostatum / Serpentinum boundary within the Early Toarcian where declines in diversity have been observed in ammonites, bivalves, brachiopods and other previously mentioned taxonomic groups (Harries and Little, 1999; V\u00C3\u00B6r\u00C3\u00B6s, 2002; Caswell et al., 2009). Neither has the multi-phased extinction been shown to be global in extent. Except for the decline across the Ibex / Davoei Zone boundary in the Early Pliensbachian, Dera et al. (2010) speculate that the multi-pulsed volcanic activity of the Karoo\u00E2\u0080\u0093Ferrar LIP could have been a trigger for the extinction phases in the Pliensbachian\u00E2\u0080\u0093Toarcian time. However, in order to better assess this control mechanism, it is necessary to first discuss the many changes in isotope geochemistry that also occur. 1.3.1 Geochemical data from northwest Europe and the Mediterranean Tethys Within the epicontinental seaway of NW Europe and the western Tethys Ocean there is a well-documented Early Toarcian disruption in seawater geochemistry that is thought to co-occur with widespread black shale deposition and extinction of marine organisms (Jenkyns, 1988; Jones et al., 1994; Hesselbo 21 et al., 2000; McArthur et al., 2000; Jenkyns et al., 2001; Jones and Jenkyns, 2001; Jenkyns et al., 2002; Cohen et al., 2004; McElwain et al., 2005; Kemp et al., 2005; Caswell et al., 2009). Carbon-isotope data (Figure 1.9) reveal a small negative shift at the Pliensbachian / Toarcian boundary (Hesselbo et al., 2007; Suan et al., 2008; Littler et al., 2010; Bodin et al., 2010) that is subsequently followed by a broad positive shift that extends into the Serpentinum Zone. It is then interrupted at the Tenuicostatum/Serpentinum Zone boundary by a pronounced negative \u00CE\u00B413C excursion of ~5\u00E2\u0080\u00937\u00E2\u0080\u00B0 (Hesselbo et al., 2000, 2007; Kemp et al., 2005; Suan et al., 2008; Sabatino et al., 2009; Hermoso et al., 2009a). 22 Figure 1.9 - Late Pliensbachian\u00E2\u0080\u0093Early Toarcian carbon-isotope records from northwest Europe and the Mediterranean illustrating correlative negative excursions at the Pliensbachian / Toarcian boundary and within the Early Toarcian at the correlative Tenuicostatum / Serpentinum Zone boundary. Figure is adapted from a compilation in Caruthers et al. (2011). Locality numbers at top refer to paleogeographic location in Figure 1.11. This negative CIE has been recorded in many geologic materials including: bulk organic carbon (Hesselbo et al., 2000; Kemp et al., 2005; Sabatino et al., 2009), carbonate (Hesselbo et al., 2007, Hermoso et al., 2009a; Sabatino et al., 2009; Bodin et al., 2010), phytoplanktonic organic matter and organic biomarkers (Schouten et al., 2000; van Breugel et al., 2006), benthic 23 shells (Suan et al., 2008, 2010), fossil wood and phytoclasts (Hesselbo et al., 2000; McElwain et al., 2005; Hesselbo et al., 2007; Hesselbo and Pienkowski, 2011) suggesting a close relationship between the carbon-isotope composition of the carbon reservoirs in the ocean and atmosphere. At the onset of this present study, the geographic extent of this large negative excursion in carbon-isotopes was challenged. In a study of \u00CE\u00B413C from belemnite calcite in Early Toarcian sedimentary successions in England (Yorkshire) and Germany, van de Schootbrugge et al., (2005) were unable to identify this negative CIE and thereby challenged the geographic extent of this excursion by suggesting it is a regional phenomena that was principally derived from the upwelling of 12C enriched bottom water into a saline-stratified body of water within the epicontinental seaway of northwest Europe. However, this current research and recent contributions by other groups has shown that the negative CIE is recorded in many coeval Early Toarcian successions that are geographically outside the Tethys Ocean area (Al-Suwaidi et al., 2010, Caruthers et al., 2011; Suan et al., 2011; Gr\u00C3\u00B6cke et al., 2011; Izumi et al., 2012). This new evidence of an Early Toarcian negative CIE that is outside the Tethys Ocean area strongly indicates a globally extensive perturbation to the carbon reservoir requiring a global control mechanism. In the epicontinental seaway of NW Europe and parts of the Mediterranean Tethys, the negative CIE coincides with dramatic increases in organic carbon burial (Jenkyns, 1988; Jenkyns and Clayton, 1997; Hesselbo et al., 2000) that are also correlative with perturbations in many geochemical 24 systems (Figure 1.10). These include: positive shifts in nitrogen-isotopes (Jenkyns et al., 2001), osmium-isotopes (Cohen et al., 2004), manganese concentrations (Sabatino et al., 2011), sulfur-isotopes (Gill et al., 2011) and sulfur concentrations (McArthur et al., 2008), turning points in strontium-isotope ratios (McArthur et al., 2000), and a minimum in molybdenum (Mo ppm, Mo/total organic carbon or TOC, and \u00CE\u00B498Mo) values. In northwest Europe, the minimum in molybdenum values is subsequently followed by a positive shift in Mo ppm, Mo/TOC, and \u00CE\u00B498Mo that coincides with a positive shift in \u00CE\u00B413C in the later part of the Early Toarcian (Pearce et al., 2008; McArthur et al., 2008). Furthermore, a recent comparison of carbon- and oxygen-isotope ratios in the Mediterranean Tethys region shows that there was a sharp negative shift in \u00CE\u00B418Obrachiopod values at the Pliensbachian / Toarcian boundary and a subsequent broad decrease during the Early Toarcian negative CIE interval (Suan et al., 2008, 2010). As previously explained by the Volcanic Greenhouse Scenario (Figure 1.2), this nearly simultaneous disruption in seawater geochemistry is compelling evidence for climate change and ocean anoxia in this region (Jenkyns et al., 2001; Jenkyns, 2003; 2010; McArthur et al., 2008). 25 26 Figure 1.10 - Correlative chemostratigraphy throughout the Late Pliensbachian Middle Toarcian successions of the Cleveland and Cardigan Bay Basins (northwest Europe) showing geochemical profiles and concentrations of (from left to right): organic carbon, molybdenum, sulfur, nitrogen-isotopes, osmium- isotopes, carbon-isotopes, and strontium-isotope values (modified from McArthur et al., 2008). Spin. = Spinatum, Tenui. = Tenuicostatum, Cleveland. = Clevelandicum. Locality numbers at top refer to paleogeographic location in Figure 1.11. Note that temporal correlation of manganese concentrations from Monte Mangart (Italy / Slovenia border) was not given at the zone-level in Sabatino et al. (2011). Black arrow denotes extent of Jet Rock. 1.3.2 Controlling mechanisms Currently there are two leading hypotheses as to the cause of the CIE in the Early Toarcian, invoking global controls in one case and regional controls in the other. The global hypothesis is synonymous with the Volcanic Greenhouse Scenario (Figure 1.2). It is argued that in the Early Toarcian, rift-related tectonic activity during the break-up of Gondwana initiated the eruption of the Karoo\u00E2\u0080\u0093 Ferrar LIP (Figure 1.11). Outgassing of volcanogenic CO2 from this eruption started a prolonged period of global warming (P\u00C3\u00A1lfy and Smith, 2000) that subsequently destabilized and released ~5000 Gt of methane hydrate from continental shelf sediments (Hesselbo et al., 2000; Beerling et al., 2002). The rapid injection and oxidation of methane reduced oceanic O2 levels and increased organic carbon burial, creating euxinic conditions (known as the Toarcian Ocean Anoxic Event or T\u00E2\u0080\u0093OAE). The T\u00E2\u0080\u0093OAE is largely evidenced by correlative perturbations in seawater geochemistry including: organic carbon, nitrogen, sulfur, molybdenum, and manganese (Figure 1.2; Jenkyns, 2010 and references therein; Gill et al., 2011). This severe environmental perturbation is further thought to have: 1) increased seawater temperatures by ~7\u00E2\u0080\u009310\u00CB\u009AC, as 27 evidenced by a sudden decrease in brachiopod oxygen-isotope ratios (Suan et al., 2010); 2) caused a significant increase in continental weathering rates, evident from increases in strontium- and osmium-isotope ratios (Jones and Jenkyns, 2001; McArthur et al., 2000; Cohen et al., 2004); 3) caused a major reduction in biocalcification (Mattioli et al., 2004; Tremolada et al., 2005; Suan et al., 2010); and 4) escalated the extinction intensity of marine organisms at the Tenuicostatum / Serpentinum ammonite Zone boundary (see Caswell et al., 2009 and references therein; Bilotta et al., 2010; Dera et al., 2010). Astronomical precession may have also helped to reinforce the effects of long-term global warming, resulting in the cyclic (at least three separate pulses) release of methane hydrate lasting an estimated ~60\u00E2\u0080\u0093350 kyr per cycle (Kemp et al., 2005; Cohen et al., 2007; Hesselbo and Pienkowski, 2011; Kemp et al., 2011). 28 Figure 1.11 - A and B) Global paleogeographic maps of the Early Toarcian depicting the location of the Tethys and Paleo\u00E2\u0080\u0093Pacific Oceans, the Karoo\u00E2\u0080\u0093Ferrar Large Igneous Province (LIP), and location of various localities in northwest Europe (after Hesselbo et al., 2007; Caswell et al., 2009), the Mediterranean (after Sabatino et al., 2010), South America (after Al-Suwaidi et al., 2010) and this study. 1 = Yorkshire (UK), 2 = Peniche (Portugal), 3 = Mochras (Wales), 4 = Sancerre (France), 5 = Monte Mangart (Italian\u00E2\u0080\u0093Slovenian border), 6 = Neuqu\u00C3\u00A9n Basin (Argentina), 7 = Wrangellia (Smith, 2006), 8 = northern Alaska (approximate position based on Smith, 2006; Imlay, 1981). As discussed earlier, the timing and geographic extent of the T\u00E2\u0080\u0093OAE has been disputed. Questions have been raised concerning: 1) differences in the timing of black shale deposition, mass extinction, carbon-isotope perturbation, and sea level change between northern (Boreal) and southern (Tethyan) Europe (Figure 1.5; Wignall et al., 2005; Bilotta et al., 2010); 2) the geographic extent and concentration of organic carbon in what is regarded as \u00E2\u0080\u0098globally distributed black shale units\u00E2\u0080\u0099, inferred to have been deposited in euxinic conditions (McArthur et al., 2008); and 3) the reported geographic extent for a disruption in molybdenum (Mo ppm) and molybdenum-isotope ratios that are thought to be an indication of euxinic conditions in Early Toarcian marine sediments of NW Europe (Pearce et al., 2008; McArthur et al., 2008). These questions have prompted an alternative hypothesis suggesting the T\u00E2\u0080\u0093OAE may have been caused by local controls and was therefore restricted to the epicontinental seaway of NW Europe (McArthur et al., 2008). This hypothesis proposes that the European epicontinental seaway was the site of several silled basins (Figure 1.12) where black shales, 12C-enriched bottom waters and the previously discussed perturbations in isotope geochemistry formed in response to euxinic conditions brought on by a salinity-driven pycnocline that developed as 29 a result of increased fresh water input (K\u00C3\u00BCspert, 1982; Jenkyns, 1988; S\u00C3\u00A6len et al., 1996; Schouten et al., 2000; McArthur et al., 2008). In this model, the negative CIE occurred by diachronous upwelling of 12C-rich bottom water in separate marine basins in the Tethys Ocean area and was therefore restricted geographically. Figure 1.12 - Schematic diagram showing environmental dis-tress, saline stratification, the development of anoxic (or euxinic) redox conditions, and subsequent deposition of organic rich black shale in restricted basins of northwest Europe during the Early Toarcian (modified from McArthur et al., 2008). It is possible that both hypotheses are applicable and that regional and global controls could have occurred simultaneously thereby reinforcing each other. As mentioned, at the onset of this research the negative CIE was unknown outside the area of the Tethys Ocean. The first discovery of this excursion in carbon-isotopes (outside the Tethys Ocean) came from the Neuqu\u00C3\u00A9n Basin in 30 Argentina, where it was described from two sections (Al-Suwaidi et al., 2010; Mazzini et al., 2010). Its presence was interpreted as evidence that regional controls were not operating alone. However, these data are somewhat problematic. Al-Suwaidi et al. (2010) demonstrate a 6\u00E2\u0080\u00B0 negative shift in \u00CE\u00B413Corg and \u00CE\u00B413Cwood that is truncated by slumping and erosion at the top of the section so that only the lower part of the Early Toarcian CIE is evident. The data presented by Mazzini et al. (2010) are poorly constrained biostratigraphically but show an 8\u00E2\u0080\u00B0 negative excursion in \u00CE\u00B413Corg that, in contrast to conclusions drawn in Al-Suwaidi et al. (2010), seems to occur below the correlative of the Tenuicostatum / Serpentinum Zone boundary. If correct, this would suggest a diachronous, rather than isochronous, negative CIE when compared to European data. Although geographically far-removed from the Tethys Ocean, the Neuqu\u00C3\u00A9n Basin was similarly somewhat restricted and therefore could have been subjected to coeval regional and global controls as postulated for Europe. It is therefore necessary to seek correlative and preferably more complete data from sediments that were deposited in a marine environment outside the Tethys Ocean area that was not part of a restricted basin. 1.4 Study objectives The study presented herein forms a critical evaluation of the Pliensbachian\u00E2\u0080\u0093Toarcian extinction and its hypothesized control mechanisms. By examining the calibrated biostratigraphy and chemostratigraphy in western North America and comparing it with correlative data in Europe, it will assess: 1) the geographic extent of the multi-phased ammonite extinction recorded by Dera et 31 al. (2010); 2) if other taxonomic groups, such as foraminifera, showed similar declines in biodiversity that are correlative with those observed in ammonites; 3) the geographic extent of the well-documented European perturbation in carbon- isotopes (negative CIE); and 4) the long-term (Pliensbachian\u00E2\u0080\u0093Toarcian) carbon- and nitrogen-isotope records that are related to a potentially global T\u00E2\u0080\u0093OAE. Carbon-isotope chemostratigraphy in this research is generated from two temporally constrained (to the zonal level) and well-studied areas of western North America, representing deposition along terrane and craton margins. A primary goal is to calibrate the stratigraphic record of biodiversity changes (from the Late Pliensbachian \u00E2\u0080\u0093 Late Toarcian strata) with a newly generated North American carbon isotope curve for the Pliensbachian\u00E2\u0080\u0093Toarcian interval. Results will be compared with established records in the Tethys Ocean in order to assess this extinction and its controlling mechanisms on a global scale. 32 Chapter 2 The Pliensbachian\u00E2\u0080\u0093Toarcian time scale In order to compare paleontological and geochemical signals of western North America with the well-established records from Europe and parts of the Mediterranean, it is imperative that coeval strata be identified. Therefore, it is important to establish a geologic time scale for the Pliensbachian and Toarcian that can be applied globally. Jurassic ammonites have been utilized for this purpose and are commonly held as one of the most useful index fossils for zone and subzone level temporal constraint. Throughout the Early Jurassic, ammonites are known to have evolved rapidly producing a broad variety of shell morphologies. Their pelagic mode of life also facilitates a wide geographic dispersal which included both deep- and shallow-water marine habitats. Ammonites of the Jurassic have had a long history of research in Europe dating back to the mid 1800\u00E2\u0080\u0099s with the work of Quenstedt (1845\u00E2\u0080\u00931849; 1856\u00E2\u0080\u0093 1858; 1882\u00E2\u0080\u00931885) and Oppel (1853; 1862). These studies laid the framework for a comprehensive Early Jurassic ammonite zonal scheme that was applicable to many successions throughout northwest Europe (Dean et al., 1961). Since this scheme was established, it has undergone refinement (Mouterde et al., 1971; Schlatter, 1980; Howarth, 1992; Page, 2003) and is currently used as a global reference standard with which to compare and calibrate other schemes (Braga et al., 1982; Hillebrandt, 1987; Smith et al., 1988; Jakobs et al., 1994; Zakharov et al., 1997; Page, 2003). 33 In western North America, Early Jurassic interbedded fossiliferous marine sedimentary and volcaniclastic successions are extremely useful for calibrating biochronologic and geochronologic time scales. Previous studies have established an ammonite-based zonal scheme for Pliensbachian and Toarcian successions (Smith et al., 1988; Jakobs et al., 1994; Smith and Tipper, 1996; Jakobs, 1997) that is calibrated with absolute age determinations from interbedded zircon-bearing ash beds (P\u00C3\u00A1lfy et al., 1997; 2000). This calibrated time scale has been further correlated with the ammonite zonal schemes of NW Europe, parts of the Mediterranean, and South America (Figure 2.1). This zonal scheme integrates associations of ammonite species that have similar, overlapping, stratigraphic ranges and provides the basis for zonal boundaries (Smith et al., 1988; Jakobs et al., 1994; Smith and Tipper, 1996). Within this scheme, the stratigraphic range of a particular index (or characterizing, or name bearing) species is not necessarily confined to a zone and, furthermore, its presence is not always necessary in order to recognize the zone (Smith et al., 1988). The study presented herein uses the ammonite based zone schemes of Smith et al. (1988) and Jakobs et al. (1994) to correlate paleontological and geochemical results from western North America with well established coeval records from northwest Europe and parts of the Mediterranean. 34 Figure 2.1 - Correlative Pliensbachian\u00E2\u0080\u0093Toarcian time scales for Northwest Europe and the Mediterranean (Dean et al., 1961; Schlatter, 1980; Braga et al., 1982; Howarth, 1992; Page, 2003); High-Arctic (Zakharov et al., 1997; Nikitenko et al., 2008); western North America (Smith et al., 1988; Jakobs et al., 1994) and South America (von Hillebrandt, 2006). Absolute U\u00E2\u0080\u0093Pb and Ar\u00E2\u0080\u0093Ar age data from P\u00C3\u00A1lfy et al. (1997; 2000). Absolute age dates in bold-face font have an error range that is less than 5 Ma and are interpreted as good quality. Note that the Pliensbachian zone level biostratigraphy in the Mediterranean Province refers to areas of Spain (Page, 2003). 35 2.1 Pliensbachian zonal scheme The Pliensbachian of western North America is divided into five separate zones that include, from the earliest to latest, the Imlayi, Whiteavesi, Freboldi, Kunae, and Carlottense zones (Figure 2.1). The Early Pliensbachian includes the Imlayi, Whiteavesi, and Freboldi Zones and is derived primarily on the stratigraphic ranges of the families Polymorphitidae and Eoderoceratidae (Smith et al., 1988). The combined duration of the chronozones is just short of being equivalent to the combined time frame of the Jamesoni, Ibex and Davoei Zones of northwest Europe. The later part of the Pliensbachian includes the Kunae and Carlottense Zones and, at present, its definition is based primarily on the gradual evolution (and stratigraphic ranges) of two ammonite superfamilies: the Hildoceratoidea and Eoderoceratoidea (Meister, 2010). However, the occurrence of these superfamilies are restricted to specific paleogeographic domains and therefore show limited overlap. The transition between the Sinemurian and Pliensbachian remains poorly defined. Smith et al. (1988) recognize that throughout much of western North America, the definition of the boundary is hampered by unfossiliferous intervals that separate ammonite species of Sinemurian affinity from those of the Pliensbachian. A more recent study, however, has identified a somewhat transitional ammonite fauna from areas of Haida Gwaii (meaning \u00E2\u0080\u009Cislands of the Haida people\u00E2\u0080\u009D and formerly the Queen Charlotte Islands), identified as the Tetraspidoceras Assemblage (P\u00C3\u00A1lfy et al., 1994). This so-called Tetraspidoceras Assemblage contains a mix of ammonites that are known from the uppermost Sinemurian and lowermost Pliensbachian. The presence of this assemblage may 36 be an indication that the base of the Imlayi Zone is not the base of the Pliensbachian. However, in western North America the Tetraspidoceras Assemblage has also not yet been demonstrated to occur outside Haida Gwaii and therefore it was not included in Figure 2.1. In the present study, successions yielding ammonite species from the Tetraspidoceras Assemblage were not observed. The Imlayi Zone is characterized by the zonal index species Pseudoskirroceras imlayi and a variety of other taxa (Figure 2.2). The Whiteavesi Zone is marked by the disappearance of Pseudoskirroceras imlayi and the appearance of Tropidoceras masseanum (Smith et al., 1988). Also within this zone, Acanthopleuroceratids such as the zonal index A. whiteavesi occur in abundance. The overlying Freboldi Zone is marked by the appearance of Dubariceras freboldi above the last acanthopleuroceratid (Smith et al., 1988). D. freboldi extends stratigraphically throughout the zone and is described as being widely distributed and endemic to the East Pacific (Frebold, 1970; Imlay, 1968; 1981; Smith, 1983; Dommergues et al., 1984; Smith et al., 1988). 37 Figure 2.2\u00E2\u0080\u00932.4 - Pliensbachian and Toarcian ammonite biostratigraphy of western North America compiled from previously published data of Imlay (1955; 1981), Smith et al. (1988), Thomson and Smith (1992), Jakobs (1992; 1995; 1997), Jakobs et al. (1994), Smith and Tipper (1996), Jakobs and Smith (1996), Smith et al. (2001), and Caruthers and Smith (2012). 38 In western North America the first appearance of the genus Fanninoceras, primarily above the species Dubariceras freboldi is used to define the base of the Kunae Zone (Smith et al., 1988). In the uppermost Freboldi Zone there is a brief overlap in the stratigraphic ranges of D. freboldi and F. bodegae (Figure 2.2, 2.3) and therefore occurrences of Fanninoceras above this interval of overlap are used to define the Kunae Zone. Although the Kunae Zone lies predominantly within the Late Pliensbachian, its base is situated within the uppermost part of the Early Pliensbachian (Figure 2.1). This zone is characterized further by a variety of species from numerous genera (Figures 2.2, 2.3). The Carlottense Zone is the uppermost zone of the Pliensbachian. Its base is marked by the first appearance of the index species Fanninoceras (Fanninoceras) carlottense (Smith et al., 1988) and is further characterized by various other species of Lioceratoides (Lioceratoides), Protogrammoceras, Amaltheus, Fieldingiceras, Lioceratoides (Pacificeras), and Tiltoniceras (Figure 2.3). In western North America the species Lioceratoides (Pacificeras) propinquum, Protogrammoceras paltum, and Tiltoniceras antiquum occur well below the first appearance of Dactylioceras, which is known globally as the earliest Toarcian ammonite genus (Smith et al., 1988; Smith and Tipper, 1996). However, in Europe and other parts of the world the first appearance of L. propinquum, P. paltum, and T. antiquum parallel the first appearance of Dactylioceras and therefore suggest a younger age. This could suggest that these species evolved in the paleo Pacific Ocean in the latest Pliensbachian and then radiated to other parts of the world by the Early Toarcian. 39 40 2.2 Toarcian zonal scheme In western North America the Toarcian is divided into five ammonite zones (Jakobs et al., 1994; Jakobs, 1997). In ascending stratigraphic order these include the Kanense, Planulata, Crassicosta, Hillebrandti, and Yakounensis Zones (Figure 2.1). Within this scheme, the Kanense Zone comprises the entire Early Toarcian time period and is equivalent to the combined time frame of both the Tenuicostatum and Serpentinum Zones of northwest Europe. The Middle Toarcian contains the Planulata and Crassicosta Zones and correlates with the Bifrons and Variabilis Zones of northwest Europe, and the Late Toarcian contains the Hillebrandti and Yakounensis Zones and correlates with the combined time frame of the Thouarsense, Dispansum, Pseudoradiosa, and Aalensis Zones of northwest Europe (Figure 2.1). As previously discussed, in northwest Europe and in parts of the Mediterranean, the main-phase of the Pliensbachian\u00E2\u0080\u0093Toarcian mass extinction and the accompanying large negative CIE occurs at the Tenuicostatum / Serpentinum Zone boundary within the Early Toarcian (Figures 1.7; 1.9). Also mentioned, in western North America, the Kanense Zone is approximately equivalent to the combined Tenuicostatum and Serpentinum Zones of northwest Europe (Figure 2.1; Jakobs et al., 1994; Jakobs, 1997). Therefore, if the Early Toarcian mass extinction and the negative CIE are isochronous global events, they should occur within the Kanense Zone and would therefore indicate a stratigraphic position that is approximately coeval with the Tenuicostatum / Serpentinum Zone boundary. 41 The base of the Kanense Zone also marks the base of the Toarcian. It is characterized by the first appearance of Dactylioceras above the last occurrences of Amaltheus and Fanninoceras species (Smith et al., 1988; Jakobs et al., 1994; Jakobs, 1997). The Kanense Zone also has several other characterizing species from the genera Dactylioceras, Taffertia, Hildaites, Harpoceras, and Cleviceras (Figure 2.3). Of these, D. kanense, D. aff. comptum, D. cf. compactum, D. alpestre, and T. taffertensis are restricted to the lower part of the zone, whereas Hildaites spp. Hildaites cf. serpentinum, Hildaites murleyi, and Cleviceras cf. chrysanthemum are known from its upper portion (Figures 2.3, 2.4). 42 43 The Middle Toarcian Planulata Zone is marked by the first appearance of Rarenodia planulata above the last occurrence of Hildaites murleyi (Jakobs et al., 1994; Jakobs, 1997). Rarenodia planulata is the index species of the Planulata Zone and unlike D. kanense whose taxonomic range is restricted to the lower part of the underlying Kanense Zone, the taxonomic range of R. planulata is known to extend throughout the zone (Figure 2.4). The base of the Crassicosta Zone is defined by the first appearance of Phymatoceras crassicosta and Peronoceras cf. moerickei above the last Rarenodia planulata (Jakobs et al., 1994). Phymatoceras crassicosta, the zonal index, extends throughout much of the zone and disappears near its top (Figure 2.4). The Late Toarcian Hillebrandti Zone is marked by the appearance of Grammoceras thouarsense, Podagrosites latescens, and the index species Phymatoceras hillebrandti above the last occurrence of P. crassicosta (Jakobs et al., 1994; Jakobs and Smith, 1996). This zone is regarded as having one of the lowest species-level diversities in the Toarcian (Jakobs and Smith, 1996), with the majority of its fauna confined to the zone. The Yakounensis Zone is the uppermost Toarcian ammonite zone in western North America. Its base is defined by the first appearance of Yakounia silvae, Pleydellia maudensis, Hammatoceras speciosum, and Pleydellia spp. above the last occurrence of Phymatoceras hillebrandti (Jakobs et al., 1994). In general, ammonite species of the Yakounensis Zone are largely confined to the zone (Figure 2.4). Species include the index Yakounia yakounensis and a host of other characterizing species (Jakobs et al., 1994). To date, only certain species belonging to the 44 family hammatoceratinae (in Jakobs et al., 1994; Jakobs, 1997) and possibly Holcophylloceras calypso (in Jakobs and Smith, 1996) appear to extend into the Aalenian Stage of the Middle Jurassic. 2.3 Calibration with U\u00E2\u0080\u0093Pb ages The Jurassic time scale has been calibrated using a database of U\u00E2\u0080\u0093Pb and 40Ar/39Ar ages to integrate bio- and geochronologic time scales in many successions throughout western North America, providing age estimates for stage and zonal boundaries throughout this period of time (P\u00C3\u00A1lfy et al., 2000). Two methods were largely employed: 1) A direct dating method, where the age of a biochronologically defined boundary is determined through isotopic dating of a particular volcanogenic layer that is in close proximity to the boundary itself, and 2) the chronogram method, where the age of the boundary in question is estimated using age determinations from adjacent units (discussed in detail in P\u00C3\u00A1lfy et al., 2000). Age determinations are considered \u00E2\u0080\u0098good quality\u00E2\u0080\u0099 if they contained a 2\u00CF\u0083 error range that is less than 5 Ma (age determinations in bold font in Figure 2.1). Although the direct dating method is preferred, within the Early Jurassic time scale it could only be used to date the Triassic\u00E2\u0080\u0093Jurassic boundary and the initial boundary of the Crassicosta Zone of the Middle Toarcian (P\u00C3\u00A1lfy et al., 2000). U\u00E2\u0080\u0093Pb and 40Ar/39Ar age estimates for the Pliensbachian and Toarcian are as follows. The base of the Pliensbachian is poorly constrained at 190.7 +2.7\u00E2\u0080\u00933.9 Ma; the base of the Freboldi Zone is constrained at 186.7 +1.8\u00E2\u0080\u00931.6 Ma; the base of the Kunae Zone is constrained at 185.7 +0.5\u00E2\u0080\u00930.6 Ma; the base of the Carlottense 45 Zone is constrained at 184.1 +1.2\u00E2\u0080\u00931.6 Ma; the base of the Toarcian (base of the Kanense Zone) is constrained at 183.6 +1.7\u00E2\u0080\u00931.1 Ma; the base of the Planulata Zone is poorly constrained at 182 +3.3\u00E2\u0080\u00934.9 Ma; the base of the Crassicosta is constrained at 181.4 +1.2\u00E2\u0080\u00931.2 Ma (using data from P\u00C3\u00A1lfy et al., 1997); the base of the Yakounensis Zone is constrained at 180.1 +0.7\u00E2\u0080\u00933.0 Ma; and the base of the Aalenian is constrained at 177.6 +1.4\u00E2\u0080\u00931.1 Ma (Figure 2.1). This data indicates that the Early Pliensbachian lasted for ~5.8 Ma, the later part of the Pliensbachian lasted for ~2.1 Ma, the Early Toarcian lasted for ~1.6 Ma, and the Middle and Late Toarcian combined lasted for ~4.4 Ma. 2.4 Zonal standards for the high-Arctic Early Jurassic fossiliferous marine successions are widespread and well studied throughout much of northeast Russia, Siberia, Barents Sea, and northern Alaska (Nikitenko 1994; 2008; Zakharov et al., 1997; Shurygin et al., 2000; Nikitenko and Mickey, 2004; Nikitenko et al., 2008; Shurygin et al., 2011). To date, ammonite biostratigraphy within these fossiliferous successions provides the basis for a Pliensbachian and Toarcian Zonal scheme for the Arctic that is applicable to many circum Polar Regions (Zakharov et al., 1997). The Pliensbachian portion of this zone scheme is composed of four ammonite-based zones that include, in ascending stratigraphic order, the Polymorphites, Stokesi, Margaritatus, and Viligaensis Zones (Figure 2.1). In this scheme the Early Pliensbachian consists only of the Polymorphites Zone and a gap interval for which zone-level constraint is currently unavailable. The basal Pliensbachian Polymorphites Zone overlies the Upper Sinemurian Colymicum 46 Zone; its base is consistent with the base of the Pliensbachian (Zakharov et al., 1997). Although the top of the Polymorphites Zone is currently not known due to the overlying gap interval, it has been illustrated as correlative with the Jamesoni Zone of northwest Europe (Zakharov et al., 1997). The Late Pliensbachian part of the zone scheme consists of the Stokesi, Margaritatus, and Viligaensis Zones of which the Stokesi Zone is temporally equivalent with the Stokesi Subzone of northwest Europe (Figure 2.1) and the top of the overlying Margaritatus Zone is consistent with that of northwest Europe (Zakharov et al., 1997). The Viligaensis Zone is the uppermost zone of the Late Pliensbachian and it is considered to be correlative with the Spinatum Zone of northwest Europe (Figure 2.1). The Toarcian part of this circum Polar zonal scheme is somewhat different than in European schemes, in that the Middle Toarcian is not formally recognized (as illustrated in Zakharov et al., 1997; Nikitenko and Mickey, 2004; Zakharov et al., 2006; Nikitenko et al., 2008). This scheme consists of eight zones that are subdivided into the Early and Late Toarcian (Zakharov et al., 1997; Nikitenko et al., 2008). These include, in ascending stratigraphic order, the Antiquum, Falciferum, Commune, Monestieri, and Spinatum Zones of the Early Toarcian and the Compactile, Wuerttenbergeri, and Falcodiscus Zones of the Late Toarcian (Figure 2.1). The basal zone of the Toarcian was originally identified as the Propinquum Zone in Zakharov et al. (1997), however a more recent work by Nikitenko et al. (2008) re-names this zone as the Antiquum Zone; its base is consistent with the base of the Toarcian (Figure 2.1). Within this Early Toarcian 47 zone scheme, the Antiquum and overlying Falciferum Zones are correlative with the Early Toarcian Tenuicostatum and Falciferum Zones of northwest Europe. The overlying Commune, Monestieri, and Spinatum Zones together are temporally equivalent to the Middle Toarcian Bifrons Zone of northwest Europe (Zakharov et al., 1997; Nikitenko et al., 2008). In the Late Toarcian part of the zone scheme, the Compactile Zone correlates with the Middle Toarcian Variabilis Zone of northwest Europe and the Wuerttenbergeri and Falcodiscus Zones are temporally equivalent to the Late Toarcian Thouarsense, Dispansum, Pseudoradiosa, and Aalensis Zones of northwest Europe (Figure 2.1). 2.4.1 Foraminiferal zone scheme for the high Arctic More recent studies have constructed separate zonal schemes that are specific to high Arctic regions and are based on stratigraphic ranges of a variety of macro and microfossils, such as bivalves, foraminifera, and ostracodes (Nikitenko and Mickey, 2004; Nikitenko et al., 2008; Shurygin et al., 2011). The research presented here utilizes the foraminifera-based zone scheme (illustrated in Figure 2.1) in conjunction with ammonite biostratigraphy, to temporally constrain well core data from northern Alaska. This foraminifera-based zone scheme is based, in part, on previously identified specimens that were recovered from various well cores in the National Petroleum Reserve of northern Alaska (Tappan, 1955; Bergquist, 1966). The foraminiferal zone scheme is applicable to the Early\u00E2\u0080\u0093Middle (Aalenian only) Jurassic of the circum Polar Region and is composed of 17 zones that are based on assemblages of foraminifera that have similar overlapping stratigraphic ranges (Nikitenko and Mickey, 2004; Nikitenko et al., 2008). Each zonal assemblage is identified as a JF# (e.g. JF9) with zonal 48 boundaries, discussed herein, referring to the northern Alaska portion of the scheme only. They are calibrated with the ammonite zone scheme (Figure 2.1) and are presented in detail by Nikitenko and Mickey (2004). It should be noted, however, that this foraminiferal zone scheme contains a regionally extensive component; in that each geographic region (in the circum- polar area) contains a slightly different representation of individual zones and their boundaries. This is represented when comparing the Upper Pliensbachian part of the zonal scheme from northeast Siberia and Russia versus northern Alaska (Figures 2 and 4 in Nikitenko and Mickey, 2008). Within these two geographic areas, northeastern Siberia and Russia contain the JF5 and JF6 zones, whereas in northern Alaska these zones do not seem to appear. Furthermore, this foraminiferal zone scheme often contains zones that are temporally divided into both large- and fine-scales, which is represented by a particular zone that has a long stratigraphic range and contains shorter ranging zones within its boundaries (e.g. the long ranging JF4 Zone with shorter ranging JF7\u00E2\u0080\u0093JF9 zones that are contained within its boundary in Figure 2.1). This could potentially indicate the presence of a regionally extensive subzone-level temporal constraint. The Pliensbachian and Toarcian portion of this zone scheme for northern Alaska contains nine foraminiferal zones (Figure 2.1). Within this scheme, the base of the Pliensbachian does not occur at a zone boundary but rather within the Trochammina inusitata\u00E2\u0080\u0093Turritellella volubilis JF2 Zone. The JF2 Zone is long ranging and extends from the Upper Sinemurian to the lower part of the Upper 49 Pliensbachian (Nikitenko and Mickey, 2004). The upper boundary of the JF2 Zone occurs within the Late Pliensbachian Stokesi Zone (Figure 2.1). Overlying the JF2 Zone is the Trochammina lapidosa JF4 Zone (Figure 4 in Nikitenko et al., 2008). The JF4 Zone occurs in northwest Siberia where it contains the species Ammodiscus siliceus, Hyperammina ex gr. odiosa, Saccammina sp. and Jaculella jacutica (Nikitenko and Mickey, 2004). This zone is long ranging, extending from near the base of the Upper Pliensbachian to the top of the Antiquum Zone of the Early Toarcian (Figure 2.1). The Anmarginulina arctica\u00E2\u0080\u0093A. gerkei JF7\u00E2\u0080\u0093JF8 Zone occurs within the boundaries of the JF4 Zone and is known to be strictly Late Pliensbachian in age (Figure 4 in Nikitenko et al., 2008). The JF7\u00E2\u0080\u0093JF8 Zone extends from the upper part of the Stokesi Zone to the middle of the Viligaensis Zone (Figure 2.1) and includes a wide variety of taxa (Nikitenko and Mickey, 2004). Overlying the JF7\u00E2\u0080\u0093 JF8, and still within the stratigraphically broader JF4 Zone, is the Recurvoides taimyrensis JF9 Zone. This zone spans the Pliensbachian / Toarcian boundary and extends to the top of the Antiquum ammonite Zone, which is also coincident with the top of the JF4 zone (Figure 2.1). In particular, they note the disappearance of JF7\u00E2\u0080\u0093JF8 Zone taxa in the lower part of the zone and the appearance of Early Toarcian taxa in its upper portion, which is taken to denote the Pliensbachian / Toarcian boundary (Nikitenko and Mickey, 2004). The stratigraphically lowest Toarcian taxa, as noted by Nikitenko and Mickey (2004), include Trochammina kisselmani, Triplasia kingakensis and Ammodiscus glumaceus. At the upper boundary of the JF9 Zone there is a distinct change in 50 the taxonomic composition of foraminifera which marks the beginning of the overlying Trochammina kisselmani JF10 and the Ammobaculites lobus\u00E2\u0080\u0093 Trochammina kisselmani JF11 zones (Nikitenko and Mickey, 2004; Nikitenko et al., 2008). The JF10 Zone is another long ranging zone and extends to the top of the Lower Toarcian, correlating with the Middle Toarcian of northwest Europe (Figure 4 in Nikitenko et al., 2008). It contains the entire JF11 Zone and the lower portion of the Astacolus praefoliaceus\u00E2\u0080\u0093Lenticulina multa JF12 Zone (Figure 2.1). The JF11 Zone extends to a stratigraphic level that is equivalent with the lower part of the Commune Zone. Lastly, the JF12 Zone extends to the Lower Aalenian and includes the species Lenticulina toarcense, and Nodosaria pulchra (Nikitenko and Mickey, 2004). 51 Chapter 3 Methods: extinction & geochemical analyses 3.1 Extinction analysis Pliensbachian and Toarcian ammonoid and foraminiferal species-level diversities in western North America were measured from compiled stratigraphic range charts derived from many previous accounts (Figures 2.2\u00E2\u0080\u00932.4; 3.1\u00E2\u0080\u00933.4). New biostratigraphy from the Talkeetna Mountains (Alaska) and Haida Gwaii (British Columbia), presented in this study, helped to supplement and modify previously known stratigraphic ranges of several Late Pliensbachian and Early Toarcian ammonoids including the species Fanninoceras (Charlotticeras) cf. maudense, Leptaleoceras? sp., Lioceratoides (Lioceratoides) cf. L. involutum, Dactylioceras cf. compactum, Cleviceras spp., Hildaites spp. and Hildaites murleyi (Caruthers et al., 2011; Caruthers and Smith, 2012). Stratigraphic range charts, used herein, incorporate occurrences of fossils from many areas of western North America. Species ranges were then analyzed using various metrics established by Foote (2000) as well as Hammer and Harper (2006) and include: Total diversity, total diversity minus singletons, estimated mean standing diversity, per-taxon rate, Van Valen rate, and estimated per-capita rate (defined on Figure 3.7). The resulting values are a measure of diversity and extinction/origination rates for these two taxonomic groups throughout the Pliensbachian\u00E2\u0080\u0093Toarcian interval. 52 53 Figures 3.1\u00E2\u0080\u00933.4 - Combined stratigraphic ranges of foraminifera species for two areas of western North America throughout the Pliensbachian and Toarcian. Stratigraphic ranges of foraminifera species are from various sections in Kottachchi et al. (2002, 2003) and well core samples in Tappan (1955). Localities used here are presented in Figures 3.5 and 3.6 of this study. Ranges are plotted with respect to the ammonite zone scheme of Smith et al (1988) and Jakobs et al. (1994). Note that species in bold occur in both areas. 54 55 56 Throughout western North America, Pliensbachian and Toarcian ammonoid faunas are common and are widely distributed throughout many stratigraphic successions (red circles in Figures 3.5, 3.6; compiled from data in Smith et al., 1988; Jakobs, 1992; Jakobs et al., 1994; Jakobs, 1995; Smith and Tipper, 1996; Jakobs and Smith, 1996; and Jakobs, 1997). In comparison, Pliensbachian and Toarcian foraminiferal faunas are much less well understood. They have only been documented from two areas of western North America, 57 Haida Gwaii and northern Alaska (green circles in Figures 3.5; Figure 3.6; compiled from data in Tappan, 1955; Kottachchi et al., 2002, 2003). Ammonoids were pelagic organisms whereas Early Jurassic foraminifera had a benthic mode of life. Planktonic foraminifera did not evolve until the Middle Jurassic (Hart, 1999; Hart et al., 2002). Therefore, together, these two groups occupied many parts of the marine ecosystem and were subjected to changes in environment (including bottom-water anoxia) that may have occurred in both the water column and on the sea floor. They therefore play a key role in understanding the extinction event, which is a primary focus of the study presented herein. 58 Figure 3.5A\u00E2\u0080\u0093C - Maps showing the location of previously published Pliensbachian and Toarcian stratigraphic sections in western North America. Sections contain occurrences of ammonites (red circles) and foraminifera (green circles) that are used in this study. Locality numbers 1\u00E2\u0080\u009322 refer to the section correlation chart (Figure 3.6). YK = Yukon, NWT = Northwest Territories, BC = British Columbia, AK = Alaska, OR = Oregon, NV = Nevada, WA = Washington, MT = Montana. Figure 3.6 - Pliensbachian and Toarcian stratigraphic sections in western North America. Sections contain occurrences of ammonites and foraminifera that are used in this study to compile stratigraphic range charts. SB#3 = South Barrow #3 well core, HC = Hicks Creek, CC = Camp Creek, YR Sec. = Yakoun River Section. Occurrences of foraminifera were placed in context of the Pliensbachian and Toarcian ammonite zone schemes in order to develop comparable 59 stratigraphic range charts, which seems to be the first reported attempt of this type of integration in Early Jurassic successions of western North America. Foraminifera were only used from sections that contain coeval ammonite faunas. In the case of Early Jurassic foraminiferal occurrences from Haida Gwaii, Pliensbachian and Toarcian successions could only be used if previous accounts had established a zone-level ammonite biostratigraphy for the section (green circles in Figures 3.5, 3.6). In northern Alaska, foraminifera are known primarily from a variety of well cores originally extracted from the National Petroleum Reserve on the North Slope (Tappan, 1955; Bergquist, 1966). Within these well cores, co-occurring Pliensbachian\u00E2\u0080\u0093Toarcian ammonite and foraminifera faunas are exceedingly rare with the exception of a single well core known as the United States Navy South Barrow #3 Core (identified herein as SB #3 core). The SB #3 core contains a diverse foraminifera fauna that is temporally constrained to the Late Pliensbachian\u00E2\u0080\u0093Early Toarcian part of the ammonoid zone scheme for the high Arctic of Zakharov et al. (2006) Nikitenko and Mickey (2004) and Nikitenko et al. (2008). Stratigraphic ranges of foraminifera from the SB #3 core (orange bars in Figures 3.1\u00E2\u0080\u00933.4) are calibrated with the ammonite zone scheme of western North America (Figure 2.1). Open nomenclature taxa (Bengston, 1988) are included, herein, in the stratigraphic range charts of this study. The term \u00E2\u0080\u009Caff.\u00E2\u0080\u009D denotes a new species that could not be named due to a small sample size, stratigraphic uncertainties, or other ambiguities. The term \u00E2\u0080\u009Ccf.\u00E2\u0080\u009D, short for \u00E2\u0080\u009CConfer\u00E2\u0080\u009D, is used to denote a 60 tentative or provisional level of identification, mostly used when preservation is poor. The terms \u00E2\u0080\u009Csp. A\u00E2\u0080\u009D, \u00E2\u0080\u009Csp. B\u00E2\u0080\u009D, \u00E2\u0080\u009Csp. C\u00E2\u0080\u009D (etc.) are used to indicate separate but as yet unnamed species within a particular genus, while \u00E2\u0080\u009Cspecies indet.\u00E2\u0080\u009D refers to an indeterminate species of a particular genus. The term \u00E2\u0080\u009Cex gr.\u00E2\u0080\u009D or \u00E2\u0080\u009Cgr.\u00E2\u0080\u009D refers to \u00E2\u0080\u009Cof the group\u00E2\u0080\u009D and, lastly, a question mark is used to denote uncertainty, mostly at the genus level. Despite the uncertainty, these somewhat tentatively identified species are nonetheless included within the stratigraphic range charts. This is largely because they represent taxa that have specific stratigraphic ranges and may in- fact be an individual (new?) species that could be separated from the well-known species with which it is being compared. They should be included in diversity and rate estimates. When plotting stratigraphic ranges of foraminiferal species (Figures 3.1\u00E2\u0080\u0093 3.4), updated taxonomic identifications of foraminifera were incorporated into the stratigraphic range charts. Systematic descriptions of foraminiferal species from Haida Gwaii were not provided in Kottachchi (2001) or Kottachchi et al. (2002; 2003). Consequently, the taxonomic relationships of this fauna to the coeval fauna from northern Alaska (e.g. Tappan, 1955) are not certain. However, updated taxonomic identification by Nagy and Johansen (1991) and Nikitenko and Shurygin (1992) re-assign certain species from northern Alaska (in Tappan, 1955) to different genera. This re-assignment is maintained here, not only within the foraminiferal fauna from northern Alaska, but also with similarly identified species from Haida Gwaii. Updated taxonomic identifications (re-assigned taxa) 61 include: Kutsevella barrowensis, Laevidentalina pseudocommunis, Grigelis apheilolocula, Reussoolina aphela, Ammovertellina irregularis, and Ammoglobigerina canningensis, Ammodiscus asper, Reophax metensis, Ammobaculites lobus and Ammodiscus siliceus. 3.1.1 Diversity measurements Ammonoid and foraminiferal species diversity is measured using methodology presented in Foote (2000) as well as Hammer and Harper (2006). Diversity is measured from the recently compiled species level stratigraphic range charts for the Pliensbachian and Toarcian stages (Figures 2.2\u00E2\u0080\u00932.4; 3.1\u00E2\u0080\u0093 3.4), first at the zone level and then re-measured with an informally divided zone scheme where zones were grouped into three intervals (lower, middle, and upper; Figure 3.7A). Using these intervals, the stratigraphic range of each species was divided into one of four mutually exclusive categories (Figure 3.7B) following Barry et al. (1995) and Foote (2000). These include: 1) taxa whose first and last appearance are confined to the interval (Nfl), 2) taxa that cross the bottom boundary and disappear within the interval (Nbl), 3) taxa that appear within the interval and cross the upper boundary (Nft), and 4) taxa that range through the entire interval and cross both stratigraphic boundaries (Nbt). Taxa were then assigned a unitary weight depending on their stratigraphic range within the interval (Sepkoski, 1975; Foote, 2000; Hammer, 2003; Hammer and Harper, 2006). Taxa that range through the interval (Nbt) were counted as one unit each, taxa that disappeared (Nbl) or appeared (Nft) within the interval were only counted as a half of a unit, and taxa that were confined to the interval (Nfl) were counted as a third of a unit (Hammer, 2003). Hammer and Harper (2006) justify these 62 unitary weights by the rational that the first appearance (Nft) and last appearance (Nbl) datum are uniformly distributed throughout the measured interval and that the average portion of the interval length that is occupied by both of these types of single-ended taxa is approximately 0.5. Similarly, taxa that are confined to the interval (Nfl) are thought to occupy approximately one-third of the interval length and therefore should be counted as such (Hammer, 2003; Hammer and Harper, 2006). Categories were then totaled with respect to each stratigraphic interval (zone or informally divided zonal unit) and then analyzed using various metrics to calculate taxonomic diversity and extinction/origination rates (Figure 3.7C). Lastly, the term singleton has been used to denote either: 1) species that are represented by a single specimen (Buzas and Culver, 1994; 1998) or 2) taxa that are confined to a single stratigraphic interval, essentially any species that is in the Nfl category (Foote, 2000). In this study, the definition of \u00E2\u0080\u0098singleton taxa\u00E2\u0080\u0099 by Foote (2000) is used, suggesting taxa that that are confined to a particular interval. 63 Figure 3.7 - A) Example of stratigraphic ranges of hypothetical species, divided at the zone- and informal subzone-levels, used in this study for measuring taxonomic diversity and extinction/origination rates. Taxa from within each informal subzone are identified as being from the \u00E2\u0080\u0098lower\u00E2\u0080\u0099, \u00E2\u0080\u0098middle\u00E2\u0080\u0099, or \u00E2\u0080\u0098upper\u00E2\u0080\u0099 part of the zone. B) Four fundamental classes of taxa (after Foote, 2000) used in this study to quantify stratigraphic ranges of species in western North America. C) Definitions of diversity and rate measures used in this study (after Foote, 2000). O = Origination rate, E = Extinction rate. 3.2 Geochemical analysis Over the 2008\u00E2\u0080\u00932010 field seasons, samples were collected for stable isotope geochemistry (\u00CE\u00B413Corg, \u00CE\u00B413Cwood, \u00CE\u00B415Norg) total organic carbon, and total nitrogen analysis from two well studied and temporally constrained stratigraphic sections on Haida Gwaii (British Columbia, Canada) and the SB #3 core (Figure 64 3.5). On Haida Gwaii, samples were collected from Whiteaves Bay on northern Moresby Island and along the Yakoun River on central Graham Island (Figure 3.5C1,C2). In northern Alaska (Figure 3.5A), the SB #3 core has been split (lengthwise) with half of the split being stored at the United States Geological Survey (USGS) in Denver, Colorado and the other half stored at the Geological Materials Center (GMC) in Eagle River, Alaska. The split that is stored at the GMC in Eagle River was sampled in the current study. In the field on Haida Gwaii, each section was measured and samples were collected at regular stratigraphic intervals ranging from 5cm to 40cm spacing. Co-occurring ammonites were also collected. Samples were prepared at UBC for isotopic analysis in the paleontology laboratory using methods that are consistent with the Stable Isotope Biogeochemistry Laboratory (SIBL) protocol at Durham University (UK). All samples were air-dried in their sample bag with the top opened. Shale samples were prepared separately from wood samples using slightly different preparation methods. Shale samples were ground to a powder and homogenized using an agate mortar and pestle and then placed into glass 25ml screw-top vials according to sample number. In between samples, the agate mortar and pestle were ground using quartz sand and wiped clean using acetone to avoid cross contamination of samples. The powdered sample was then placed into a 50ml centrifuge tube with enough powder to cover the conical portion of the tube. Three Molar HCl was added up to the 45ml mark on each tube and left to decalcify for ~8 hours (or overnight). To avoid overflow, HCl was sometimes 65 added slowly if a high concentration of calcite was present. After the HCL was fully added, the lid of the centrifuge tube was tightened and the vial shaken by hand until all of the sediment was suspended in solution. The lid was slowly uncapped to allow the CO2 gas to escape, and then this step was repeated. The lid was then loosened and left slightly ajar overnight. After ~8 hours or overnight, the 3M HCl supernatant was decanted, the centrifuge tube refilled with de- ionized water (thus rinsing the sample), and then the sample was centrifuged at ~3000 rpm for 6 minutes. This rinsing step was performed four times with the de- ionized water being decanted off after each centrifuge step. After the last rinse, the sample was placed in a drying oven set at 60\u00C2\u00BAC for 24 hours or until dry. Once dry, the sample was then re-ground to a powder using an agate mortar and pestle and cleaned in-between samples with quartz sand and acetone (same as above). The decalcified, powdered sample was then transferred to a small 1- dram glass vial and transported to the SIBL laboratory in Durham (UK) for stable isotope analysis. Preparation of wood samples involved a slightly different methodology in comparison to the shale. After the previously described drying step, fossil wood was separated from the surrounding sediment using tweezers under a dissecting microscope. The wood sample was ground using an agate mortar and pestle, which was cleaned in between each sample using the same methodology as above. Once ground and homogenized, the wood sample was placed into a 15ml centrifuge tube with as much sample wood as possible, or until the sample reached the top of the conical portion of the tube. Three Molar HCl was then 66 added up to the 12ml in the tube, shaken (same as above), and then left to decalcify overnight. The supernatant was then decanted, the sample rinsed with de-ionized water, and centrifuged using the same methodology as above. Rinsing was completed four times and then the sample was placed into a drying oven, set at 60\u00C2\u00BAC for 24 hours. Once dry, the sample was re-ground (same as above) and then transferred to a small 1-dram glass vial and shipped to Durham (UK) for stable isotope analysis. At the SIBL laboratory each sample was weighed on a mass balance according to the approximated concentration of organic carbon in each sample. For this study, between 2\u00E2\u0080\u00934 mg of powder was weighed per decalcified shale sample and < 1 mg of powder was weighed per decalcified wood sample. Weighed samples were then placed into tin sample cups and analyzed by mass spectrometry to obtain \u00CE\u00B413C, total organic carbon, \u00CE\u00B415N and total nitrogen values. Anomalous isotope values were re-analyzed if they were either very different (more negative or positive) in comparison to stratigraphically adjacent values or were considered suspect from any technical/mechanical problems that may have occurred in the analysis. 67 Chapter 4 Regional geology The Cordilleran region of western North America (Figure 4.1) consists of allochthonous, fault-bounded, tectonostratigraphic terranes that are composed of sedimentary, igneous and metamorphic rock units (Coney et al., 1980; Plafker and Berg, 1994; Miller et al., 2002; Haggart et al., 2006; Ridgway et al., 2007; Blodgett and Stanley, 2008; Colpron and Nelson, 2009). Currently it is believed that many of the more inboard terranes (e.g. Yukon\u00E2\u0080\u0093Tanana, Quesnellia, and Stikinia) originated in the mid Paleozoic along the western margin of western North America as a series of rifted continental fragments (Colpron and Nelson, 2009 and references therein). This differs from the more outboard terranes (e.g. Wrangellia, Alexander, Peninsular, and Arctic Alaska\u00E2\u0080\u0093Chukotka) containing Early Paleozoic and older basement sequences that are thought to have originated far from western North America in a variety of geographically distant areas that include Baltica, Siberia, and the northern Caledonides (Colpron and Nelson, 2009 and references therein). Although the timing of terrane accretion to the western margin of North America has been debated, it is currently believed that many of the displaced terranes of the North American Cordillera began to collide with the western margin of North America in the Late Paleozoic as a series of belts that continued accreting throughout the Late Cretaceous (Coney et al., 1980; Plafker and Berg, 1994; Trop et al., 2002; Colpron and Nelson, 2009). 68 Figure 4.1 - Tectonic map of western North America showing major fault systems and terrane boundaries (modified from Colpron and Nelson, 2009; Beranek, 2009). Terranes that are discussed in this study are coloured. Boundaries of the Peninsular terrane are from Trop et al. (2002; 2005). Red boxes refer to areas discussed herein. 69 Of particular importance is the western-most outboard crustal fragment known as the Wrangellia composite terrane (Plafker and Berg, 1994; Nokleberg et al., 1994) and the northern-most Arctic Alaska\u00E2\u0080\u0093Chukotka terrane (Till and Dumoulin, 1994; Amato et al., 2004; 2009). The Wrangellia composite terrane (Figure 4.1) is a subcontinent sized crustal fragment that consists of three large tectonostratigraphic terranes, which include the Wrangellia, Alexander, and Peninsular terranes (Plafker and Berg, 1994; Nokleberg et al., 1994). The Arctic Alaska\u00E2\u0080\u0093Chukotka terrane, also illustrated on Figure 4.1, encompasses much of present-day northern Alaska (Brooks Range and Seward Peninsula), Russia (Chukotka Peninsula), East Siberia and Bering continental shelves (Amato et al., 2004; 2009; Colpron et al., 2007). During the Pliensbachian and Toarcian, it is thought that the Wrangellia composite terrane was located in the northeast paleo Pacific Ocean at a paleolatitude that is comparable with the modern US / Canada border (Smith, 2006) and the Arctic Alaska\u00E2\u0080\u0093Chukotka terrane was located adjacent to Laurentia close to its present-day position in the high Arctic (Miller et al., 2006; Amato et al., 2009). 4.1 Haida Gwaii The Haida Gwaii, are located off mainland British Columbia, Canada in the eastern Pacific Ocean (Figure 3.5C). This group of islands is part of Wrangellia, which is currently distributed along the western margin of the North American Cordillera from Vancouver Island to southern Alaska (Jones et al., 1977; Coney et al., 1980). It has been informally divided into two portions based on modern geographical placement (Green et al., 2010): 1) a northern portion, which 70 includes The Wrangell Mountains (Alaska) and parts of the Yukon (Canada) and 2) a southern portion, which includes Vancouver Island and Haida Gwaii (British Columbia). Wrangellia was originally defined as a fault-bounded section of upper crust containing correlative successions of Middle\u00E2\u0080\u0093Late Triassic flood basalt with overlying Late Triassic carbonate that, altogether, overlie Paleozoic units (Jones et al., 1977; Greene et al., 2010). However, it has been determined that deposition began in the Devonian and continued intermittently throughout the Cretaceous until its collision with Western North America by Latest Cretaceous time (Csejtey et al., 1982; McClelland et al., 1992; Nokleberg et al., 1994; Plafker and Berg, 1994; Hillhouse and Coe, 1994; Trop et al., 2002; Umhoefer and Blakey, 2006; Trop and Ridgway, 2007; Greene et al., 2010). On Haida Gwaii, Upper Triassic to Lower Jurassic strata (Figure 4.2) were deposited continuously in a backarc, deep water, environment that produced the Karmutsen Formation (not included in Figure 4.2) overlain by the Kunga and Maude Groups (Andrew and Goodwin, 1989; Thompson et al., 1991; Tipper et al., 1991). The Maude Group encompasses the Pliensbachian\u00E2\u0080\u0093Aalenian part of the stratigraphy and is composed predominantly of volcaniclastic and pyroclastic sediments that were deposited at varying water depths in a shelf environment (Cameron and Tipper, 1985; Jakobs, 1997). In ascending stratigraphic order the Maude Group includes the Ghost Creek, Fannin, Whiteaves, and Phantom Creek formations (Figure 4.2). Outcrops of Maude Group sedimentary rocks are best exposed along the intertidal zone of the Skidegate Inlet area (Cameron and 71 Tipper, 1985), as well as along many road- and stream-cuts of the Yakoun River drainage in central Graham Island (Figure 3.5C1, C2). Figure 4.2 - Lower\u00E2\u0080\u0093Middle Jurassic stratigraphy and correlative ammonite zonal scheme for Pliensbachian\u00E2\u0080\u0093Toarcian units on Haida Gwaii, British Columbia (Modified from Jakobs and Smith, 1996; Jakobs, 1997). Pliens. = Pliensbachian, Yakoun = Yakoun Group, R.J.M. = Rennell Junction Member. 4.1.1 Lithostratigraphy The Ghost Creek Formation is an approximately 60 m thick, recessive dark grey/black, organic-rich, shale and silty shale unit that contains abundant 72 pyrite and occasional sandstones, nodular limestones and rare ash or tuffaceous interbeds (Cameron and Tipper, 1985; Smith and Tipper, 1996). It rests conformably on the Sandilands Formation of the Kunga Group (Figure 4.2; Cameron and Tipper, 1985) and has been temporally constrained to the Imlayi and Whiteavesi Zones by Smith and Tipper (1996). The Ghost Creek Formation contrasts with the underlying Sandilands Formation in being softer with more shale, contains less volcanic material, and lacks lamination (Cameron and Tipper, 1985; Smith and Tipper, 1996). In the Skidegate Inlet area, the basal part of the Ghost Creek Formation shows highly contorted beds, calcite veins, and shale rip-up clasts which Cameron and Tipper (1985) suggest are the result of higher energy deposition but do not specify a specific environment of deposition. Near the top of the formation, in areas of Central Graham Island, the Ghost Creek Formation is more regularly bedded and contains glauconite which is interpreted as evidence for a slower rate of sedimentation (Cameron and Tipper, 1985). Benthic fossils and bioturbation are uncommon, but are noted to increase in abundance near the upper part of the unit (Cameron and Tipper, 1985; Smith and Tipper, 1996). Lying conformably above the Ghost Creek Formation is the Early Pliensbachian (Whiteavesi Zone) to Early Toarcian (Kanense Zone) Fannin Formation (Figure 4.2; Tipper et al., 1991; Smith and Tipper, 1996; Jakobs, 1997). Original description of this sequence by Cameron and Tipper (1985) identifies a lower unit called the Rennell Junction Formation. However, this was subsequently redefined as the Rennell Junction Member of the Fannin Formation 73 because the unit was unmappable (Tipper et al., 1991; Smith and Tipper, 1996). The Rennell Junction Member is described as a fine grained sandstone, siltstone, and shale unit that contains irregular limestone beds and nodules and ranges in thickness from 15 to 40 m (Smith and Tipper, 1996). In the Skidegate Inlet area, the base of the Rennell Junction Member occurs at the base of the Fannin Formation. The contact of the Rennell Junction Member and the overlying upper Fannin Formation occurs near the Freboldi / Kunae Zone boundary (Smith and Tipper, 1996). The upper part of the Fannin Formation is generally coarser-grained, more resistant, and contains thicker beds of crossbedded sandstone that are occasionally convoluted. The upper Fannin Formation also contains limestone interbeds, chamosite ooliths, locally developed concretions, tuffs, breccias, and conglomerate lenses (Cameron and Tipper, 1985; Smith and Tipper, 1996). The Fannin Formation also contains a fairly diverse benthic and pelagic fauna including: ammonites, nautiloids, bivalves, brachiopods, radiolarians, foraminifera, and bioturbation (Cameron and Tipper, 1985; Smith and Tipper, 1996; Kottachchi et al., 2002). The Early to Late Toarcian (Kanense\u00E2\u0080\u0093Hillebrandti Zone) Whiteaves Formation is a recessive unit that conformably overlies the Fannin Formation (Figure 4.2; Jakobs, 1997). It consists predominantly of grey/green medium to fine grained siltstone and mudstone with medium to fine grained sandstone interbeds of variable thickness, zircon-bearing ash beds, pale grey/light brown calcareous nodules, and concretionary limestones (Cameron and Tipper, 1985; 74 Jakobs, 1997). The abundance of siltstone and mudstone throughout the succession and no evidence of transport indicators within sandstone interbeds, suggest the Whiteaves Formation was deposited slowly in a moderately deep- water environment (Cameron and Tipper, 1985). The presence of glauconite in some sandstones suggest a depositional paleo-depth between 200\u00E2\u0080\u0093300 m (Cameron and Tipper, 1985 and references therein). However, the occasional local presence of fragmented bivalve shells in small sandy pockets suggests a minor degree of intermittent higher energy re-sedimentation (Cameron and Tipper, 1985) which was not observed in stratigraphic sections of the Whiteaves Formation within the study presented herein. The Late Toarcian (Yakounensis Zone) to Aalenian Phantom Creek Formation rests conformably on the Whiteaves Formation and is separated from the overlying Yakoun Group by an angular unconformity (Figure 4.2; Cameron and Tipper, 1985; Jakobs, 1997). It is an approximately 30 m thick, resistant, brown to buff weathering, fine\u00E2\u0080\u0093coarse grained, sandstone unit that contains partly calcareous and common argillaceous grains, thin shale interbeds in its lower part, massively bedded sandstone units in its upper part, and a rich fossil fauna throughout that includes ammonites, bivalves and belemnites (Cameron and Tipper, 1985). It has been informally subdivided into two members that are separated by an erosional hiatus (Cameron and Tipper, 1985; Jakobs, 1990; Jakobs, 1997). The lower member is known as the coquinoid sandstone member (Cameron and Tipper, 1985; Jakobs, 1997). It is a 3-4 m thick, well-bedded, grey 75 to greenish-brown sandstone unit that contains an abundant bivalve and ammonite fauna (Cameron and Tipper, 1985). Fossils within this unit are tightly packed, randomly oriented, and commonly broken. Overlying this unit is the belemnite sandstone member (Cameron and Tipper, 1985; Jakobs, 1997). It is an approximately 20 m thick unit composed of massively bedded, carbonaceous, brown weathering sandstone. This unit is characterized by an abundance of belemnites, pectinoid bivalves, gastropods and nautiloids, with rare occurrences of ammonites (Cameron and Tipper, 1985; Jakobs, 1997). The Phantom Creek Formation is interpreted as being deposited in a high-energy shallow water environment based on: 1) the disappearance of shale in its lower part, 2) the appearance of massively bedded coarse-grained sandstone in its upper part, and 3) appearance of coquinas that are densely packed, randomly oriented and contain highly fragmented fossils (Cameron and Tipper, 1985; Jakobs, 1997). 4.2 Northern Alaska The Arctic Alaska\u00E2\u0080\u0093Chukotka terrane (Figure 4.1) includes many different fragments of the present-day circum-Arctic region and is generally composed of continental rocks in northern Alaska and northeastern Russia (Moore et al., 1994; Amato et al., 2004; 2009; Miller et al., 2006; Colpron et al., 2007). It is an extremely large terrane that covers an area of approximately 3 million km2 which is more than 85% the area of Greenland (Amato et al., 2009). It is bounded to the North by the edge of the Arctic Alaskan outer shelf, and to the south by arc and ophiolite rocks of the southern Brooks Range (Moore et al., 1994; Miller et al., 2006). 76 The northern Alaska portion of the Alaska\u00E2\u0080\u0093Chukotka terrane lies beneath much of the present-day continental shelves of the U.S. Chukchi and Beaufort Seas and the North Slope of Alaska (Hubbard et al., 1987a,b; Sherwood et al., 2002; Houseknecht and Bird, 2004). It is comprised of sedimentary sequences from two connected basins, the Hanna Trough and Arctic Alaska Basin, that rest on a basement sequence of highly deformed Devonian and older crystalline rocks known as the Franklinian sequence (Moore et al., 1994; Sherwood et al., 2002; Miller et al., 2006; Amato et al., 2009). Today, these basins are bordered by the Barrow arch to the north, Brooks Range to the south, Herald Thrust to the West, and continental Canada (Laurentia) to the East (Figures 3.5A and 4.1). In northern Alaska, deposition of the Alaska\u00E2\u0080\u0093Chukotka terrane began in the Late Devonian (?) and continued intermittently until the Late Jurassic / Cretaceous, with sequences of marine carbonate and nonmarine\u00E2\u0080\u0093shallow marine siliciclastic strata that are approximately 6 km thick (Hubbard et al., 1987a,b; Bird, 1988). This sequence is widely known as the Ellesmerian sequence (Moore et al., 1994; Moore et al., 1997; Natal\u00E2\u0080\u0099in et al., 1999; Dumoulin et al., 2002; Sherwood et al., 2002; Amato et al., 2003; Houseknecht and Bird, 2004; Miller et al., 2006; Amato et al., 2009) and is thought to represent rift- (lower Ellesmerian) and subsidence- (upper Ellesmerian) phases of basin development (Sherwood et al., 2002). During the Jurassic (Late Jurassic in Sherwood et al. (2002) and Early\u00E2\u0080\u0093Middle Jurassic in Hubbard et al. (1987a,b)) the northeast part of the basin experienced a rifting event that was associated with the opening of the Arctic Ocean. This rifting is known as the Beaufortian sequence and it is thought to 77 have covered the Ellesmerian rocks with thick deposits of rift-related clastics, sandstones, siltstones and shales (Hubbard et al. 1987a,b; Sherwood et al., 2002; Houseknecht and Bird, 2004). Subsurface sediments in the South Barrow area of northern Alaska contain strata of the Beaufortian sequence (Figure 4.3) and were sampled in this study. Several other studies have addressed the paleogeographic history of the Arctic region (Lawver and Scotese, 1990; Blodgett et al., 2002; Lawver et al., 2002; Miller et al., 2006; Amato et al., 2009), some suggest that during the Neoproterozoic, the Arctic Alaska\u00E2\u0080\u0093Chukotka terrane was located in a geographic position northeast of Laurentia and had faunal ties with Siberia and Laurentia by the Early Ordovician (Blodgett et al., 2002; Amato et al., 2009). Other studies have suggested that sediments of the Ellesmerian and Beaufortian sequences were derived from an exposed landmass to the North, and were then subsequently deposited on a south-facing continental shelf (Moore et al., 1994; Sherwood et al., 2002; Houseknecht and Bird, 2004). However, it is thought that by the Late Paleozoic\u00E2\u0080\u0093Jurassic, the Arctic Alaska\u00E2\u0080\u0093Chukotka terrane as a whole was located in the high Arctic (above 50\u00C2\u00B0N) near its present location (Amato et al., 2009). 78 Figure 4.3 - Schematic diagrams showing a summary of the inferred depositional sequence of the Kingak Shale through the National Petroleum Reserve in Alaska (modified from Houseknecht and Bird, 2004). A) shows the litho- and chronostratigraphy of the various formations that comprise the Kingak Shale and B) depicts a generalized South\u00E2\u0080\u0093North cross section of the strata through the Barrow Arch. 4.2.1 Lithostratigraphy The Barrow area of northern Alaska contains a thick sequence of Beaufortian rock generally identified as the Kingak Shale (Houseknecht and Bird, 2004 and references therein). It is a geographically extensive unit that is described from many surface outcrops along the North Slope and in subsurface 79 wells of the National Petroleum Reserve (NPRA) of northern Alaska. As illustrated by Houseknecht and Bird (2004), this all-inclusive unit contains three formations of Jurassic\u00E2\u0080\u0093Cretaceous age (Figure 4.3) and is stratigraphically confined by the Shublik and Sag River Formations (below) and a geographically extensive \u00E2\u0080\u0098pebble shale\u00E2\u0080\u0099 unit (above). The Kingak Shale is truncated to the North by a Lower Cretaceous unconformity which is thought to represent, among others, erosion of the unit by the uplift of the Barrow arch (Houseknecht and Bird, 2004). The formally defined Kingak Formation is the basal unit of the Kingak Shale and contains upper and lower units that are separated by a disconformity (Figure 4.3). The Lower Kingak Formation includes sedimentary sequences that are of Early Jurassic (Hettangian\u00E2\u0080\u0093Toarcian) age, while the Upper Kingak Formation contains mainly Late Jurassic (Oxfordian\u00E2\u0080\u0093Kimmeridgian) deposits (Imlay, 1981; Sherwood et al., 2002; Houseknecht and Bird, 2004; Mickey et al., 2006). The Lower Kingak Formation is a 900\u00E2\u0080\u00931200 m thick fissile, light to dark grey, argillaceous siltstone and claystone with minor sandstone interbeds and glauconite (Collins, 1961; Imlay, 1981; Sherwood et al., 2002; Houseknecht and Bird, 2004). Temporal constraint of the Lower Kingak Formation is provided mainly through ammonite and foraminifera that were collected in subsurface well cores within the National Petroleum Reserve (Imlay, 1955; 1981; Tappan, 1955; Bergquist, 1966). However, collections of ammonites from surface outcrops along the North Slope have also provided temporal constraint for the formation (Imlay, 1955; 1981; Smith et al., 2001). 80 4.3 Southern Alaska The Peninsular terrane of southern Alaska (Figure 4.1) is one of the largest outboard tectonostratigraphic terranes in western North America (Coney et al., 1980; Wilson et al., 1985; Wang et al., 1988). It comprises much of the modern-day Alaskan Peninsula extending ~1200 km throughout the Talkeetna Mountains and is bordered by the Chugach terrane to the southeast and the Wrangell terrane to the east (Csejtey et al., 1978; Jones and Silberling, 1979; Csejtey and St. Aubin, 1981; Wilson et al., 1985). Sequences deposited on the Peninsular terrane are thought to have initially accumulated in the paleo Pacific Ocean far from the North American margin (Wilson et al., 1985; Wang et al., 1988; Plafker et al., 1989, 1994; Clift et al., 2005a,b; Blodgett and Stralla, 2006; Rioux et al., 2007). Deposition began in a shallow-water, tropical, backarc environment during Permian\u00E2\u0080\u0093Triassic time (Wang et al., 1988; Blodgett and Stralla, 2006) and shifted toward an intraoceanic, volcanic island arc-type environment throughout the Early Jurassic and into the early Middle Jurassic (Barker and Grantz, 1982). Island arc related magmas have recently been dated using U\u00E2\u0080\u0093Pb zircon and whole-rock isotope analysis indicating that magmatism occurred primarily between 202.1 and 181.4 Ma (Clift et al., 2005a,b; Rioux et al., 2007). By Middle\u00E2\u0080\u0093Late Jurassic time, it is thought that the depositional environment had shifted toward a forearc including deep-water fan-delta deposition, shelf sedimentation, post-depositional uplift, and strike-slip displacement (Trop et al., 2005). Current tectonic models postulate that the Peninsular terrane was amalgamated with Wrangellia and the Alexander terrane, forming the Wrangellia 81 composite terrane, which then accreted to the western margin of North America (Wang et al., 1988; Plafker et al., 1989; Ridgway et al., 2002; Trop et al., 2005). However, the timing of amalgamation versus the timing of accretion is a complex issue that is still unclear. Originally it was thought that the composite terrane was assembled by Triassic time (Wang et al., 1988; Plafker et al., 1989) and accreted to the continental margin of western North America between the Late Jurassic and Late Cretaceous (Csejtey et al., 1982; Jones et al., 1982; Pavlis, 1982; McClelland et al., 1992; Cole et al., 1999; Ridgway et al., 2002; Trop et al., 2002). Other studies have indicated that collision of the composite terrane with western North America may have occurred as early as Middle Jurassic\u00E2\u0080\u0093Early Cretaceous time (McClelland and Gehrels, 1990; McClelland et al., 1992; van der Heyden, 1992; Kapp and Gehrels, 1998; Gehrels, 2001). More recent work concludes that Late Jurassic deformation and synorogenic sedimentation could reflect either the initial collision of Wrangellia and the Peninsular terranes or their collision with western North America (Trop et al., 2005). 4.3.1 Lithostratigraphy There are no known crystalline basement rocks underlying the Peninsular terrane (Wang et al., 1988; Blodgett and Stralla, 2006). Previous work has shown that the oldest strata crop out at Puale Bay along the Alaska Peninsula. They consist of a small (~11 m thick) unnamed unit of volcanic agglomerate, volcaniclastic sandstone, and fossiliferous limestone that is thought to be of Middle Permian age (Hansen, 1957; Jeffords, 1957; Wang et al., 1988; Blodgett and Stralla, 2006). Lying unconformably above the Permian unit is the Upper Triassic (Norian) Kamishak Formation (Capps, 1923; Smith, 1926; Silberling in 82 Detterman et al., 1985; Wang et al., 1988; Newton, 1989; Silberling et al., 1997; P\u00C3\u00A1lfy et al., 1999; Blodgett and Stralla, 2006; Blodgett, 2008). This unit crops out throughout much of the Alaska Peninsula and comprises ~700 m of shallow fossiliferous marine and biogenic carbonate interbedded with volcanic rocks in its upper part (Wang et al., 1988; Blodgett, 2008). The Early Jurassic Talkeetna Formation rests conformably on the Late Triassic Kamishak Formation (Newton, 1989; P\u00C3\u00A1lfy et al., 1999). The Talkeetna Formation is a widespread unit that is exposed in many outcrops and cores throughout the Alaska Peninsula extending northeast to the Talkeetna Mountains (Barker and Grantz, 1982; Nokleberg et al., 1994; Sandy and Blodgett, 2000). It is primarily composed of volcanic (flows and pyroclastics) rocks with minor interbedded marine sedimentary sequences (volcaniclastics, sandstones, siltstones, and mudstones) that reach a thickness of ~3 km (Barker and Grantz, 1982; Nokleberg et al., 1994; Trop et al., 2005). The Talkeetna Formation is thought to be Hettangian\u00E2\u0080\u0093Late Toarcian in age (Imlay and Detterman, 1973; Imlay, 1981; Newton, 1989) with zone-level age constraints published for a Hettangian\u00E2\u0080\u0093Sinemurian sequence exposed in Puale Bay (P\u00C3\u00A1lfy et al., 1999). For the Talkeetna Mountain sequences there are currently no age constraints at the zone-level. Imlay (1981) indicates the presence of Late Sinemurian\u00E2\u0080\u0093Late Toarcian ammonites from several isolated localities but does not provide detailed stratigraphic distributions. Stratigraphically above the Talkeetna Formation are the Middle Jurassic\u00E2\u0080\u0093 Late Cretaceous Tuxedni (Bajocian-Bathonian), Chinitna (Bathonian-Callovian), 83 Naknek (Oxfordian-Kimmeridgian), Nelchina limestone (Valanginian-Hauterivian), and Matanuska formations (Albian-Maastrichtian) representing deposition in a shallow water marine shelf, forearc environment (Grantz, 1960a,b; Winkler, 1992; Trop et al., 2005). The lithology of these units is quite variable in that the Tuxedni and Chinitna formations are composed primarily of mudstone and fine-grained sandstone; the Naknek Formation of coarse-grained cobble conglomerate, sandstone, and mudstone; the Nelchina limestone of massively bedded limestone, and the Matanuska Formation of sandstone, conglomerate and mudstone (Trop et al., 2005 and references therein). The coarse-grained cobble conglomerate lithofacies of the Upper Jurassic Naknek Formation is thought to record collision of the Peninsular terrane with either adjacent Wrangellia or the North American margin (Trop et al., 2005 and references therein). 4.3.1.1 Biostratigraphy of the Hicks and Camp Creek sections Herein, 64 Late Pliensbachian ammonoid specimens are described from two sections of the Talkeetna Formation in the southeastern Talkeetna Mountains, Anchorage D-3 quadrangle (number 2 in Figures 3.5B, 3.6) representing the Kunae and Carlottense Zones of the Late Pliensbachian (Figure 4.4). Biostratigraphy of these two stratigraphic sections helped supplement previously known stratigraphic ranges of Late Pliensbachian ammonoids in western North America (Figures 2.2\u00E2\u0080\u00932.4, 3.1\u00E2\u0080\u00933.4) that were utilized in the extinction analysis, and also provided several new occurrences in the coeval strata of Alaska; including the species Fanninoceras (Charlotticeras) cf. maudense, Leptaleoceras? sp., and Lioceratoides (Lioceratoides) cf. L. involutum (Figures 4.4 and 4.5). 84 Figure 4.4 - Lithostratigraphy and biostratigraphy of the Late Pliensbachian Camp Creek (A) and Hicks Creek (B) sections in the Talkeetna Mountains, southern Alaska (after Caruthers and Smith, 2012). Ammonites from the Camp Creek section are from the Kunae Zone, while ammonites recovered from the section along Hicks Creek are indicative of the Carlottense Zone. f.s. = fine sand; m.s. = medium sand; c.s. = coarse sand; peb. = pebble; cg. = conglomerate; cob. = cobble; volc. = volcaniclastic; CCA-1 = volcanic ash sample (Figure 4.6, this study). 85 4.3.1.1.1 Camp Creek section This section is located along an unnamed tributary of Camp Creek on the southeastern side of Gunsight Mountain in the Talkeetna Mountains (Figure 3.5B). It is a ~250 m thick sequence of the Talkeetna Formation that is truncated above and below by faulting (Figure 4.4A). It consists predominantly of medium\u00E2\u0080\u0093 coarse grained volcaniclastic sandstone with interbedded conglomerate and minor volcanic and pyroclastic rocks. Conglomerate facies include rounded and sub-rounded basalt and felsic magma clasts with rare occurrences of chert and sandstone. The basal 41 m consists of mostly siltstone with minor sandstone and conglomerate interbeds. Above 120 m, scoured and graded beds become more frequent, generally beginning with coarser-grained facies that grade into a fine sand or siltstone, suggesting deposition by turbidity currents. Low angle cross beds are recorded at ~37 and 89 m and oscillatory ripples at ~14 and 97 m. 86 87 Figure 4.5 - Distribution of Pliensbachian and Toarcian ammonite taxa that are systematically described in Chapter 7 of this study, according to locality number at the Hicks Creek and Camp Creek sections (Figure 4.4), the Whiteaves Bay and Yakoun River sections (Figures 5.4, 5.6) and the South Barrow #3 Core (Figure 5.11). Ch. = Chapter, A.A. = Arctic Alaska, and SB3 = South Barrow #3 well core. Previous work at this section suggested that the marine sedimentation that produced the Talkeetna Formation was restricted within the Sinemurian to Toarcian interval. This was based on the occurrence of the bivalves Weyla unca, Ostrea sp., and Cardinia sp., and several specimens of the brachiopod Callospiriferina tumida (Imlay, 1981; Sandy and Blodgett, 2000). Ammonites identified herein indicate that deposition of the Talkeetna Formation (at this section) occurred over a much shorter time interval than previously thought. Forty-five specimens of Late Pliensbachian ammonoid were identified, representing five genera that occur at ten localities distributed throughout the section (Figures 4.4A, 4.5). A Kunae Zone age determination is indicated by the presence of Leptaleoceras? sp., Fanninoceras (Charlotticeras) cf. maudense, and Amaltheus sp. from localities 6, 8, 13 and 14; these taxa are known to be restricted to this zone (Smith and Tipper, 1996). The presence of Arieticeras aff. domarense and Fanninoceras (Fanninoceras) fannini from localities 6\u00E2\u0080\u009315 are also indicative of the Kunae Zone, however, these taxa also rarely occur in the lower Carlottense Zone (Smith and Tipper, 1996). Although a Kunae Zone age is determinable for this section, its stratigraphic relationship to the overlying Carlottense Zone is uncertain due to a lack of age-specific ammonites from the 88 top of the section. A U\u00E2\u0080\u0093Pb TIMS age date has been obtained from an ash bed occurring near the top of the section at 242.2 m (CCA-1 on Figure 4.4A). The age of the ash bed, derived from the youngest cluster of zircon grains (F\u00E2\u0080\u0093J), is calculated to be 184.12 + 0.17 Ma (Figure 4.6A). Based on current time scale calibrations (P\u00C3\u00A1lfy et al., 2000), this indicates a late Kunae Zone age Figure 4.6B which is consistent with the ammonite fauna from the Camp Creek section. The older population of zircon grains (A\u00E2\u0080\u0093E), yielding a date of ~188 m.y., may consist of either xenocrysts or crystals with inherited cores. These zircon crystals are most likely related to arc-related magmatism of the Peninsular terrane that occurred over a time span of 202.1 to 181.4 Ma (Clift et al., 2005a,b; Rioux et al., 2007). 89 Figure 4.6 - A) U-Pb concordia diagram for ash sample CCA-1 in the Camp Creek section of the Talkeetna Mountains, Alaska (Figure 4.4A). Diagram shows 206Pb/238U TIMS ages for ten zircon grains showing two distinct age populations, one at ~188 Ma and the other at ~184 Ma. The age of magmatism is calculated to be 184.12 + 0.17 Ma, based on five zircon grains (F\u00E2\u0080\u0093J). The population of five zircon grains A\u00E2\u0080\u0093E at ~188 Ma consists of xenocrysts or crystals containing inherited cores. B) A comparative diagram showing the relationship of the calculated age for grains F\u00E2\u0080\u0093J of CCA-1 to the calibrated geochronologic Time Scale for the Kunae / Carlottense zonal boundary (P\u00C3\u00A1lfy et al., 2000). Figure from Caruthers and Smith (2012). 4.3.1.1.2 Hicks Creek section The section cropping out along Hicks Creek in the southern Talkeetna Mountains ~10 km from the Glenn Highway (Figure 3.5B) consists of ~205 m of fossiliferous marine volcaniclastic and volcanic rocks (Figure 4.4B). The lower 43 m consist of coarse-grained volcaniclastic sandstones, interbedded conglomerate lenses and minor siltstone. Conglomerate lenses contain rounded\u00E2\u0080\u0093 sub-rounded, pebble-sized clasts of basalt and other igneous rocks. Siltstones and sandy siltstones occur from 43\u00E2\u0080\u009387 m and are directly overlain by a thick (~60 m) unit of undifferentiated volcanic/volcaniclastic rocks. The succession is capped by ~58 m of coarse-grained sandstone and sandy siltstone. In general, the stratigraphic succession at Hicks Creek is lithology similar to that of Camp Creek but differs by the predominance of finer grained sedimentary rocks and the absence of scoured bedding surfaces and graded bedding. The fauna from the Hicks Creek section is taxonomically diverse and includes ammonoids, bivalves, brachiopods, gastropods, corals, plant fossils, fish scales and trace fossils. Bivalves and brachiopods are most abundant and diverse throughout the section, frequently occurring in coquinas but also as isolated individuals in siltstone intervals. The bivalves contain a mix of free-lying 90 and encrusting forms. Other taxonomic groups such as ammonoids, gastropods and corals are rare occurring mostly in the siltstone intervals. Two intervals in the lower 48 m of the section contain abundant fish scales that occur in a thin- bedded (1 cm), light gray, volcanic ash (Figure 4.4B). The diverse benthic fauna from this section will be the subject of further research. Twenty specimens of Late Pliensbachian ammonoids were identified representing five genera that occur at five localities within this stratigraphic section (Figures 4.4B, 4.5). Fanninoceras (Fanninoceras) carlottense, and Lioceratoides (Lioceratoides) cf. involutum are identified from localities 1\u00E2\u0080\u00934 and indicate the Carlottense Zone (Smith and Tipper, 1996). Fanninoceras (Fanninoceras) fannini, which is recorded from locality 3, is known primarily from the Kunae Zone of western North America but it also occurs rarely in the lower Carlottense Zone (Smith and Tipper, 1996). Lastly, Lytoceras sp. is identified from locality 5. Lytoceras is uncommon in the Lower Jurassic of North America and the specimen from Hicks Creek is the first figured specimen from the Pliensbachian. Ammonite biostratigraphy of this section indicates that deposition of the Talkeetna Formation at Hicks Creek occurred during the Carlottense Zone of the Late Pliensbachian, a slightly younger age than that of the Camp Creek section. 91 Chapter 5 Results 5.1 Extinction data and patterns of diversity Stratigraphic ranges of 206 ammonite and 242 foraminifera species were compiled and analyzed at an informally defined \u00E2\u0080\u0098sub zonal\u00E2\u0080\u0099 level for Pliensbachian\u00E2\u0080\u0093Toarcian strata in western North America. Compiled ammonite species were analyzed as one dataset and foraminiferal species were first analyzed according to region (e.g. Haida Gwaii, and northern Alaska) and then as a unified dataset. This is largely because the studied localities containing ammonite species were located in the paleo Pacific Ocean during the Early Jurassic, whereas the two areas containing benthic foraminiferal species are thought to have been located on the margins of two separate ocean basins during the Early Jurassic, namely the paleo Pacific Ocean and the paleo Arctic Ocean. Patterns in species diversity for ammonite and foraminiferal species are presented in Figure 5.1. Ammonite species diversity (Figure 5.1A) is measured as a total sum (Ntot), without singletons, and as an estimated mean. Regional foraminiferal species diversity data (Figure 5.1B) are only presented as Ntot values. In both cases, diversity levels in all three metrics generally follow the same overall pattern (within each taxonomic group independently) and show very similar values, which indicates that taxonomic singletons do not have a strong influence in this analysis. Results are therefore presented in terms of the total diversity (Ntot) with occasional reference to singleton-free data. Diversity data 92 pertaining to the singleton-free and estimated mean metrics for foraminiferal species are presented in Figure 5.2. Figure 5.1 - Pliensbachian\u00E2\u0080\u0093Toarcian ammonite (A) and foraminiferal (B) species level diversity in western North America. Data is derived from compiled stratigraphic range charts (Figures 2.2\u00E2\u0080\u00932.4; 3.1\u00E2\u0080\u00933.4). Ammonite diversity data is compiled from many localities throughout western North America, whereas foraminiferal diversity data is from Haida Gwaii (BC) and the South Barrow #3 core (Alaska). Major declines in ammonite species diversity (A) are evident at six distinct intervals, whereas foraminiferal diversity data (B) only shows two major declines (Kanense Zone and in the later part of the Toarcian). Wh. = Whiteavesi, Fre. = Freboldi, Carl. = Carlottense, Kan. = Kanense, Plan. = Planulata, Crass. = Crassicosta, Hill. = Hillebrandti, Yak. = Yakounensis. 93 Results for the ammonite species diversity analysis (Figure 5.1A) show a steady increase in species diversity throughout the Imlayi\u00E2\u0080\u0093middle Whiteavesi Zones of the Early Pliensbachian. Diversity reaches a high point in the middle Whiteavesi Zone (19 species) and then declines steadily throughout the remaining part of the Early Pliensbachian, reaching a low point in the middle Freboldi Zone (12 species). From the late Freboldi Zone to the middle Kunae Zone, there is a sharp increase and diversity reaches its maximum value of 30 species. Also, over this time frame during the early Kunae Zone, there is an observed offset between the total diversity and singleton-free curves of ~5 species. Throughout the remaining part of the Late Pliensbachian and into the Early Toarcian, diversity undergoes a major decline that consists of two distinct steps. The initial step is an abrupt decline from the middle Kunae to the early Carlottense Zones when diversity drops from 30 to 14 species. The second step is a decline that begins after the middle Carlottense Zone, crosses the Pliensbachian / Toarcian boundary, and reaches a minimum in species diversity of 6 species in the middle Kanense Zone. From the late Kanense Zone and into the Middle Toarcian there is a gradual increase in species diversity that reaches a high point of 15 species in the middle Planulata Zone. Following this peak in diversity in the Middle Toarcian, there is a gradual decline from the late Middle Toarcian into the early Late Toarcian. During this interval, diversity falls to 8 species in the lower Crassicosta Zone and then gradually declines again, 94 reaching low levels (of ~5 species) throughout the Hillebrandti Zone of the Late Toarcian. Throughout the remaining part of the Late Toarcian there is an abrupt rise in diversity in the Yakounensis Zone that reaches a high point of 14 species, with a subsequent decline in the later part of the zone. Regional foraminiferal species diversity data (Figure 5.1B), analyzed independently according to geographic region, shows a much different pattern throughout the Pliensbachian\u00E2\u0080\u0093Toarcian time. The record from Haida Gwaii indicates an initial increase in diversity from ~10 species in the Imlayi Zone to ~40 in the Whiteavesi Zone and then increases, very gradually, to ~47 species throughout the rest of the Pliensbachian. Throughout the Early\u00E2\u0080\u0093Middle Toarcian there is a steady rise in diversity that reaches ~72 species in the late Planulata- middle Crassicosta Zones. At this point, there is a steady decline to the Late Toarcian where species diversity in the late Yakounensis Zone drops to ~33. Data from the Alaska core (orange line in 5.1B) shows consistent foraminiferal species diversity of ~27 species throughout the Imlayi\u00E2\u0080\u0093late Freboldi Zones of the Early Pliensbachian. Throughout the Kunae Zone there is a distinct rise in diversity that reaches a maximum of 51 species in the late Kunae Zone. In the later part of the Pliensbachian extending across the Pliensbachian / Toarcian boundary and into the early Middle Toarcian, there is a major decline in diversity that is evident as two distinct steps. The first decline is into the Carlottense Zone, where diversity maintains fairly steady levels, decreasing only slightly from 43 to 39 species. The second decline is much more significant dropping from 39 in the late Carlottense Zone to 5 in the lower Planulata Zone. It should be noted that 95 above this interval in the SB #3 Alaska core there is a gap in the core-recovery and therefore it is not possible to obtain foraminiferal species diversity data from the middle Planulata Zone onward. The combined foraminiferal species diversity analysis (Figure 5.2A) constitutes the entire combined stratigraphic range chart from both British Columbia and Alaska. This analysis shows that foraminiferal species diversity maintains a steady increase throughout the Early Pliensbachian, reaching ~70 species in the late Freboldi Zone. Following this there is a small, sharp, decline in the early Kunae Zone to 65 species with a subsequent abrupt increase in diversity throughout the remaining part of the zone where diversity reaches a maximum of 88 species. In the later part of the Pliensbachian\u00E2\u0080\u0093Early Toarcian there is a gradual decline to 73 species in the middle part of the Kanense Zone. From the middle part of the Early Toarcian to end of the Middle Toarcian, species diversity fluctuates slightly but remains at this level. In the Late Toarcian diversity declines, gradually, and reaches ~30 species in the Late Toarcian. 96 Figure 5.2 - Combined Pliensbachian\u00E2\u0080\u0093Toarcian foraminiferal species diversity from Haida Gwaii and the South Barrow #3 core in Arctic Alaska, western North America. Data in (A) is the total combined species from the compiled stratigraphic range chart (Figures 2.2\u00E2\u0080\u00932.4; 3.1\u00E2\u0080\u00933.4). Data in (B) does not include species whose stratigraphic ranges are greater than 6 ammonite zones within the Pliensbachian\u00E2\u0080\u0093Toarcian interval. In (A) only small declines in biodiversity are noted in the Kunae, Kanense, and Planulata Zones and a gradual decline in species diversity that occurs throughout the later part of the Toarcian. In (B) declines in species diversity are evident over five potential intervals throughout the Pliensbachian\u00E2\u0080\u0093Toarcian time (red numbers). See Figure 5.1 for abbreviations. 97 An important factor of this diversity analysis is the proportion of long- ranging taxa compared with the single ended taxa (taxa that contain stratigraphic ranges that end or begin within the measured time interval). As mentioned, taxa that cross both measured boundaries (i.e. Nbt value in Figure 3.7B) are counted as a whole unit, single ended taxa counted as a half, and singletons counted as a third. If there is an over abundance of Nbt taxa, then changes in diversity within a particular interval may be diminished. In the combined foraminiferal species diversity analysis, there is an abundance of foraminifera whose stratigraphic ranges are exceedingly long in comparison with those of ammonite species (Figures 2.2\u00E2\u0080\u00932.4; 3.1\u00E2\u0080\u00933.4). Therefore the combined foraminiferal species stratigraphic range chart was re-analyzed without the long-ranging species (omitting species whose stratigraphic ranges are greater than six ammonite zones within the Pliensbachian\u00E2\u0080\u0093Toarcian interval, or are greater than half of the studied interval). The results are presented in Figure 5.2B. The combined foraminiferal species analysis without long-ranging species (Figure 5.2B) shows an increase in diversity from the Imlayi to Whiteavesi Zones, a very slight decrease across the transition between the Whiteavesi and Freboldi Zones, with a subsequent gradual increase in diversity from the Freboldi Zone to the early Kunae Zone. Throughout the Kunae Zone, there is a sharp rise in foraminifera, reaching a maximum of ~40 species in the late part of the zone. Following this apparent peak in foraminiferal diversity there is a large decline into the Early Toarcian that is divided into three phases, an initial drop in the early Carlottense Zone, a subsequent gradual decline throughout the remaining part of 98 the zone and a sharp decline across the Pliensbachian / Toarcian boundary where diversity reaches a low point of 24 species in the middle part of the Kanense Zone. From the late Early Toarcian, there is a steady rise in diversity with one very tenuous decline across the Planulata / Crassicosta zone boundary (number 4? in Figure 5.2). Diversity levels reach a Toarcian peak in the Crassicosta Zone at 37 species. Foraminiferal diversity then declines gradually throughout the Late Toarcian. 5.1.1 Extinction and Origination patterns Rate metrics established by Foote (2000) were used to assess patterns of extinction and origination in ammonoid and foraminiferal species throughout the Pliensbachian and Toarcian. For this analysis, long-ranging foraminiferal species were excluded for reasons previously discussed. Results show that, in general, the derived rates for all four metrics used (i.e. Van Valen with singletons, Van Valen without singletons, per-taxon rate, and estimated per capita rate) had similar values, except for the lower Kanense Zone. Consequently only the Van Valen metric with singletons is discussed (Figure 5.3). Ammonite and foraminiferal species had similar patterns of extinction and origination although the scales were quite different in that extinction rate among ammonites is much higher than in the foraminifera. Throughout much of the Pliensbachian, ammonites showed similar extinction and origination rates that were fairly low (blue line in Figure 5.3). There is an increase in origination in the lower Kunae Zone as well as a slight elevation in both rates that is offset in the late Kunae\u00E2\u0080\u0093early Carlottense Zonal interval. Across the Pliensbachian / Toarcian boundary, covering an interval from the middle of the Carlottense Zone to the 99 middle of the Kanense Zone, there is a very dramatic increase in the extinction rate. Throughout the Middle\u00E2\u0080\u0093Late Toarcian the extinction and origination rates do fluctuate, but generally maintain lower levels. Foraminiferal species data shows low extinction and origination rates throughout much of the Pliensbachian with offset increases in both metrics throughout the entire Kunae Zone interval (Figure 5.3B). Across the Pliensbachian / Toarcian boundary, covering an interval from the middle of the Carlottense Zone to the middle of the Kanense Zone, there is a sharp and marked increase in the extinction rate (green line in Figure 5.3B). Within this interval, origination rate is somewhat elevated but it lags behind the extinction rate considerably. Throughout the Middle Toarcian, rates are of similar magnitude as those experienced in the Late Pliensbachian. In the Late Toarcian, extinction rates are predominantly higher than origination and reach maximum values in the late Hillebrandti and middle Yakounensis zones. 100 Figure 5.3 - Extinction and Origination rates for Pliensbachian\u00E2\u0080\u0093Toarcian ammonite (A) and foraminiferal (B) species in western North America. Rate metrics for foraminifera do not contain species whose stratigraphic ranges are greater than 6 ammonite zones within the Pliensbachian\u00E2\u0080\u0093Toarcian time interval. Data shows accelerated extinction rates in the Kanense Zone for both taxonomic groups that correlates with major declines in species diversity at interval 3 (Figures 5.1, 5.2). See Figure 5.1 for abbreviations. 101 5.2 Geochemical data A total of 1596 samples were collected and analyzed for isotope geochemistry analysis from two temporally constrained stratigraphic sections on Haida Gwaii and the SB #3 core (Appendix B). Samples were analyzed for carbon-isotope (\u00CE\u00B413Corg, reported vs. \u00E2\u0080\u00B0 Vienna PeeDee Belemnite), total organic carbon (reported as wt%), nitrogen-isotope (\u00CE\u00B415N, reported vs. \u00E2\u0080\u00B0 air), and total nitrogen (reported as wt%) analyses. Previous and new work has established the temporal framework within which the geochemical data are presented. 5.2.1 Whiteaves Bay section The Whiteaves Bay section is located on the northern part of Moresby Island in the Skidegate Inlet area of Haida Gwaii (number 16 in Figure 3.5A2) and consists of approximately 115 m of strata, which includes the Ghost Creek, Fannin, and Whiteaves Formations of the Early Jurassic Maude Group. It was originally measured and described by Cameron and Tipper (1985) and subsequently re-measured and calibrated with the Early Pliensbachian\u00E2\u0080\u0093Middle Toarcian part of the zone scheme (Figure 5.4; using data in Smith et al., 1988; Jakobs et al., 1994; Smith and Tipper, 1996; Jakobs, 1997). 102 103 Figure 5.4 - Lithostratigraphy and ammonite biostratigraphy of the Whiteaves Bay section on Central Graham Island, Haida Gwaii (Modified from Smith and Tipper, 1996 and Jakobs, 1997). Fm. = Formation, C. = Carlottense, Pr. = Protogrammoceras, L. = Lioceratoides, P. = Pacificeras. In total, 375 samples were collected from the Whiteaves Bay section (Figure 5.5A, B). Of these samples, twelve from the base of the section were also analyzed for nitrogen-isotope (\u00CE\u00B415N) and total nitrogen data (Figure 5.5C, D). From 0 to 5 m in the section \u00CE\u00B413C values average \u00E2\u0080\u009329\u00E2\u0080\u00B0 and then show a gradual shift to ~ \u00E2\u0080\u009326\u00E2\u0080\u00B0 at 22 m in the section. From 22\u00E2\u0080\u009328 m, \u00CE\u00B413C values shift to ~ \u00E2\u0080\u0093 28\u00E2\u0080\u00B0 with a small positive shift to \u00E2\u0080\u009325\u00E2\u0080\u00B0 at 25m in the section. Throughout this interval, organic carbon concentrations peak at 3.36% at 14.7 m in the section and average 1.5%. From 29 to 52 m in the section \u00CE\u00B413C values fluctuate markedly, but generally show a positive excursion in carbon-isotopes throughout the interval. Throughout this interval there is often a noticeable difference between adjacent samples, where \u00CE\u00B413C values have a difference that is greater than 1\u00E2\u0080\u00933\u00E2\u0080\u00B0. Also within this interval there are several samples that contain carbon-isotope values that are anomalously high and range from \u00E2\u0080\u009318 to \u00E2\u0080\u009321\u00E2\u0080\u00B0 (red circles in Figure 5.5A, B). Organic carbon throughout this interval has lower variability, averaging 0.5% with higher amounts reaching ~2% at 30 m in the section. From 54 to 65 m in the section there is a positive excursion, where \u00CE\u00B413C values shift from \u00E2\u0080\u009328\u00E2\u0080\u00B0 (at 54 m) to approximately \u00E2\u0080\u009326\u00E2\u0080\u00B0 (at 56\u00E2\u0080\u009357 m) and then return to approximately \u00E2\u0080\u009328\u00E2\u0080\u00B0 (at 61\u00E2\u0080\u009364 m). From 66 to 76 m in the section there is a slight negative shift, but \u00CE\u00B413C values generally average \u00E2\u0080\u009327\u00E2\u0080\u00B0 in this interval. 104 From 76 to 77 m in the section \u00CE\u00B413Corg values are slightly more positive and average \u00E2\u0080\u009326\u00E2\u0080\u00B0. From 54 to 77 m in the section, TOC is consistent and averages 0.5% throughout the interval. From 77\u00E2\u0080\u009383 m there is an abrupt \u00E2\u0080\u00932\u00E2\u0080\u00B0 negative shift in \u00CE\u00B413Corg to values that averages \u00E2\u0080\u009328\u00E2\u0080\u00B0 for 6 m. This negative shift coincides with a slight increase in TOC to approximately 0.6% near the top of the excursion interval. From 83\u00E2\u0080\u0093102 m (\u00CE\u00B413Corg values become more positive averaging \u00E2\u0080\u009325\u00E2\u0080\u00B0 throughout the interval with a small negative excursion at ~92 m to \u00E2\u0080\u009326 or \u00E2\u0080\u009327\u00E2\u0080\u00B0. Organic carbon has lower variability and averages 0.4% throughout the interval. Low resolution \u00CE\u00B415N and TN data from the basal 21 m of the section are shown in Figure 5.5C, D. In general \u00CE\u00B415N values are positive and range between 0 and 1\u00E2\u0080\u00B0 throughout the interval. However three intervals (from 6\u00E2\u0080\u00938 m, at 13.5 m and at 21.10 m) show signs of a negative trend where values are slightly negative and do not exceed \u00E2\u0080\u00930.25\u00E2\u0080\u00B0. Total nitrogen (TN) throughout this interval ranged between 0.12% and 0.20% with two instances of lower TN values that were 0.08% and 0.07% at 1.8 and 17.4 m respectively. 105 106 Figure 5.5 - Geochemistry of the Whiteaves Bay section. (A) \u00CE\u00B413Corg (\u00E2\u0080\u00B0 VPDB: Vienna PeeDee Belemnite) data; (B) Total Organic Carbon TOC (wt%) data; (C) \u00CE\u00B415N (\u00E2\u0080\u00B0 Air) data; (D) Total Nitrogen TN (wt%) data. Red circles are samples that contain anomalous \u00CE\u00B413Corg (greater than -22\u00E2\u0080\u00B0) values. CIE = Carbon isotope excursion, C. = Carlottense. 5.2.1.1 Temporal constraints The basal 27 m of section is assigned to the Ghost Creek Formation and contains species of Tropidoceras, Acanthopleuroceras and Metaderoceras, which are indicative of the Imlayi and Whiteavesi Zones (Figure 5.4). The Fannin Formation conformably overlies the Ghost Creek Formation and is ~51 m thick (Smith and Tipper, 1996). It contains species from many genera that are indicative of the Freboldi, Kunae, Carlottense, and lower Kanense Zones (Smith and Tipper, 1996; Jakobs, 1997). Dactylioceras kanense occurs at 71 m in the section (Figure 5.4) and, along with other species including Dactylioceras aff. comptum, Dactylioceras cf. alpestre, Lioceratoides allifordense, Lioceratoides propinquum, Protogrammoceras cf. paltum, Tiltoniceras antiquum (Figures 4.5, 5.4), indicates the earliest Toarcian (lowermost part of the Kanense Zone). As mentioned previously in Chapter 2.1, the incoming of Dactylioceras marks the base of the Toarcian generally and the species D. comptum occurs at this level in NE Russia (Dagys, 1968). Protogrammoceras cf. P. paltum, Lioceratoides propinquum and Tiltoniceras antiquum occur in Europe and the North America Cordillera, ranging from the Upper Pliensbachian to the Lower Toarcian (Howarth, 1992; Smith and Tipper, 1996). The co-occurrence of these taxa indicates the lower part of the Kanense Zone. 107 Although there are a number of negative excursions to lighter \u00CE\u00B413Corg values throughout the section, there is a sustained interval at 77-83 m where values are approximately \u00E2\u0080\u009329\u00E2\u0080\u00B0 for 6 m. Above the negative CIE interval, from 98\u00E2\u0080\u0093103 m in the section, several ammonite species have been identified which are indicative of the Planulata Zone (Figure 5.4). These include Rarenodia planulata, Leukadiella ionica, Leukadiella aff. helenae and an unidentified species of Cleviceras. At the top of the Whiteaves bay section the Whiteaves Formation is truncated by a fault, which separates it from the overlying Middle Jurassic Yakoun Group. It seems evident that the negative CIE at 77 m in the section occurs within the Early Toarcian at a correlative interval with the negative CIE in Europe. 5.2.2 Yakoun River section The Yakoun River section crops out along a stream-cut of the Yakoun River in Central Graham Island (number 10 in Figure 3.5A1) and is known as one of the most complete Toarcian sections in Haida Gwaii. This section, originally measured and described by Cameron and Tipper (1985) and re-measured by Jakobs (1997), extends from the lower part of the Whiteaves Formation through the Phantom Creek Formation and into Graham Formation (of the Middle Jurassic Yakoun Group). Jakobs et al. (1994) and Jakobs (1997) place zone- level biostratigraphic constraint for the Whiteaves Formation and the lower part of the Phantom Creek Formation (Figure 5.6). Original measurement of the section indicates that the Whiteaves Formation is ~60 m thick (Jakobs, 1997). However, recent erosion has significantly increased exposure of the lower part (of the section), which now totals ~100 stratigraphic meters. 108 109 Figure 5.6 - Lithostratigraphy and ammonite biostratigraphy of the Yakoun River section on Central Graham Island, Haida Gwaii (Modified from Jakobs, 1997). Ph. = Phantom Creek Formation, Yak. = Yakounensis Zone, Pr. = Protogrammoceras, D. = Dactylioceras, Cl. = Cleviceras, H. = Harpoceras, P. = Peronoceras, Phy. = Phymatoceras, Den. = Denckmannia, Ps. = Pseudomercaticeras, Gr. = Grammoceras, Sph. = Sphaerocoeloceras, Pseu. = Pseudolioceras, Ha. = Hammotoceras. From the initial 300 samples that were collected in 2008 for geochemical analysis from the Yakoun River section, \u00CE\u00B413Corg and TOC were determined on 274 samples, and \u00CE\u00B413C data were obtained on 26 wood samples (Figure 5.7A, B). A sub-set of 169 samples was also used for \u00CE\u00B415N and total nitrogen determinations (Figure 5.7C, D). From 0 to 32 m in the section, \u00CE\u00B413Corg values average \u00E2\u0080\u009324\u00E2\u0080\u00B0 with a small negative shift to ~ \u00E2\u0080\u009327\u00E2\u0080\u00B0 at 17 m and a small positive shift to ~ \u00E2\u0080\u0093 22\u00E2\u0080\u00B0 at 23 m. From 0\u00E2\u0080\u009322.5 m, TOC values average 0.9% (wt %) with two single- point maxima values of 2% and 2.3% at 7 m and 11 m respectively. From 22.5\u00E2\u0080\u0093 32 m, TOC values decrease, averaging ~0.4%. From 32 m to 43 m in the section there is an abrupt \u00E2\u0080\u00937\u00E2\u0080\u00B0 negative shift in \u00CE\u00B413Corg to values that average \u00E2\u0080\u009331\u00E2\u0080\u00B0 for 11 m. This pronounced negative shift coincides with an increase in TOC values from 0.4% to 0.7% (from 32\u00E2\u0080\u009338 m in the section), eventually reaching values that average 1.2% (from 39\u00E2\u0080\u009341 m in the section), and subsequently show a sharp decrease in TOC to ~0.5% (from 41\u00E2\u0080\u009343 m). After a return to more positive values of approximately \u00E2\u0080\u009325\u00E2\u0080\u00B0 at 43 m in the section, values show a steady negative trend From 43 m to 105 m in the section; averaging \u00E2\u0080\u009325\u00E2\u0080\u00B0 (from 43\u00E2\u0080\u009355 m), \u00E2\u0080\u009326\u00E2\u0080\u00B0 (from 55\u00E2\u0080\u009370 m), \u00E2\u0080\u009327\u00E2\u0080\u00B0 (from 70\u00E2\u0080\u009390 m), and \u00E2\u0080\u009329\u00E2\u0080\u00B0 (from 90\u00E2\u0080\u0093105 m). TOC values remain low throughout this interval, averaging 0.5%. Although \u00CE\u00B413Cwood samples are sparse there is broad similarity to 110 the \u00CE\u00B413Corg curve, including a large 8.5\u00E2\u0080\u00B0 negative shift to \u00E2\u0080\u009330\u00E2\u0080\u00B0 at 33 m in the section and a similar pattern in carbon-isotope fluctuation from 47\u00E2\u0080\u009353 m, 60\u00E2\u0080\u009361 m, and 71\u00E2\u0080\u009380 m in the section (Figure 5.7A, B). Nitrogen-isotope and total nitrogen data was generated for 169 samples of the initial sample-set that was collected in 2008 (Figure 5.7C, D). Throughout the entire section, \u00CE\u00B415N values range from ~ \u00E2\u0080\u00931.5 to 0.8\u00E2\u0080\u00B0 with the lightest values occurring in the negative CIE interval. From 0 to 32 m in the section \u00CE\u00B415N values average \u00E2\u0080\u00930.28\u00E2\u0080\u00B0 with occasional samples containing slightly positive values of 0.08\u00E2\u0080\u00930.5\u00E2\u0080\u00B0. From 32 to 42 m in the section \u00CE\u00B415N values are more negative, peaking at \u00E2\u0080\u00931.5\u00E2\u0080\u00B0 at 32.8 m and averaging \u00E2\u0080\u00930.6\u00E2\u0080\u00B0 for the interval. From 43\u00E2\u0080\u009369 m \u00CE\u00B415N values are closer to 0, averaging \u00E2\u0080\u00930.2\u00E2\u0080\u00B0 and peaking at \u00E2\u0080\u00930.6\u00E2\u0080\u00B0 at 45.4 m. From 70\u00E2\u0080\u0093103 m values average very close to 0.0\u00E2\u0080\u00B0 (at \u00E2\u0080\u00930.001\u00E2\u0080\u00B0) with occasional small-scale positive and negative trends. Throughout the entire section, total nitrogen values consistently average 0.05%. 111 112 Figure 5.7 - Geochemistry of the Yakoun River section. (A) \u00CE\u00B413Corg and \u00CE\u00B413Cwood (\u00E2\u0080\u00B0 VPDB: Vienna PeeDee Belemnite) data; (B) Total Organic Carbon TOC (wt%) data; (C) \u00CE\u00B415N (\u00E2\u0080\u00B0 Air) data; (D) Total Nitrogen TN (wt%) data; (E) 87Sr/86Sr ratios (from Gr\u00C3\u00B6cke et al. (2007). Ph. = Phantom Creek Formation, Yak. = Yakounensis Zone, CIE = Carbon isotope excursion. See Figure 5.6 for lithological legend. Averaged \u00CE\u00B413Cwood data from similar stratigraphic horizons are reported with bars showing standard deviation. Hexagonal-shaped symbol denotes stratigraphic position of a zircon-bearing ash dated by P\u00C3\u00A1lfy et al. (1997). Figure adapted from Caruthers et al. (2011). 5.2.2.1 Temporal constraints Geochemical data for the Yakoun River stratigraphic section are temporally constrained by previous and new work (Figures 4.5; 5.6). In total, all five Toarcian ammonite zones are identified in this section (Jakobs, et al., 1994; Jakobs, 1997). At the base of the section, poorly preserved specimens of Hildaites and Cleviceras occur above, below and within the CIE interval which indicates a Kanense Zone (lowest Toarcian) age for this part of the sequence. Below the excursion interval, Dactylioceras cf. compactum (Dagys), and Protogrammoceras cf. paltum Buckman occur at the base of the section (Figure 5.6). Dactylioceras cf. compactum is known from the earliest Toarcian in NE Russia, at a similar stratigraphic level to D. kanense and D. comptum (Dagys, 1968). Its presence, along with Protogrammoceras cf. paltum, at the base of the Yakoun River section indicates the lowermost part of the Kanense Zone. Unfortunately, the ash bed below the CIE interval, at ~6 m in the section, did not yield enough zircons for U\u00E2\u0080\u0093Pb analysis. Above the negative CIE, from 45\u00E2\u0080\u009370 m above the base of the section (Figure 5.6), Cleviceras cf. exaratum (Young and Bird), Cleviceras cf. chrysanthemum (Yokoyama), and Hildaites murleyi (Moxon) occur just below 113 Rarenodia planulata Venturi and Lukadiella ionica Renz and Renz (Jakobs, 1997). This is indicative of the Upper Kanense and Planulata Zones, which correlate approximately with the Serpentinum and Bifrons ammonite Zones of NW Europe (Figure 2.1). Previous analysis of belemnite rostra from the Yakoun River section (Figure 5.7E) show 87Sr/86Sr ratios ranging from 0.707138 to 0.707164 in three samples that were recovered from 53 m above the base of the section, ~ 10 m above the CIE (Gr\u00C3\u00B6cke et al., 2007). These values plot on the steeply rising (basal Serpentinum Zone) portion of the 87Sr/86Sr curve for NW Europe (star-shaped symbol in Figure 5.8), which is above the extinction event and in the waning part of the negative CIE (McArthur et al., 2000). A U\u00E2\u0080\u0093Pb age date of 181.4 + 1.2 Ma (2\u00CF\u0083) from a zircon-bearing ash bed occurring ~29 m above the termination of the CIE interval in the Yakoun River section (hexagonal- shaped symbol in Figure 5.7) has contributed to the geochronological time scale for the Planulata/Crassicosta Zone boundary (P\u00C3\u00A1lfy et al., 1997). In summary, the large negative CIE from 32 m to 43 m within the Kanense Zone is concomitant with the negative CIE interval of NW Europe. 114 115 Figure 5.8 - Seawater 87Sr/86Sr for the Late Pliensbachian\u00E2\u0080\u0093Toarcian of Yorkshire, UK (after McArthur et al., 2000). Closed circles are data from McArthur et al. (2000) and grey squares are from Jones et al. (1994). The green star shows the position of three 87Sr/86Sr values at 53 m in the Yakoun River section (from Figure 5.7E). Dashed light-grey line shows stratigraphic position of the extinction interval documented by Caswell et al. (2009). Ammonite zones and subzones (vertical text) after Howarth (1992). Th. = Thouarsense Zone, Tenui. = Tenuicostatum Zone, tenuicostat. = Tenuicostatum Subzone. 5.2.2.2 Higher resolution sampling In 2010, an additional 318 samples were collected at a higher sampling resolution from the large negative CIE interval in the Yakoun River section and analyzed for \u00CE\u00B413Corg (Figure 5.9). From 27 to 30.75 m in the section \u00CE\u00B413Corg values show a steady gradual positive shift from approximately \u00E2\u0080\u009326 to \u00E2\u0080\u009324\u00E2\u0080\u00B0. From 30.80 to 31.15 m there is a large negative shift in carbon-isotopes, in which \u00CE\u00B413Corg values decrease at three separate intervals (steps I, II and III in Figure 5.9B) of approximately 2\u00E2\u0080\u00B0, 3\u00E2\u0080\u00B0, and 1\u00E2\u0080\u00932\u00E2\u0080\u00B0 respectively. This is potentially comparable with similar declines within the Early Toarcian successions of Europe, which are attributed to astronomical precession (Kemp et al., 2005; Hermoso et al., 2009a; Hesselbo and Pienkowski, 2011; Kemp et al., 2011). From 31.2 to 35.6 m in the section \u00CE\u00B413Corg values remain very negative, ranging between \u00E2\u0080\u009330 and \u00E2\u0080\u009331\u00E2\u0080\u00B0. From 31.2\u00E2\u0080\u009332.3 m \u00CE\u00B413Corg values are the lowest, averaging \u00E2\u0080\u009331\u00E2\u0080\u00B0 and reaching a minimum value of \u00E2\u0080\u009332\u00E2\u0080\u00B0. At 35.7 m in the section there is an abrupt \u00E2\u0080\u00936\u00E2\u0080\u00B0 positive shift in carbon isotope ratios, where \u00CE\u00B413Corg values jump from \u00E2\u0080\u009330\u00E2\u0080\u00B0 to \u00E2\u0080\u009324\u00E2\u0080\u00B0 over 0.05 cm in the section. From 35.7 to 45.05 m in the section \u00CE\u00B413Corg values remain more positive, averaging \u00E2\u0080\u009325\u00E2\u0080\u00B0 with an occasional negative shift to \u00E2\u0080\u009328 or \u00E2\u0080\u009329\u00E2\u0080\u00B0. 116 In comparing the two sets of carbon-isotope data across the negative CIE interval in the Yakoun River section (Figure 5.9), it is now evident that the initial negative shift in carbon-isotopes occurred over three separate steps of \u00E2\u0080\u00931 to \u00E2\u0080\u0093 3\u00E2\u0080\u00B0, as opposed to the \u00E2\u0080\u00936\u00E2\u0080\u00B0 shift that is evident in the low resolution data (Figure 5.9A). It is also evident that the negative CIE interval in the high-resolution data is ~6 m thick, while the low-resolution data shows a thickness of ~11 m. The apparent truncation of the negative CIE interval at 35.7 m occurred, coincidentally, with a change in sampling strategy from the cut-bank of the Yakoun River to the riverbed. This retrospectively indicates that sampling occurred in section that was not in place. Within this \u00E2\u0080\u0098out of place\u00E2\u0080\u0099 portion of the 2010 data (36\u00E2\u0080\u009345 m in Figure 5.9B), \u00CE\u00B413Corg values are most similar to those in the uppermost Kanense Zone from the lower resolution data (Figure 5.9A), which suggests a small fault or bend in the section between the cut-bank and riverbed. 117 Figure 5.9 - Carbon-isotope data for the negative CIE interval of the Yakoun River comparing: A) the 2008 sample set (Figure 5.7) and B) samples collected at a higher-resolution sample interval in 2010. Higher-resolution sampling suggests an initial negative shift in \u00CE\u00B413Corg of ~3\u00E2\u0080\u00B0 that occurs over three separate intervals (B), as opposed to the much larger shift of ~6\u00E2\u0080\u00B0 that is evident in (A). The entire negative CIE interval in (B) was not sampled, values in red are out of place. See Figure 5.6 for lithologic key. 5.2.3 South Barrow #3 core The SB #3 core (number 1 in Figures 3.5A, 3.6) is located on the North Slope of Alaska near the town of Barrow and is part of the National Petroleum Reserve. Drilling of the core commenced and was completed in 1949 reaching a total depth of ~884 m or 2,900 ft (Collins, 1961). Several lithostratigraphic units have been identified in the core, ranging from Triassic\u00E2\u0080\u0093Pleistocene in age. The 118 Lower Jurassic portion of the core is ~275 m (902 ft) thick and encompasses sediments of the Lower Kingak Formation. It is thought to be Hettangian\u00E2\u0080\u0093 Toarcian in age (Collins, 1961). Samples collected for isotope geochemistry from a portion of the core that extends from ~675 to 530 m in depth has yielded several Late Pliensbachian and Early Toarcian ammonites and foraminifera (Imlay, 1955; 1981; Tappan, 1955). In total, 603 samples were collected for geochemical analysis from the SB #3 Alaska core and were analyzed for \u00CE\u00B413Corg and TOC (Figure 5.10). From 676 to 652 m in the core there is a gradual negative excursion in \u00CE\u00B413Corg, where values reach a minimum of \u00E2\u0080\u009328\u00E2\u0080\u00B0 at 666 m and then steadily increase to \u00E2\u0080\u009325\u00E2\u0080\u00B0 at 652 m in the core. TOC data for this interval shows heightened values that average 1.12% and reach a maximum of 2.8% at 667 m. At 652 m there is a gap interval of ~7 m where there was no recovery of core from the well. From 645\u00E2\u0080\u0093 629 m there is a steady decrease in \u00CE\u00B413Corg with minimum values of \u00E2\u0080\u009329\u00E2\u0080\u00B0 at 626.9 m and an average TOC of 0.7%. From 624\u00E2\u0080\u0093603 m \u00CE\u00B413Corg values are slightly more positive and average \u00E2\u0080\u009327\u00E2\u0080\u00B0. TOC is slightly higher throughout this interval and averages 1.06% from 624\u00E2\u0080\u0093620 m, 1.10% from 618\u00E2\u0080\u0093611 m, and 0.8% from 608\u00E2\u0080\u0093603 m in the core. At 603 m there is another gap of ~55 m where there was no recovery. From 548 to 534 m there is a slightly negative trend in carbon- isotopes, but \u00CE\u00B413Corg values average \u00E2\u0080\u009327\u00E2\u0080\u00B0 throughout the interval. TOC is consistent throughout this interval, averaging 0.7%. 119 120 Figure 5.10 - A) Carbon-isotope chemostratigraphy and B) Total Organic Carbon (wt% TOC) for the U.S. Navy South Barrow #3 well core of Northern Alaska. See Figure 3.5 for location of the well core and Figure 5.6 for lithologic key. ? = ammonite zone-level constraint uncertain, but the earliest Toarcian foraminiferal zone JF9b is probable based on the presence of the species Triplasia kingakensis. Ku\u00E2\u0080\u0093Car = Kunae\u00E2\u0080\u0093Carlottense Zones, Kan\u00E2\u0080\u0093Plan = Kanense\u00E2\u0080\u0093 Planulata Zones. 3.2.3.1 Temporal constraints Geochemical data for the SB #3 core are temporally constrained by previous work summarized in Figure 5.11. Imlay (1955; 1981) identifies the species Amaltheus cf. stokesi, Amaltheus cf. margaritatus, Amaltheus engelhardti and Amaltheus sp. from 669 to 630 m in the core, which indicates the Kunae and Carlottense Zones of the Late Pliensbachian (Figure 2.1). From 615 to 540 m in the core Dactylioceras sp., Dactylioceras kanense, and Dactylioceras cf. commune were identified. These species are known to occur in the Kanense and Planulata Zones of the Early and Middle Toarcian. This thereby leaves an interval of approximately 15 m in the core that could be either latest Pliensbachian or earliest Toarcian age. Within this interval, above the last occurrence of the Late Pliensbachian genus Amaltheus and before the first Early Toarcian genus Dactylioceras, there is a small negative CIE where \u00CE\u00B413Corg reach a minimum value of \u00E2\u0080\u009329\u00E2\u0080\u00B0 at 626 m in the core (Figure 5.10). At 628 m in the core the ammonite species Catacoeloceras? sp. juv. is identified and at 618\u00E2\u0080\u0093624 m the foraminifera Triplasia kingakensis is identified at (Imlay, 1955, 1981; Tappan, 1955). This foraminifera species is known as one of the earliest Toarcian foraminifera (Tappan, 1955; Nikitenko and Mickey, 2004) and therefore constrains this portion of the core 121 considerably, indicating that ~624 m in the core is stratigraphically just above the Pliensbachian / Toarcian boundary. Therefore, the small negative CIE at 626 m could potentially correlate with a small negative CIE at the Pliensbachian / Toarcian boundary from NW Europe (Littler et al., 2010). Above the large ~55 m gap in the core, Dactylioceras cf. commune is identified (Imlay, 1955; 1981). This species is known to occur in the uppermost Kanense Zone to Planulata Zone strata in western North America, which suggests that this interval is above the negative CIE interval of NW Europe, and Haida Gwaii (Figures 5.5, 5.7, 5.9). It therefore stands to reason that the Early Toarcian negative CIE interval, if present, would be located within the GAP interval from 603 to 548 m in the SB #3 core. 122 Figure 5.11 - Lithostratigraphy and biostratigraphy for the U.S. Navy South Barrow #3 well core of Northern Alaska. See Figure 3.5 for location of the well core. Stratigraphic position and identification of ammonites is from Imlay (1955; 1981) and Foraminiferal Zones are courtesy of B. Nikitenko (using data in Tappan, 1955). Ammonite Zone scheme is from Smith et al. (1988) and Jakobs et al. (1994). 123 Chapter 6 Discussion 6.1 A multi-phased Pliensbachian\u00E2\u0080\u0093Toarcian mass extinction One of the more pertinent issues that surround this extinction event is gauging its overall magnitude, duration and geographic extent on a global scale. It is currently suggested that the Pliensbachian\u00E2\u0080\u0093Toarcian extinction event in Europe and parts of the Arctic is multi-phased with the two most significant diversity declines occurring at the Pliensbachian / Toarcian boundary and at the Tenuicostatum / Serpentinum zonal boundary in the Early Toarcian (Dera et al., 2010). To date, this type of multi-phased scenario for the Pliensbachian\u00E2\u0080\u0093 Toarcian extinction has only been assessed in ammonoid faunas of the northwest Tethys and Arctic domains and, aside from the main-phase of extinction, has not yet been demonstrated as occurring in other taxonomic groups. New paleontological data from western North America provides insight into the magnitude and duration of this multi-phased event. Ammonite and foraminiferal species diversity in western North America suggests a similar multi-phased event (Figure 6.1A). Species diversity in both taxonomic groups declined and reached low points in six separate intervals that correspond to the: 1) middle Whiteavesi\u00E2\u0080\u0093middle Freboldi Zones, 2) late Kunae\u00E2\u0080\u0093 early Carlottense Zones, 3) late Carlottense\u00E2\u0080\u0093middle Kanense Zones, 4) late Planulata\u00E2\u0080\u0093early Crassicosta Zones, 5) middle Crassicosta\u00E2\u0080\u0093Hillebrandti Zones and 6) lower\u00E2\u0080\u0093middle Yakounensis Zone. These episodes of decreasing species 124 diversity correlate well with the multi-phased event recorded in combined ammonite data from the northwest European and Arctic domains as documented by Dera et al. (2010) (Figure 6.1B). Figure 6.1 - Ammonite and foraminiferal species level biodiversity in (A) Western North America (this study) and (B) NW Tethys and Arctic domains (Dera et al., 2010). Figure shows a multi-phased event with major declines occurring over six correlative intervals. In both datasets (A and B), the main phase of extinction is a large progressive decline that begins just before the Pliensbachian / Toarcian boundary and extends into the Early Toarcian where diversity reaches its lowest levels at an interval coeval with the negative CIE. Note: the Middle Toarcian event in Dera et al. (2010) is illustrated here as two separate events (#4, #5). Intervals i & ii = approximate extinction intervals previously identified by Harries and Little (1999) and Caswell et al. (2009), Wh. = Whiteavesi, Fre. = Freboldi, Carl. = Carlottense, Kan., = Kanense, Plan. = Planulata, Crass. = Crassicosta, Hill. = Hillebrandti, Yak. = Yakounensis, Jam. = Jamesoni, Marg. = Margaritatus, Spin. = Spinatum, T. = Tenuicostatum, S. = Serpentinum, Var. = Variabilis, Th. = Thouarsense, D. = Dispansum, P. = Pseudoradiosa, A. = Aalensis, negative CIE = negative carbon-isotope excursion interval in Caruthers et al. (2011). 125 The new dataset from western North America shows a modest decline in ammonite diversity and a correlative, but smaller, decline in foraminiferal diversity in the Early Pliensbachian that begins in the middle of the Whiteavesi Zone and continues to the middle of the Freboldi Zone (number 1 in Figure 6.1A). Throughout this interval, the extinction rates in both taxonomic groups are slightly elevated in comparison with origination (Figure 5.3). This modest decline in species diversity occurs over a similar time frame, and is of similar magnitude, to the decline in ammonite species diversity across the Ibex\u00E2\u0080\u0093Davoei Zone boundary in the dataset from northwest Europe and the Arctic (number 1 in Figure 6.1B). This could therefore suggest a controlling mechanism that is global in extent. Of the five episodes of diversity decline that constitute the multi-phased Pliensbachian\u00E2\u0080\u0093Toarcian extinction described by Dera et al. (2010), all are distinguishable in western North America (numbers 2 to 6 in Figure 6.1). In both datasets, the main phase of extinction consists of two successive phases of decline (number 3 in Figure 6.1): one at the Pliensbachian / Toarcian boundary and the other within the middle part of the Kanense Zone (at a correlative interval with the Tenuicostatum / Serpentinum Zone boundary of northwest Europe; Figure 2.1). At the onset of the main-phase of extinction in western North America, ammonite and foraminiferal diversity reached maximum values of ~30 and 40 species respectively in the Kunae Zone and then began to decline gradually into the Early Toarcian (numbers 2 and 3 in Figure 6.1A). In the combined northwest European and Arctic data (Figure 6.1B), species diversity 126 rebounds in the latest Pliensbachian following the decline in the Margaritatus\u00E2\u0080\u0093 Spinatum Zones and then subsequently collapses across the Pliensbachian\u00E2\u0080\u0093 Toarcian boundary into the Early Toarcian where ammonite species diversity reaches a minimum value. During this main-phase of extinction in western North America, ammonite and foraminiferal diversity reach minimum values of 5 and 25 species respectively in the middle part of the Kanense Zone (Figure 6.1A). These low levels of Early Toarcian diversity probably resulted from extraordinarily high extinction rates observed in both taxonomic groups in the lowest part of the Kanense Zone (Figure 5.3). One of the major discrepancies related to a global control mechanism for this event is the suggestion that the negative CIE only occurred in the Tethys Ocean area at an interval that is above the main extinction horizon (Wignall et al., 2005). However, more recent work has shown that the negative CIE is recognizable in many correlative Early Toarcian successions that are far- removed from the Tethys Ocean (Al-Suwaidi et al., 2010; Caruthers et al., 2011; Suan et al., 2011; Gr\u00C3\u00B6cke et al., 2011; Izumi et al., 2012). This suggests that this phenomenon was most likely global in extent and strongly supports influence by the methane hydrate reservoir. Furthermore, our study shows that the main phase of extinction is a progressive decline in species diversity that begins just before the Pliensbachian / Toarcian boundary and extends into the Early Toarcian where diversity reaches minimum values in three ocean basins (namely the paleo Pacific, Arctic and Tethys oceans) at an interval that is precisely correlative with the negative CIE (Figure 6.1). This apparent synchronicity 127 between diversity minimums during the main-phase of extinction and the negative CIE is a good indication that methane release played an important role in this extinction event, working to escalate its effects. Following the main phase of extinction in the Kanense Zone, ammonite species diversity in both datasets shows a rise into the Planulata Zone of the Middle Toarcian and its European correlative with a subsequent decline into the Hillebrandti Zone of the Late Toarcian. This decline in ammonite species is composed of two separate phases (numbers 4 and 5 in Figure 6.1). Although the extinction rate during phase number 4 is somewhat higher than origination (Figure 5.3), the observed values are considerably lower than those in the Kanense Zone. Within the later phase (number 5 in Figure 6.1), ammonite species diversity maintains lower levels throughout the Hillebrandti Zone of the Late Toarcian with low levels also observed in the correlative zone of Europe. During this phase, extinction and origination rates were nearly identical which resulted in no observable change in diversity. Lastly in the Late Toarcian, the ammonite dataset from northwest Europe and the Arctic shows an abrupt rise in diversity in the lower part of the Dispansum Zone with a final decline in the upper part of the zone; maintaining lower levels throughout the Pseudoradiosa and Aalensis zones (number 6 in Figure 6.1). In western North America, this final event is recognizable and is constrained based on the appearance of two ammonite species, Hammatoceras insigne ranging from the upper Hillebrandti\u00E2\u0080\u0093lower Yakounensis zones and Dumortieria cf. levesquei in the upper part of the Yakounensis Zone (Figure 2.4). 128 In the zone scheme of Northwest Europe (Page, 2003), these two species are noted as the index for the Insigne and Levesquei Subzones encompassing the lower parts of the Dispansum and Pseudoradiosa zones respectively (Figure 2.1). Therefore, the diversity peak occurring in the Yakounensis Zone (Figure 6.1A) is potentially correlative with the peak in the lower part of the Dispansum Zone (Figure 6.1B). The subsequent decline in diversity in the Yakounensis Zone is correlative with the decline in the lower part of the Pseudoradiosa Zone. This thereby constitutes a sixth potential correlative phase of diversity decline evident in North America, Europe and parts of the Arctic in the Late Toarcian. Foraminiferal species diversity, following the main-phase of extinction, shows a similar steady increase from the late Kanense to middle Crassicosta zones with a very small (negligible?) decline across the Planulata\u00E2\u0080\u0093Crassicosta Zone boundary that is correlative with phase number 4 in Figure 6.1. Diversity then declines gradually throughout the Late Toarcian. During the small decline in diversity at the Planulata / Crassicosta Zone boundary there is no observable trend in extinction rate (Figure 5.3) and therefore this does not appear to be a recognizable event (in the foraminiferal data). However, in the later part of the Toarcian, extinction rates are generally higher than origination which could account for the observed decline in diversity. 6.1.1 Correlation with the Karoo\u00E2\u0080\u0093Ferrar magmatism Currently this multi-phased extinction event is attributed to a variety of paleoenvironmental changes that could be related to the Volcanic Greenhouse Scenario but, except for the main-phase of extinction (number 3 in Figure 6.1), are believed to be restricted to the Tethys Ocean area (Dera et al., 2010). These 129 include: sea-level fluctuation affecting restricted basins, saline stratification, warming (or cooling) of seawater, changing water current dynamics and variation in geochemical cycles (McArthur et al., 2000; 2008; Bailey et al., 2003; Suan et al., 2008; 2010; Dera et al., 2010; Dera and Donnadieu, 2012). However, if the six extinction phases are global in extent, it is plausible that the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP is the underlying controlling factor. When comparing the approximate timing of this multi-phased event to the previously established, and calibrated, time scale for the Pliensbachian and Toarcian Stages (Figure 2.1), the middle Whiteavesi\u00E2\u0080\u0093middle Freboldi Zone event occurred at 186 Ma, the late Kunae\u00E2\u0080\u0093early Carlottense Zone event occurred at 184 Ma, the late Carlottense\u00E2\u0080\u0093middle Kanense Zone event occurred at 183 Ma, the late Planulata\u00E2\u0080\u0093early Crassicosta Zone event occurred between 182 and 181 Ma, the middle Crassicosta\u00E2\u0080\u0093Hillebrandti Zone event occurred between 181 and 180 Ma, and the Yakounensis Zone event occurred between 179\u00E2\u0080\u0093178 Ma. A study by Jourdan et al. (2008) shows eruption ages for magmatism in the Karoo Basin that occur over ~10 My, between 186 and 176 Ma, with the main pulses of magmatism occurring over a window of ~3 Ma from 184 to 181 My (Figure 6.2); with the main volume of basalt emplacement taking place over a 3 to 4.5 Ma window around 180 Ma. A comparison of eruption ages from the Karoo Basin and the multi-phased extinction event reveals a correlation, in that: 1) four of the six pulses of extinction occur within the main-phase of magmatism, 2) the initial Early Pliensbachian decline in species diversity occurs within error range of the onset of Karoo magmatism and 3) the latest Toarcian decline correlates with the 130 later stages of Karoo magmatism (Figure 6.2). This suggests that the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP was a major underlying factor in the long-term environmental change that resulted in the multi-phased extinction. During the Early Toarcian, a number of compounding factors could have simultaneously created a significant amount of environmental stress which induced episodes of methane release and marine anoxia, and escalated this (main) phase of extinction (as indicated in Hesselbo et al., 2000; 2007; Beerling et al., 2002; Jenkyns et al., 2001; Jenkyns, 2003; 2010; Jourdan et al., 2008; Dera et al., 2010 and references therein). 131 132 Figure 6.2 - Diagram comparing the timing of the multi-phased Pliensbachian\u00E2\u0080\u0093 Toarcian extinction with emplacement of the Karoo volcanic province. The 40Ar/39Ar age probability density distribution diagram and frequency histogram of volcanic rocks in the Karoo Basin are modified from Jourdan et al., (2008) to show its correlation in timing with the known phases of extinction. Figure suggests that all six phases of diversity decline occur within the duration of Karoo magmatism. Baj. = Bajocian. 133 6.2 The negative CIE interval & the long-term carbon- isotope record The principal line of evidence that supports the suspected global release of methane hydrate along continental margins is a ~5\u00E2\u0080\u00937\u00E2\u0080\u00B0 \u00CE\u00B413C change, which points to a large perturbation in the global carbon reservoir, identifiable as a negative CIE (Hesselbo et al., 2000; 2007; Kemp et al., 2005; Sabatino et al., 2009; Hermoso et al., 2009a; Suan et al., 2008). At the onset of this study presented herein, there was considerable debate surrounding the global extent and temporal correlation of this excursion. This debate is centered around whether or not the negative CIE, observable in parts of the NW European and Mediterranean Tethys Ocean, is suggestive of: 1) the sudden and global release of methane hydrate within a runaway greenhouse scenario (Hesselbo et al., 2000; 2007), or 2) regional upwelling of 12C-rich bottom water in various parts of the Tethys Ocean (van de Schootbrugge et al., 2005; Wignall et al., 2005; McArthur et al., 2008). The data presented herein from western North America will contribute to this debate by directly comparing the coeval geochemical records from an ocean basin that is geographically far removed from the Tethys Ocean. Haida Gwaii is part of the tectonically displaced Wrangellia terrane and as such its geographic position has varied through time. Ammonite data suggest that during the Late Pliensbachian, Wrangellia was located in the northeastern part of the paleo-Pacific Ocean (number 7 in Figure 1.11A) near the present Canada/US border (at roughly 49\u00C2\u00BA N latitude) at an unknown distance west of the North American margin (Smith, 2006). In the study presented herein, large 134 negative CIEs are identified within the Early Toarcian stratigraphy from the two sampled sections in Haida Gwaii, along the Yakoun River and at Whiteaves Bay (Figures 5.5, 5.7, 6.3). The discovery of chronologically well-constrained negative CIEs from the paleo-Pacific Ocean with similar profiles and magnitude to correlative European successions clearly demonstrates a global control. Of the two postulated mechanisms for the negative CIE, only methane release could reasonably cause such a global response. A defining characteristic, in the European data, is a broad positive shift in \u00CE\u00B413C that begins near the Pliensbachian / Toarcian boundary and extends into the Serpentinum Zone, but is interrupted by a pronounced negative excursion at the Tenuicostatum / Serpentinum Zone boundary (Figure 6.3C\u00E2\u0080\u0093E). Data from Haida Gwaii show a somewhat similar perturbation, particularly in the Whiteaves Bay section where there is a more definitive positive trend in carbon-isotope values that begins at the Pliensbachian / Toarcian boundary and is interrupted, with a large negative CIE, about midway through the Kanense Zone strata (Figure 5.5). In the Yakoun River section the Pliensbachian / Toarcian boundary is covered and therefore the onset of a possible positive carbon isotope excursion cannot be determined. 135 136 Figure 6.3 - Correlative Early to Middle Toarcian carbon-isotope data from western North America, NW Europe, and the Mediterranean (modified from Caruthers, 2011). (A and B) Kanense\u00E2\u0080\u0093Planulata Zone data from Whiteaves Bay and the Yakoun River sections on Haida Gwaii, British Columbia Canada; (C) Spinatum\u00E2\u0080\u0093Bifrons Zone data from the Mochras Borehole, Wales; (D) Spinatum\u00E2\u0080\u0093Falciferum Zone data from Sancerre\u00E2\u0080\u0093Couy Borehole, France; (E) Spinatum\u00E2\u0080\u0093Levisoni Zone data from Peniche, Portugal (after Hesselbo et al., 2007); (F) Spinatum\u00E2\u0080\u0093Falciferum Zone data from Yorkshire, UK. Pl. = Pliensbachian, Car. = Carlottense Zone, Spin. = Spinatum Zone. Furthermore, Haida Gwaii isotope data support the broad correlation of the Kanense Zone with the Tenuicostatum and Serpentinum ammonite Zones of the standard NW European scheme and suggests the approximate correlative position of the Tenuicostatum / Serpentinum boundary-equivalent within the Kanense Zone. Ammonites in western North America generally include endemic, pandemic, Tethyan, and Boreal taxa. However Boreal taxa, upon which the NW European ammonite zonal scheme is based, are somewhat rare and are confined to either the North American craton or the northern parts of displaced terranes (Smith and Tipper, 1986; Smith et al., 1988; Jakobs et al., 1994; Smith, 2006). This can cause difficulties when correlating between NW Europe and western North America. The presence of what appears to be an isochronous CIE in both areas is therefore an extremely useful aid in calibrating the North American ammonite zonal scheme with the primary standard scheme of NW Europe. Since publication of the initial Yakoun River carbon-isotope and TOC data from the current project (Caruthers et al., 2011), the negative CIE interval has been identified from other areas that are far-removed from the Tethys Ocean. To date, the negative CIE interval is now identifiable from correlative successions in 137 South America (Al-Suwaidi et al., 2010), Siberia (Suan et al., 2011), and Japan (Gr\u00C3\u00B6cke et al., 2011; Izumi et al., 2012). Together, these data strongly support the hypothesis of a sudden and catastrophic perturbation in the global carbon reservoir during the Early Toarcian (Hesselbo et al., 2000); most likely caused by the release of methane hydrate from continental margins. 6.2.1 The long-term carbon-isotope record In order to better assess the correlation of this extinction event with the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP, it is necessary to consider the effect that this eruption had on the carbon-isotope record. This is partially addressed by the methane release hypothesis in Hesselbo et al. (2000). In this work, Hesselbo et al. (2000) argue that the magnitude of the Early Toarcian negative CIE is too great to have been derived solely from volcanogenic outgassing of CO2 (with a \u00CE\u00B413C signature \u00E2\u0089\u0088 \u00E2\u0080\u00937\u00E2\u0080\u00B0), instead they suggest that this negative CIE more likely resulted from an additional release from the continental margin methane clathrate reservoir (\u00CE\u00B413C \u00E2\u0089\u0088 \u00E2\u0080\u009360\u00E2\u0080\u00B0) due to volcanogenic CO2-driven warming. However, in a more recent study, Deines (2002) suggests that mantle derived CO2 contains a bimodal \u00CE\u00B413C signature that peaks at \u00E2\u0080\u00935 and \u00E2\u0080\u009325\u00E2\u0080\u00B0, thereby re- affirming the possibility that volcanogenic outgassing of CO2 could create large perturbations in the \u00CE\u00B413C record. Furthermore, as we now understand, there are multiple events of species level diversity decline throughout the Pliensbachian\u00E2\u0080\u0093 Toarcian stages that also correlate (temporally) very well with the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP (Figures 6.1, 6.2). In order to gain a better perspective on the hypothesized environmental change, it is necessary to expand the scope of the 138 carbon-isotope record by observing its trend throughout the entire duration of the Karoo\u00E2\u0080\u0093Ferrar magmatism. To help understand the potential long-term affect of Karoo\u00E2\u0080\u0093Ferrar magmatism, new carbon-isotope data in western North America was compiled along with pre-existing published data (also from Haida Gwaii in western North America) in order to generate a long-term carbon-isotope profile that spans much of the Late Triassic (Norian) to Toarcian time (Figures 6.4, 6.5). Composite data from Williford et al. (2007) shows that in the Late Triassic, \u00CE\u00B413C is mostly constant at ~ \u00E2\u0080\u009329\u00E2\u0080\u00B0, at the Triassic / Jurassic boundary there is a large ~2\u00E2\u0080\u00B0 negative shift that is immediately followed by a large positive ~6\u00E2\u0080\u00B0 excursion that extends throughout the middle part of the Hettangian (Figure 6.4). Following this large positive excursion in carbon-isotope ratios, from the later part of the Hettangian to the Sinemurian, \u00CE\u00B413C returns to more negative values that are stabilized at ~ \u00E2\u0080\u0093 31\u00E2\u0080\u00B0 (Figure 6.4). 139 140 Figure 6.4 - Late Triassic to Early Jurassic (Sinemurian) carbon-isotope data for a section at Kennecott Point, Haida Gwaii (after Williford et al., 2007). In the Early Pliensbachian, at the bottom part of the composite Pliensbachian\u00E2\u0080\u0093Toarcian carbon-isotope profile (Figure 6.5), \u00CE\u00B413C values are similarly more negative (to Sinemurian values in Figure 6.4) at \u00E2\u0080\u009329\u00E2\u0080\u00B0 before showing a steady shift to heavier values. From the middle part of the Early Pliensbachian a very interesting pattern develops, in that \u00CE\u00B413C values fluctuate abruptly and frequently with a somewhat inconsistent magnitude that ranges ~3\u00E2\u0080\u0093 7\u00E2\u0080\u00B0 (Figure 6.5). This pattern of fluctuation in carbon-isotope values continues throughout the majority of the Pliensbachian, into the middle part of the Toarcian (upper part of the Planulata Zone) before returning to more negative values of ~ \u00E2\u0080\u009329\u00E2\u0080\u00B0. Interestingly this type of perturbation appears to be recognizable at a global scale. A recent study by Silva et al. (2011) shows a carbon-isotope profile from the Lusitanian Basin of Portugal that contains a similar and correlative pattern of isotopic perturbation within the Early to Late Pliensbachian stratigraphy (Figure 6.6). When comparing this dataset from Portugal to the correlative compiled dataset from western North America (Figure 6.5), a series of small positive and negative shifts are evident. Correlative data shows three positive and five negative CIEs throughout the studied interval (Figure 6.6). Positive CIEs occur within the: Whiteavesi (Ibex) Zone, lower part of the Freboldi (Davoei) Zone, and upper part of the Freboldi Zone. Negative CIEs occur at the Whiteavesi / Freboldi Zonal boundary, within the Freboldi Zone and within the 141 Kunae Zone (Figure 6.6). During the Pliensbachian Portugal was located in the western part of the Tethys Ocean, Haida Gwaii was located in the eastern part of the paleo Pacific Ocean and northern Alaska was located in the paleo Arctic Ocean (Figure 1.11). Therefore this remarkably similar pattern of perturbation in the carbon reservoir is recorded in the coeval strata from three distant and geographically far-removed ocean basins and should be considered a global phenomenon. Perturbation, such as this, in the global carbon reservoir would consequently require a global controlling mechanism. 142 143 Figure 6.5 - Combined \u00CE\u00B413Corg values for the Pliensbachian and Toarcian from Haida Gwaii and Arctic Alaska, western North America. Data shows a series of small-scale excursions that extend throughout much of the interval, with a well- defined negative CIE in the Kanense Zone of the Early Toarcian. Figure 6.6 - A comparison of Early to Late Pliensbachian \u00CE\u00B413C values between Peniche, Portugal (from Silva et al., 2011) and western North America. Data shows similar correlative trends in carbon-isotope values between the western Tethys, paleo Pacific Ocean and paleo Arctic Ocean carbon reservoirs. Whit. = Whiteavesi, Car. = Carlottense, Sp. = Spinatum, N.A. = North America, Eur. = Europe. Payne and Kump (2007) examine such large perturbations in \u00CE\u00B413C with respect to the severe Permian\u00E2\u0080\u0093Triassic extinction event. In their study they assess the influence of LIP volcanism, oxidation of organic carbon, and oxidation of biogenic methane on the Late Permian\u00E2\u0080\u0093Middle Triassic global carbon record. This record shows a series of large shifts (of ~8\u00E2\u0080\u00B0 \u00CE\u00B413Ccarbonate in Figure 6.7) 144 through the Late Permian and Early Triassic that are similar to the variations throughout the Pliensbachian\u00E2\u0080\u0093Toarcian reported here (Figures 6.5, 6.6). This comparison is justified because both intervals have been associated with erupting flood basalt provinces, namely the Siberian Traps in the Late Permian\u00E2\u0080\u0093 Middle Triassic and the Karoo\u00E2\u0080\u0093Ferrar LIP in the Pliensbachian\u00E2\u0080\u0093Toarcian, which are hypothesized to have initiated events of enhanced global warming and environmental change (Hallam and Wignall, 1997; P\u00C3\u00A1lfy and Smith, 2000). In their carbon cycle box model, Payne and Kump (2007) used a computer simulation to investigate the potential effect that various environmental disturbances have on the Late Permian\u00E2\u0080\u0093Middle Triassic carbon-isotope record. Throughout this interval, the carbon-isotope record shows a series of large-scale shifts, whereby small negative CIEs are followed by larger positive excursions (Figure 6.7). They argue that, although the release of methane (in the Volcanic Greenhouse Scenario in Figure 1.2) cannot be ruled out for any one excursion, this record of repeated negative CIEs is incompatible with this causal factor for all of the observed negative CIEs (Payne and Kump, 2007). This is because there is insufficient time available between events with which to regenerate the clathrate reservoir (Payne and Kump, 2007). Instead they look at other factors within the Volcanic Greenhouse Scenario (Figure 1.2) as potential causes for these shifts in the carbon-isotope record. They suggest that volcanogenic outgassing of 30,000 GT CO2 during Siberian Trap volcanism would produce the observed pattern (Figure 6.7). As mentioned above, methane hydrate has a \u00CE\u00B413C value of approximately \u00E2\u0080\u009360\u00E2\u0080\u00B0 whereas volcanogenic outgassing of CO2 carries a bimodal 145 \u00CE\u00B413C signature of approximately \u00E2\u0080\u00935 and \u00E2\u0080\u009325\u00E2\u0080\u00B0 (Deines, 2002). Therefore a negative CIE from volcanogenic outgassing of CO2 (during LIP eruption) would have less of an effect on the carbon-isotope record, and would produce a smaller magnitude negative CIE, than would a release of methane hydrate. Furthermore, Siberian Trap volcanism is thought to occur over a similar interval of time (Figure 1.3B), as with the observed disturbance in the carbon-isotope record, and therefore supports this theory. The Volcanic Greenhouse Scenario also helps to explain the multiple positive excursions that are observed within the Early Triassic record. Regarding these large positive excursions, Payne and Kump (2007) suggest that they were caused by an increase in productivity within the marine environment. As discussed in Chapter 1.1, global warming (from increased volcanogenic CO2 during LIP eruption) causes: 1) increased continental weathering and runoff, which increases the supply of nutrients into the marine environment and, in turn, increases primary productivity (Wignall, 2005), 2) anoxic marine water which enhances the regeneration of phosphate and, this too, intensifies the amount of primary productivity in the water column (Payne and Kump, 2007). This corresponding production of increased biomass preferentially absorbs C12 from the water column and causes the observed positive excursion in the carbon- isotope record (Payne and Kump, 2007). However, in their model, they were unable to produce multiple positive and negative excursion couplets, as observed in the Late Permian\u00E2\u0080\u0093Middle Triassic carbon-isotope record (Figure 6.7). Therefore they conclude that the observed pattern in carbon-isotope profile can 146 only be accounted for through several pulses of carbon release from volcanogenic sources. Figure 6.7 - Composite Late Permian\u00E2\u0080\u0093Middle Triassic carbon-isotope record from southern China (generalized from Payne and Kump, 2007). Data shows a series of excursions, thought to have been largely the result of large-scale outgassing of volcanogenic CO2 during Siberian Trap magmatism. The eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP, on the other hand, is thought to have produced ~9,000 GT CO2 (Beerling and Brentnall, 2007), which is much less than the estimated 30,000 GT CO2 (Payne and Kump, 2007) produced during Siberian Trap volcanism. However the change in isotope ratios in the Early Pliensbachian to Middle Toarcian successions is much less than those recorded in the Late Permian\u00E2\u0080\u0093Middle Triassic (Figures 6.5, 6.6, 6.7). Therefore it is suggested that volcanogenic outgassing of CO2 during the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP could have produced a series of small perturbations in the 147 carbon-isotope record, reflecting a smaller volume of CO2 release, that began in the middle of the Early Pliensbachian and continued until the Planulata Zone of the Middle Toarcian (Figure 6.5). The large negative CIE occurring in the Early Toarcian Kanense Zone is much larger than the other, smaller, fluctuations in \u00CE\u00B413C that occur before and after this event (Figure 6.5). As previously discussed, this negative CIE is thought to have resulted from a catastrophic release from the methane hydrate reservoir. Therefore, a suspected mechanism chain for this event would infer that volcanogenic outgassing of CO2 from the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP initiated greenhouse conditions and global warming during the Pliensbachian\u00E2\u0080\u0093 Toarcian interval. In the Early Toarcian a tipping point was reached, whereby ~5,000 GT of methane hydrate was released along the continental margins. This large injection of methane hydrate greatly accelerated greenhouse conditions and further escalated the extinction of marine organisms. 6.3 Global vs. regional marine anoxia The sudden release of methane hydrate along continental margins is also suspected to have contributed toward a globally extensive body of anoxic marine water, dubbed the T\u00E2\u0080\u0093OAE (Jenkyns, 1988). As discussed previously in Chapter 1, the primary line of evidence supporting the T\u00E2\u0080\u0093OAE stems from geochemical studies in northwest Europe and in parts of the Mediterranean. In these studies, primarily in northwest Europe, isotopic perturbation (Figures, 1.9, 1.10) typically co-occurs with organically enriched shale that, in places, reach as high as 15% TOC (Jenkyns, 1985; 1988; Baudin et al., 1990a,b) and is largely thought to be 148 the result of marine anoxia (Jenkyns, 1988; 2010). In particular, it has been argued that excess burial of organic carbon, resulting from marine anoxia (T\u00E2\u0080\u0093 OAE), created a broad positive excursion in the carbon-isotope profile throughout the correlative Tenuicostatum / Serpentinum Zones (Jenkyns and Clayton, 1986; Jenkyns, 1988; Jenkyns et al., 2001). This broad positive excursion is interrupted at a level that is correlative with the Tenuicostatum / Serpentinum zone boundary by the large negative CIE interval (Hesselbo et al., 2000), which is also correlative with the main-phase of marine extinction (number 3 in Figure 6.1). Interestingly, when looking at the averaged carbon-isotope values from sections in western North America (Figure 6.5), a similar trend emerges with a broad positive excursion extending from latest Pliensbachian to the middle of the Planulata Zone that is interrupted by a large negative CIE in the Early Toarcian. However, as previously discussed, a positive excursion in carbon-isotopes may also be derived through a rebounding effect from increased productivity during volcanogenic outgassing of CO2, derived primarily from outgassing during LIP eruption (Payne and Kump, 2007). Therefore, with respect to this particular trend in the carbon-isotope record, the ultimate cause may be difficult to differentiate between the observed effects from outgassing during LIP eruption versus the effects from marine anoxia. Another, more definitive, indication of changing redox conditions in the Tethys Ocean area (to a more anoxic environment) is perturbation in isotopic systems such as nitrogen, manganese, sulfate and molybdenum, also discussed in Chapter 1 (Figures 1.2, 1.10). However, a comprehensive record for these 149 isotope systems outside the Tethys Ocean area is not yet available. Moreover, a clear record of Early Toarcian carbon-isotope perturbation co-occurring with high TOC in areas outside the Tethys Ocean has not yet been established, except for two sections from northern Siberia where TOC reached 2% and 6% within the negative CIE interval (Suan et al., 2011). Consequently, questions remain concerning the actual geographic extent of the T\u00E2\u0080\u0093OAE. Total organic carbon, nitrogen-isotope and TN data compiled for the Pliensbachian and Toarcian of western North America sheds some light on this question. 6.3.1 Total organic carbon Pederson and Calvert (1990) address the importance of organic matter supply to the sea floor in the context of bottom water anoxia, suggesting that increased primary productivity is the first-order controlling factor for the accumulation of organic-rich facies in the modern sediments of the Black Sea and that these sediments are not particularly enriched with organic carbon. This is particularly relevant with respect to the hypothesized global geographic extent of the T-OAE, which is evidenced primarily by a suspected coeval record showing elevated TOC within the Early Toarcian stratigraphy (Jenkyns, 1988; 2010 and references therein). Throughout much of the Pliensbachian and Toarcian (excluding the Early Pliensbachian) in western North America, TOC values are consistently low and range generally between 0.5% and 1.5% (Figure 6.8). These values are more in line with modern hemipelagic shelf sediments (McIver, 1975; Table 1 in Tyson and Pearson, 1991). However, at certain intervals throughout the Pliensbachian\u00E2\u0080\u0093 Toarcian, TOC values appear elevated and reach concentrations of ~1.5% to 150 3.5%. These include the: 1) Early Pliensbachian (Imlayi and Whiteavesi Zones), 2) Late Pliensbachian (Kunae Zone), 3) the earliest Toarcian (below the negative CIE interval), and 4) within the negative CIE interval (Figure 6.8). Within the negative CIE interval, concentrations of TOC are much lower than the coeval data from northwest Europe (at ~15% TOC, as referenced above), and are also much lower than the other intervals (of elevated TOC) in western North America. Additionally, many of these intervals with elevated TOC (in Figure 6.8) are often only comprised of a couple data points that contain higher concentrations of organic carbon. This contrasts with the TOC data from the Early Pliensbachian of western North America where TOC values are consistently higher (Figure 6.8). Sabatino et al. (2009) report TOC values in pelagic shales of the western Tethys that are elevated, but much lower (< 2%) than those from northern Europe. They suggest that in deeper water environments a variety of factors could have led to a degradation of organic matter in the water column prior to deposition, resulting in lower organic carbon values. Furthermore, a recent comparison of carbon-isotope data from the Paris (France) and Lusitanian (Portugal) basins demonstrates a negative CIE that is decoupled from black shale deposition (Hermoso et al., 2009b). These studies along with this new TOC data from the paleo-Pacific Ocean therefore suggest that organic carbon concentration seems to be a factor that is controlled locally and should not be used to characterize global marine anoxia during the Early Toarcian. Data from Haida Gwaii does not show significant organic carbon enrichment occurring in 151 the same interval as the negative CIE. While this does not rule-out marine anoxia in the paleo-Pacific Ocean, it also does not support it. 152 153 Figure 6.8 - Combined Total Organic Carbon for Pliensbachian and Toarcian samples from Haida Gwaii and Arctic Alaska, western North America. 6.3.2 Nitrogen-isotope data Early Toarcian geochemical data from northern Europe (British Isles) and southern Europe (Italy) show a pronounced positive excursion in \u00CE\u00B415N that generally ranges in magnitude ~3\u00E2\u0080\u00936\u00E2\u0080\u00B0 (Figure 1.10) and correlates with elevated TOC in the negative CIE interval (Jenkyns et al., 2001). This enrichment in nitrogen-isotope ratios is interpreted by Jenkyns et al. (2001) to represent upwelling of a deoxygenated water mass that had undergone partial denitrification. As discussed previously in Chapter 1.1, degradation (oxidation) of organic matter is normally accomplished through the consumption of oxygen, which produces abundances of carbon dioxide, nitrate, and phosphate (Froelich et al., 1979). However, in times of environmental stress such as in an event of global warming, oxygen becomes less abundant or is absent from the environment (Figure 1.2). The microbially mediated system then begins to consume nitrate in order to continue this process (Froelich et al., 1979; Jenkyns et al., 2001). Its consumption from the water column is referred to as \u00E2\u0080\u0098denitrification\u00E2\u0080\u0099. Jenkyns et al. (2001) provides further evidence of the advancement of this process by arguing that in the correlative Exaratum subzone of northwest Europe and Italy, denitrification occurred just before the presence of isorenieratane (Schouten et al., 2000), which is a biomarker compound derived from phototrophic sulfur bacteria. The presence of small pyrite framboids within laminated carbon-rich shales is also indicated from Yorkshire (UK) and the 154 Belluno Trough (Italy), which is further indicative of euxinic conditions in the water column (Wilkin et al., 1996; Bellanca et al., 1999; Jenkyns et al., 2001; Wignall et al., 2005; Jenkyns et al., 2007; Jenkyns, 2010). Jenkyns et al. (2001) therefore conclude that denitrification (and subsequent sulfate reduction) probably resulted from enhanced productivity and that the observed positive excursion in nitrogen-isotopes required the upwelling of 15N-rich water, which was then consumed by organic-walled phytoplankton. This suggestion of euxinic conditions in the northern European epeiric and marginal Tethyan carbonate- platform setting is further supported by sulfate-S isotope data in Gill et al. (2011); showing increases in sulfur isotope ratios of sulfate that are coeval with positive excursions in carbon-isotope values across the suspected anoxic interval. In comparison with European data, nitrogen-isotope data from western North America does not show a positive excursion in \u00CE\u00B415N within the Early Toarcian negative CIE interval (Figure 6.9). In fact, \u00CE\u00B415N values show a general fluctuation between approximately \u00E2\u0080\u00931.0\u00E2\u0080\u00B0 and 0.25\u00E2\u0080\u00B0 throughout much of the Toarcian with occasional shifts of +/- 1\u00E2\u0080\u00B0 magnitude at various intervals. Throughout the negative CIE interval, \u00CE\u00B415N values show two successive (albeit of small magnitude) negative excursions that reach minimum values of \u00E2\u0080\u00931.50\u00E2\u0080\u00B0 and \u00E2\u0080\u00931.27\u00E2\u0080\u00B0 respectively (Figure 6.9). As discussed by Jenkyns et al. (2001), typical Jurassic and Cretaceous marine shale has \u00CE\u00B415N values that range between \u00E2\u0080\u00932.5 to 4\u00E2\u0080\u00B0, which encompasses the observed values from western North America for the Toarcian. In modern marine sediments, \u00CE\u00B415N ranges approximately from 4 to 10\u00E2\u0080\u00B0 (Altabet 155 and Francois, 1994; Figure 2 in Robinson et al., 2012) with denitrification (\u00CE\u00B415N of NO3-) producing mean values within anoxic water masses that reach 19\u00E2\u0080\u00B0 in the eastern equatorial Pacific Ocean (Tesdal et al., 2013). Therefore it seems to be evident with respect to the Toarcian \u00CE\u00B415N data that the northeast paleo-Pacific Ocean was not experiencing denitrification. If denitrification was not occurring, then it is possible that the Toarcian sediments were potentially deposited under more normal redox conditions, which is somewhat contrary to hypotheses suggesting a globally extensive T\u00E2\u0080\u0093OAE. If correct then it could be surmised that the release of methane hydrate and the subsequent acceleration in greenhouse conditions (during the Early Toarcian) created anoxic water in only certain marine environments (Figure 6.10), through mechanisms related to the basin-restriction model of McArthur et al. (2008). These restricted basins, containing anoxic marine water, could have been distributed globally and were most-likely not solely tied to the Tethys Ocean area. 156 157 Figure 6.9 - Combined \u00CE\u00B415N values for the Early Pliensbachian and Toarcian from Haida Gwaii, western North America. Figure shows nitrogen-isotope values that are more positive in the Early Pliensbachian and values that are slightly negative to near 0.0\u00E2\u0080\u00B0 throughout much of the Toarcian. Figure 6.10 - Adapted flow chart for the Volcanic Greenhouse Scenario (modified from Wignall, 2001 and 2005) showing the potential effects of global warming in restricted vs. unrestricted basins (orange diamond). \t\r \u00C2\u00A0 6.3.2.1 Regional denitrification during the Early Pliensbachian Cameron and Tipper (1985) suggest that the Ghost Creek Formation (including the Early Pliensbachian stratigraphy in Whiteaves Bay) was deposited in a euxinic deeper water basin. This is supported by its black colour, the 158 abundance of pyrite and organic carbon, as well as the presence of oil stains and bitumen. They also noted that benthic fossils and bioturbation were uncommon. Near the top of the Ghost Creek Formation, biodiversity and bioturbation increase, organic carbon seems to decrease (evidenced by a lighter coloured lithology), and it is therefore thought that sediments were deposited under more normal redox conditions (Cameron and Tipper, 1985). These interpretations are further corroborated by the low-resolution geochemical data from the Ghost Creek Formation at Whiteaves Bay. In contrast to the previously discussed data from the Toarcian stratigraphy, in the Early Pliensbachian: 1) the nitrogen-isotope dataset yield \u00CE\u00B415N values that are more positive and range generally between 0.0 and 1\u00E2\u0080\u00B0 (Figure 6.9), 2) TOC values are also consistently high and decrease substantially throughout the remaining part of the Pliensbachian\u00E2\u0080\u0093Toarcian interval (Figure 6.8) and 3) organic nitrogen concentrations (measured as Wt % total nitrogen or TN) are also consistently much higher, ranging ~0.075 and 0.2%, in comparison with the Toarcian values that are consistent at ~0.05% (Figure 6.11). Although these Early Pliensbachian \u00CE\u00B415N values are more positive than those recorded in the Early Toarcian, it should be noted that they are not as light as Toarcian values in European successions which are thought to indicate denitrification (~3\u00E2\u0080\u00B0 in Jenkyns et al., 2001). However these more positive Early Pliensbachian \u00CE\u00B415N values, in conjunction with the lithologic interpretations of Cameron and Tipper (1985), may infer that the sediments of the Ghost Creek Formation were potentially deposited in a suboxic water column. 159 As currently understood, there do not seem to be any accounts of global anoxic marine water in the Early Pliensbachian and therefore this instance might only be regionally extensive within part of the Wrangellia composite terrane. It seems reasonable to hypothesize that this episode of marine anoxia could have resulted from a stratified body of water within a locally restricted basin within the terrane. Perhaps, within this basin, the dynamics were such that water currents were weak or absent and could not adequately circulate the water column, thus producing denitrification and anoxia. Gradually, throughout the Early Jurassic, this basin was filled with coarser sedimentation of the Fannin Formation, during the later part of the Early Pliensbachian, and seawater stratification was no longer possible. The water column then returned to a normal oxidizing environment. 160 161 Figure 6.11 - Combined total nitrogen (TN) values for the Early Pliensbachian and Toarcian from Haida Gwaii, western North America. Data shows much greater levels of nitrogen in the Early Pliensbachian, in comparison with the Toarcian data. 162 Chapter 7 Systematic paleontology The following ammonoid taxa are described using the open nomenclature following Bengston (1988) and Smith and Tipper (1996) and are used to supplement: 1) the previously known biostratigraphy from the Talkeetna Mountains and Haida Gwaii and 2) the known stratigraphic ranges of Pliensbachian\u00E2\u0080\u0093Toarcian ammonoids in western North America, which are used in the extinction analysis of this study. See Figure 4.5 (reproduced below) for the distribution of ammonite species, identified here, according to geographic locality. Many of these taxa are well known to western North America and have been previously described in detail with extensive synonymy lists. Therefore the synonymy list, provided herein for each species, is abbreviated with reference for a more complete list given. Specimens described herein are curated at the University of Montana Paleontology Center (UMPC) and the Pacific Museum of the Earth (part of the Beaty Biodiversity Museum, University of British Columbia). 163 164 Figure 4.5 - Distribution of Pliensbachian and Toarcian ammonite taxa that are systematically described in Chapter 7 of this study, according to locality number at the Hicks Creek and Camp Creek sections (Figure 4.4), the Whiteaves Bay and Yakoun River sections (Figures 5.4, 5.6) and the South Barrow #3 Core (Figure 5.11). Ch. = Chapter, A.A. = Arctic Alaska, and SB3 = South Barrow #3 well core. Class Cephalopoda CUVIER, 1797 Order Ammonoidea ZITTEL, 1884 Suborder Ammonitina HYATT, 1889 Family Oxynoticeratidae HYATT, 1875 Genus Fanninoceras MCLEARN, 1930 Subgenus Fanninoceras MCLEARN, 1930 Type species: Fanninoceras fannini MCLEARN, 1930 Fanninoceras (Fanninoceras) carlottense MCLEARN, 1930 Pl. 7.1, figs 1\u00E2\u0080\u00936 1884. Sphenodiscus requienianus? D\u00E2\u0080\u0099ORBIGNY.\u00E2\u0080\u0093 WHITEAVES, p. 248, pl. 22, fig. 4. 1996. Fanninoceras (Fanninoceras) carlottense MCLEARN.\u00E2\u0080\u0093 SMITH & TIPPER, pl. 2, figs. 3\u00E2\u0080\u00937, text-figs. 30j, 31d (and synonymy therein). Description.\u00E2\u0080\u0093 Involute, rapidly expanding form with a very narrow umbilicus, compressed whorl section and acute venter without a keel. Inner whorls have gently sinuous prorsiradiate ribs that project onto the venter and become weak and disappear as shell diameter increases. Material.\u00E2\u0080\u0093 Nine specimens from localities 1, 2, & 4 in the Hicks Creek section, Talkeetna Mountains Alaska. 165 Discussion.\u00E2\u0080\u0093 F. carlottense is the most involute and stratigraphically highest ranging species of Fanninoceras in western North America and is distinguishable by its characteristically narrow umbilicus. F. carlottense has previously been reported from two isolated localities in the Talkeetna Mountains by IMLAY (1981) but it was not figured or placed within stratigraphic context. Occurrence.\u00E2\u0080\u0093 Reported from the East Pacific Realm from South America (Argentina and Chile) to western North America (Nevada, Oregon, British Columbia Canada, and Alaska). Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian). Fanninoceras (Fanninoceras) fannini MCLEARN, 1930 Pl. 7.1, figs 7\u00E2\u0080\u009319 1930. Fanninoceras fannini MCLEARN, p. 4, pl. 1, fig. 3. 1996. Fanninoceras (Fanninoceras) fannini MCLEARN.\u00E2\u0080\u0093 SMITH & TIPPER, pl. 3, figs. 1\u00E2\u0080\u009312; pl. 5, figs. 1, 2, text-figs. 27, 30d\u00E2\u0080\u0093e, 31a\u00E2\u0080\u0093c (and synonymy therein). Description.\u00E2\u0080\u0093 An involute form with a moderately wide umbilicus for the genus. Whorls are depressed with a broadly arched venter early in ontogeny becoming more compressed with an acute venter in later whorls. Ribs are well spaced, simple\u00E2\u0080\u0093slightly sinuous, projected onto the venter in early whorls and becoming weak with growth. Outer whorls are smooth. Material.\u00E2\u0080\u0093 Twenty five specimens from localities 6, & 8\u00E2\u0080\u009314 in the Camp Creek section and three specimens from locality 3 in the Hicks Creek section, Talkeetna Mountains Alaska. 166 Discussion.\u00E2\u0080\u0093 As discussed by SMITH and TIPPER (1996) there is a strong resemblance between F. fannini and F. carlottense but F. fannini has a larger umbilicus. Occurrence.\u00E2\u0080\u0093 F. fannini is common in the East Pacific Realm from South America to southern Alaska. Age.\u00E2\u0080\u0093 Primarily known from the Kunae Zone with rare occurrences in the Carlottense Zone (Late Pliensbachian). Subgenus Charlotticeras SMITH & TIPPER, 1996 Type species: Fanninoceras (Charlotticeras) carteri SMITH & TIPPER, 1996 Fanninoceras (Charlotticeras) cf. maudense SMITH & TIPPER, 1996 Pl. 7.1, figs 20\u00E2\u0080\u009322 cf. 1996. Fanninoceras (Charlotticeras) maudense SMITH & TIPPER, p. 32, pl. 6, figs. 6\u00E2\u0080\u009311, text-figs. 30a\u00E2\u0080\u0093b. Description.\u00E2\u0080\u0093 Midvolute form with a compressed whorl section, flat flanks, and incipient keel. Umbilical wall is low, sharp, and vertical. Ornamentation is well defined with coarse, sinuous, ribs that are prorsiradiate and project onto the venter to the incipient keel. Material.\u00E2\u0080\u0093 Two specimens from the Camp Creek section, one from locality 8 and another from locality 13, Talkeetna Mountains Alaska. Occurrence.\u00E2\u0080\u0093 Only known from British Columbia Canada (Haida Gwaii, formerly the Queen Charlotte Islands), and Alaska (Talkeetna Mountains). Age.\u00E2\u0080\u0093 Kunae Zone (Late Pliensbachian). 167 Family Dactylioceratidae HYATT, 1867 Genus Dactylioceras HYATT, 1867 Type species: Ammonites communis SOWERBY, 1815 Dactylioceras cf. compactum (DAGIS, 1968) Pl. 7.2, figs 1, 2 cf. 1968. Kedonoceras compactum DAGIS, pl. XI, figs 1\u00E2\u0080\u00933. Description.\u00E2\u0080\u0093 An evolute cadicone form with a depressed sub-circular whorl and a broad venter baring no keel. Rectiradiate ribs project from the umbilicus to the ventrolateral shoulder and become paired across the venter. Material.\u00E2\u0080\u0093 One specimen from Locality #2 in the Yakoun River section, Haida Gwaii British Columbia. Discussion.\u00E2\u0080\u0093 This species is originally identified from the Propinquum Zone of northeast Russia. Its occurrence in the Yakoun River section is just above the species Protogrammoceras cf. paltum, which suggests a similar age to the Russian specimen. Occurrence.\u00E2\u0080\u0093 Currently this species is only known from Russia and western Canada. Age.\u00E2\u0080\u0093 Kanense Zone (Early Toarcian). Family Amaltheidae HYATT, 1867 Genus Amaltheus DE MONTFORT, 1808 168 Type species: Amaltheus margaritatus DE MONTFORT, 1808 Amaltheus sp. Pl. 7.1, figs 23\u00E2\u0080\u009325 Description.\u00E2\u0080\u0093 An involute form with a compressed whorl section and convex flanks. The acute venter bears a low keel that consists of distinct forwardly directed chevrons. A sharp umbilical wall is slightly undercut. Sinuous primary ribs are prorsiradiate and fade toward the venter. Secondary ribs are much less defined and fade quickly toward the venter. Material.\u00E2\u0080\u0093 Two poorly preserved specimens from Locality #14 in the Camp Creek section, Talkeetna Mountains Alaska. Two poorly preserved specimens from Localities #1 & #2 in the South Barrow #3 well core. Discussion.\u00E2\u0080\u0093 Specimens resemble the well known Boreal Realm species Amaltheus stokesi (SOWERBY), discussed in detail by HOWARTH (1958). However poor preservation precludes positive identification. IMLAY (1981) also identifies A. stokesi from an isolated locality in the Talkeetna Mountains without stratigraphic context as well as in other intervals within the South Barrow #3 well core. Occurrence.\u00E2\u0080\u0093 Amaltheus is distributed throughout the Boreal Realm including: Canada, Russia, northwest Europe, and Alaska. It is also known from localities with mixed Tethyan/Boreal faunas including: Canada, Alaska, Italy, Hungary, Japan, and Russia. Age.\u00E2\u0080\u0093 Kunae Zone (Late Pliensbachian). Family Hildoceratidae HYATT, 1867 169 Genus Arieticeras SEGUENZA, 1885 Type species: Ammonites algovianus OPPEL, 1862. Lectotype and paralectotype (the latter refigured from SCHR\u00C3\u0096DER, 1927) designated and figured by WIEDENMAYER (1977). Arieticeras aff. domarense (MENEGHINI, 1867) Pl. 7.1, figs 26\u00E2\u0080\u009331 1867. Ammonites (Harpoceras) domarensis MENEGHINI, p. 7. 1996. Arieticeras aff. domarense (MENEGHINI).\u00E2\u0080\u0093 SMITH & TIPPER, pl. 20, fig. 4; text-fig. 37a (and synonymy therein). 2007. Arieticeras gr. domarense (MENEGHINI).\u00E2\u0080\u0093 MOUTERDE, DOMMERGUES, MEISTER & ROCHA, p. 88, pl. 6, fig. 5. Description.\u00E2\u0080\u0093 Evolute and slowly expanding with a wide umbilicus, compressed, subquadrate, whorl section and sharp keel. Simple to slightly flexuous rectiradiate ribs are most prominent along the flank and disappear at the shoulder. Material.\u00E2\u0080\u0093 Seventeen specimens from localities 7, 8, 14, & 15 in the Camp Creek section, Talkeetna Mountains Alaska. Occurrence.\u00E2\u0080\u0093 Known from the northern Mediterranean area, Oregon, Nevada, British Columbia Canada (Haida Gwaii), and Alaska (Talkeetna Mountains). Age.\u00E2\u0080\u0093 Kunae Zone (Late Pliensbachian). Genus Leptaleoceras BUCKMAN, 1918 Type species: Leptaleoceras leptum BUCKMAN, 1918 170 Leptaleoceras? sp. Pl. 7.1, figs 32, 33 Description.\u00E2\u0080\u0093 Midvolute to evolute with a compressed whorl section bearing a keel. Ribs are rectiradiate, sinuous and are densely spaced along the flank. The specimen figured on Pl. 7.1 fig. 33, shows incipient bundling of ribs near the umbilicus. Material.\u00E2\u0080\u0093 Three specimens from locality 6 in the Camp Creek section, Talkeetna Mountains Alaska. Discussion.\u00E2\u0080\u0093 All the specimens are small. The densely spaced ribbing suggests an assignment of Leptaleoceras but the bundling of ribs on one specimen suggests that Canavaria may be represented. Occurrence.\u00E2\u0080\u0093 Only known from British Columbia (Haida Gwaii), and Alaska (Talkeetna Mountains). Age.\u00E2\u0080\u0093 Kunae Zone (Late Pliensbachian). Genus Protogrammoceras SPATH, 1913 Subgenus Protogrammoceras SPATH, 1913 Type species: Grammoceras bassanii FUCINI, 1901 Protogrammoceras cf. paltum (Buckman, 1922) Pl. 7.2, figs 3\u00E2\u0080\u00936 cf. 1922. Paltarpites paltus BUCKMAN, pl. 362A. 1981. Protogrammoceras cf. paltum (Buckman) IMLAY, pl. 12, figs 11, 12. 171 cf. 1992. Protogrammoceras (Protogrammoceras) paltum (Buckman) HOWARTH, pl. 1, figs 1\u00E2\u0080\u00933; pl. 2, fig 1 (and synonymy therein). 1996. Protogrammoceras (Protogrammoceras) cf. paltum (Buckman), SMITH & TIPPER, pl. 24, figs 1\u00E2\u0080\u00934 (and synonymy therein). Description.\u00E2\u0080\u0093 A poorly preserved, evolute (midvolute?) form with a compressed ellipsoidal or ogivale whorl section, wide umbilicus and a keel. Evenly spaced falcoid ribs project along the flank to the ventrolateral shoulder, where they disappear before reaching the keel. Ribs are well pronounced and evenly spaced. Material.\u00E2\u0080\u0093 Four specimens in total, three from Localities #5 & #6 in the Whiteaves Bay section and one from Locality #1 in the Yakoun River section on Haida Gwaii. Discussion.\u00E2\u0080\u0093 The small size and poor preservation of these specimens precludes positive identification. Occurrence.\u00E2\u0080\u0093 This species occurs throughout much of Europe and parts of North America where it has been identified from British Columbia (Frebold, 1970; Thompson and Smith, 1992; Smith and Tipper, 1996), Arctic Canada (Hall and Howarth, 1983) and Alaska (Imlay, 1981). Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian) to Kanense Zone (Early Toarcian). Genus Lioceratoides SPATH, 1919 172 Subgenus Lioceratoides SPATH, 1919 Type species: Leioceras? Grecoi FUCINI, 1901, p. 91, pl. 11, figs. 4, 5, by original designation (SPATH, 1919, p. 174). Lioceratoides (Lioceratoides) cf. involutum SMITH & TIPPER, 1996 Pl. 7.1, figs 34\u00E2\u0080\u009336 cf. 1996. Lioceratoides (Lioceratoides) involutum SMITH and TIPPER, pl. 26, figs. 2\u00E2\u0080\u00934; text-fig. 39p. Description.\u00E2\u0080\u0093 A midvolute to involute, rapidly expanding form with a prominent keel, compressed whorl section with flat flanks, and an abrupt umbilical wall. Volution tends to be more midvolute on outer whorls. Ribs are densely spaced, sinuous, and occasionally bundled forming irregularities in appearance. Material.\u00E2\u0080\u0093 Two specimens from locality 3 in the Hicks Creek section, Talkeetna Mountains Alaska. Occurrence.\u00E2\u0080\u0093 British Columbia Canada (Haida Gwaii) and Alaska (Talkeetna Mountains). Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian). Lioceratoides (Lioceratoides) cf. allifordense (MCLEARN, 1930) Pl. 7.2, fig 7 cf. 1930. Harpoceras allifordense, MCLEARN, pl. 2, fig 1. cf. 1932. Harpoceras allifordense, MCLEARN, pl. 5, figs 1\u00E2\u0080\u00933. cf. 1964. Harpoceras allifordense, FREBOLD, pl. 8, fig 5 (holotype refigured). 173 cf. 1996. Lioceratoides (L.) allifordense, SMITH & TIPPER, pl. 26, figs 5, 6. Description.\u00E2\u0080\u0093 Moldic impression of a midvolute form with an oval whorl, wide umbilicus, convex flanks, and narrow venter baring a keel. Fasciculate ribs are coarse with rounded tops on the inner whorls and become more sinuous and flat topped with fine intercalated secondary ribs on the outer whorl. Material.\u00E2\u0080\u0093 One specimen from Locality #4 in the Whiteaves Bay, Haida Gwaii. Discussion.\u00E2\u0080\u0093 This specimen resembles L. allifordense from a section on Maude Island, Haida Gwaii, in SMITH & TIPPER (1996 pl. 26, fig. 5) with respect to the sinuous fasciculate ribs that become flat topped and intercalated with growth. However, this specimen from Whiteaves Bay is poorly preserved and incomplete and precludes a positive identification. Occurrence.\u00E2\u0080\u0093 This species is only known from Haida Gwaii, British Columbia Canada. Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian) to Kanense Zone (Early Toarcian). Subgenus Pacificeras REPIN, 1970 Type species: Schloenbachia propinqua WHITEAVES, 1884, p. 274, pl. 33, fig. 2. Lioceratoides (Pacificeras) cf. propinquum (WHITEAVES, 1884) Pl. 7.2, figs 8\u00E2\u0080\u009310 cf. 1884. Schloenbachia propinqua WHITEAVES, pl. 33, figs 2, 2a. 174 cf. 1996. Lioceratoides (Pacificeras) propinquum (Whiteaves) SMITH & TIPPER, pl. 28, figs 1\u00E2\u0080\u009311; pl. 29, fig. 1 (and synonymy therein). Description.\u00E2\u0080\u0093 Evolute with an ogivale whorl section, flat or slightly convex flanks, abrupt umbilical shoulder and a simple keel. Coarse widely spaced ribs are sparse and occasionally paired on inner whorls, becoming more evenly distributed and slightly sinuous with growth. Material.\u00E2\u0080\u0093 Two specimens from Localities #1 & #2 in the Whiteaves Bay section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 The coarse widely spaced ribs on the inner whorls distinguishes this species, as indicated by SMITH & TIPPER (1996; p. 72). This feature is also evident on similar specimens from Haida Gwaii (SMITH & TIPPER, 1996; pl. 28 figs 5b, 7b, 9a, and 11a). Occurrence.\u00E2\u0080\u0093 As indicated in SMITH & TIPPER (1996), this species is only known from the northeast Pacific. Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian) to Kanense Zone (Early Toarcian). Genus Tiltoniceras BUCKMAN, 1913 Type species: Tiltoniceras costatum BUCKMAN, 1913, p. 8. Tiltoniceras cf. antiquum (WRIGHT, 1882) Pl. 7.2, figs 11, 12 cf. 1882. Harpoceras antiquum WRIGHT, pl. 57, figs 1, 2 (non figs 3, 4) 175 cf. 1992. Tiltoniceras antiquum (Wright) HOWARTH, Text-fig. 13, pl. 6, figs 1\u00E2\u0080\u00938 (and synonymy therein). cf. 1996. Tiltoniceras antiquum (Wright) SMITH & TIPPER, pl. 30, figs 1\u00E2\u0080\u00934; Text-fig. 39 l, m (and synonymy therein). Description.\u00E2\u0080\u0093 An evolute form with a compressed shell, oval whorl, flat flanks and a narrow venter baring a prominent keel that extends to the ventrolateral shoulder. Inner whorls have coarse ribs that are slightly sinuous and evenly spaced, and become very fine to non-existent quickly with growth. Material.\u00E2\u0080\u0093 Two specimens from Localities #3 & #6 in the Whiteaves Bay section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 Poor preservation of these specimens precludes positive identification. However in the larger specimen (Pl. 7.2, Fig. 12), the compressed whorl and prominent keel is similar to specimens in HOWARTH (1992; text-fig. 13, pl. 6, figs. 1, 7) and SMITH & TIPPER (1996; pl. 30, fig. 1). The sutures that appear above the outer whorl fragment are thought to be from another poorly preserved ammonite and not part of this specimen. Occurrence.\u00E2\u0080\u0093 This species has been described from northwest Europe, Russia, and Haida Gwaii. Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian) to Kanense Zone (Early Toarcian). Genus Harpoceras WAAGEN, 1869 176 Type species: Ammonites falcifer SOWERBY, 1820, p. 99, pl. 254, fig. 2. Harpoceras cf. nitescens (SCHLEGELMILCH, 1976) Pl. 7.8, figs 1 cf. 1976. Harpoceras nitescens SCHLEGELMILCH, pl. 46, fig 1. Description.\u00E2\u0080\u0093 An evolute, compressed whorl fragment with an ellipsoidal whorl section, flat flanks and a narrow carinate venter baring a keel. Ornamentation consists of coarse, gently flexuous and evenly spaced, primary ribs that are surrounded on either side by fine secondaries. Ribs project along the flank, coarsening slightly toward the venter and terminating at the ventrolateral shoulder. Material.\u00E2\u0080\u0093 One specimen from Locality #10 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 This specimen also resembles H. eseri in SCHLEGELMILCH (1976; pl. 46, fig. 2). However H. eseri has ribs that are occasionally paired at the umbilical shoulder, which is not evident in this specimen and in the material in SCHLEGELMILCH (1976). Occurrence.\u00E2\u0080\u0093 This species is known from northwest Europe and Haida Gwaii in western North America. Age.\u00E2\u0080\u0093 Planulata Zone (Middle Toarcian). Genus Cleviceras HOWARTH, 1992 177 Type species: Ammonites exaratus YOUNG & BIRD, 1828, p. 266. Cleviceras exaratum (YOUNG & BIRD, 1828) Pl. 7.3, figs 1, 4, 5 See HOWARTH (1992) and JAKOBS (1997) for a complete list of synonymies. 1828. Ammonites exaratus YOUNG & BIRD, p. 266. 1992. Cleviceras exaratum (Young & Bird) HOWARTH p. 90, pl. 9, figs 2\u00E2\u0080\u00936; pl. 10, figs 1\u00E2\u0080\u00938; pl. 11, figs 1\u00E2\u0080\u009317; pl. 12, figs 1\u00E2\u0080\u00935; pl. 13, figs. 1, 2; text figs 10, 16, 18C, 19C, 20, 21 (and synonymy therein). 1997. Cleviceras cf. exaratum (Young & Bird) JAKOBS pl. 3, figs. 6, 7, 12, 13; pl. 4, figs. 3, 4 (and synonymy therein). Description.\u00E2\u0080\u0093 An evolute (midvolute?) form with an ellipsoidal whorl section, rounded umbilical and ventrolateral shoulders, wide umbilicus and convex flanks. Falcoid ribs are evenly distributed and project onto the venter fading just prior to the sharp keel. Ribs are prorsiradiate and fine close to the umbilical shoulder and then change direction at approximately mid-flank and become gently flexuous and coarse with rounded tops as they project toward the venter. Material.\u00E2\u0080\u0093 Two specimens from Localities #8 & #9 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 These specimens strongly resemble C. cf. exaratum in JAKOBS (1997) which (in part) was also collected from the same section. The falcoid ribbing that changes direction, abruptly, at approximately mid-flank strongly resembles that of C. exaratum in HOWARTH (1992, pl. 10, fig. 1\u00E2\u0080\u00938) and differs 178 from Harpoceras serpentinum and H. falciferum (HOWARTH, 1992, pl. 19, figs 1\u00E2\u0080\u0093 4; pl. 20, figs. 1\u00E2\u0080\u009311) in that the mid-flank flexure in rib direction (in C. exaratum) is not as pronounced as in H. serpentinum and H. falciferum (e.g. falcoid ribbing as opposed to falcate ribbing). Occurrence.\u00E2\u0080\u0093 This species is known from northwestern Europe, Siberia and Japan. In western North America it is known from the Canadian Rockies as well as parts of the Cordilleran region including Haida Gwaii. Age.\u00E2\u0080\u0093 upper part of the Kanense Zone (Early Toarcian) to Planulata Zone (Middle Toarcian). Cleviceras spp. Pl. 7.3, figs 2, 3, 6, 7 Description.\u00E2\u0080\u0093 A midvolute form with an ellipsoidal\u00E2\u0080\u0093sub-quadrate whorl, a wide umbilicus, abrupt umbilical shoulder, convex flanks and a broad venter baring a sharp keel. Ribs on inner whorls are coarse, evenly spaced, and sinuous. On the outer whorl ribs become falcoid and are fine from the umbilical shoulder to \u00C2\u00BC flank, where they abruptly change direction become flexuous and coarse with round tops, projecting towards the venter and terminating at the shoulder. Material.\u00E2\u0080\u0093 Two specimens from Localities #3 & #6 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 The fragmentary nature and poor preservation of this specimen precludes further identification within the genus. This specimen is similar to C. 179 exaratum with respect to the falcoid ribbing on the outer whorl, but differs with respect to the position of the bend in rib direction along the flank. In C. exaratum, this occurs at approximately \u00C2\u00BD flank and in this specimen the flexure in rib direction occurs at approximately \u00C2\u00BC flank. When comparing with specimens in HOWARTH (1992, pl. 10, figs 1\u00E2\u0080\u00938), specimens from Haida Gwaii seem to have a wider umbilicus with coarser ribs on inner whorls. Age.\u00E2\u0080\u0093 Kanense Zone (Early Toarcian). Genus Hildaites BUCKMAN, 1921 Type species: Hildaites subserpentinus BUCKMAN, 1921, p. 217. Hildaites murleyi (MOXON, 1841) Pl. 7.4, figs 1\u00E2\u0080\u00934 1841. Ammonites murleyii MOXON, pl. 24, figs. 6. 1992. Hildaites murleyi (Moxon) HOWARTH, pl. 30, figs 9, 10; pl. 31, figs 1\u00E2\u0080\u00938; pl. 32, fig 4 (and synonymy therein). 1994. Hildaites murleyi (Moxon) JAKOBS, SMITH & TIPPER, pl. 1, figs 15\u00E2\u0080\u009318. 1997. Hildaites murleyi (Moxon) JAKOBS, pl. 5, figs 1\u00E2\u0080\u00939. Description.\u00E2\u0080\u0093 An evolute\u00E2\u0080\u0093midvolute (?) form with a sub-rectangular to ellipsoidal whorl, round umbilical shoulder, convex flanks, and a broad carinate venter baring a large keel. Coarse primary ribs are sinuous with round tops and extend from the umbilical shoulder across the flank and terminate at the 180 ventrolateral shoulder. Fine sinuous secondary ribs also occur between the primaries, and project onto the venter to the keel. Material.\u00E2\u0080\u0093 Three specimens from Localities #5 & #7 in the Yakoun River section, Haida Gwaii. Occurrence.\u00E2\u0080\u0093 This species has been described from the Mediterranean area, northwest Europe and South America. In western North America it is known from several localities in the Cordilleran region including Haida Gwaii. Age.\u00E2\u0080\u0093 upper part of the Kanense Zone (Early Toarcian). Hildaites spp. Pl. 7.2, figs 13, 14 Description.\u00E2\u0080\u0093 Evolute with a sub-rectangular to sub-quadrate whorl section, convex flanks and a broad carinate or tricarinate (?) venter baring a keel. Ribs are gently sinuous, become more prominent and coarse across the flank toward the venter and project onto the ventrolateral shoulder. Material.\u00E2\u0080\u0093 Two specimens from Locality #4 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 Poor preservation of these specimens precludes a more positive identification. These specimens resemble H. murleyi with their coarse ribs that are gently sinuous, however, there does not seem to be evidence of fine secondary ribs that project to the keel. Age.\u00E2\u0080\u0093 Kanense Zone (Early Toarcian). 181 Hildaites? sp. Pl. 7.2, fig 15 Description.\u00E2\u0080\u0093 An evolute form with a sub-quadrate whorl section, convex flank, and a broad carinate venter that bears a keel. Ribs are coarse, well spaced and sinuous along the flank, terminating at the ventrolateral shoulder. Material.\u00E2\u0080\u0093 One specimen from Locality #8 in the Whiteaves Bay section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 The poor preservation of this specimen preclude a positive identification. The coarse, well-spaced and gently sinuous ribs, suggest its placement within this genus. Age.\u00E2\u0080\u0093 Kanense Zone (Early Toarcian). Genus Leukadiella RENZ, 1913 Type species: Leukadiella helenae RENZ, 1913, p. 587\u00E2\u0080\u0093590. Leukadiella ionica RENZ & RENZ, 1947 Pl. 7.4, figs 5, 6 1947. Leukadiella ionica RENZ & RENZ, pl. 12, figs 5, 7. 1966. Leukadiella ionica Renz & Renz WENDT, pl. 2, fig 3. 1966. Leukadiella ionica Renz & Renz KOTTEK, pl. 13, fig 1. 1969. Leukadiella ionica Renz & Renz GALLITELLI\u00E2\u0080\u0093WENDT, pl. 3, fig 8. 1994. Leukadiella ionica Renz & Renz JAKOBS, SMITH & TIPPER, pl. 2, figs 9, 10. 182 1995. Leukadiella ionica Renz & Renz JAKOBS, figs. 5.1\u00E2\u0080\u00935.5, 6.1\u00E2\u0080\u00936.6, 6.9, 6.10 (and synonymy therein). 1997. Leukadiella ionica Renz & Renz JAKOBS, pl. 6, figs 1, 2, 4, 7, 8. Description.\u00E2\u0080\u0093 Evolute with a sub-quadrate whorl section, round umbilical shoulder containing prorsiradiate bullae, slightly convex flank, and a broad venter baring a keel with tubercles lining the ventrolateral shoulder. Ribs are coarse, gently sinuous, and well spaced along the flank. Prorsiradiate ribs extend from the umbilical shoulder at poorly preserved tubercles and bend gently rursiradiately at \u00C2\u00BC to \u00C2\u00BD flank, and possibly join at the ventrolateral tubercles. Material.\u00E2\u0080\u0093 One specimen from Locality #7 in the Whiteaves Bay section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 A characterizing feature of Leukadiella ionica is the presence of forked ribs that rejoin at the ventrolateral tubercles (RENZ and RENZ, 1947; JAKOBS, 1995), which is present in this specimen from Haida Gwaii (first noted in other specimens from Haida Gwaii in JAKOBS, 1995). Occurrence.\u00E2\u0080\u0093 This species is primarily known from the Mediterranean area, and is also known from Haida Gwaii. Age.\u00E2\u0080\u0093 Planulata Zone (Middle Toarcian). Genus Podagrosites GUEX, 1973 Type species: Pseudogrammoceras podagrosum MONESTIER, 1921, neotype figured by Guex, 1973, pl. 1, fig. 9. 183 Podagrosites latescens (SIMPSON, 1843) Pl. 7.4, figs 7, 8 1843. Ammonites latescens SIMPSON, p. 54, 55. 1997. Podagrosites latescens (Simpson) JAKOBS, pl. 7, figs 1\u00E2\u0080\u00934, 7, 8; pl. 8, figs 1, 2, 8\u00E2\u0080\u009315 (and synonymy therein). Description.\u00E2\u0080\u0093 Evolute with a wide umbilicus and an oval whorl containing gently rounded umbilical shoulder, flat to slightly convex flanks, and a carinate venter baring a sharp prominent keel. Ribs are coarse, gently flexuous, and evenly spaced with fine secondary ribs evident on the outer two whorls. Material.\u00E2\u0080\u0093 One specimen from Locality #14 in the Yakoun River section, Haida Gwaii. Occurrence.\u00E2\u0080\u0093 This species is known from northwest Europe and Haida Gwaii. Age.\u00E2\u0080\u0093 Hillebrandti Zone (Late Toarcian). Family Phymatoceratidae HYATT, 1867 Genus Phymatoceras HYATT, 1867 Type species: Phymatoceras robustum HYATT, 1867, p. 88. Phymatoceras cf. erbaense (HAUER, 1856) Pl. 7.4, figs 11, 12 cf. 1856. Ammonites erbaensis HAUER, pl. 11, fig. 10\u00E2\u0080\u009314. cf. 1976. Phymatoceras erbaense (Hauer) SCHLEGELMILCH, pl. 44, fig 6. cf. 1997. Phymatoceras cf. erbaense (Hauer) JAKOBS, pl. 13, figs 10, 11. 184 Description.\u00E2\u0080\u0093 An evolute whorl fragment with a sub-quadrate whorl section and broad venter baring a keel. Gently prorsiradiate ribs are very coarse and well spaced along the flank, projecting slightly onto the venter. Material.\u00E2\u0080\u0093 One specimen from Locality #13 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 Poor preservation precludes positive identification, however, the very coarse prorsiradiate ribs that project onto the venter is similar to specimens in JAKOBS (1997). Occurrence.\u00E2\u0080\u0093 This species is known from northwest Europe and the Cordilleran region of western North America including Haida Gwaii. Age.\u00E2\u0080\u0093 Crassicosta Zone (Middle Toarcian) to Hillebrandti Zone (Late Toarcian). Phymatoceras hillebrandti JAKOBS, 1992 Pl. 7.4, figs 9, 10 1992. Phymatoceras hillebrandti JAKOBS, pl. 13, fig 4; pl. 14, figs 1\u00E2\u0080\u00935; pl. 15, figs 1\u00E2\u0080\u00932. 1994. Phymatoceras hillebrandti Jakobs JAKOBS, SMITH & TIPPER, pl. 4, figs. 13, 14, 18\u00E2\u0080\u009323. 1997. Phymatoceras hillebrandti Jakobs JAKOBS, pl. 14, figs 1\u00E2\u0080\u00936; pl. 15, figs 1\u00E2\u0080\u0093 12; pl. 16, figs 7\u00E2\u0080\u00938. Description.\u00E2\u0080\u0093 An evolute fragment with an ellipsoidal whorl, flat flanks with a gently rounded ventrolateral shoulder and a carinate venter baring a keel. Ribs 185 are paired, gently flexuous, and consist of strong primaries with weaker secondaries that are joined at a set of weak tubercles near the umbilical shoulder. Paired ribs project prorsiradiately from the umbilicus, curve gently at mid-flank and diminish in size as they project across the flank, terminating at the ventrolateral shoulder. Material.\u00E2\u0080\u0093 One specimen from Locality #15 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 The paired ribbing and presence of tubercles in this specimen matches well with material illustrated in JAKOBS (1992; 1997). Occurrence.\u00E2\u0080\u0093 This species is only described from Haida Gwaii. Age.\u00E2\u0080\u0093 Hillebrandti Zone (Late Toarcian). Phymatoceras sp. Pl. 7.5, figs 1, 2 Description.\u00E2\u0080\u0093 A slowly expanding evolute form with an ellipsoidal\u00E2\u0080\u0093oval whorl section, convex flanks and a broad venter without a keel. Ribs are simple and very coarse. They project rectiradiately along the flank and terminate abruptly at a set of poorly preserved ventrolateral tubercles located at the shoulder. Material.\u00E2\u0080\u0093 One specimen from Locality #10 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 This specimen is similar to Phymatoceras cf. erbaense in JAKOBS (1997; pl. 13, figs 10\u00E2\u0080\u009311) with respect to whorl shape and size, as well as 186 the coarse rectiradiate ribs. However, this specimen differs in that it does not possess a keel and has ribs that do not project onto the venter. It also occurs in strata of a slightly older age than previous descriptions of P. cf. erbaense from Haida Gwaii (described from the Crassicosta Zone of the Middle Toarcian by JAKOBS, 1997). Age.\u00E2\u0080\u0093 Planulata Zone (Middle Toarcian). Phymatoceras? sp. Pl. 7.6, figs 1, 2 Description.\u00E2\u0080\u0093 An evolute form with a wide umbilicus, compressed ellipsoidal whorl, flat flanks and carinate venter. Poorly preserved rursiradiate bullae project inward from the ventrolateral shoulder. Fine sinuous ribs are well spaced with rounded tops and project from the umbilicus along the flank and terminate at the ventrolateral shoulder. Material.\u00E2\u0080\u0093 One specimen from Locality #12 in the Yakoun River section, Haida Gwaii. Discussion.\u00E2\u0080\u0093 Poor preservation in this specimen precludes a more definitive identification. Age.\u00E2\u0080\u0093 Crassicosta Zone (Middle Toarcian). Genus Rarenodia VENTURI, 1975 Type species: Rarenodia planulata VENTURI, 1975, p. 13, pl. 1, fig. 7. 187 Rarenodia planulata (VENTURI, 1975) Pl. 7.6, figs 3\u00E2\u0080\u00936; Pl. 7.7, fig 1; Pl. 7.8, fig. 2 1975. Rarenodia planulata VENTURI, pl. 1, figs 3, 6\u00E2\u0080\u00939, text-figs 3\u00E2\u0080\u00935. 1992. Rarenodia planulata Venturi JAKOBS, pl. 3, fig. 9; pl. 4, fig. 1; pl. 5, fig. 1; pl. 7, fig. 1. 1994. Rarenodia planulata Venturi JAKOBS, SMITH & TIPPER, pl. 2, figs 21, 22. 1997. Rarenodia planulata Venturi JAKOBS, pl. 10, figs 1, 2; pl. 11, figs. 1, 2; pl. 12, figs. 1, 2, 5\u00E2\u0080\u00938 (and synonymies therein). Description.\u00E2\u0080\u0093 An evolute, slowly expanding form with an ellipsoidal\u00E2\u0080\u0093sub rectangular whorl section, an abrupt umbilical shoulder, convex flanks and a broad carinate venter containing a sharp and prominent keel. Simple ribs are paired and project rectiradiately along the flank from a set of tubercles, located along the umbilical shoulder. Tubercles are more pronounced along the inner whorls. Ribs terminate on the ventrolateral shoulder. Material.\u00E2\u0080\u0093 Three specimens from Localities #3, #10 & #11 in the Yakoun River section, Haida Gwaii. Occurrence.\u00E2\u0080\u0093 This species is known from Italy in the Mediterranean area and the Cordilleran region of western North America, including Haida Gwaii. Age.\u00E2\u0080\u0093 Planulata Zone (Middle Toarcian). Genus Lytoceras SUESS, 1865 Type species: Ammonites fimbriatus SOWERBY 1817. 188 Lytoceras sp. Pl. 7.1, figs 37\u00E2\u0080\u009339 Description.\u00E2\u0080\u0093 An evolute form with a depressed sub-quadrate whorl section, a flattened venter, flat flanks and a well-defined umbilical wall. Prorsiradiate constrictions are well defined along the flank and become rectiradiate to gently rursiradiate across the venter. Ribbing is not present. A poorly preserved but complex suture is evident on the adapical part of the specimen. Material.\u00E2\u0080\u0093 One fragment from locality 5 in the Hicks Creek section, Talkeetna Mountains Alaska. Discussion.\u00E2\u0080\u0093 The prorsiradiate constrictions that become gently rursiradiate across the venter resemble L. apertum in GEYER (1893; Pl. 8, figs. 4, 5) and L. fimbriatum (SOWERBY) in RAKUS & GUEX (2002; Pl. 3, fig. 2), however the whorl shape is not the same. Our specimen also bears slight resemblance in whorl shape to L. fuggeri (GEYER, 1893; Pl. 8, fig. 8 & 9) but it has a depressed sub- quadrate whorl shape rather than the more compressed whorl shape of L. fuggeri. Occurrence.\u00E2\u0080\u0093 This constitutes the first figured occurrence of Lytoceras from the Pliensbachian of western North America. Age.\u00E2\u0080\u0093 Carlottense Zone (Late Pliensbachian). 189 Plate 7.1 190 Plate 7.1 \u00E2\u0080\u0093 All figures in (Plates I\u00E2\u0080\u0093VIII) are natural size unless otherwise indicated. UMPC = University of Montana Paleontology Center, PME = Pacific Museum of the Earth (UBC), loc. = ammonite locality. Figs. 1-6: Fanninoceras (Fanninoceras) carlottense MCLEARN 1-2, UMPC 13356; Carlottense Zone; loc. 4, Hicks Creek section. 3, 5, UMPC 13353; Carlottense Zone; loc. 2, Hicks Creek section. 4, UMPC 13351; Carlottense Zone; loc. 1, Hicks Creek section. 6, UMPC 13352; Carlottense Zone; loc. 2, Hicks Creek section. Figs. 7-19: Fanninoceras (Fanninoceras) fannini MCLEARN 7, UMPC 13347; 8, UMPC 13360; 9-10, 13361; 17-18, UMPC 13362; Carlottense Zone; loc. 3, Hicks Creek section. 11, UMPC 13345; Kunae Zone; loc. 8, Camp Creek section. 12, UMPC 13342; Kunae Zone; loc. 8, Camp Creek section. 13-14, UMPC 13343; Kunae Zone; loc. 13, Camp Creek section. 15-16, UMPC 13340; Kunae Zone; loc. 8, Camp Creek section. 19, UMPC 13341; Kunae Zone; loc. 8, Camp Creek section (latex cast). Figs. 20-22: Fanninoceras (Charlotticeras) cf. maudense SMITH & TIPPER 20-21, UMPC 13331; Kunae Zone; loc. 13, Camp Creek section (21 is X2). 22, UMPC 13332; Kunae Zone; loc. 8, Camp Creek section. Figs. 23-25: Amaltheus sp. 23, UMPC 13308; Kunae Zone; loc. 14, Camp Creek section. 24- 25, UMPC 13309; Kunae Zone; loc. 14, Camp Creek section (25 is X2). 191 Figs. 26-31: Arieticeras aff. domarense (MENEGHINI) 26-27, UMPC 13318; Kunae Zone; loc. 7, Camp Creek section. 28- 29, UMPC 13319; Kunae Zone; loc. 8, Camp Creek section. 30, UMPC 13310; Kunae Zone; loc. 15, Camp Creek section. 31, UMPC 13313; Kunae Zone; loc. 14, Camp Creek section (latex cast). Figs. 32-33: Leptaleoceras? sp. 32-33, UMPC 13330; Kunae Zone; loc. 6, Camp Creek section (specimen in Fig. 33 is a cast of the specimen in Fig. 32). Figs. 34-36: Lioceratoides (Lioceratoides) cf. involutum SMITH & TIPPER 34-35, UMPC 13357; Carlottense Zone; loc. 3, Hicks Creek section. 36, UMPC 13358; Carlottense Zone; loc. 3, Hicks Creek section. Figs. 37-39: Lytoceras sp. 37-39, UMPC 13359; Carlottense Zone; loc. 5, Hicks Creek section. 192 Plate 7.2 193 Plate 7.2 Figs. 1-2: Dactylioceras cf. compactum (DAGIS) 1-2, PME 000111; Kanense Zone; loc. 2, Yakoun River section. (x2) Figs. 3-6: Protogrammoceras cf. paltum (BUCKMAN) 3-4, PME 000112; Kanense Zone; loc. 1, Yakoun River section. 5, PME 000113; Carlottense Zone; loc. 6, Whiteaves Bay section. 6, PME 000114; Carlottense Zone; loc. 6, Whiteaves Bay section. (4 is x1.10) Fig. 7: Lioceratoides (Lioceratoides) cf. allifordense (MCLEARN) 7, PME 000115; Carlottense Zone; loc. 4, Whiteaves Bay section. (X2) Figs. 8-10: Lioceratoides (Pacificeras) cf. propinquum (WHITEAVES) 8-9, PME 000116; Carlottense Zone; loc. 1, Whiteaves Bay section. 10, PME 000117; Carlottense Zone; loc. 2, Whiteaves Bay section. (8-9 are X2) Figs. 11-12: Tiltoniceras cf. antiquum (WRIGHT) 11, PME 000118; Carlottense Zone; loc. 3, Whiteaves Bay section. 12, PME 000119; Carlottense Zone; loc. 6, Whiteaves Bay section. Figs. 13-14: Hildaites spp. 13-14, PME 000120; Kanense Zone; loc. 4, Yakoun River section. (x2) Figs. 15: Hildaites? sp. 194 15, PME 000121; Kanense Zone; loc. 8, Whiteaves Bay section. 195 Plate 7.3 196 Plate 7.3 Figs. 1, 4-5: Cleviceras exaratum (YOUNG & BIRD) 1, PME 000122; Kanense Zone; loc. 9, Yakoun River section. 4-5, PME 000123; Kanense Zone; loc. 8, Yakoun River section. (1, 4 are x0.75) Figs. 2-3, 6-7: Cleviceras spp. 2-3, PME 000124; Kanense Zone; loc. 3, Yakoun River section. 6-7, PME 000125; Kanense Zone; loc. 6, Yakoun River section. 197 Plate 7.4 198 Plate 7.4 Figs. 1-4: Hildaites murleyi (MOXON) 1, PME 000126; Kanense Zone; loc. 5, Yakoun River section. 2-3, PME 000127; Kanense Zone; loc. 7, Yakoun River section. 4, PME 000128; Kanense Zone; loc. 7, Yakoun River section. Figs. 5-6: Leukadiella ionica (RENZ & RENZ) 5-6, PME 001090; Planulata Zone; loc. 7, Whiteaves Bay section. Fig. 7-8: Podagrosites latescens (SIMPSON) 7-8, PME 001091; Hillebrandti Zone; loc. 14, Yakoun River section. Figs. 9-10: Phymatoceras hillebrandti (JAKOBS) 9-10, PME 001092; Hillebrandti Zone; loc. 15, Yakoun River section. Figs. 11-12: Phymatoceras cf. erbaense (HAUER) 11-12, PME 001093; Crassicosta Zone; loc. 13, Yakoun River section. 199 Plate 7.5 200 Plate 7.5 Figs. 1-2: Phymatoceras sp. 1-2, PME 001094; Planulata Zone; loc. 10, Yakoun River section. (1-2 are x0.75) 201 Plate 7.6 202 Plate 7.6 Figs. 1-2: Phymatoceras? sp. 1-2, PME 001095; Crassicosta Zone; loc. 12, Yakoun River section. (2 is x2) Figs. 3-6: Rarenodia planulata (VENTURI) 3-4, PME 001096; Planulata Zone; loc. 11, Yakoun River section. 5- 6, PME 001097; Planulata (?) Zone; float loc. near 3, Yakoun River section. 203 Plate 7.7 204 Plate 7.7 Fig. 1: Rarenodia planulata (VENTURI) 1, PME 001098; Planulata Zone; loc. 10, Yakoun River section. 205 Plate 7.8 206 Plate 7.8 Fig. 1: Harpoceras cf. nitescens (SCHLEGELMILCH) 1, PME 001099; Planulata Zone; loc. 10, Yakoun River section. Fig. 2: Rarenodia planulata (VENTURI) 2, PME 002000; Planulata Zone; loc. 10, Yakoun River section. 207 Chapter 8 Conclusions This study uses paleontology and isotope geochemistry to investigate the causes and effects of a well-known second order extinction of marine organisms during the Pliensbachian\u00E2\u0080\u0093Toarcian stages of the Early Jurassic. A primary objective of this work is to compare new data from western North America with previously established records in European successions in order to test hypotheses related to its duration, magnitude and controlling mechanisms. Analysis of the stratigraphic ranges of 206 ammonite and 242 foraminiferal species in North America indicates declines in diversity during six separate intervals throughout Pliensbachian\u00E2\u0080\u0093Toarcian time. These intervals of extinction can be correlated with declines in ammonite species and generic data from a combined dataset of the northwest European and Arctic domains (Dera et al., 2010). This suggests that the multi-phased species level extinction during the Pliensbachian\u00E2\u0080\u0093Toarcian interval was potentially global. Phases of extinction correspond with the: 1) middle of the Early Pliensbachian (middle Whiteavesi\u00E2\u0080\u0093 middle Freboldi Zones), 2) middle of the Late Pliensbachian (late Kunae\u00E2\u0080\u0093early Carlottense Zone), 3) Pliensbachian / Toarcian boundary into the Early Toarcian (late Carlottense\u00E2\u0080\u0093middle Kanense Zones), 4) Middle Toarcian (late Planulata\u00E2\u0080\u0093 early Crassicosta Zones) and 5) late Middle\u00E2\u0080\u0093early Late Toarcian (middle Crassicosta\u00E2\u0080\u0093Hillebrandti Zones), and 6) Late Toarcian (lower\u00E2\u0080\u0093middle Yakounensis Zone). 208 Species level data from western North America also support previous conclusions suggesting that the main-phase of extinction is a global event during which diversity declined in a gradual fashion beginning at the Pliensbachian / Toarcian boundary and extending into the Early Toarcian (Harries and Little, 1999; Caswell et al., 2009). In western North America, species diversity in both ammonite and foraminifera reach minimum values in the middle part of the Early Toarcian Kanense Zone which coincides with the negative CIE interval described by Caruthers et al. (2011). Recognition of a multi-phased extinction across the paleo Pacific, Arctic and Tethys Ocean basins greatly expands the known geographic extent of this event. Previously, at the onset of this study, it was suggested that only the main- phase of extinction might be global but it is now apparent that other phases of decline could potentially be global requiring a controlling mechanism that is also global in its reach. Previously it was argued that volcanogenic outgassing of CO2 during the eruption of the Karoo\u00E2\u0080\u0093Ferrar LIP initiated greenhouse conditions and caused the extinction of marine organisms (P\u00C3\u00A1lfy and Smith, 2000). Data presented herein support this interpretation. There is a good overlap between the eruption ages of main volume of basalt emplacement in the Karoo Basin (Jourdan et al., 2008) and calibrated ages of declines in species diversity that constitute this multi- phased event. Four of the six pulses of extinction occur within the main-phase of magmatism, the Early Pliensbachian decline occurs within error range of the 209 onset of Karoo magmatism, and the Late Toarcian decline coincides with the later stages of magmatism. New carbon-isotope data provided in this research strongly supports the Volcanic Greenhouse hypothesis, which implicates global warming (from volcanogenic outgassing of CO2 during Karoo\u00E2\u0080\u0093Ferrar magmatism) as a primary control mechanism for the Pliensbachian\u00E2\u0080\u0093Toarcian extinction (P\u00C3\u00A1lfy and Smith, 2000). The long-term carbon-isotope record (using a comprehensive dataset of 1596 samples analyzed herein as well as a previously published dataset by Williford et al., 2007) shows isotopic fluctuations over much of the Late Triassic (Norian) to Toarcian interval on Haida Gwaii. Within this record, there is a distinct pattern that develops from the Early Pliensbachian to the Middle Toarcian. Within the entire profile, \u00CE\u00B413C values are consistent (at approximately \u00E2\u0080\u009331\u00E2\u0080\u00B0) throughout much of the Hettangian and into the Sinemurian. Throughout the middle part of the Early Pliensbachian into the middle part of the Middle Toarcian, \u00CE\u00B413C values become much more positive and fluctuate abruptly and frequently with a magnitude that ranges ~3\u00E2\u0080\u00937\u00E2\u0080\u00B0. The long-term carbon-isotope profile is consistent with the carbon cycle box model originally proposed by Payne and Kump (2007) to demonstrate the influence of Siberian Trap magmatism during the Late Permian\u00E2\u0080\u0093Middle Triassic. Payne and Kump\u00E2\u0080\u0099s (2007) model suggests that volcanogenic outgassing of CO2 during LIP eruption produces small negative CIEs that are followed by larger (higher magnitude) positive excursions brought on by increased primary productivity. Several pulses of magmatism would 210 therefore produce the observed pattern in the carbon-isotope profile of the Early Pliensbachian\u00E2\u0080\u0093Middle Toarcian interval. Carbon- and nitrogen-isotope geochemistry is also utilized in this study to address two other mechanisms that are thought to have intensified the extinction during part of its main-phase. It is argued that during the Early Toarcian, correlative with the diversity minimum at the Tenuicostatum / Serpentinum Zone boundary, there was a rapid release from the global methane hydrate reservoir which created a large negative CIE in marine and terrestrial carbon reservoirs (Hesselbo et al., 2000). Previously, the geographic extent of this large negative CIE was uncertain, as it was only known from parts of Northwest Europe and Mediterranean area of the Tethys Ocean. New data from two sections on Haida Gwaii demonstrates an Early Toarcian negative CIE that is of similar magnitude and duration to the excursion evident in Europe. During the Early Toarcian, Haida Gwaii was located in the northeast part of the paleo-Pacific Ocean and therefore this data provides good evidence of a globally controlled perturbation in carbon-isotopes. The two datasets from Haida Gwaii show a similar pattern of perturbation, in that a broad positive shift in \u00CE\u00B413Corg is interrupted by a sharp and pronounced negative excursion of a similar magnitude and duration as in the European data. At the onset of the negative CIE, high-resolution data in the Yakoun River section shows a large decline in \u00CE\u00B413Corg that is composed of three distinct pulses (ranging 1\u00E2\u0080\u00933\u00E2\u0080\u00B0). The entire negative CIE interval averages \u00E2\u0080\u009331\u00E2\u0080\u00B0 for 11 m within the Early Toarcian Kanense Zone. This negative CIE in the Yakoun River section also 211 coincides with TOC values that increase from ~0.4% to ~1.2% and subsequently decrease to ~0.5% for the remainder of the section. In the Whiteaves Bay section, the initial part of the negative CIE occurs in at least two separate pulses of ~ \u00E2\u0080\u00931 to 2\u00E2\u0080\u00B0, and the entire negative CIE interval averages \u00E2\u0080\u009328\u00E2\u0080\u00B0 for 6 m in the section. This interval also has TOC values that increase slightly to ~ 0.6%. The apparent difference in stratigraphic thickness of the negative CIE interval between these two sections could potentially have resulted from condensing in the Whiteaves Bay section. Carbon-isotope data was also collected from the South Barrow #3 core of northern Alaska. However, there was an unfortunate gap in core recovery that spanned ~55 m within the Early Toarcian stratigraphy where it is believed that the negative CIE (if present) would occur. The new carbon-isotope data from Haida Gwaii: 1) supports a global control rather than local upwelling mechanism for carbon-isotope perturbation, 2) provides further evidence of cyclic pulsing within the negative CIE interval that has been attributed to astronomical precession (Kemp et al., 2005) and 3) also supports calibration of the Early Toarcian ammonite zonal schemes of western North America, NW Europe and parts of the Mediterranean area. Lastly, the study presented herein uses TOC, nitrogen-isotope and TN data to test the highly debated global extent of an anoxic body of marine water during the Early Toarcian (Jenkyns, 1988), dubbed the T\u00E2\u0080\u0093OAE. The T\u00E2\u0080\u0093OAE is thought to have greatly escalated the observed effects of the main-phase of extinction at the level that is correlative with the Tenuicostatum / Serpentinum zone boundary (Caswell et al., 2009 and references therein). Evidence for the T\u00E2\u0080\u0093 212 OAE stems primarily from northwest and parts of southern Europe where isotopic perturbations in many systems, including nitrogen- (Jenkyns et al., 2001), typically co-occur with organically enriched shale with TOC levels as high as 15% (Jenkyns, 1985; 1988; Baudin et al., 1990a,b). A positive excursion in nitrogen- isotope values is thought to signify denitrification in an anoxic water column. High concentrations of TOC in European successions support this by indicating that organic matter was not being degraded or oxidized at the time of deposition. In the Pliensbachian\u00E2\u0080\u0093Toarcian of western North America, concentrations of TOC did not exceed 2% within the negative CIE interval, values that are more typical of modern hemipelagic shelf sediments (McIver, 1975). Also, in the CIE interval and throughout the majority of the Toarcian, \u00CE\u00B415N values show no positive excursion but are generally consistent, ranging between 0.0 and \u00E2\u0080\u00931.0\u00E2\u0080\u00B0. In fact, \u00CE\u00B415N values seem to show two successive, small-scale, negative excursions within the negative CIE interval. Therefore this TOC and nitrogen-isotope data does not suggest denitrification within the water column for this part of northeast paleo-Pacific Ocean during the Early Toarcian. Unlike the carbon-isotope excursion, which seems to be globally isochronous, TOC concentration is dependent on local factors and therefore should not be used to characterize global marine anoxia. An alternative model for the T\u00E2\u0080\u0093OAE suggests that the anoxic water was locally restricted to silled basins within northwest Europe (McArthur et al., 2008). This hypothesis is based on the over-abundance of organic carbon that co- occurs with perturbations in many isotopic systems within the Early Toarcian 213 stratigraphy (McArthur et al., 2008 and references therein). New TOC, nitrogen- isotope and TN data from western North America support this hypothesis, in that there does not seem to be an over abundance of organic carbon nor evidence of denitrification within the Early Toarcian stratigraphy. Therefore, it is unlikely that there was an anoxic water mass within this part of the paleo Pacific Ocean. It is possible that during this time of intense climate change, anoxic water of the T\u00E2\u0080\u0093 OAE could only develop under certain conditions and environments (e.g. mechanisms described in the restricted basin model by McArthur et al., 2008). These restricted basins, containing anoxic masses of water were not necessarily restricted to the Tethys Ocean area. Curiously, however, there does seem to be evidence for denitrification and marine anoxia in the Early Pliensbachian Ghost Creek Formation on Haida Gwaii. When comparing with the entire Toarcian interval, the Ghost Creek Formation at Whiteaves Bay has TOC and TN concentrations that are much higher and also has nitrogen-isotope values that are more positive. This new geochemical data corroborates previous interpretations, which suggest a euxinic basin during deposition of this formation. However, currently there are no accounts of global marine anoxia in the Early Pliensbachian and therefore this instance most likely record a regional occurrence. It is possible that a body of anoxic marine water could have developed from a stratified water column in a restricted basin setting within the Wrangellia composite terrane. Gradually this basin was in-filled and the water column then returned to more normal redox conditions by late Early Pliensbachian time. 214 Therefore, to summarize the results of this work from western North America, several major conclusions can be drawn concerning the magnitude, duration and potential controlling mechanisms of the Pliensbachian\u00E2\u0080\u0093Toarcian mass extinction: 1) It now seems evident that species-level diversity in two taxonomic groups declined globally over six intervals throughout Pliensbachian\u00E2\u0080\u0093 Toarcian time; 2) There is a strong correlation in timing between this multi- phased event, long-term carbon-isotope perturbation and eruption ages of the Karoo magmatic province. This suggests that volcanogenic outgassing of CO2 during LIP eruption was the preeminent factor driving global warming and mass extinction; 3) The most severe, main-phase, extinction interval begins at the Pliensbachian / Toarcian boundary and continues into the Early Toarcian where global species diversity reach minimum values in the middle part of the Kanense Zone; 4) This main-phase (middle Kanense Zone) extinction and diversity minimum was reinforced significantly by a global release from the methane hydrate reservoir, evidenced by a large coeval negative excursion in carbon- isotope values from two successions on Haida Gwaii (the first well-documented record of this excursion outside of the Tethys Ocean area); and 5) The northeast paleo Pacific Ocean was not experiencing denitrification and significant organic carbon enrichment during the main-phase of extinction in the middle part of the Kanense Zone, and therefore cannot be considered anoxic at this time. Although the T\u00E2\u0080\u0093OAE does not appear to have affected every (global) marine environment, it is possible that anoxic water masses occurred in restricted basins outside the Tethys Ocean area. 215 References Al-Suwaidi, A.H., Angelozzi, G.N., Baudin, F., Damborenea, S.E., Hesselbo, S.P., Jenkyns, H.C., Mance\u00C3\u00B1ido, M.O., and Riccardi, A.C., 2010. First record of the Early Toarcian Oceanic Anoxic Event from the Southern Hemisphere, Neuqu\u00C3\u00A9n Basin, Argentina. Journal of the Geological Society, London, 167, 1\u00E2\u0080\u00934. doi:10.1144/0016-76492010-025. Altabet, M.A., and Francois, R. 1994. Sedimentary nitrogen isotopic ratio as a recorder for surface nitrate utilization. Global Biogeochemical Cycles, 8, 103\u00E2\u0080\u0093 116, doi: 10.1029/93GB03396. Amato, J.M., Miller, E.L., and Gehrels, G.E., 2003. Lower Paleozoic through Achaean detrital zircon ages from metasedimentary rocks of the Nome Group, Seward Peninsula, Alaska, Eos Trans. AGU, 84, Fall Meeting Supplemental Abstract T31F0891. Amato, J.M., 2004. Crystalline basement ages, detrital zircon ages, and metamorphic ages from Seward Peninsula: Implications for Proterozoic and Cambrian-Ordovician paleogeographic reconstructions of the Arctic-Alaska terrane. Geological Society of America Abstracts with Programs, 36, p. 22. Amato, J.M., Toro, J., Miller, E.L., Gehrels, G.E., Farmer, G.L., Gottlieb, E.S., and Till, A.B., 2009. Late Proterozoic\u00E2\u0080\u0093Paleozoic evolution of the Arctic Alaska\u00E2\u0080\u0093 Chukotka terrane based on U-Pb igneous and detrital zircon ages: Implications for Neoproterozoic paleogeographic reconstructions. Geological Society of America Bulletin 121, p. 1219\u00E2\u0080\u00931235. Andrew, A., and Goodwin, C.I., 1989. Lead- and strontium-isotope geochemistry of the Karmutsen Formation, Vancouver, Island, British Columbia. Canadian Journal of Earth Sciences, 26, p. 908-919. Arndt, N.T., Czamanske, G.K., Wooden, J.L., and Fedorenko, V.A., 1993. Mantle and crustal contributions to continental flood basalt volcanism. Tectonophysics, 223, 39-52. Baksi, A.K., 1990. Timing and duration of Mesozoic\u00E2\u0080\u0093Tertiary flood basalt volcanism. EOS: Transactions, American Geophysical Union, 71, 1835\u00E2\u0080\u00931836. Bambach, R.K., 2006. Phanerozoic Biodiversity Mass Extinctions. Annual Review of Earth and Planetary Sciences, 34, 125-155. Barker, F., and Grantz, A.,1982. Talkeetna Formation in the southeastern Talkeetna Mountains, southern Alaska: An Early Jurassic andesitic intraoceanic 216 island arc. Geological Society of America Abstracts with Programs, Boulder, USA, 14, p. 147. Barry, J.C., Morgan, M.E., Flynn, L.J., Pilbeam, D., Jacobs, L.L., Lindsay, E.H., Raza, S.M., and Solounias, N., 1995. Patterns of faunal turnover and diversity in the Neogene Siwaliks of northern Pakistan. Palaeogeography, Palaeoclimatology, Palaeoecology, 115, 209\u00E2\u0080\u0093226. Baudin, F., Herbin, J.P., Bassoulet, J.P., Dercourt, J., Lachkar, G., Manivit, M. and Renard, M., 1990a. Distribution of organic matter during the Toarcian in the Mediterranean and Middle East. In: Huc, A.Y. (Ed.), Deposition of Organic Facies. American Association of Petroleum Geologists Studies in Geology, 30, 73\u00E2\u0080\u009391. Baudin, F., Herbin, J.P., and Vandenbroucke, M., 1990b. Mapping and geochemical characterization of the Toarcian organic matter in the Mediterranean Tethys and Middle East. Organic Geochemistry, 16, 677\u00E2\u0080\u0093687. Beerling, D.J., and Brentnall, S.J., 2007. Numerical evaluation of mechanisms driving Early Jurassic changes in global carbon cycling. Geology, 35, 247\u00E2\u0080\u0093250, doi: 10.1130/G23416A.1. Beerling, D.J., Lomas, M.R., and Gr\u00C3\u00B6cke, D.R., 2002. On the nature of methane gas-hydrate dissociation during the Toarcian and Aptian Oceanic anoxic events. American Journal of Science, 302, 28\u00E2\u0080\u009349. Bellanca, A., Masetti, D., Neri, R., and Venezia, F., 1999. Geochemical and Sedimentological evidence of productivity cycles recorded in Toarcian black shales from the Belluno Basin, Southern Alps, northern Italy. Journal of Sedimentary Research 69, 466\u00E2\u0080\u0093476. Bengston, P., 1988. Open nomenclature. Palaeontology, London, UK, 31: 223\u00E2\u0080\u0093 227. Benton, M.J. (Ed.), 1993. The Fossil Record 2. London, UK. Chapman and Hall. Benton, M.J., 1995. Diversification and extinction in the history of life. Science 268, 53\u00E2\u0080\u009358. Benton, M.J., and Storrs, G.W., 1994. Testing the quality of the fossil record: paleontological knowledge is improving. Geology 22, 111\u00E2\u0080\u0093114. Beranek, L.P., 2009. Provenance and Paleotectonic setting of North American Triassic Strata in Yukon: The Sedimentary Record of Pericratonic terrane accretion in the northern Canadian Cordillera. PhD. Thesis, University of British Columbia, Vancouver, British Columbia, 338 p. 217 Bergquist, H.R., 1966. Micropaleontology of the Mesozoic Rocks of Northern Alaska. United States Geological Survey Professional Paper 302\u00E2\u0080\u0093D, 227p. Bilotta, M., Venturi, F., and Sassaroli, S., 2010. Ammonite faunas, OAE and the Pliensbachian\u00E2\u0080\u0093Toarcian boundary (Early Jurassic) in the Apennines. Lethaia, 43, 357\u00E2\u0080\u0093380. Bird, K.J., 1988. Alaska North Slope stratigraphic nomenclature and data summary for government-drilled wells. In: Gryc, G. (Ed.), Geology and exploration of the National Petroleum Reserve in Alaska, 1974\u00E2\u0080\u00931982. United States Geological Survey Professional Paper 1399, 317\u00E2\u0080\u0093353. Blodgett, R.B., 2008. Paleontology and stratigraphy of the Upper Triassic Kamishak Formation in the Puale Bay-Cape Kekurnoi-Alinchak Bay area, Karluk C-4 and C-5 quadrangle, Alaska Peninsula. In: Reifenstuhl, R.R. & P.L. Decker (Eds.), Bristol Bay-Alaska Peninsula region, overview of 2004-2007 geologic research. Alaska Division of Geological & Geophysical Surveys Report of Investigation 2008-1H, Anchorage, 131-160. Blodgett, R.B., and Stralla, B., 2006. A major unconformity between Permian and Triassic strata at Cape Kekurnoi, Alaska Peninsula: Old and new observations on stratigraphy and hydrocarbon potential. In: Haeussler, P.J. and Galloway, J.P., (Eds.), Studies by the United States Geological Survey in Alaska 2006. United States Geological Survey Professional Paper 1739\u00E2\u0080\u0093E, Denver, 13p. Blodgett, R.B., and Stanley, G.D. Jr., (Eds.), 2008. The terrane puzzle: New perspectives on paleontology and stratigraphy from thee North American Cordillera. Geological Society of America Special Paper 442, 326p. Blodgett, R.B., Rohr, D.M., and Boucot, A.J., 2002, Paleozoic links among some Alaskan accreted terranes and Siberia based on megafossils, in Miller, E.L., et al., eds., Tectonic Evolution of the Bering Shelf-Chukchi Sea- Arctic Margin and Adjacent Landmasses: Geological Society of America Special Paper 360, p. 273- 290. Bodin, S., Mattioli, E., Fr\u00C3\u00B6hlich, S., Marshall, J.D., Boutib, L., Lahsini, S., and Redfern, J., 2010. Toarcian carbon isotope shifts and nutrient changes from the Northern margin of Gondwana (High Atlas, Morocco, Jurassic): Palaeoenvironmental implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 297, 377\u00E2\u0080\u0093390. Braga, J.C., Comas-Rengifo, M.J., Goy, A., and Rivas, P., 1982. Comparaciones faunisticas y correlaciones en el Pliensbachiense de la Zona Subb\u00C3\u00A9tica y Cordillera Ib\u00C3\u00A9rica. Bolletin de la Reale Sociedad Espa\u00C3\u00B1ola de Historia Naturale, Seccion Geologica, 80, 133\u00E2\u0080\u0093152. 218 Bryan, S.E., Peate, I.U., Peate, D.W., Self, S., Jerram, D.A., Mawby, M.R., Marsh (Goonie), J.S., and Miller, J.A., 2010. The largest volcanic eruptions on Earth. Earth Science Reviews 102, 207\u00E2\u0080\u0093229. Buckman, S.S., 1909\u00E2\u0080\u00931930 (1913, 1921, 1922). Yorkshire type ammonites, I-II, and type ammonites, III-VII. Reprinted 1976, Wheldon and Wesley Ltd. and Verlag von Cramer, Herts., England. vol. I, pls. 24\u00E2\u0080\u009344 (1911); vol. II, 16 pp., pls. 68\u00E2\u0080\u009390 (1913); pls. 130A\u00E2\u0080\u0093130C (1919); vol. III, pls. 148\u00E2\u0080\u0093194 (1920); vol. IV, 67 pp., pls. 267B\u00E2\u0080\u0093362 (1922); pls. 363\u00E2\u0080\u0093422 (1923); vol. V, 88p., pls. 458\u00E2\u0080\u0093536 (1924). Buckman, S.S., 1918. Jurassic chronology, I-Lias. Quarterly Journal of the Geological Society of London, UK, 73: 257\u00E2\u0080\u0093327. Buzas, M.A., and Culver, S.J., 1994. Species pool and dynamics of marine paleocommunities. Science, 264, 1439\u00E2\u0080\u00931441. Buzas, M.A., and Culver, S.J., 1998. Assembly, disassembly, and balance in marine communities. Palaios, 13, 263\u00E2\u0080\u0093275. Cameron, B.E.B., and Tipper, H.W., 1985. Jurassic stratigraphy of the Queen Charlotte Islands, British Columbia. Geological Survey of Canada, Bulletin 365, 1\u00E2\u0080\u009349. Capps, S.R., 1923. The Cold Bay district. In: Mineral resources of Alaska; report on progress of investigations in 1921. United States Geological Survey Bulletin 739-C, Denver, 77\u00E2\u0080\u0093116. Caruthers, A.H., and Smith, P.L., 2012. Pliensbachian ammonoids from the Talkeetna Mountains (Peninsular Terrane) of Southern Alaska. Review de Pal\u00C3\u00A9obiologie Volume sp\u00C3\u00A9cial 11, 365\u00E2\u0080\u0093378. Caruthers, A.H., Gr\u00C3\u00B6cke, D.R., and Smith, P.L., 2011. The significance of an Early Jurassic (Toarcian) carbon-isotope excursion in Haida Gwaii (Queen Charlotte Islands), British Columbia, Canada. Earth and Planetary Science Letters, 307, 19\u00E2\u0080\u009326. Caswell, B.A., Coe, A.L., and Cohen, A.S., 2009. New range data for marine invertebrate species across the Early Toarcian (Early Jurassic) mass extinction. Journal of the Geological Society, London, 166, 859\u00E2\u0080\u0093872. Cecca, F., and Macchioni, F., 2004. The two Early Toarcian (Early Jurassic) extinction events in ammonoids. Lethaia, 37, 35\u00E2\u0080\u009356. 219 Clift, P.D., Draut, A.E., Kelemen, P.B., Blusztajn, J., and Greene, A., 2005a. Stratigraphic and geochemical evolution of an oceanic arc upper crustal section: The Jurassic Talkeetna Volcanic Formation, south-central Alaska. Geological Society of America Bulletin, 117, 902\u00E2\u0080\u0093925. Clift, P.D., Pavlis, T., DeBari, S.M., Draut, A.E., Rioux, M., and Kelemen, P.B., 2005b. Subduction erosion of the Jurassic Talkeetna-Bonanza arc and the Mesozoic accretionary tectonics of western North America. Geology, 33, 881\u00E2\u0080\u0093 884. doi: 10.1130/G21822.1. Coffin, M.F., and Eldholm, O., 1993. Scratching the surface: estimating dimensions of large igneous provinces. Geology, 21, 515\u00E2\u0080\u0093518. Coffin, M.F., and Eldholm, O., 1994. Large igneous provinces: crustal structure, dimensions, and external consequences. Reviews of Geophysics, 32, 1\u00E2\u0080\u009336. Cohen, A.S., and Coe, A.L., 2002. New geochemical evidence for the onset of volcanism in the Central Atlantic Magmatic Province and environmental change at the Triassic\u00E2\u0080\u0093Jurassic boundary. Geology, 30, 267\u00E2\u0080\u0093270. Cohen, A.S., and Coe, A.L., 2007. The impact of the Central Atlantic Magmatic Province on climate and on the Sr- and Os-isotope evolution of seawater. Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 374\u00E2\u0080\u0093390. Cohen, A.S., Coe, A.L., Harding, S.M., and Schwark, L., 2004. Osmium isotope evidence for the regulation of atmospheric CO2 by continental weathering. Geology, 32, 157\u00E2\u0080\u0093160. Cohen, A.S., Coe, A.L., and Kemp, D.B., 2007. The Late Palaeocene\u00E2\u0080\u0093Early Eocene and Toarcian (Early Jurassic) carbon isotope excursions: a comparison of their time scales, associated environmental changes, causes and consequences. Journal of the Geological Society, London, 164, 1093\u00E2\u0080\u00931108. Cole, R.B., Ridgway, K.D., Layer P.W., and Drake, J., 1999. Kinematics of basin development during the transition from terrane accretion to strike-slip tectonics, Late Cretaceous early Tertiary Cantwell Formation, south central Alaska. Tectonics, 18, 1224\u00E2\u0080\u00931244. Collins, F.R., 1961. Core Tests and Test Wells Barrow Area, Alaska. United States Geological Survey Professional Paper 305\u00E2\u0080\u0093K, 569\u00E2\u0080\u0093644. Colpron, M., and Nelson, J.L., 2009. A Palaeozoic Northwest Passage: incursion of Caledonian, Baltican and Siberian terranes into eastern Panthalassa, and the early evolution of the North American Cordillera. In: Cawood, P.A., and Kr\u00C3\u00B6ner, A. (Eds.), Earth Accretionary Systems in Space and Time. The Geological Society, London, Special Publications 318, 273\u00E2\u0080\u0093307. 220 Colpron, M., Nelson, J.L., and Murphy, D.C., 2007. Northern Cordilleran terranes and their interactions through time. Geological Society of America Today, 17, 4\u00E2\u0080\u0093 10. Coney, P.J., Jones, D.L., and Monger, J.W.H., 1980. Cordilleran suspect terranes. Nature, 288, 329\u00E2\u0080\u0093333. Courtillot, V., 1999. Evolutionary Catastrophes: The Science of Mass Extinction. Cambridge University Press, Cambridge, 237 p. Courtillot, V., and Renne, P.R., 2003. On the ages of flood basalt events. Compte Rendus\u00E2\u0080\u0093 Acad\u00C3\u00A9mie des sciences, 335, 113\u00E2\u0080\u0093140. Courtillot, V., Jaupert, C., Manighetti, I., Tapponier, P., and Besse, J., 1999. On Causal links between flood basalts and continental break-up. Earth Planetary Science Letters 166, 177\u00E2\u0080\u0093195. Csejtey, G., Jr., and St. Aubin, D.R., 1981. Evidence for northwestward thrusting of the Talkeetna superterrane, and its regional significance. In: Albert, N.R.D. and Hudson, T., (Eds.), The United States Geological Survey in Alaska: Accomplishments during 1979. United States Geological Survey Circular 823-B, Denver, B49\u00E2\u0080\u0093B51. Csejtey, B., Jr., Nelson, W.H., Jones, D.L., Silberling, N.J., Dean, R.M., Morris, M.S., Lanphere, M.A., Smith, J.G., and Silberman, M.L., 1978. Reconnaissance geologic map and geochronology, Talkeetna Mountains quadrangle, northern part of Anchorage Quadrangle, and southwest corner of Healy quadrangle, Alaska. United States Geological Survey Open-File Report 78-558-A, Denver, 1:250,000 scale. Csejtey, B., Jr., Cox, D.P., Evarts, R.C., Stricker, G.D., and Foster, H.L., 1982. The Cenozoic Denali fault system and the Cretaceous accretionary development of southern Alaska. Journal of Geophysical Research, 87: 3741\u00E2\u0080\u00933754. Cuvier, G., 1797. Tableau \u00C3\u00A9l\u00C3\u00A9mentaire de l\u00E2\u0080\u0099histoire naturelle des animaux. Paris, France, 710p. Dagys, A.A., 1968. Toarskie ammonity (Dactylioceratidae) Severa Sibiri. Trudy Instituta Geologii i Geofiziki, Novosibirsk 40, 1\u00E2\u0080\u0093108. Dean, W.T., Donovan, D.T., and Howarth, M.K., 1961. The Liassic Ammonite zones and subzones of northwest European province; Bulletin of the British Museum (Natural History), Geology, 4, 435\u00E2\u0080\u0093505. 221 Deines, P., 2002. The carbon isotope geochemistry of mantle xenoliths. Earth- Science Reviews, 58, 247\u00E2\u0080\u0093278. Dera, G., and Donnadieu, Y., 2012. Modeling evidences for global warming, Arctic seawater freshening, and sluggish oceanic circulation during the Early Toarcian anoxic event. Paleoceanography, 27, PA2211, doi: 10.1029/2012PA002283. Dera, G., Neige, P., Dommergues, J.-L., Fara, E., Laffont, R., and Pellenard, P., 2010. High-resolution dynamics of Early Jurassic marine extinctions: the case of Pliensbachian\u00E2\u0080\u0093Toarcian ammonites (Cephalopoda). Journal of the Geological Society, London, 167, 21\u00E2\u0080\u009333, doi: 10.1144/0016-76492009-068. Detterman, R.L., Miller, J.W., and Case, J.E., 1985. Megafossil locality map, checklists, and pre-Quaternary stratigraphic sections of Ugashik, Bristol Bay, and part of Karluk quadrangles, Alaska: Secondary Megafossil locality map, checklists, and pre-Quaternary stratigraphic sections of Ugashik, Bristol Bay, and part of Karluk quadrangles, Alaska: United States Geological Survey Miscellaneous Field Studies Map MF-1539-E, 1:250,000 scale. Dommergues, J.-L., Mouterde, R., and Rivas, P., 1984. Un faux Polymophitin\u00C3\u00A9: Dubariceras, nouveau genre d\u00E2\u0080\u0099Ammonitina du Carixien mesog\u00C3\u00A9en. G\u00C3\u00A9obios, 17, 831\u00E2\u0080\u0093839. Dommergues, J.-L., Fara, E., and Meister, C., 2009. Ammonite diversity and its palaeobiogeographical structure during the early Pliensbachian (Jurassic) in the western Tethys and adjacent areas. Palaeogeography, Palaeoclimatology, Palaeoecology 280, 64\u00E2\u0080\u009377. Duncan, R.A., 2002. A Time frame for Construction of the Kerguelen Plateau and Broken Ridge. Journal of Petrology, 43, 1109\u00E2\u0080\u00931119. Dumoulin, J.A., Harris, A.G., Gagiev, M., Bradley, D.C., and Repetski, J.E., 2002. Lithostratigraphic, conodont, and other faunal links between lower Paleozoic strata in northern and central Alaska and northeastern Russia. In: Miller, E.L., Grantz, A., and Klemperer, S.L., (Eds.) Tectonic Evolution of the Bering Shelf\u00E2\u0080\u0093 Chukchi Sea\u00E2\u0080\u0093Arctic Margin and Adjacent Landmasses. Geological Society of America Special Paper 360, 291\u00E2\u0080\u0093312. Erickson, B.E., and Helz, G.R., 2000. Molybdenum (VI) speciation in sulfidic waters: Stability and lability of thiomolybdates. Geochim. Geochimica et Cosmochimica Acta 64, 1149\u00E2\u0080\u00931158. Foote, M., 2000. Origination and extinction components of taxonomic diversity: General problems. Paleobiology supplement, 26, 74\u00E2\u0080\u0093102. 222 Frebold, H., 1964. Illustrations of Canadian Fossils Jurassic of Arctic and Western Canada. Geological Survey of Canada, Paper 63\u00E2\u0080\u00934, 107 p. Frebold, H., 1970. Pliensbachian ammonoids from British Columbia and southern Yukon. Canadian Journal of Earth Sciences, 7, 435\u00E2\u0080\u0093456. Froelich, P.N., Klinkhammer, G.P., Bender, M.L., Luedtke, N.A., Heath, G.R., Cullen, D., and Dauphin, P., 1979. Early oxidation of organic matter in pelagic sediments of the eastern equatorial Atlantic: Suboxic diagenesis. Geochimica et Cosmochimica Acta, 43, 1075\u00E2\u0080\u00931090. Fucini, A., 1901. Ammoniti del Lias medio dell\u00E2\u0080\u0099Appennino centrale. Palaeontographia Italica, Pisa, 6, 43\u00E2\u0080\u0093104. Gallitelli\u00E2\u0080\u0093Wendt, M.F., 1969. Ammonitie stratigrafia del Toarciano umbro- marchigiano (Appennino centrale); Bollettino della Societ\u00C3\u00A0 Paleontologica Italiana, 8, 11\u00E2\u0080\u009362. Garcia Joral, F., Gomez, J.J., and Goy, A., 2011. Mass extinction and recovery of the Early Toarcian (Early Jurassic) brachiopods linked to climate change in northern and central Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 302, 367\u00E2\u0080\u0093380. Gehrels, G.E., 2001. Geology of the Chatham Sound region, Southeast Alaska and coastal British Columbia. Canadian Journal of Earth Sciences, Ottawa, 38, 1579\u00E2\u0080\u00931599. Geyer, G., 1893. Die mittel-liasische Cephalopoden-Fauna des Hinter- Schafberges in Ober\u00C3\u00B6sterreich. Abhandlungen der Kaiserlich-k\u00C3\u00B6niglichen geologischen Reichsanstalt, Austria, 15, 1\u00E2\u0080\u009376. Gill, B.C., Lyons, T.W., and Jenkyns, H.C., 2011. A global perturbation to the sulfur cycle during the Toarcian Oceanic Anoxic Event. Earth and Planetary Science Letters, 312, 484\u00E2\u0080\u0093496. G\u00C3\u00B3mez, J.J., Goy, A., and Canales, M.L., 2008. Seawater temperature and carbon isotope variations in belemnites linked to mass extinction during the Toarcian (Early Jurassic) in Central and Northern Spain. Comparison with other European sections. Palaeogeography, Palaeoclimatology, Palaeoecology, 258, 28\u00E2\u0080\u009358. Grantz, A., 1960a. Geologic map of Talkeetna Mountains (A-2) quadrangle, Alaska and the contiguous area to the north and northwest. United States Geological Survey Miscellaneous Geologic Investigations Map I-313, 1:48,000 scale. 223 Grantz, A., 1960b. Geologic map of Talkeetna Mountains (A-1) quadrangle, Alaska and the contiguous area to the north and northwest. United States Geological Survey Miscellaneous Geologic Investigations Map I-314, Denver, 1:48,000 scale. Greene, A.R., Scoates, J.S., Weis, D., Katvala, E.C., Israel, S., and Nixon, G.T., 2010. The architecture of oceanic plateaus revealed by the volcanic stratigraphy of the accreted Wrangellia oceanic plateau. Geosphere, 6, 47\u00E2\u0080\u009373. Gr\u00C3\u00B6cke , D.R., Hesselbo, S.P., and Jenkyns, H.C., 1999. Carbon-isotope composition of Lower Cretaceous fossil wood: Ocean-atmosphere chemistry and relation to sea-level change. Geology, 27, 155\u00E2\u0080\u0093158. Gr\u00C3\u00B6cke, D.R., Hesselbo, S.P., and Findlay, D.J., 2007. Atypical diagenetic effects on strontium-isotope composition of Early Jurassic belemnites, Queen Charlotte Islands, British Columbia, Canadian Journal of Earth Sciences, 44, 181\u00E2\u0080\u0093197. Gr\u00C3\u00B6cke, D.R., Hori, R.S., Trabucho-Alexandre, J., Kemp, D.B., and Schwark, L., 2011. An Open Ocean Record of the Toarcian Oceanic Anoxic Event. Solid Earth, 2, 245\u00E2\u0080\u0093257. Guex, J., 1973. Observations sur la r\u00C3\u00A9partition biostratigraphique des ammonites du Toarcien sup\u00C3\u00A9rieur del\u00E2\u0080\u0099Aveyron (France); Bulletin des Laboratoires de G\u00C3\u00A9ologie, Min\u00C3\u00A9ralogie, G\u00C3\u00A9ophysique et du Mus\u00C3\u00A9e G\u00C3\u00A9ologique de l\u00E2\u0080\u0099Universit\u00C3\u00A9 de Lausanne, 207, 7\u00E2\u0080\u009314. Haggart, J.W., Enkin, R.J., and Monger, J.W.H., (Eds.), 2006. Paleogeography of the North American Cordillera; Evidence for and against Large-Scale Displacements. Geological Association of Canada Special Paper 46, 420 p. Hall, R.L., and Howarth, M.K., 1983. Protogrammoceras paltum (Buckman), a Late Pliensbachian (Jurassic) ammonite from Axel Heiberg Island, Canadian Arctic Archipelago. Canadian Journal of Earth Sciences, 20, 1470\u00E2\u0080\u00931475. Hallam, A., 1986. The Pliensbachian and Tithonian extinction events. Nature, 319, 765\u00E2\u0080\u0093768. Hallam, A., 1987. Radiations and extinctions in relation to environmental change in the marine Jurassic of north west Europe. Paleobiology, 13, 152\u00E2\u0080\u0093168. Hallam, A., 1995. Oxygen-restricted facies of the basal Jurassic of north west Europe. Historical Biology, 10, 247\u00E2\u0080\u0093257. Hallam, A., and Wignall, P.B., 1997. Mass Extinctions and Their Aftermath. Oxford University Press, Oxford, 320 p. 224 Hammer, \u00C3\u0098., 2003. Biodiversity curves for the Ordovician of Baltoscandia. Lethaia, 36, 305\u00E2\u0080\u0093313. Hammer, \u00C3\u0098., and Harper, D., 2006. Paleontological Data Analysis. Blackwell Publishing, Massachusetts, 351 p. Hansen, B.M., 1957. Middle Permian limestone on Pacific side of Alaska Peninsula. American Association of Petroleum Geologists Bulletin, 41, 2376\u00E2\u0080\u0093 2378. Harries, P.J., and Little, C.T.S., 1999. The early Toarcian (Early Jurassic) and the Cenomanian\u00E2\u0080\u0093Turonian (Late Cretaceous) mass extinctions: similarities and contrasts. Palaeogeography, Palaeoclimatology, Palaeoecology, 154, 39\u00E2\u0080\u009366. Hart, M.B., 1999. The evolution and biodiversity of Cretaceous planktonic Foraminiferida. Geobios, 32, 247\u00E2\u0080\u0093255. Hart, M.B., Oxford, M.J., and Hudson, W., 2002. The early evolution and palaeobiogeography of Mesozoic planktonic foraminifera. Geological Society, London, Special Publications, 194, 115\u00E2\u0080\u0093125. von Hauer, F.R., 1856. \u00C3\u009Cber die Cephalopoden aus dem Lias der nord\u00C3\u00B6stlichen Alpen. Denkschriften der Mathematisch-Naturwissenschaftlichen Klasse der Kaiserlichen Akademie der Wissenschaften, Wien, 11, 86 p. Helz, G.R., Miller, C.V., Chamock, J.M., Mosselmans, J.F.M., Pattrick, R.A.D., Gardner, C.D., and Vaughan, D.J., 1996. Mechanism of molybdenum removal from the sea and its concentration in black shales: EXAFS evidence. Geochimica et Cosmochimica Acta, 60, 3631\u00E2\u0080\u00933642. Hermoso, M., Le Callonnec, L., Minoletti, F., Renard, M. and Hesselbo, S.P., 2009a. Expression of the Early Toarcian negative carbon-isotope excursion in separated carbonate microfractions (Jurassic, Paris Basin). Earth and Planetary Science Letters, 277, 194\u00E2\u0080\u0093203. Hermoso, M., Minoletti, F., Le Callonnec, L., Jenkyns, H.C., Hesselbo, S.P., Rickaby, R.E.M., Renard, M., de Raf\u00C3\u00A9lis, M., and Emmanuel, L, 2009b. Global and local forcing of Early Toarcian seawater chemistry: A comparative study of different paleoceanographic settings (Paris and Lusitanian basins). Paleoceanography, 24, PA4208, doi:10.1029/2009PA001764. Hesselbo, S.P., and Pienkowski, G., 2011. Stepwise atmospheric carbon-isotope excursion during the Toarcian Oceanic Anoxic Event (Early Jurassic, Polish 225 Basin). Earth and Planetary Science Letters, 301, 365\u00E2\u0080\u0093372, doi:10.1016/j.epsl.2010.11.021. Hesselbo, S.P., Gr\u00C3\u00B6cke, D.R., Jenkyns, H.C., Bjerrum, C.J., Farrimond, P., Morgans Bell, H.S., and Green, O.R., 2000. Massive dissociation of gas hydrate during a Jurassic oceanic anoxic event. Nature, 406, 392\u00E2\u0080\u0093395. Hesselbo, S.P., Robinson, S.A., Surlyk, F., and Piasecki, S., 2002. Terrestrial and marine extinction at the Triassic\u00E2\u0080\u0093Jurassic boundary synchronized with major carbon-cycle perturbation: a link to initiation of massive volcanism? Geology, 30, 251\u00E2\u0080\u0093254. Hesselbo, S.P., Robinson, S.A., and Surlyk, F., 2004. Sea-level change and facies development across potential Triassic\u00E2\u0080\u0093Jurassic boundary horizons, SW Britain. Journal of the Geological Society, London, 161, 365\u00E2\u0080\u0093379. Hesselbo, S.P., Jenkyns, H.C., Duarte, L.V., and Oliveira, L.C.V., 2007. Carbon- isotope record of the Early Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine carbonate (Lusitanian Basin, Portugal). Earth and Planetary Science Letters, 253, 455\u00E2\u0080\u0093470. von Hillebrandt, A., 1987. Liassic ammonite zones of South America and correlations with other provinces, with descriptions of new genera and species of ammonites. In: Biostratigrafia de los sistemas regionales del Jur\u00C3\u00A1sico y Cret\u00C3\u00A1cico en Am\u00C3\u00A9rica del Sur. Mendoza, Argentina, 111\u00E2\u0080\u0093157. von Hillebrandt, A., 2006. Ammoniten aus dem Pliensbachium (Carixium und Domerium) von S\u00C3\u00BCdamerika. Revue de Pal\u00C3\u00A9obiologie, 25, 1\u00E2\u0080\u0093403. Hillhouse, J.W., and Coe, R.S., 1994. Paleomagnetic data from Alaska. In: Plafker, G. and Berg, H.G., (Eds.), The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder, G\u00E2\u0080\u00931, 797\u00E2\u0080\u0093812. Holser, W.T., Sch\u00C3\u00B6nlaub, H.-P., Attrep, M. Jr., Boeckelmann, K., Klein, P., Magaritz, M., Pak, E., Schramm, J.-M., Stattgegger, K., and Schm\u00C3\u00B6ller, R., 1989. A unique geochemical record at the Permian/Triassic boundary. Nature, 337, 39\u00E2\u0080\u0093 44. Houseknecht, D.W., and Bird, K.J., 2004. Sequence stratigraphy of the Kingak Shale (Jurassic\u00E2\u0080\u0093Lower Cretaceous), National Petroleum Reserve in Alaska. American Association of Petroleum Geologists Bulletin, 88, 279\u00E2\u0080\u0093302. Howarth M.K., 1958. A monograph of the ammonites of the Liassic family Amaltheidae in Britain. Palaeontographical Society, pt. 1, UK, 53p. 226 Howarth, M.K., 1992. The ammonite family Hildoceratidae in the Lower Jurassic of Britain. Palaeontographical Society Monograph 145, 1\u00E2\u0080\u0093200. Hubbard, R.J., Edrich, S.P., and Rattey, R.P., 1987a. Geologic evolution and hydrocarbon habitat of the \u00E2\u0080\u009CArctic Alaska microplate.\u00E2\u0080\u009D In: Tailleur, I., and Weimer, P., (Eds.), Alaskan North Slope Geology. SEPM Pacific Section and Alaska Geological Society, 50, 797\u00E2\u0080\u0093830. Hubbard, R.J., Edrich, S.P., and Rattey, R.P., 1987b. Geologic evolution and hydrocarbon habit of the \u00E2\u0080\u009CArctic Alaska microplate.\u00E2\u0080\u009D Marine and Petroleum Geology, 4, 2\u00E2\u0080\u009334. Hyatt, A., 1867. The fossil Cephalopods of the Museum of Comparative Zoology. Bulletin of the Museum of Comparative Zoology, Cambridge, 1, 71\u00E2\u0080\u0093102. Hyatt, A., 1875. Remarks on two new genera of ammonites: Agassiceras and Oxynoticeras. Boston Society of Natural History, Proceedings, 17, 225\u00E2\u0080\u0093235. Hyatt, A., 1889. Genesis of the Arietidae. Smithsonian Contributions to Knowledge, No. 673, Washington D.C., 238p. Imlay, R.W., 1955. Characteristic Jurassic mollusks from northern Alaska. United States Geological Survey Professional Paper 274-D, 69\u00E2\u0080\u009396. Imlay, R.W., 1968. Lower Jurassic (Pliensbachian and Toarcian) ammonites from eastern Oregon and California. United States Geological Survey Professional Paper 593-C, C1\u00E2\u0080\u0093C51. Imlay, R.W., 1981. Early Jurassic ammonites from Alaska. United States Geological Survey Professional Paper 1190, Denver, 40p. Imlay, R.W. and Detterman, R.L., 1973. Jurassic paleobiogeography of Alaska. United States Geological Survey Professional Paper 801, 34p. Izumi, K., Miyaji, T., and Tanabe, K., 2012. Early Toarcian (Early Jurassic) oceanic anoxic event recorded in the shelf deposits in the northwestern Panthalassa: Evidence from the Nishinakayama Formation in the Toyora area, west Japan. Palaeogeography, Palaeoclimatology, Palaeoecology, 315\u00E2\u0080\u0093316, 100\u00E2\u0080\u0093108. Jablonski, D., 1995. Extinction in the fossil record. In: May, R.M., and Lawton, J.H., (Eds.), Extinction rates. Oxford: Oxford University Press, 25\u00E2\u0080\u009344. Jakobs, G.K., 1990. A discussion of the Phantom Creek Formation of the Maude Group, Queen Charlotte Islands, British Columbia. In: Current Research, Part F, Geological Survey of Canada, Paper 90\u00E2\u0080\u00931F, 57\u00E2\u0080\u009360. 227 Jakobs, G.K., 1992. Toarcian (Lower Jurassic) ammonite biostratigraphy and ammonite fauna of North America. PhD. Thesis, University of British Columbia, Vancouver, British Columbia, 682p. Jakobs, G.K., 1995. New occurrences of Leukadiella and Paroniceras (Ammonoidea) from the Toarcian (Lower Jurassic) of the Canadian Cordillera. Journal of Paleontology, 69, 89\u00E2\u0080\u009398. Jakobs, G.K., 1997. Toarcian (Early Jurassic) Ammonoids from Western North America. Geological Survey of Canada Bulletin 428, 1\u00E2\u0080\u0093137. Jakobs, G.K., and Smith, P.L., 1996. Latest Toarcian ammonoids from the North American Cordillera. Palaeontology, 39, 97\u00E2\u0080\u0093147. Jakobs, G.K., Smith, P.L., and Tipper, H.W., 1994. An ammonite zonation for the Toarcian (Lower Jurassic) of the North American Cordillera. Can. J. Earth Sci. 31, 919\u00E2\u0080\u0093942. Jeffords, R.M., 1957. Permian fossils from an island in Puale Bay, Alaska. Humble Oil and Refining Company Geologic Research Report 57-24, Houston, USA, 6p. Jenkyns, H.C., 1985. The Early Toarcian and Cenomanian-Turonian anoxic events in Europe: comparisons and contrasts. International Journal of Earth Sciences (Geol. Rundsch), 74, 505\u00E2\u0080\u0093518. Jenkyns, H.C., 1988. The Early Toarcian (Jurassic) event: stratigraphy, sedimentary and geochemical evidence. American Journal of Science, 288, 101\u00E2\u0080\u0093 151. Jenkyns, H.C., 2003. Evidence for rapid climate change in the Mesozoic\u00E2\u0080\u0093 Palaeogene greenhouse world. Philosophical Transactions of the Royal Society of London Series A., 361, 1885\u00E2\u0080\u00931916. Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochemistry Geophysics, Geosystems, 11, Q03004, doi:10.1029/2009GC002788. Jenkyns, H.C., and Clayton, C.J., 1986. Black Shales and carbon isotopes from the Tethyan Lower Jurassic. Sedimentology, 33, 87\u00E2\u0080\u0093106, doi: 10.1111j.1365- 3091.1986.tb00746.x. Jenkyns, H.C., and Clayton, C.J., 1997. Lower Jurassic epicontinental carbonates and mudstones from England and Wales: Chemostratigraphic signals and the early Toarcian anoxic event. Sedimentology, 44, 687\u00E2\u0080\u0093706, doi: 10.1046/j.1365-3091.1997.d01-43.x. 228 Jenkyns, H.C., Gr\u00C3\u00B6cke, D.R., and Hesselbo, S.P., 2001. Nitrogen isotope evidence for water mass denitrification during the early Toarcian (Jurassic) ocean anoxic event. Paleoceanography 16, 593\u00E2\u0080\u0093603. Jenkyns, H.C., Jones, C.E., Gr\u00C3\u00B6cke, D.R., Hesselbo, S.P., and Parkinson, D.N., 2002. Chemostratigraphy of the Jurassic System: Applications, limitations and implications for palaeoceanography. Journal of the Geological Society, London, 159, 351\u00E2\u0080\u0093378. Jenkyns, H.C., Matthews, A., Tsikos, H., and Erel, Y., 2007. Nitrate reduction, sulfate reduction, and sedimentary iron isotope evolution during the Cenomanian-Turonian oceanic anoxic event. Paleoceanography, 22, PA3208, doi:10.1029/2006PA001355. Jones, C.E., and Jenkyns, H.C., 2001. Seawater strontium isotopes, oceanic anoxic events, and seafloor hydrothermal activity in the Jurassic and Cretaceous. American Journal of Science, 301, 112\u00E2\u0080\u0093149, doi:10.2475/ajs.301.2.112. Jones, D.L. and Silberling, N.J., 1979. Mesozoic stratigraphy\u00E2\u0080\u0094The key to tectonic analysis of southern and central Alaska. United States Geological Survey Open-File Report 79-1200, 41p. Jones, D.L., Silberling, N.J., and Hillhouse, J., 1977. Wrangellia\u00E2\u0080\u0093A displaced terrane in northwestern North America. Canadian Journal of Earth Sciences, 14, 2565\u00E2\u0080\u00932577, doi:10.1139/e77-222. Jones, D.L., Cox, A., Coney, P., and Beck, M., 1982. The growth of Western North America. Scientific American, London, 247, 70\u00E2\u0080\u009384. Jones, C.E., Jenkyns, H.C., and Hesselbo, S.P., 1994. Strontium isotopes in Early Jurassic seawater. Geochimica et Cosmochimica Acta, 58, 1285\u00E2\u0080\u00931301, doi:10.1016/0016-7037(94)90382-4. Jourdan, F., F\u00C3\u00A9raud, G., Bertrand, H., Watkeys, M.K., Renne, P.R., 2008. The 40Ar/39Ar ages of the sill complex of the Karoo large igneous province: Implications for the Pliensbachian\u00E2\u0080\u0093Toarcian climate change. Geochemistry, Geophysics, Geosystems, 9, 1\u00E2\u0080\u009320. Q06009, doi:10.1029/2008GC001994. Kapp, P.A. and Gehrels, G.E., 1998. Detrital zircon constraints on the tectonic evolution of the Gravina belt, southeastern Alaska. Canadian Journal of Earth Sciences, 35, 253\u00E2\u0080\u0093268. Keller, G., 2008. Cretaceous climate, volcanism, impacts, and biotic effects. Cretaceous Research, 29, 754\u00E2\u0080\u0093771. 229 Kemp, D.B., Coe, A.L., Cohen, A.S., and Schwark, L., 2005. Astronomical pacing of methane release in the Early Jurassic period. Nature 437, 396\u00E2\u0080\u0093399. Kemp, D.B., Coe, A.L., Cohen, A.S., and Weedon, G.P., 2011. Astronomical forcing and chronology of the early Toarcian (Early Jurassic) oceanic anoxic event in Yorkshire, UK. Paleoceanography, 26, PA4210. doi:10.1029/2011PA002122. Kottachchi, N., 2001. Jurassic Foraminifera from the Queen Charlotte Islands, British Columbia: biostratigraphy, paleoenvironments and paleogeographic implications. M.S. Thesis, Carleton University, Ottawa, Ontario, 122p. Kottachchi, N., Schr\u00C3\u00B6der-Adams, C.J., Haggart, J.W., and Tipper, H.W., 2002. Jurassic foraminifera from the Queen Charlotte Islands, British Columbia, Canada: biostratigraphy, paleoenvironments and paleogeographic implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 180, 93\u00E2\u0080\u0093127. Kottachchi, N., Schr\u00C3\u00B6der-Adams, C.J., Haggart, J.W., and Page, J.E., 2003. Lower and Middle Jurassic Foraminifera of Queen Charlotte Islands, British Columbia: raw data and preliminary results. Geological Survey of Canada Open File 1739, 47p. Kottek, A.V., 1966. Die Ammonitenabfolge des griechischen Toarcium. Annales g\u00C3\u00A9ologiques des Pays hell\u00C3\u00A9niques, Premi\u00C3\u00A8re S\u00C3\u00A9rie, 17, 1\u00E2\u0080\u0093157. Kuroda, J., Hori, R.S., Suzuki, K., Gr\u00C3\u00B6cke, D.R., Ohkouchi, N., 2010. Marine osmium isotope record across the Triassic-Jurassic boundary from a Pacific pelagic site. Geology, 38, 1095\u00E2\u0080\u00931098. doi: 10.1130/G31223.1. K\u00C3\u00BCspert, W., 1982. Environmental changes during oil shale deposition as deduced from stable isotope ratios. In: Einsele, G., and Seilacher, A. (Eds.), Cyclic and Event Stratification. Springer, Berlin, 482\u00E2\u0080\u0093501. Lawton, J.H., and May, R.M., 1995. Extinction rates. Oxford University Press, Oxford. Lawver, L.A., and Scotese, C.R., 1990. A review of tectonic models for the evolution of the Canada Basin. In: Grantz, A., Johnson, G.L., and Sweeney, J.F., (Eds.), the Arctic Ocean region. Geological Society of America, Geology of North America, L, 593\u00E2\u0080\u0093618. Lawver, L.A., Grantz, A., and Gahagan, L.M., 2002. Plate kinematic evolution of the present Arctic region since the Ordovician. In: Miller, E.L., Grantz, A., Klemperer, S.L., (Eds.), Tectonic Evolution of the Bering Shelf\u00E2\u0080\u0093Chukchi Sea\u00E2\u0080\u0093 Arctic Margin and Adjacent Landmasses. Geological Society of America Special Paper 360, 333\u00E2\u0080\u0093358. 230 Leakey, R.E., and Lewin, R., 1995. The sixth extinction: patterns of life and the future of humankind. Doubleday, New York. Little, C.T.S., 1995. The Pliensbachian\u00E2\u0080\u0093Toarcian (Lower Jurassic) Extinction Event. PhD. Thesis, University of Bristol, Bristol, England. Little, C.T.S., 1996. the Pliensbachian\u00E2\u0080\u0093Toarcian (Lower Jurassic) extinction event. In: Ryder, G., Fastovsky, D., Gartner, S., (Eds.), the Cretaceous\u00E2\u0080\u0093Tertiary Event and Other Catastrophes in Earth History. Geological Society of America Special Paper 307, 505\u00E2\u0080\u0093512. Little, C.T.S., and Benton, M.J., 1995. Early Jurassic mass extinction: A global long-term event. Geology, 23, 495\u00E2\u0080\u0093498. Littler, K., Hesselbo, S.P., and Jenkyns, H.C., 2010. A carbon-isotope perturbation at the Pliensbachian\u00E2\u0080\u0093Toarcian boundary: evidence from the Lias Group, NE England. Geological Magazine, 147, 181\u00E2\u0080\u0093192. Magaritz, M., Krishnamurthy, R.V., and Holser, W.T., 1992. Parallel trends in organic and inorganic carbon isotopes across the Permian/Triassic boundary. American Journal of Science, 292, 727\u00E2\u0080\u0093739. Mattioli, E., and Pittet, B., Bucefalo Palliani, R., R\u00C3\u00B6hl, H.J., Schmid-R\u00C3\u00B6hl, A., and Morettini, E., 2004, Phytoplankton evidence for timing and correlation of palaeoceanographical changes during the Early Toarcian oceanic anoxic event (Early Jurassic), Journal of the Geological Society of London, 161, 685\u00E2\u0080\u0093693. Mazzini, A., Svensen, H., Leanza, H.A., Corfu, F., and Planke, S., 2010. Early Jurassic shale chemostratigraphy and U\u00E2\u0080\u0093Pb ages from the Neuqu\u00C3\u00A9n Basin (Argentina): Implications for the Toarcian Oceanic Anoxic Event. Earth and Planetary Science Letters, 297, 633\u00E2\u0080\u0093645, doi: 10.1016/j.epsl.2010.07.017. McArthur, J.M., Donovan, D.T., Thirlwall, M.F., Fouke, B.W., and Mattey, D., 2000. Strontium isotope profile of the early Toarcian (Jurassic) oceanic anoxic event, the duration of ammonite biozones, and belemnite palaeotemperatures. Earth and Planetary Science Letters, 179, 269\u00E2\u0080\u0093285. McArthur, J.M., Algeo, T.J., van de Schootbrugge, B., Li, Q., and Howarth, R.J., 2008. Basinal restriction, black shales, Re-Os dating, and the Early Toarcian (Jurassic) oceanic anoxic event. Paleoceanography, 23, 1\u00E2\u0080\u009322. McElwain, J.C., Wade-Murphy, J., and Hesselbo, S.P., 2005. Changes in carbon dioxide during an oceanic anoxic event linked to intrusion into Gondwana coals. Nature, 435, 479\u00E2\u0080\u0093482. 231 McClelland, W.C. and Gehrels, G.E., 1990. The Duncan Canal shear zone: A right-lateral shear zone of Jurassic age along the inboard margin of the Alexander terrane. Geological Society of America Bulletin, 102, 1378\u00E2\u0080\u00931392. McClelland, W.C., Gehrels, G.E., and Saleeby, J.B., 1992. Upper Jurassic\u00E2\u0080\u0093 Lower Cretaceous basinal strata along the Cordilleran margin: Implications for the accretionary history of the Alexander-Wrangellia-Peninsular terrane. Tectonics, 11, 823\u00E2\u0080\u0093835. McIver, R., 1975. Hydrocarbon occurrences from JOIDES Deep Sea Drilling Project. Proceedings of the 9th World Petroleum Congress (Tokyo), 2. Applied Science Publishers, Barking, 269\u00E2\u0080\u0093280. McLearn, F.H., 1930. Notes on some Canadian Mesozoic faunas. Royal Society of Canada, Proceedings and Transactions, Series 3, 24, 1\u00E2\u0080\u00937. McLearn, F.H., 1932. Contributions to the stratigraphy and palaeontology of Skidegate Inlet, Queen Charlotte Islands, British Columbia. Transactions of the Royal Society of Canada, series 3, 24, 1\u00E2\u0080\u00937. Meister, C., 2010. Worldwide ammonite correlation at the Pliensbachian Stage and Substage Boundaries (Lower Jurassic). Stratigraphy, 7, 83\u00E2\u0080\u0093101. Meneghini, J., 1867. Monographie des fossiles du calcaire rouge ammonitique (Lias sup\u00C3\u00A9rieur) de Lombardie et de l\u00E2\u0080\u0099Appennin Centrale. In: Stoppani, A. (Ed.), Pal\u00C3\u00A9ontologie Lombarde, Series 4, Milan, Italy, 1\u00E2\u0080\u009324. Mickey, M.B., Haga, H., and Bird, K.J., 2006. Micropaleontology of Selected Wells and Seismic Shot Holes, Northern Alaska. U.S. Geological Survey Open- File Report 2006-1055, 11p. Miller, E.L., Grantz, A., Klemperer, S.L., (Eds.), 2002. Tectonic Evolution of the Bering Shelf\u00E2\u0080\u0093Chukchi Sea\u00E2\u0080\u0093Arctic Margin and Adjacent Landmasses. Geological Society of America Special Paper 360, 387p. Miller, E.L., Toro, J., Gehrels, G., Amato, J.M., Prokopiev, A., Tuchkova, M.I., Akinin, V.V., Dumitru, T.A., Moore, T.E., and Cecile, M.P., 2006. New insights into Arctic paleogeography and tectonics from U-Pb detrital zircon geochronology. Tectonics, 25, TC3013, 19p. Monestier, J., 1921. Ammonites rares ou peu connues et ammonites nouvelles du Toarcien sup\u00C3\u00A9rieur du sud-est de l\u00E2\u0080\u0099Aveyron. M\u00C3\u00A9moires de la Soci\u00C3\u00A9t\u00C3\u00A9 de France, Pal\u00C3\u00A9ontologie, M\u00C3\u00A9moire 54, 44p. de Montfort, D., 1808. Conchyliologie syst\u00C3\u00A9matique et classification m\u00C3\u00A9thodique des Coquilles. Paris, 1, 409p. 232 Moore, T.E., Wallace, W.K., Bird, K.J., Karl, S.M., Mull, C.G., and Dillon, J.T., 1994. Geology of northern Alaska. In: Plafker, G. & H.G. Berg (Eds.), The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder, 49\u00E2\u0080\u0093140. Moore, T.E., Wallace, W.K., Mull, C.G., Adams, K.E., Plafker, G., and Nokleberg, W.J., 1997. Crustal implications of bedrock geology along the Trans-Alaska Crustal Transect (TACT) in the Brooks Range, northern Alaska. Journal of Geophysical Research, 102, 20,645\u00E2\u0080\u009320,684. doi: 10.1029/96JB03733. Morgan, W.J., 1971. Convection plumes in the lower mantle. Nature, 230, 42\u00E2\u0080\u009343. Morgan, W.J., 1972. Deep mantle convection plumes and plate motion. American Association of Petroleum Geologists Bulletin, 56, 203\u00E2\u0080\u0093213. Morgan, W.J., 1981. Hotspot tracks and the opening of the Atlantic and Indian oceans. In: Emiliani, C., (Ed.), The Sea, The Oceanic Lithosphere, vol. 7, Wiley- Interscience, New York, 443\u00E2\u0080\u0093487. Morrison, A.D., and Reay, A., 1995. Geochemistry of Ferrar Dolerite sills and dykes at Terra Cotta Mountain, south Victoria Land, Antarctica. Antarctic Science 7, 73\u00E2\u0080\u009385. Mouterde, R., Enay, R., Cariou, E., Contini, D., Elmi, S., Gabilly, J., Mangold, C., Mattei, J., Rioult, M., Thierry, J., and Tintant, H., 1971. Les zones du Jurassique en France. Compte rendu sommaire des s\u00C3\u00A9ances de la Soci\u00C3\u00A9t\u00C3\u00A9 g\u00C3\u00A9ologique de France, Fasc. 6, s\u00C3\u00A9ance du premiere mars 1971, extract. Mouterde, R., Dommergues, J.-L., Meister, C., and Rocha, R.B., 2007. Atlas des fossiles caract\u00C3\u00A9ristiques du Lias portugais III a) Dom\u00C3\u00A9rien (Ammonites). Ci\u00C3\u00AAncias da Terra (Universidade Nova de Lisboa), 16, 67\u00E2\u0080\u0093111. Moxon, C., 1841. Illustration of the Characteristic Fossils of British Strata, London, 46p., 34 pl. Nagy, J., and Johansen, H.O., 1991. Delta-Influenced Foraminiferal Assemblages from the Jurassic (Toarcian-Bajocian) of the Northern North Sea. Micropaleontology, 37, 40p. Natal\u00E2\u0080\u0099in, B.A., Amato, J.M., Toro, J., and Wright, J.E., 1999. Paleozoic rocks of northern Chukotka Peninsula, Russian Far East: Implications for the tectonics of the Arctic region. Tectonics, 18, 977\u00E2\u0080\u00931003. 233 Newton, C.R., 1989. Triassic-Jurassic boundary section at Puale Bay, Alaska Peninsula: Comparative diversity pattern of skeletal faunas and trace fossils. American Association of Petroleum Geologists Bulletin, 74, 730p. Nikitenko, B.L., 1994. Early and Middle Jurassic Ostracodes of Northern Siberia: Principal Regularities in Evolution and the Zonal Scale. Stratigraphy and Geological Correlation, 4, 340\u00E2\u0080\u0093357. Nikitenko, B.L., 2008. The Early Jurassic to Aalenian Paleobiogeography of the Arctic Realm: Implication of Microbenthos (Foraminifers and Ostracodes). Stratigraphy and Geological Correlation, 16, 59\u00E2\u0080\u009380. Nikitenko, B.L., and Mickey, M.B., 2004. Foraminifera and Ostracodes across the Pliensbachian\u00E2\u0080\u0093Toarcian Boundary in the Arctic Realm (stratigraphy, palaeobiogeography and biofacies). In: Beaudoin, A.B., and Head, M.J., (Eds.) The Palynology and Micropalaeontology of Boundaries. Geological Society, London, Special Publications 230, 137\u00E2\u0080\u0093173. Nikitenko, B.L., and Shurygin, B.N., 1992. Lower Toarcian Black Shales and Pliensbachian\u00E2\u0080\u0093Toarcian Crisis of the Biota of Siberian Paleoseas. ICAM proceedings 39\u00E2\u0080\u009344. Nokleberg, W.L., Plafker, G., and Wilson, F.H., 1994. Geology of south-central Alaska. In: Plafker, G. & H.G. Berg (Eds.), The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder, 311\u00E2\u0080\u0093365. Oppel, A., 1853. Der mittlere Lias schwabens neu bearbeitet. W\u00C3\u00BCrttemberger Naturwissenschaffen Jahreshefte, Stuttgart, 10, 92p. Oppel, A., 1862. \u00C3\u009Cber jurassische Cephalopoden. Pal\u00C3\u00A4ontologische Mitteilungen, Museum k\u00C3\u00B6niglich bayerischen Staates, Munich, Germany, 1, 127\u00E2\u0080\u0093162. Page, K.N., 2003. The Lower Jurassic of Europe: its subdivision and correlation. Geological Survey of Denmark and Greenland Bulletin, 1, 23\u00E2\u0080\u009359. P\u00C3\u00A1lfy, J., 2003. Volcanism of the Central Atlantic Magmatic Province as a Potential Driving Force in the End-Triassic Mass Extinction. In: Hames, W.E., McHone, J.G., Renne, P.R., and Ruppel, C., (Eds.), the Central Atlantic Magmatic Province: Insights from Fragments of Pangea. American Geophysical Union Geophysical Monograph Series, 136, 255\u00E2\u0080\u0093267. P\u00C3\u00A1lfy, J., and Smith, P.L., 2000. Synchrony between Early Jurassic extinction, oceanic anoxic event, and the Karoo\u00E2\u0080\u0093Ferrar flood basalt volcanism. Geology, 28, 747\u00E2\u0080\u0093750. 234 P\u00C3\u00A1lfy, J., Smith, P.L., and Tipper, H.W., 1994. Sinemurian (Lower Jurassic) Ammonoid Biostratigraphy of the Queen Charlotte Islands, Western Canada. Geobios, 17, 385\u00E2\u0080\u0093393. P\u00C3\u00A1lfy, J., Parrish, R.R., and Smith, P.L., 1997. A U\u00E2\u0080\u0093Pb age from the Toarcian (Lower Jurassic) and its use for time scale calibration through error analysis of biochronologic dating. Earth and Planetary Science Letters, 146, 659\u00E2\u0080\u0093675. P\u00C3\u00A1lfy, J., Smith, P.L., Mortensen, J.K., and Friedman, R.M., 1999. Integrated ammonite biochronology and U\u00E2\u0080\u0093Pb geochronometry from a basal Jurassic section in Alaska. Geological Society of America Bulletin, Boulder, 111, 1537\u00E2\u0080\u0093 1549. P\u00C3\u00A1lfy, J., Smith, P.L., and Mortensen, J.K., 2000. A U-Pb and 40Ar-39Ar time scale for the Jurassic. Canadian Journal of Earth Sciences, 37, 923\u00E2\u0080\u009344. P\u00C3\u00A1lfy, J., Dem\u00C3\u00A9ny, A., Haas, J., Het\u00C3\u00A9nyi, M., Orchard, M.J., and Vet\u00C3\u00B6, I., 2001. Carbon isotope anomaly and other geochemical changes at the Triassic/Jurassic boundary from a marine section in Hungary. Geology, 29, 1047\u00E2\u0080\u00931050. Pavlis, T.L., 1982. Origin and age of the Border Ranges fault of southern Alaska and its bearing on the late Mesozoic tectonic evolution of Alaska. Tectonics, 1, 343\u00E2\u0080\u0093368. Payne, J.L., and Kump, L.R., 2007. Evidence for recurrent Early Triassic massive volcanism from quantitative interpretation of carbon isotope fluctuations. Earth and Planetary Science Letters, 256, 264\u00E2\u0080\u0093277. Payne, J.L., Lehrmann, D.J., Wei, J.Y., Orchard, M.J., Schrag, D.P., and Knoll, A.H., 2004. Large Perturbations of the carbon cycle during recovery from the end-Permian extinction. Science, 305, 506\u00E2\u0080\u0093509. Pearce, C.R., Cohen, A.S., Coe, A.L., and Burton, K.W., 2008. Molybdenum isotope evidence for global ocean anoxia coupled with perturbations to the carbon cycle during the Early Jurassic. Geology, 36, 231\u00E2\u0080\u0093234. doi: 10.1130/G24446A.1. Pederson, T.F., and Calvert, S.E., 1990. Anoxia vs. Productivity: What Controls the Formation of Organic-Carbon-Rich Sediments and Sedimentary Rocks? The American Association of Petroleum Geologists Bulletin, 74, 454\u00E2\u0080\u0093466. Peters, S.E., 2008. Environmental determinants of extinction selectivity in the fossil record. Nature, 454, 626\u00E2\u0080\u0093630. Pimm, S.L., Russel, G.J., Gittleman, J.L., and Brooks, T.M., 1995. The future of biodiversity. Science, 269, 347\u00E2\u0080\u0093350. 235 Plafker, G. & H.G. Berg (Eds.), 1994. The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder. Plafker, G., Nokleberg, W.J., and Lull, J.S., 1989. Bedrock geology and tectonic evolution of the Wrangellia, Peninsular, and Chugach terranes along the Trans- Alaskan Crustal Transect in the northern Chugach Mountains and southern Copper River basin, Alaska. Journal of Geophysical Research, 94, 4255\u00E2\u0080\u00934295. Plafker, G., Moore, J.C., and Winkler, G.R., 1994. Geology of the southern Alaska margin. In : Plafker, G. & H.G. Berg (Eds.), The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder, 389\u00E2\u0080\u0093 450. Quenstedt, F.A., 1845\u00E2\u0080\u00931849. Petrefactenkunde Deutschlands, I: Cephalopoden. T\u00C3\u00BCbingen, 581p., 36pl.; (1845), 1\u00E2\u0080\u0093104, 1\u00E2\u0080\u00936pl. Quenstedt, F.A., 1856\u00E2\u0080\u00931858. Der Jura. Verlag der H. Laupp\u00E2\u0080\u0099shen Buchhandlung, T\u00C3\u00BCbingen, 842p., 100pl.; (1856), 1\u00E2\u0080\u0093576,1\u00E2\u0080\u009372pl. Quenstedt, F.A., 1882\u00E2\u0080\u00931885. Die Ammoniten des schw\u00C3\u00A4bischen Jura, Vol. 1, Der schwarze Jura (Lias). Schweizerbart, Stuttgart, 440p., 54pl.; (1885), 241\u00E2\u0080\u0093440, 31\u00E2\u0080\u009354pl. Rakus, M. and Guex, J., 2002. Les ammonites du jurassiques inf\u00C3\u00A9rieur et moyen de la dorsale tunisienne. M\u00C3\u00A9moires de G\u00C3\u00A9ologie (Lausanne) No. 39, 217p. Raup, D.M., and Sepkoski, J.J., 1982. Mass Extinctions in the Marine Fossil Record. Science, 215, 1501\u00E2\u0080\u00931503. Raup, D.M., 1994. The role of extinction in evolution. Proceedings of the National Academy of Science, 91, 6758\u00E2\u0080\u00936763. Renz, C., 1913. Neuere Fortschritte in der Geologie und Pal\u00C3\u00A4ontologie Griechenlands. Zeitschrift der Deutschen Geologischen Gesellschaft (Abhandlungen und Monatsberichte), 64, 530\u00E2\u0080\u0093630. Renz, C., and Renz, O., 1947. Einige seltene Ammoniten aus dem griechischen Mesozoikum. Ecologae geologicae Helvetiae, 39, 169\u00E2\u0080\u0093176. Repin, Y.S., 1970. K voprosny o zonalinom raschlenenii toarskikh otlozhenii Severo-vostoka SSSR. Kolyma, Magadan, 5, 41\u00E2\u0080\u009344. Ridgway, K.D., Trop, J.M., Nokleberg, W.J., Davidson, C.M., and Eastham, K.D., 2002. Mesozoic and Cenozoic tectonics of the eastern and central Alaska 236 Range: Progressive basin development and deformation within a suture zone. Geological Society of America Bulletin, 114, 1480\u00E2\u0080\u00931504. Ridgway, K.D., Trop, J.M., Glen, J.M.G., and O\u00E2\u0080\u0099Neill, J.M., (Eds), 2007. Tectonic growth of a collisional continental margin: crustal evolution of southern Alaska. Geological Society of America Special Paper 431, 658p. Rioux, M., Hacker, B., Mattinson, J., Kelemen, P.B., Blusztajn, J., and Gehrels, G.E., 2007. Magmatic development of an intra-oceanic arc: High-precision U\u00E2\u0080\u0093Pb zircon and whole-rock isotopic analyses from the accreted Talkeetna arc, south- central Alaska. Geological Society of America Bulletin, 119, 1168\u00E2\u0080\u00931184. Robinson, R.S., Kienast, M., Albuquerque, A.L., Altabet, M., et al., 2012. A review of nitrogen isotopic alteration in marine sediments. Paleoceanography, 27, PA4203, doi: 10.1029/2012PA002321. Sabatino, N., Neri, R., Bellanca, A., Jenkyns, H.C., Baudin, F., Parisi, G. and Masetti, D., 2009. Carbon-isotope records of the Early Jurassic (Toarcian) oceanic anoxic event from the Valdorbia (Umbria\u00E2\u0080\u0093Marche Apennines) and Monte Mangart (Julian Alps) sections: palaeoceanographic and stratigraphic implications. Sedimentology, 56, 1307\u00E2\u0080\u00931328. doi:10.1111/j.1365- 3091.2008.01035.x. Sabatino, N., Neri, R., Bellanca, A., Jenkyns, H.C., Masetti, D., and Scopelliti, G., 2011. Petrography and high-resolution geochemical records of Lower Jurassic manganese-rich deposits from Monte Mangart, Julian Alps. Palaeogeography, Palaeoclimatology, Palaeoecology, 299, 97\u00E2\u0080\u0093109. S\u00C3\u00A6len, G., Doyle, P., and Talbot, M.R., 1996. Stable isotope analyses of belemnite rostra from the Whitby Mudstone Fm., England: Surface water conditions during deposition of a marine black shale. Palaios, 11, 97\u00E2\u0080\u0093117, doi:10.2307/3515065. Sandy, M.R., and Blodgett, R.B., 2000. Early Jurassic spiriferid brachiopods from Alaska and their paleogeographic significance. Geobios, 33, 319\u00E2\u0080\u0093328. Schlatter, R.V., 1980. Biostratigraphie und Ammonitenfauna des Unter- Pliensbachium im Typusgebiet (Pliensbach, Holzmaden und N\u00C3\u00BCrtingen; W\u00C3\u00BCrttemberg, SW.-Deutschland). Stuttgarter Beitr\u00C3\u00A4ge zur Naturkunde, Serie B (Geologie und Pal\u00C3\u00A4ontologie), 65, 251p. Schlegelmilch, R.S., 1976. Die Ammoniten des Suddeutschen Lias. Gustav Fischer Verlag, Stuttgart, 212p. 237 Schouten, S., Van Kaam-Peters, H.M.E., Rijpstra, W.I.C., Schoell, M., Sinninghe Damste, J.S., 2000. Effects of an oceanic event on the stable carbon isotopic composition of Early Toarcian carbon. American Journal of Science, 300, 1\u00E2\u0080\u009322. Schr\u00C3\u00B6der, J., 1927. Die Ammoniten der jurassischen Fleckenmergel in den Bayrischen Alpen. Palaeontographica, Stuttgart, 69, 110p. Seguenza, A., 1885. I minerali della provincia di Messina. I, Rocche Messinesi, 245\u00E2\u0080\u0093254. Sepkoski, J.J., Jr., 1975. Stratigraphic biases in the analysis of taxonomic survivorship. Paleobiology, 1, 343\u00E2\u0080\u0093355. Sepkoski, J.J., Jr., 1982. Mass extinction in the Phanerozoic oceans: a review. In: Silver, L.T., and Schulz, P.H., (Eds.), Geological implications of large asteroids and comets on the Earth. Geological Society of America Special Paper 190, 283\u00E2\u0080\u0093290. Sepkoski, J.J., Jr., 1986. Phanerozoic overview of mass extinctions. In: Raup, D.M., and Jablonski, D., (Eds.), Patterns and processes in the history of life. Springer-Verlag, Berlin. Sepkoski, J.J., Jr., 1992. A compendium of fossil marine families. Milwaukee Public Museum Contributions to Biology and Geology, 83, 156p. Sepkoski, J.J., Jr., 1993. Ten years in the library: new data confirm paleontological patterns. Paleobiology, 19, 43\u00E2\u0080\u009351. Sepkoski, J.J., Jr., 1996. Patterns of Phanerozoic extinction: a perspective from global data bases. In: Walliser, O.H., (Ed.), Global events and event stratigraphy. Springer-Verlag, Berlin, 35\u00E2\u0080\u009352. Sherwood, K.W., Johnson, P.P., Craig, J.D., Zerwick, S.A., Lothamer, R.T., Thurston, D.K., and Hurlbert, S.B., 2002. Structure and stratigraphy of the Hanna Trough, U.S., Chukchi Shelf, Alaska. In: Miller, E.L., Grantz, A., Klemperer, S.L., (Eds.), 2002. Tectonic Evolution of the Bering Shelf\u00E2\u0080\u0093Chukchi Sea\u00E2\u0080\u0093Arctic Margin and Adjacent Landmasses. Geological Society of America Special Paper 360, 39\u00E2\u0080\u009366. Shurygin, B.N., Nikitenko, B.L., Devyatov, V.P., Il\u00E2\u0080\u0099ina, V.I., Meledina, S.V., Gaideburova, E.A., Dzyuba, O.S., Kazakov, A.M., Mogucheva, N.K., 2000. Stratigraphy of Oil and Gas Basins of Siberia: The Jurassic System. Izd. SO RAN, Filial Geo, Novosibirsk [in Russian]. Shurygin, B.N., Nikitenko, B.L., Meledina, S.V., Dzyuba, O.S., Knyazev, V.G., 2011. Comprehensive zonal subdivisions of Siberian Jurassic and their 238 significance for Circum-Arctic correlations. Russian Geology and Geophysics, 52, 825\u00E2\u0080\u0093844. Silberling, N.J., Grant-Mackie, J.A., and Nichols, K.M., 1997. The Late Triassic bivalve Monotis in accreted terranes of Alaska. United States Geological Survey Bulletin 2151, 21p. Silva, R.L., Duarte, L.V., Comas-Rengifo, M.J., Mendon\u00C3\u00A7a Filho, J.G., and Azer\u00C3\u00AAdo, A.C., 2011. Update of the carbon and oxygen isotopic records of the Early\u00E2\u0080\u0093Late Pliensbachian (Early Jurassic, ~187 Ma): Insights from the organic- rich hemipelagic series of the Lusitanian Basin (Portugal). Chemical Geology, 283, 177\u00E2\u0080\u0093184. Simpson, M., 1843. A monograph of the ammonites of the Yorkshire Lias. London, 60p. Smith, P.L., 1983. The Pliensbachian ammonite Dayiceras dayiceroides and Early Jurassic paleogeography. Canadian Journal of Earth Sciences, 20, 86\u00E2\u0080\u009391. Smith, P.L., 2006, Paleobiogeography and Early Jurassic molluscs in the context of terrane displacement in western Canada. In: Haggart, J.W., Enkin, R.J., and Monger, J.W.H., (Eds.), Geological Association of Canada Special Paper 46, p. 81\u00E2\u0080\u009394. Smith, P.L., and Tipper, H.W., 1986, Plate tectonics and paleobiogeography: Early Jurassic (Pliensbachian) endemism and diversity. Palaios, 1, 399\u00E2\u0080\u0093412. Smith, P.L., and Tipper, H.W., 1996, Pliensbachian (Lower Jurassic) Ammonites of the Queen Charlotte Islands, British Columbia: Bulletins of American Paleontology, v. 108, p. 1\u00E2\u0080\u0093122. Smith, P.L., Tipper, H.W., Taylor, D.G., and Guex, J., 1988. An ammonite zonation for the Lower Jurassic of Canada and the United States: the Pliensbachian. Canadian Journal of Earth Sciences, 25, 1503\u00E2\u0080\u00931523. Smith, P.L., Tipper, H.W., and Ham, D.M., 2001. Lower Jurassic Amaltheidae (Ammonitina) in North America: paleobiogeography and tectonic implications. Canadian Journal of Earth Sciences, 38, 1439\u00E2\u0080\u00931449. Smith, W.R., 1926. Geology and oil development of the Cold Bay district. In: Mineral Resources of Alaska; Report on progress of investigations in 1924. United States Geological Survey Bulletin 783, 63\u00E2\u0080\u009388. Sowerby, J., 1812\u00E2\u0080\u00931822. The Mineral Conchology of Great Britain, London. (1812-1815) v. 1, pl. 1\u00E2\u0080\u0093102; (1815-1818) v. 2, pl. 103\u00E2\u0080\u0093203; (1818-1821) v. 3, pl. 204\u00E2\u0080\u0093306; (1821, 1822) v. 4, pl. 307\u00E2\u0080\u0093383. 239 Sowerby, J. and Sowerby, J. de C., 1817. The mineral conchology of Great Britain; or coloured figures and descriptions of those remains of testaceous animals or shells, which have been preserved at various times and depths in the earth. London, 7 volumes, 1\u00E2\u0080\u0093186. Spath, L.F., 1913. On Jurassic ammonites from Jebel Zaghuan (Tunisia). Quarterly Journal of the Geological Society of London, 69, 540\u00E2\u0080\u0093580. Spath, L.F., 1919. Notes on Ammonites. Geological Magazine, 6, 170\u00E2\u0080\u0093177. Suan, G., Mattioli, E., Pittet, B., Mailliot, S., and L\u00C3\u00A9cuyer, C., 2008, Evidence for major environmental perturbation prior to and during the Toarcian (Early Jurassic) oceanic anoxic event from the Lusitanian Basin. Paleoceanography, v. 23, p. 1\u00E2\u0080\u009314. Suan, G. Mattioli, E., Pittet, B., L\u00C3\u00A9cuyer, C., Such\u00C3\u00A9ras-Marx, B., Duarte, L.V., Philippe, M., Reggiani, L., and Martineau, F., 2010. Secular environmental precursors to Early Toarcian (Jurassic) extreme climate changes. Earth and Planetary Science Letters, 290, 448\u00E2\u0080\u0093458, doi:10.1016/j.epsl.2009.12.047. Suan, G., Nikitenko, B.L., Robov, M.A., Baudin, F., Spangenberg, J.E., Knyazev, V.G., Glinskikh, L.A., Goryacheva, A.A., Adatte, T., Riding, J.B., F\u00C3\u00B6llmi, K.B., Pittet, B., Mattioli, E., and L\u00C3\u00A9cuyer, C., 2011. Polar record of Early Jurassic massive carbon injection. Earth and Planetary Science Letters, 312, 102\u00E2\u0080\u0093113. Suess, E., 1865. \u00C3\u009Cber Ammoniten. Sitzungsberichte der Mathematische Naturwissenschaftlichen Classe der Kaiserlichen Akademie der Wissenschaftn, Vienna, Austria, 52, 71\u00E2\u0080\u009389. Sun, Y., Lai, X., Wignall, P.B., Widdowson, M., Ali, J.R., Jiang, H., Wang, W., Yan, C., Bond, D.P.G., and V\u00C3\u00A9drine, S., 2010. Dating the onset and nature of the Middle Permian Emeishan large igneous province eruptions in SW China using conodont biostratigraphy and its bearing on mantle plume uplift models. Lithos, 119, 20\u00E2\u0080\u009333. Tappan, H., 1955. Jurassic Foraminifera, pt. 2, of Foraminifera from the Arctic slope of Alaska. U.S. Geological Survey Professional Paper 236\u00E2\u0080\u0093B, 21\u00E2\u0080\u009390. Tesdal, J.E., Galbraith, E.D., and Kienast, M., 2013. Nitrogen isotopes in bulk marine sediment: linking seafloor observations with subseafloor records. Biogeosciences, 10, 101\u00E2\u0080\u0093118. Thomas, C.D., Cameron, A., Green, R.E., Bakkenes, M., Beaumont, L.J., Collingham, Y.C., Erasmus, B.F.N., De Siqueira, M.F., Grainger, A., Hannah, L., Hughes, L., Huntley, B., A.S., V.J., G.F., M., Miles, L., Ortega-Huerta, M.A., 240 Peterson, A.T., Phillips, O.L., and Williams, S.E., 2004a, Extinction risk from climate change. Nature, 427, 145\u00E2\u0080\u0093148. Thomas, J.A., Telfer, M.G., Roy, D.B., Preston, C.D., Greenwood, J.J.D., Asher, J., Fox, R., Clarke, R.T., and Lawton, J.H., 2004b. Comparative losses of British butterflies, birds, and plants and the global extinction crisis. Science, 303, 1879\u00E2\u0080\u0093 1881. Thompson, R.C., and Smith, P.L., 1992. Pliensbachian (Lower Jurassic) biostratigraphy and ammonite fauna of the Spatsizi area, north-central British Columbia. Geological Survey of Canada Bulletin 437, 87p. Thompson, R.I., Haggart, J.W., and Lewis, P.D., 1991. Late Triassic through early Tertiary evolution of the Queen Charlotte Basin, British Columbia, with a perspective on hydrocarbon potential. In: Woodsworth, G.J., (Ed.), Evolution and Hydrocarbon Potential of the Queen Charlotte Basin, British Columbia. Geological Survey of Canada, Paper 90\u00E2\u0080\u009310, 3\u00E2\u0080\u009329. Till, A.B., and Dumoulin, J.A., 1994. Geology of Seward Peninsula and Saint Lawrence Island. In : Plafker, G. and Berg, H.G., (Eds.), The geology of Alaska. Geological Society of America, the Geology of North America G-1, Boulder, 141\u00E2\u0080\u0093 152. Tipper, H.W., Smith, P.L., Cameron, B.E.B., Carter, E.S., Jakobs, G.K., and Johns, M.J., 1991. Biostratigraphy of the Lower Jurassic formations of the Queen Charlotte Islands, British Columbia. In: Woodsworth, G.J., (Ed.), Evolution and Hydrocarbon Potential of the Queen Charlotte Basin, British Columbia. Geological Survey of Canada, Paper 90\u00E2\u0080\u009310, 203\u00E2\u0080\u0093235. Tremolada, F., Van de Schootbrugge, B., and Erba, E., 2005. Early Jurassic schizosphaerellid crisis in Cantabria, Spain: Implications for calcification rates and phytoplankton evolution across the Toarcian oceanic anoxic event. Paleoceanography, 20, PA2011, doi:10.1029/2004PA001120. Tribovillard, N., Riboulleau, A., Lyons, T., and Baudin, F., 2004a, Enhanced trapping of molybdenum by sufurized organic matter of marine origin as recorded by various Mesozoic formations. Chemical Geology, 213, 385\u00E2\u0080\u0093401. doi: 10.1016/j.chemgeo.2004.08.011. Tribovillard, N., Trentesaux, A., Ramdani, A., Baudin, F., and Riboulleau, A., 2004b. Contr\u00C3\u00B4les de l\u00E2\u0080\u0099accumulation de mati\u00C3\u00A8re organique dans la Kimmeridge Clay Formation (Jurassique sup\u00C3\u00A9rieur, Yorkshire, G.B.) et son \u00C3\u00A9quivalent lat\u00C3\u00A9ral du Boulonnais: L\u00E2\u0080\u0099apport des \u00C3\u00A9l\u00C3\u00A9ments traces m\u00C3\u00A9talliques. Bulletin of the Geological Society of France, 175, 491\u00E2\u0080\u0093506. 241 Tribovillard, N., Algeo, T.J., Lyons, T., and Riboulleau, A., 2006. Trace metals as paleoredox and paleoproductivity proxies: An update. Chemical Geology, 232, 12\u00E2\u0080\u009332. doi: 10.1016/j.chemgeo.2006.02.012. Tribovillard, N., Lyons, T.W., Riboulleau, A., and Bout-Roumazeilles, V., 2008. A possible capture of molybdenum during early diagenesis of dysoxic sediments. Bulletin of the Geological Society of France, 179, 3\u00E2\u0080\u009312. doi: 10.2113/gssgfbull.179.1.3. Trop, J.M., and Ridgway, K.D., 2007. Mesozoic and Cenozoic tectonic growth of southern Alaska: A sedimentary basin perspective. In: Ridgway, K.D., Trop, J.M., Glen, J.M.G., and O\u00E2\u0080\u0099Neill, J.M., (Eds), 2007. Tectonic growth of a collisional continental margin: crustal evolution of southern Alaska. Geological Society of America Special Paper 431, 658p. Trop, J.M., Ridgway, K.D., Manuszak, J.D., and Layer, P.W., 2002. Sedimentary basin development on the allochthonous Wrangellia composite terrane, Mesozoic Wrangell Mountains basin, Alaska: a long-term record of terrane migration and arc construction. Geological Society of America Bulletin, 114, 693\u00E2\u0080\u0093717. Trop, J.M., Szuch, D.A., Rioux, M., and Blodgett, R.B., 2005. Sedimentology and provenance of the Upper Jurassic Naknek Formation, Talkeetna Mountains, Alaska: Bearings on the accretionary tectonic history of the Wrangellia composite terrane. Geological Society of America Bulletin, 117, 570\u00E2\u0080\u0093588. Tyson, R.V., and Pearson, T.H., Eds., 1991. Modern and ancient continental shelf anoxia: an overview. London, the Geological Society Special Publication, 1\u00E2\u0080\u009324. Umhoefer, P.J., and Blakey, R.C., 2006. Moderate (1600 km) northward translation of Baja British Columbia from southern California: An attempt at reconciliation of paleomagnetism and geology. In: Haggart, J.W., Enkin, R.J., and Monger, J.W.H., (Eds.), 2006. Paleogeography of the North American Cordillera; Evidence for and against Large-Scale Displacements. Geological Association of Canada Special Paper 46, 307\u00E2\u0080\u0093329. van Breugel, Y., Bass, M., Schouten, S., Mattioli, E., and Sinninghe Damst\u00C3\u00A9, J.S., 2006. Isorenieratane record in black shales from the Paris Basin, France: Constraints on recycling of respired CO2 as a mechanism for negative carbon isotope shifts during the Toarcian oceanic anoxic event. Paleoceanography, 21, PA4220, 8p. doi: 10.1029/2006PA001305. van de Schootbrugge, B., Bailey, T.R., Rosenthal, Y., Katz, M.E., Wright, J.D., Miller, K.G., Feist-Burkhardt, S., Falowski, P., 2005. Early Jurassic climate change and the radiation of organic-walled phytoplankton in the Tethys Ocean. Paleobiology, 31, 73\u00E2\u0080\u009397. 242 van der Heyden, P., 1992. A Middle Jurassic to early Tertiary Andean-Sierran arc model for the Coast Belt of British Columbia. Tectonics, 11, 82\u00E2\u0080\u009397. Venturi, F., 1975. Rarenodia nuovo genere di ammoniti (sottofam. Hammatoceratinae Buckman, 1887) del Toarciano inferiore \u00E2\u0080\u009CRosso Ammonitico\u00E2\u0080\u009D umbro-marchigiano; Bollettino della Societ\u00C3\u00A0 Paleontologica Italiana, 14, 11\u00E2\u0080\u009319. Vorlicek, T.P., and Helz, G.R., 2002. Catalysis by mineral surfaces: Implications for Mo geochemistry in anoxic environments. Geochimica et Cosmochimica Acta, 66, 3679\u00E2\u0080\u00933692. doi: 10.1016/S0016-7037(01)00837-7. V\u00C3\u00B6r\u00C3\u00B6s, A., 2002. Victims of the Early Toarcian anoxic event: the radiation and extinction of Jurassic Koninckinidae (Brachiopoda). Lethaia, 35, 345\u00E2\u0080\u0093357. Waagen, W., 1869. Die Formenreihe des Ammonites subradiatus \u00E2\u0080\u0093 Versuch einer Pal\u00C3\u00A4ontologischen Monographie. Geognostisch-Pal\u00C3\u00A4ontologische Beitr\u00C3\u00A4ge, 2, 179\u00E2\u0080\u0093256. Wang, J., Newton, C.R., and Dunne, L., 1988. Late Triassic transition from biogenic to arc sedimentation on the Peninsular terrane: Puale Bay, Alaska Peninsula. Geological Society of America Bulletin, 100, 1466\u00E2\u0080\u00931478. Ward, P.D., Haggart, J.W., Carter, E.S., Wilbur, D., Tipper, H.W., and Evans, T., 2001. Sudden productivity collapse associated with the Triassic\u00E2\u0080\u0093Jurassic boundary mass extinction. Science, 292, 1148\u00E2\u0080\u00931151. Wendt, J., 1966. Revision der Ammoniten Gattung Leukadiella Renz aus dem mediterranen Oberlias. Neues Jahrbuch f\u00C3\u00BCr Geologie und Pal\u00C3\u00A4ontologie, Abhandlungen, 125, 136\u00E2\u0080\u0093154. Whiteaves, J.F., 1884. On the fossils of the coal-bearing deposits of the Queen Charlotte Islands collected by Dr. G.M. Dawson in 1878. Geological and Natural History Survey of Canada, Mesozoic Fossils, 1, 262p. Whiteside, J.H., Olsen, P.E., Kent, D.V., Fowell, S.J., and Et-Touhami, M., 2007. Synchrony between the Central Atlantic magmatic province and the Triassic\u00E2\u0080\u0093 Jurassic mass-extinction event? Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 345\u00E2\u0080\u0093367. Whiteside, J.H., Olsen, P.E., Eglinton, T., Brookfield, M.E., and Sambrotto, R.N., 2010. Compound-specific carbon isotopes from Earth\u00E2\u0080\u0099s largest flood basalt eruptions directly linked to the end-Triassic mass extinction. Proceedings of the National Academy of Sciences, 107, 6721\u00E2\u0080\u00936725. 243 Wiedenmayer, F., 1977. Die Ammoniten des Besazio-Kalks (Pliensbachian, S\u00C3\u00BCdtessin). Schweizerische Pal\u00C3\u00A4ontologische Abhandlungen, 98, 169p. Wignall, P.B., 2001. Large igneous provinces and mass extinctions. Earth- Science Reviews 53, 1\u00E2\u0080\u009333. Wignall, P.B., 2005. The Link between Large Igneous Province Eruptions and Mass Extinctions. Elements, 1, 293\u00E2\u0080\u0093297. Wignall, P.B., and Bond, D.P.G., 2009. The end-Triassic and Early Jurassic mass extinction records in the British Isles. Proceedings of the Geologists\u00E2\u0080\u0099 Association, 119, 73\u00E2\u0080\u009384. Wignall, P.B., Newton, R.J., and Little, C.T.S., 2005. The timing of paleoenvironmental change and cause-and-effect relationships during the Early Jurassic mass extinction Europe. American Journal of Science, 305, 1014\u00E2\u0080\u00931032. Wignall, P.B., Sun, Y., Bond, D.P.G., Izon, G., Newton, R.J., V\u00C3\u00A9drine, S., Widdowson, M., Ali, J.R., Lai, X., Jiang, H., Cope, H., and Bottrell, S.H., 2009. Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China. Science, 324, 1179\u00E2\u0080\u00931182. Wilkin, R.T., Barnes, H.L., and Brantley, S.L., 1996. The size distribution of framboidal pyrite in modern sediments: An indicator of redox conditions. Geochimica et Cosmochimica Acta, 60, 3897\u00E2\u0080\u00933912. Williford, K.H., Ward, P.D., Garrison, G.H., and Buick, R., 2007. An extended organic carbon-isotope record across the Triassic\u00E2\u0080\u0093Jurassic boundary in the Queen Charlotte Islands, British Columbia, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 290\u00E2\u0080\u0093296. Wilson, E.O., 1993. The Diversity of Life. Harvard University Press, Cambridge. Wilson, F.H., Detterman, R.L., and Case, J.E.,1985. The Alaska Peninsula terrane: a definition. United States Geological Survey Open-File Report 85-450, 17p. Wilson, J.T., 1963. A possible origin of the Hawaiian Islands. Canadian Journal of Physics, 41, 836\u00E2\u0080\u0093870. Winkler, G.R., 1992. Geologic map and summary geochronology of the Anchorage 1\u00C2\u00BA x 3\u00C2\u00BA quadrangle, southern Alaska. United States Geological Survey Miscellaneous Investigations Series Map I-2283, 1:250,000 scale. 244 Wright, T., 1878\u00E2\u0080\u00931886. A monograph on the Lias ammonites of the British Islands. Monographs of the Palaeontographical Society, part 3 (1880), 165\u00E2\u0080\u0093264; part 5 (1882), 329\u00E2\u0080\u0093400; part 7 (1884), 441\u00E2\u0080\u0093480. Young, G., and Bird, J., 1828. A geological survey of the Yorkshire Coast: describing the strata and fossils occurring between the Humber and the tees, from the German Ocean to the Plain of York. 2nd ed., Whitby, London, 368p. Zakharov, V.A., Bogomolov, Y.I., Il\u00E2\u0080\u0099ina, V.I., Konstantinov, A.G., Kurushin, N.I., Lebedeva, N.K., Meledina, S.V., Nikitenko, B.L., Sobolev, E.S., and Shurygin, B.N., 1997. Boreal Zonal Standard and Biostratigraphy of the Siberian Mesozoic. Russian Geology and Geophysics, 38, 965\u00E2\u0080\u0093993. Zakharov, V.A., Shurygin, B.N., Il\u00E2\u0080\u0099ina, V.I., and Nikitenko, B.L., 2006. Pliensbachian\u00E2\u0080\u0093Toarcian Biotic Turnover in North Siberia and the Arctic Region. Stratigraphy and Geological Correlation, 14, 399\u00E2\u0080\u0093417. Zittel, K.A., 1884. Handbuch der Pal\u00C3\u00A4ontologie. Cephalopoda, Munich, Germany, 1, 329\u00E2\u0080\u0093522. 245 A ppendices A ppendix A : D iversity and R ate M etrics 246 247 248 249 Appendix B: Geochemical Data Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 01 4.36 0.79 -24.02 0.05 -0.74 0.00 YR 02 4.41 0.84 -24.05 0.40 YR 03 4.68 0.88 -24.10 0.05 -0.17 0.80 YR 04 4.25 1.00 -24.34 1.20 YR 05 4.35 0.82 -23.42 0.06 -0.85 1.60 YR 06 4.76 0.81 -24.71 2.00 YR07 4.45 0.45 -24.30 0.04 -0.14 2.40 YR 08 4.07 0.81 -24.88 2.80 YR 09 4.49 0.80 -24.64 0.05 -0.06 3.20 YR 10 4.12 0.85 -24.44 3.60 YR 11 4.26 0.79 -24.27 0.06 -0.08 4.00 YR 12 4.74 0.84 -24.49 4.40 YR 13 4.26 0.90 -24.27 0.06 0.06 4.80 YR 14 4.87 0.74 -24.54 5.20 YR 15 4.75 0.68 -24.87 0.05 -0.36 5.60 YR 16 4.18 0.78 -24.57 6.00 YR 17 4.38 0.96 -24.59 0.05 -0.15 6.40 YR 18 4.50 2.09 -24.92 6.80 YR19 4.46 0.70 -25.21 0.04 -0.59 7.60 YR 20 4.70 1.06 -24.09 8.00 YR 21 4.32 0.62 -24.44 0.05 -0.64 8.40 YR 22 4.68 0.73 -24.89 8.80 YR 23 4.28 0.86 -24.66 0.05 -0.39 9.20 YR 24 4.75 1.16 -24.16 9.60 YR 25 4.43 0.90 -24.49 0.05 -0.67 10.00 YR 26 4.32 0.90 -24.56 10.40 YR 27 4.70 2.31 -24.76 0.04 -0.08 10.80 YR 28 4.89 0.82 -24.76 11.20 YR 29 4.27 0.82 -24.93 0.05 0.15 11.60 YR 30 4.84 0.73 -24.95 12.00 YR 31 4.35 0.36 -24.66 0.05 0.15 12.40 YR 32 4.62 0.40 -24.29 12.80 YR 33 4.38 0.67 -24.61 0.06 -0.04 13.20 YR 34 4.55 0.96 -24.23 13.60 YR 35 4.29 0.90 -23.89 0.06 -0.15 14.00 YR 36 4.33 0.76 -24.19 14.40 250 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 37 4.62 0.65 -24.14 0.06 0.18 14.80 YR 38 4.32 0.81 -24.87 15.20 YR 39 4.71 0.50 -25.26 0.04 -0.44 15.60 YR 40 4.63 0.92 -25.39 16.00 YR 41 0.05 -0.02 16.40 YR42 4.46 1.11 -26.16 16.80 YR43 4.49 0.79 -26.71 0.03 -0.39 17.20 YR 44 4.86 0.92 -24.60 17.60 YR 45 4.84 0.94 -24.23 0.06 -0.26 18.00 YR 46 4.65 0.82 -23.63 18.40 YR 47 4.27 0.91 -24.06 0.06 -0.37 18.80 YR 48 4.65 0.84 -24.37 19.20 YR 49 4.77 0.86 -25.09 0.07 -0.31 19.60 YR 50 4.28 0.80 -24.00 0.05 -0.35 21.60 YR 51 4.29 0.74 -24.55 0.05 -0.92 22.00 YR 52 4.22 0.94 -23.99 22.40 YR 53 4.72 0.31 -24.68 0.08 -0.38 22.80 YR 54 4.46 0.26 -24.57 23.20 YR 055 4.36 0.37 -22.07 0.08 -0.02 23.60 YR 56 4.29 0.37 -25.10 24.00 YR 57 4.47 0.39 -23.95 0.05 -0.48 26.00 YR 58 4.39 0.46 -24.44 26.40 YR 59 4.33 0.44 -24.63 0.05 -0.26 26.80 27.00 YR 61 4.55 0.57 -24.44 0.06 0.54 27.20 YR 62 4.67 0.36 -24.73 27.60 YR 63 4.71 0.34 -24.73 0.05 -0.04 28.00 YR 64 4.35 0.32 -25.04 28.40 YR 65 4.56 0.31 -25.24 0.04 -0.29 28.80 YR 66 4.44 0.44 -25.36 0.05 -0.31 29.00 YR 67 4.54 0.47 -25.86 0.06 -0.14 29.20 YR 68 4.64 0.34 -25.52 0.05 -0.69 29.60 YR 69 4.91 0.30 -25.15 0.05 -0.32 30.00 YR 70 4.52 0.39 -25.23 30.40 YR 71 4.52 0.30 -24.81 0.05 -0.38 30.80 YR 72 4.48 0.49 -24.90 0.04 -0.46 31.20 YR 73 4.81 0.62 -24.94 0.06 -0.28 31.60 YR 74 4.83 0.71 -24.77 32.00 251 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 75 4.66 0.69 -24.59 0.06 -0.49 32.40 YR 76 4.52 0.38 -31.39 0.04 -1.50 32.80 YR 77 4.43 0.99 -31.24 0.05 -0.28 33.20 YR 78 4.40 0.83 -31.18 0.05 -1.08 33.60 YR 79 4.49 0.44 -30.77 34.00 YR 80 4.69 0.65 -30.64 0.04 -0.88 34.40 YR 81 4.68 1.04 -31.08 0.06 -0.95 34.80 YR 82 4.84 0.69 -30.27 0.04 -0.67 35.20 YR 83 4.27 0.64 -31.56 0.05 -0.05 35.60 YR 84 4.57 0.98 -31.67 0.06 -0.81 36.00 YR 85 4.39 0.62 -31.67 0.06 0.08 36.40 YR 86 4.34 0.97 -31.77 36.80 YR 87 4.30 0.61 -31.50 0.06 -0.05 37.20 YR 88 4.44 0.66 -31.58 0.06 -0.62 37.60 YR 89 4.57 0.47 -31.37 0.05 0.00 38.00 YR 90 4.58 0.53 -31.80 0.05 -0.72 38.40 YR 91 4.48 0.61 -31.61 0.05 -0.24 38.80 YR 92 4.23 1.26 -30.31 0.05 -1.08 39.20 YR 93 4.81 1.33 -29.51 0.04 -0.85 39.60 YR 94 4.90 1.96 -29.64 0.05 -1.27 40.00 YR 95 4.39 0.91 -29.89 0.05 -0.71 40.40 YR 96 4.64 0.85 -30.24 0.05 -1.10 40.80 YR 97 4.25 0.95 -30.39 0.04 -0.64 41.20 YR 98 4.54 1.28 -29.42 0.04 -0.76 41.40 YR 99 4.55 0.69 -31.39 0.05 0.06 41.60 YR 100 4.65 0.51 -31.03 0.06 -0.51 41.80 YR 101 4.80 0.60 -31.22 42.00 YR 102 4.65 0.50 -31.20 0.05 -0.76 42.20 YR 103 4.45 0.45 -31.27 0.05 -0.46 42.40 YR 104 4.71 0.41 -31.22 0.05 -0.55 42.60 YR 105 4.29 0.56 -31.11 0.05 -0.26 42.80 YR 106 4.37 0.61 -31.29 0.04 -0.87 43.00 YR 107 4.73 0.49 -31.14 0.05 -0.09 43.20 YR 108 4.60 0.44 -29.74 0.07 -0.52 43.40 YR 109 4.31 0.57 -25.71 0.06 -0.07 43.60 YR 110 4.28 0.45 -25.57 0.04 -0.35 43.80 YR 111 4.43 0.48 -24.63 0.06 -0.06 44.00 YR 112 4.72 0.46 -24.84 44.20 252 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 113 4.48 0.51 -24.67 0.06 -0.01 44.40 YR 114 4.83 1.13 -24.75 0.05 -0.11 44.60 YR 115 4.43 0.73 -24.92 0.05 -0.01 44.80 YR 116 4.67 0.94 -24.43 0.05 -0.57 45.00 YR 117 4.36 0.84 -25.11 0.04 0.15 45.20 YR 118 4.47 0.59 -25.67 0.04 -0.66 45.40 YR 119 4.27 0.59 -26.15 0.04 -0.52 45.60 YR 120 4.65 0.79 -24.96 0.05 -0.50 45.80 YR 121 4.36 0.84 -25.11 0.05 -0.62 46.00 YR 122 4.51 0.88 -25.23 46.20 YR 123 4.36 0.94 -24.93 0.05 0.24 46.40 YR 124 4.32 0.79 -25.00 46.60 YR 125 4.33 0.81 -25.04 0.05 -0.56 46.80 YR 126 4.28 0.78 -25.00 47.00 YR 127 4.62 0.38 -26.71 0.03 -0.37 47.20 YR 128 4.75 0.52 -26.58 47.40 YR 129 4.55 0.69 -25.70 0.05 -0.06 47.60 YR 130 4.45 0.77 -26.44 47.80 YR 131 4.62 0.65 -26.25 0.05 0.02 48.00 YR 132 4.29 0.73 -25.75 48.20 YR 133 4.30 0.65 -25.59 0.05 0.01 48.40 YR 134 4.95 0.76 -25.85 48.60 YR 135 4.65 0.83 -25.65 0.05 0.01 48.80 YR 136 4.50 0.80 -25.51 49.00 YR 137 4.50 0.75 -26.15 0.05 0.04 49.20 YR 138 4.38 0.86 -26.74 49.40 YR 139 4.43 0.59 -27.04 0.03 -0.07 49.60 YR 140 4.73 0.48 -26.61 49.80 YR 141 4.51 0.74 -25.75 0.05 -0.01 50.00 YR 142 4.64 0.75 -26.03 50.20 YR 143 4.83 0.57 -25.83 0.05 -0.44 50.40 YR 144 4.72 0.65 -26.06 50.60 YR 145 4.84 0.49 -26.16 0.04 -0.79 50.80 YR 146 4.63 0.47 -26.85 51.00 YR 147 4.63 0.57 -26.41 0.04 0.09 51.20 YR 148 4.68 0.65 -24.84 51.60 YR 149 4.79 0.51 -24.98 0.05 -0.39 52.00 YR 150 4.28 0.68 -25.30 52.40 253 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 151 4.31 0.73 -25.14 0.04 -0.61 52.80 YR 152 4.84 0.61 -25.40 53.20 YR 153 4.87 0.51 -24.88 0.05 -0.20 53.60 YR 154 4.75 0.80 -25.28 54.00 YR 155 4.25 0.64 -24.81 0.05 0.20 54.40 YR 156 4.64 0.43 -26.69 54.80 YR 157 4.85 0.64 -25.04 0.05 0.31 55.20 YR158 4.51 0.23 -27.02 55.60 YR 159 4.17 0.48 -26.31 0.05 -0.09 56.00 YR 160 4.46 0.60 -26.16 56.40 YR 161 4.40 0.69 -25.63 0.05 0.14 56.80 YR 162 4.47 0.67 -25.96 57.20 YR 163 4.40 0.56 -25.35 0.05 -0.25 57.60 YR 164 4.54 0.56 -25.10 58.00 YR 165 4.81 0.71 -25.90 0.05 -0.04 58.40 YR 166 4.76 0.51 -25.52 58.80 YR 167 4.77 0.45 -26.31 0.04 -0.50 59.20 YR 168 4.37 0.53 -26.11 59.60 YR 169 4.35 0.65 -26.35 0.05 -0.49 60.00 YR 170 4.38 0.44 -26.61 60.40 YR 171 4.35 0.50 -27.09 0.04 -0.11 60.80 YR 172 4.82 0.32 -27.08 61.20 YR 173 4.67 0.44 -27.46 0.04 -0.77 61.60 YR 174 4.38 0.38 -27.12 62.00 YR 175 4.14 1.46 -25.24 0.04 -0.50 62.40 YR 176 4.95 0.37 -26.48 62.80 YR 177 4.77 0.37 -26.85 0.04 0.12 63.20 YR 178 4.61 0.47 -27.19 63.60 YR 179 4.29 0.42 -26.25 0.04 0.03 64.00 YR 180 4.84 0.41 -26.87 64.40 YR 181 4.29 0.47 -26.80 0.04 0.32 64.80 YR 182 4.64 0.59 -26.77 65.20 YR 183 4.39 0.49 -26.97 0.05 -0.06 65.60 YR 184 4.65 0.35 -26.84 66.00 YR 185 4.83 0.46 -27.66 0.05 0.17 66.40 YR 186 4.21 0.40 -26.33 66.80 YR 187 4.66 0.36 -26.69 0.04 -0.31 67.20 YR 188 4.30 0.52 -27.04 67.60 254 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 189 4.55 0.41 -26.49 0.05 -0.06 68.00 YR 190 4.28 0.56 -26.05 68.40 YR 191 4.72 0.88 -25.44 0.05 -0.62 68.80 YR 192 4.01 0.84 -24.91 69.20 YR 193 4.58 0.51 -26.97 0.04 -0.27 69.60 YR 194 4.80 0.43 -26.66 70.00 YR 195 4.91 0.47 -26.84 0.05 -0.08 70.40 YR 196 4.24 0.44 -28.95 70.80 YR 197 4.34 0.47 -27.12 0.05 -0.26 71.20 YR 198 4.61 0.51 -26.62 71.85 YR 199 4.32 0.72 -26.85 0.05 -0.18 72.25 YR 200 4.44 0.69 -26.93 72.65 YR 201 4.71 0.61 -26.41 0.05 -0.21 73.05 YR 202 4.46 0.57 -26.49 73.45 YR 203 4.32 0.71 -26.38 0.06 0.37 73.85 YR 204 4.83 0.64 -26.32 74.25 YR 205 4.54 0.54 -26.33 0.05 0.30 74.65 YR 206 4.38 0.51 -26.67 75.05 YR 207 4.72 0.59 -26.28 0.06 0.22 75.45 YR 208 4.60 0.55 -26.44 75.85 YR 209 4.59 0.73 -28.11 0.05 0.13 76.25 YR 210 4.81 0.52 -27.31 76.65 YR 211 4.84 0.56 -28.30 0.05 -0.01 77.05 YR 212 4.17 0.51 -27.93 77.45 YR 213 4.90 0.45 -27.65 0.06 -0.15 78.45 YR 214 4.51 0.65 -27.78 78.85 YR 215 4.83 0.56 -27.28 0.06 0.18 79.25 YR 216 4.66 0.52 -27.47 79.65 YR 217 4.66 0.44 -27.15 80.05 YR 218 4.29 0.44 -28.89 80.45 YR 219 4.31 0.38 -29.34 0.05 -0.30 80.85 YR 220 4.48 0.25 -28.97 82.85 YR 221 4.54 0.51 -27.91 0.06 -0.30 82.88 YR 222 4.72 0.54 -27.96 83.28 YR 223 4.54 0.40 -28.13 0.05 -0.21 83.68 YR 224 4.68 0.41 -28.54 84.08 YR 225 4.68 0.49 -27.79 0.05 -0.09 84.48 255 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 226 4.37 0.49 -28.03 84.88 YR 227 4.55 0.54 -27.73 0.06 -0.14 85.28 YR 228 4.51 0.55 -28.05 85.68 YR 229 4.29 0.37 -28.43 0.05 0.24 86.08 YR 230 4.46 0.63 -28.73 86.48 YR 231 4.47 0.38 -28.54 0.06 -0.01 86.88 YR 232 4.34 0.37 -27.62 87.28 YR 233 4.33 0.35 -27.60 0.05 0.05 87.68 YR 234 4.40 0.32 -27.79 88.08 YR 235 4.85 0.38 -27.80 0.05 -0.04 88.48 YR 236 4.49 0.53 -27.55 88.88 YR 237 4.71 0.55 -27.88 0.05 0.31 89.18 89.58 YR 239 4.53 0.30 -28.49 0.05 -0.02 89.98 YR 240 4.53 0.37 -29.54 90.38 YR 241 4.53 0.41 -29.79 0.03 -0.53 90.78 YR 242 4.82 0.39 -29.84 91.18 YR 243 4.68 0.19 -28.90 0.04 -0.54 91.58 YR 244 4.45 0.27 -29.37 91.98 YR 245 0.05 -0.07 92.38 YR 246 4.75 0.52 -29.74 92.78 YR 247 4.62 0.43 -28.98 0.05 -0.03 93.18 YR 248 4.57 0.47 -29.57 93.58 YR 249 4.73 0.57 -28.96 0.05 0.05 93.98 YR 250 4.51 0.52 -29.07 94.38 YR 251 4.46 0.38 -29.02 0.05 0.15 94.78 YR 252 4.69 0.52 -28.97 95.18 YR 253 4.54 0.37 -29.86 0.05 -0.19 95.58 YR 254 4.52 0.46 -29.56 95.98 YR 255 4.25 0.58 -29.27 0.06 0.37 96.38 YR 256 4.77 0.54 -29.33 96.78 YR 257 4.31 0.62 -29.74 0.06 -0.14 97.18 YR 258 4.43 0.38 -29.32 97.58 YR 259 4.66 0.67 -29.30 0.05 -0.07 97.98 YR 260 4.26 0.52 -29.14 98.38 YR 261 4.83 0.64 -29.20 0.06 0.23 98.78 YR 262 4.41 0.45 -29.45 99.18 YR 263 4.35 0.42 -29.21 0.06 0.31 99.58 256 Haida Gwaii Yakoun River 2008 data Identifier Amt. (Mg) TOC% \u00CE\u00B413C adj TN% \u00CE\u00B415N Ht (m) YR 264 4.41 0.54 -29.10 99.98 YR 265 4.75 0.35 -29.19 0.06 0.33 100.38 YR 266 4.49 0.41 -29.17 100.78 YR 267 4.68 0.52 -29.32 0.06 0.31 101.18 YR 268 4.61 0.61 -29.27 101.58 YR 269 4.29 0.59 -29.24 0.06 0.30 101.98 YR 270 4.49 0.60 -29.06 102.38 YR 271 4.84 0.46 -29.36 0.05 -0.28 102.78 YR 272 4.20 0.52 -28.60 103.18 YR 273 4.64 0.41 -29.12 0.06 -0.07 103.58 YR 274 4.70 0.50 -29.21 103.98 257 Haida Gwaii Yakoun River 2008 wood data Identifier Amt. (Mg) \u00CE\u00B413Cadj Ht (m) YR41W 0.046 -21.19 16.40 YR41W 0.061 -20.53 16.40 YR72W 0.053 -29.10 31.20 YR77W 0.059 -29.74 33.20 YR77W 0.050 -29.94 33.20 YR01W 0.048 -26.20 47.80 YR02W 0.048 -22.62 49.20 YR08W 0.045 -22.13 49.20 YR03W 0.043 -25.20 49.90 YR04W 0.046 -22.37 50.00 YR05W 0.043 -24.90 51.00 YR06W 0.048 -22.16 51.00 YR07W 0.046 -24.46 51.40 YR10W 0.057 -23.02 51.40 YR09W 0.052 -23.80 53.40 YR11W 0.054 -23.72 60.50 YR12W 0.050 -20.55 60.80 YR13W 0.045 -22.41 61.50 YR14W 0.045 -28.84 71.20 YR15W 0.044 -25.56 71.67 YR16W 0.046 -23.52 71.72 YR17W 0.043 -25.79 71.85 YR20W 0.046 -26.09 79.05 YR18W 0.048 -23.67 80.25 YR19W 0.044 -25.72 80.50 YR21W 0.046 -24.30 90.38 258 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 275 3.454 0.35 -25.00 27.20 YR 276 3.532 0.41 -25.41 27.30 YR 277 3.739 0.30 -25.13 27.40 YR 278 3.088 0.35 -26.11 27.50 YR 279 3.078 0.44 -25.47 27.60 YR 280 3.732 0.36 -25.80 27.70 YR 281 3.001 0.35 -25.90 27.80 YR 282 3.221 0.35 -25.40 27.90 YR 283 3.418 0.35 -25.27 28.00 YR 284 3.181 0.35 -25.20 28.10 YR 285 3.555 0.33 -25.40 28.20 YR 286 3.182 0.32 -25.04 28.30 YR 287 3.057 0.31 -25.66 28.40 YR 288 3.149 0.36 -25.23 28.50 YR 289 3.582 0.33 -25.18 28.60 YR 290 3.811 0.32 -25.47 28.70 YR 291 3.825 0.35 -24.89 28.80 YR 292 3.802 0.34 -24.81 28.90 YR 293 3.758 0.41 -25.05 29.00 YR 294 3.239 0.34 -25.35 29.10 YR 295 3.419 0.38 -25.06 29.20 YR 296 3.167 0.39 -24.52 29.30 YR 297 3.259 0.48 -24.37 29.40 YR 298 3.331 0.52 -24.85 29.50 YR 299 3.761 0.51 -24.93 29.60 YR 300 3.131 0.66 -24.54 29.70 YR 301 3.561 0.58 -24.59 29.80 YR 302 3.306 0.35 -25.20 29.90 YR 303 3.529 0.59 -24.17 30.00 YR 304 3.549 0.57 -24.39 30.05 YR 305 3.557 0.53 -24.65 30.10 YR 306 3.666 0.53 -24.63 30.15 YR 307 3.292 0.51 -24.70 30.20 YR 308 3.355 0.58 -24.40 30.25 YR 309 3.342 0.65 -26.20 30.30 YR 310 3.262 0.47 -25.00 30.35 259 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 311 3.749 0.40 -25.23 30.40 YR 312 3.551 0.47 -26.29 30.45 YR 313 3.957 0.44 -26.32 30.50 YR 314 3.663 0.42 -25.79 30.55 YR 315 4.149 0.47 -25.89 30.60 YR 316 3.53 0.43 -25.53 30.65 YR 317 3.744 0.40 -26.37 30.70 YR 318 3.937 0.39 -25.95 30.75 YR 319 3.773 0.40 -26.23 30.80 YR 320 6.359 0.12 -29.73 30.85 YR 321 4.264 0.16 -29.47 30.90 YR 322 3.453 0.24 -29.65 30.95 YR 323 6.155 0.21 -29.71 31.00 YR 324 6.616 0.22 -29.72 31.05 YR 325 5.232 0.13 -29.50 31.10 YR 326 5.971 0.19 -29.83 31.15 YR 327 3.632 0.27 -30.21 31.20 YR 328 4.185 0.24 -30.85 31.25 YR 329 3.724 0.28 -31.15 31.30 YR 330 3.043 0.56 -31.48 31.35 YR 331 3.994 0.17 -30.12 31.40 YR 332 3.564 0.94 -32.07 31.45 YR 333 3.643 0.23 -30.85 31.50 YR 334 3.123 0.60 -31.69 31.55 YR 335 3.875 0.72 -31.95 31.60 YR 336 3.418 0.46 -31.12 31.65 YR 337 3.432 0.51 -31.51 31.70 YR 338 3.142 0.69 -31.56 31.75 YR 339 3.614 1.07 -31.24 31.80 YR 340 3.763 0.99 -31.21 31.85 YR 341 3.504 0.88 -31.00 31.90 YR 342 3.879 0.78 -31.06 31.95 YR 343 3.936 0.50 -30.89 32.00 YR 344 3.381 0.95 -32.07 32.05 YR 345 3.768 0.68 -31.66 32.10 YR 346 3.374 0.76 -31.94 32.15 260 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 347 3.879 0.76 -31.62 32.20 YR 348 3.721 0.59 -31.39 32.25 YR 349 3.619 0.90 -31.59 32.30 YR 350 4.056 0.56 -30.75 32.35 YR 351 3.538 0.61 -30.85 32.40 YR 352 3.435 0.67 -30.39 32.45 YR 353 3.561 0.90 -31.09 32.50 YR 354 3.473 0.73 -30.96 32.55 YR 355 3.735 0.62 -30.83 32.60 YR 356 3.117 0.75 -30.33 32.65 YR 357 3.382 0.55 -30.67 32.70 YR 358 3.775 0.52 -30.95 32.75 YR 359 3.967 0.62 -30.09 32.80 YR 360 3.967 0.43 -31.40 32.85 YR 361 4.041 0.51 -31.23 32.90 YR 362 3.374 0.49 -31.32 32.95 YR 363 3.666 0.52 -31.38 33.00 YR 364 3.356 0.36 -31.03 33.05 YR 365 3.518 0.72 -31.26 33.10 YR 367 3.46 0.58 -31.26 33.15 YR 368 3.449 0.55 -31.47 33.20 YR 369 3.063 0.56 -31.24 33.25 YR 370 3.249 0.61 -31.13 33.30 YR 371 3.652 0.76 -31.31 33.35 YR 372 3.67 0.82 -30.76 33.40 YR 373 3.439 0.91 -31.43 33.45 YR 374 3.435 0.98 -31.22 33.50 YR 375 3.171 0.63 -31.51 33.55 YR 376 4.095 0.45 -31.34 33.60 YR 377 3.165 0.55 -31.46 33.65 YR 378 3.661 0.55 -31.72 33.70 YR 379 3.311 0.37 -30.22 33.75 YR 380 3.51 0.30 -31.02 33.80 YR 381 3.658 0.29 -31.23 33.85 YR 382 3.557 0.41 -31.16 33.90 YR 383 3.774 0.39 -31.47 33.95 261 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 384 3.583 0.58 -31.69 34.00 YR 385 3.73 0.48 -31.45 34.05 YR 386 3.685 0.36 -31.12 34.10 YR 387 3.903 0.51 -31.31 34.15 YR 388 3.028 0.49 -31.10 34.20 YR 389 3.524 0.53 -31.14 34.25 YR 390 3.324 0.71 -30.06 34.30 YR 391 3.24 0.59 -31.50 34.35 YR 392 3.217 0.54 -31.20 34.40 YR 393 3.872 0.63 -30.34 34.45 YR 394 3.673 0.32 -30.81 34.50 YR 395 3.543 0.60 -31.71 34.55 YR 396 3.281 0.65 -30.99 34.60 YR 397 3.776 0.49 -30.67 34.65 YR 398 3.705 0.38 -31.21 34.70 YR 399 3.477 0.35 -30.88 34.75 YR 400 3.794 0.37 -30.61 34.80 YR 401 3.465 0.32 -30.54 34.85 YR 402 3.247 0.32 -30.12 34.90 YR 403 3.948 0.26 -30.06 34.95 YR 404 6.019 0.29 -30.04 35.00 YR 405 3.637 0.38 -30.50 35.05 YR 406 3.73 0.44 -30.24 35.10 YR 407 3.137 0.42 -30.39 35.15 YR 408 3.554 0.41 -30.39 35.20 YR 409 3.112 0.31 -30.26 35.25 YR 410 3.237 0.74 -30.50 35.30 YR 411 3.728 0.49 -30.66 35.35 YR 412 3.474 0.36 -30.65 35.40 YR 413 3.406 0.38 -30.67 35.45 YR 414 3.406 0.50 -30.46 35.50 YR 415 3.098 0.57 -30.90 35.55 YR 416 3.746 0.54 -31.37 35.60 YR 417 3.703 0.47 -30.51 35.65 YR 418 3.586 0.68 -24.21 35.70 YR 419 3.493 0.81 -24.03 35.75 262 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 420 3.898 0.79 -24.15 35.80 YR 421 3.521 0.86 -24.16 35.85 YR 422 3.591 0.98 -23.95 35.90 YR 423 3.872 1.00 -24.47 35.95 YR 424 3.792 0.92 -24.40 36.00 YR 425 3.721 0.89 -24.29 36.05 YR 426 3.141 0.65 -25.31 36.10 YR 427 3.885 0.58 -25.51 36.15 YR 428 3.836 0.66 -25.94 36.20 YR 429 3.66 0.70 -25.79 36.25 YR 430 3.452 0.84 -24.67 36.30 YR 431 3.732 0.74 -24.32 36.35 YR 432 3.06 0.85 -24.11 36.40 YR 433 3.021 0.86 -24.16 36.45 YR 434 3.631 0.81 -25.10 36.50 YR 435 3.152 0.94 -24.01 36.55 YR 436 3.059 0.96 -24.21 36.60 YR 437 3.724 0.90 -24.37 36.65 YR 438 3.384 0.91 -24.09 36.70 YR 439 3.567 0.90 -24.55 36.75 YR 440 3.742 0.88 -24.55 36.80 YR 441 3.266 0.90 -24.64 36.85 YR 442 3.685 0.81 -24.68 36.90 YR 443 3.669 0.81 -24.67 36.95 YR 444 3.737 0.76 -24.49 37.00 YR 445 3.031 0.80 -25.34 37.05 YR 446 3.112 0.85 -25.01 37.10 YR 447 3.674 0.84 -24.57 37.15 YR 448 3.278 0.57 -25.25 37.20 YR 449 6.123 0.79 -25.09 37.25 YR 450 6.029 0.70 -24.21 37.30 YR 451 3.528 0.67 -24.72 37.35 YR 452 3.734 0.96 -25.33 37.40 YR 453 5.861 0.81 -24.48 37.45 YR 454 6.333 0.26 -28.48 37.50 YR 455 5.519 0.21 -27.25 37.55 263 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 456 3.333 0.70 -23.74 37.60 YR 457 3.812 0.67 -23.86 37.65 YR 458 3.872 0.74 -23.98 37.70 YR 459 3.926 0.73 -25.50 37.75 YR 460 3.568 0.57 -25.77 37.80 YR 461 4.086 0.52 -26.49 37.85 YR 462 3.282 0.59 -25.64 37.90 YR 463 3.809 0.53 -25.70 37.95 YR 464 5.766 0.54 -25.53 38.00 YR 465 3.468 0.55 -25.53 38.05 YR 466 3.502 0.49 -25.76 38.10 YR 467 3.701 0.50 -25.05 38.15 YR 468 4.423 0.57 -25.17 38.20 YR 469 2.937 0.96 -28.99 38.25 YR 470 3.668 0.85 -24.94 38.30 YR 471 3.588 0.88 -25.25 38.35 YR 472 3.24 0.81 -24.76 38.40 YR 473 3.669 0.80 -25.49 38.45 YR 474 4.034 0.59 -25.88 38.50 YR 475 3.624 0.75 -24.90 38.55 YR 476 3.988 0.82 -25.24 38.60 YR 477 3.889 0.92 -25.91 38.65 YR 478 3.443 0.93 -24.79 38.70 YR 479 3.691 0.88 -25.10 38.75 YR 480 3.526 0.93 -25.22 38.80 YR 481 3.449 0.95 -25.33 38.85 YR 482 3.799 0.81 -25.30 38.90 YR 483 3.39 0.77 -24.51 38.95 YR 483 4.875 0.49 -25.85 39.00 YR 484 4.42 0.81 -24.85 39.05 YR 485 3.408 0.79 -24.41 39.10 YR 486 4.292 0.77 -24.77 39.15 YR 487 3.326 0.82 -24.28 39.20 YR 488 3.787 0.90 -24.71 39.25 YR 489 3.907 0.85 -24.62 39.30 YR 490 3.484 0.86 -25.02 39.35 264 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 491 3.284 0.80 -24.66 39.40 YR 492 3.736 0.67 -25.17 39.45 YR 493 3.55 0.84 -25.34 39.50 YR 494 3.785 0.79 -25.32 39.55 YR 495 3.808 0.83 -25.75 39.60 YR 496 3.355 0.85 -25.30 39.65 YR 497 3.298 0.69 -25.43 39.70 YR 498 3.39 0.49 -26.25 39.75 YR 499 3.274 0.36 -26.85 39.80 YR 500 3.116 0.55 -27.23 39.85 YR 501 3.962 0.58 -24.98 39.90 YR 501 3.804 0.50 -25.37 39.95 YR 502 4.388 0.48 -27.84 40.00 YR 502 4.639 0.42 -27.97 40.05 YR 503 3.638 0.83 -24.94 40.10 YR 503 3.462 0.72 -25.29 40.15 YR 504 4.398 1.50 -25.62 40.20 YR 505 3.26 0.96 -26.03 40.25 YR 506 3.298 0.83 -25.47 40.30 YR 507 3.676 0.86 -25.96 40.35 YR 508 3.14 0.64 -25.86 40.40 YR 509 3.077 0.75 -26.09 40.45 YR 510 3.644 0.52 -26.29 40.50 YR 511 3.433 0.58 -26.05 40.55 YR 512 3.206 0.67 -25.76 40.60 YR 513 3.662 0.79 -25.36 40.65 YR 514 3.492 0.80 -24.99 40.70 YR 515 3.395 0.64 -28.16 40.75 YR 516 3.707 0.66 -28.34 40.80 YR 517 3.168 0.52 -25.60 40.85 YR 518 3.209 0.56 -25.50 40.90 YR 519 3.603 0.66 -25.49 40.95 YR 520 3.836 0.75 -25.73 41.00 YR 522 3.287 0.71 -25.59 41.05 YR 523 3.418 0.72 -25.29 41.10 YR 524 3.448 0.65 -25.51 41.15 265 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 525 3.811 0.69 -25.38 41.20 YR 526 3.298 0.75 -25.30 41.25 YR 527 3.514 0.56 -27.28 41.30 YR 528 3.649 0.71 -25.77 41.35 YR 529 3.811 0.65 -25.72 41.40 YR 530 3.587 1.10 -26.99 41.45 YR 531 3.799 0.68 -25.46 41.50 YR 532 3.604 0.70 -25.60 41.55 YR 533 3.496 0.68 -25.58 41.60 YR 534 3.736 0.68 -25.26 41.65 YR 535 3.326 0.63 -25.39 41.70 YR 536 3.449 0.70 -25.25 41.75 YR 537 3.354 0.86 -25.07 41.80 YR 538 3.779 0.69 -25.26 41.85 YR 539 3.987 0.63 -24.89 41.90 YR 540 3.098 0.62 -24.74 41.95 YR 541 3.524 0.64 -25.22 42.00 YR 542 3.382 0.59 -24.86 42.05 YR 543 3.811 0.68 -25.65 42.10 YR 544 3.624 0.61 -25.89 42.15 YR 545 3.796 0.68 -26.26 42.20 YR 546 3.625 0.64 -25.71 42.25 YR 547 3.77 0.68 -26.15 42.30 YR 548 3.241 0.77 -25.69 42.35 YR 549 3.702 0.56 -25.55 42.40 YR 550 5.303 0.63 -25.82 42.45 YR 551 3.533 0.53 -25.93 42.50 YR 552 3.372 0.61 -25.59 42.55 YR 553 3.673 0.63 -25.17 42.60 YR 554 3.976 0.58 -25.64 42.65 YR 555 3.03 0.64 -25.65 42.70 YR 556 5.285 0.28 -25.86 42.75 YR 557 3.689 0.61 -25.95 42.80 YR 558 3.679 0.51 -26.03 42.85 YR 559 3.578 0.51 -26.44 42.90 YR 560 3.765 0.45 -26.14 42.95 266 Haida Gwaii Yakoun River 2010 data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj Ht (m) YR 561 3.42 0.47 -25.86 43.00 YR 562 3.328 0.62 -25.26 43.05 YR 563 3.246 0.62 -25.25 43.10 YR 564 3.931 0.60 -24.38 43.15 YR 565 3.536 0.81 -24.00 43.20 YR 566 4.272 0.72 -23.98 43.25 YR 567 3.342 0.62 -24.14 43.30 YR 568 3.623 0.56 -24.48 43.35 YR 569 4.042 0.71 -24.34 43.40 YR 570 3.959 0.60 -24.72 43.45 YR 571 4.077 0.76 -24.92 43.50 YR 572 3.196 0.71 -24.42 43.55 YR 573 3.835 0.66 -24.55 43.60 YR 574 3.38 0.66 -24.45 43.65 YR 575 3.298 0.68 -24.89 43.70 YR 576 3.73 0.70 -24.95 43.75 YR 577 3.996 0.62 -25.09 43.80 YR 578 3.168 0.59 -25.04 43.85 YR 579 3.386 0.58 -25.13 43.90 YR 580 3.639 0.61 -24.89 43.95 YR 581 3.526 0.79 -24.52 44.00 YR 582 6.273 0.30 -23.55 44.05 44.15 YR 584 3.808 0.55 -25.74 44.25 YR 585 3.453 0.83 -25.55 44.35 YR 586 3.72 0.62 -25.65 44.45 YR 587 4.184 0.65 -25.20 44.55 YR 588 3.061 0.59 -25.58 44.65 YR 589 3.593 0.57 -25.16 44.75 YR 590 3.099 0.64 -24.75 44.85 YR 591 3.735 0.59 -25.48 44.95 YR 592 6.65 0.62 -24.74 45.05 267 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 079 0.90 2.15 -29.36 0.19 0.00 0.00 WB 080 2.71 0.74 -29.15 0.30 WB 081 6.51 0.40 -28.72 0.60 WB 082 5.64 0.33 -28.31 0.90 WB 083 1.01 1.33 -29.08 1.20 WB 084 1.45 1.42 -29.40 1.50 WB 085 3.25 0.74 -28.39 0.08 0.67 1.80 WB 086 3.29 0.66 -29.23 2.10 WB 087 1.02 1.47 -29.39 2.40 WB 088 1.36 1.12 -29.41 2.70 WB 089 1.46 2.03 -29.45 3.00 WB 090 1.94 1.77 -28.59 3.30 WB 091 0.88 2.90 -28.65 0.16 0.70 3.60 WB 092 0.80 2.46 -28.82 3.90 WB 093 1.05 1.83 -29.27 4.20 WB 094 0.82 2.06 -28.84 4.50 WB 095 1.08 0.81 -28.56 4.80 WB 096 1.80 0.72 -27.70 5.10 WB 097 1.31 1.32 -28.36 0.16 -0.25 6.30 WB 098 0.96 1.63 -28.14 6.60 WB 099 0.83 1.31 -28.57 6.90 WB 100 2.66 0.85 -28.75 7.20 WB 101 2.53 0.66 -28.50 7.50 WB 102 2.74 0.91 -26.68 7.80 WB 103 2.61 1.02 -28.84 0.14 -0.27 8.10 WB 104 2.17 1.24 -28.33 8.40 WB 105 0.69 1.62 -28.37 8.70 WB 106 0.65 2.22 -28.36 9.00 WB 107 1.88 1.42 -28.27 9.30 WB 108 0.77 2.51 -28.52 9.60 WB 109 0.82 2.15 -27.63 0.14 0.97 9.90 WB 110 0.73 2.37 -27.99 10.20 WB 111 0.73 2.52 -28.06 10.50 WB 112 0.68 2.17 -28.10 10.80 WB 113 0.82 3.00 -27.56 11.10 WB 114 0.63 2.84 -27.98 11.40 268 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 115 0.71 2.10 -27.83 0.15 0.09 11.70 WB 116 0.62 2.28 -28.12 12.00 WB 117 0.90 1.61 -28.18 12.30 WB 118 0.91 1.68 -28.22 12.60 WB 119 0.74 1.80 -27.86 12.90 WB 120 0.78 2.03 -28.69 13.20 WB 121 0.62 1.87 -28.00 0.20 -0.18 13.50 WB 122 0.79 2.45 -28.10 13.80 WB 123 0.66 2.38 -27.64 14.10 WB 124 14.40 WB 125 0.69 3.36 -28.18 14.70 WB 126 0.61 1.79 -27.79 15.00 WB 127 0.75 1.90 -27.67 15.30 WB 128 0.77 1.87 -27.50 0.17 0.21 15.60 WB 129 0.93 2.34 -27.84 15.90 WB 130 0.82 1.56 -26.57 16.20 WB 131 0.62 1.39 -27.54 16.50 WB 132 3.10 1.14 -25.22 16.80 WB 133 0.84 1.97 -27.20 17.10 WB 134 3.00 0.70 -26.51 0.07 0.88 17.40 WB 135 0.66 1.57 -27.22 17.70 WB 136 0.90 1.71 -26.82 18.00 WB 137 0.66 1.62 -25.76 18.30 WB 138 3.46 0.38 -25.66 18.60 WB 139 0.69 1.79 -27.72 18.90 WB 140 0.92 1.70 -27.06 0.12 0.12 19.30 WB 141 1.07 1.20 -27.36 19.60 WB 142 1.29 1.13 -26.74 19.90 WB 143 6.01 0.18 -27.60 20.20 WB 144 1.86 0.65 -26.16 20.50 WB 145 0.78 1.64 -26.96 20.80 WB 146 0.66 3.33 -26.85 0.16 -0.18 21.10 WB 147 1.10 1.37 -26.66 21.40 WB 148 0.93 2.10 -26.46 21.70 WB 149 0.86 1.59 -26.51 22.00 WB 150 0.90 2.08 -26.48 22.30 269 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 151 0.84 1.32 -26.57 22.60 WB 152 0.65 2.03 -27.08 23.13 WB 153 3.71 0.34 -27.04 23.43 WB 154 3.69 0.51 -28.16 23.73 WB 155 3.44 0.47 -27.98 24.03 WB 156 3.28 0.71 -27.62 24.33 WB 157 3.41 0.74 -28.07 24.63 WB 158 0.88 1.71 -27.47 24.93 WB 159 3.63 0.51 -25.81 25.23 WB 160 3.38 0.45 -27.50 25.53 WB 161 2.90 0.55 -27.68 25.83 WB 162 3.62 0.48 -27.31 26.13 WB 163 3.73 0.36 -27.56 26.43 WB 164 3.38 0.59 -28.90 26.83 WB 165 0.71 1.38 -28.40 27.13 WB 166 1.36 1.13 -28.39 27.43 WB 167 3.32 0.43 -28.59 27.73 WB 168 4.15 0.52 -27.83 27.83 WB 169 6.39 0.16 -28.45 28.13 WB 170 0.90 0.85 -28.09 28.43 WB 171 0.73 1.56 -28.35 28.73 WB 172 0.78 1.70 -27.59 29.03 WB 173 1.08 2.05 -26.60 29.33 WB 174 1.06 1.83 -27.85 29.63 WB 175 0.62 1.99 -28.28 29.93 WB 176 3.07 0.88 -23.56 30.23 WB 177 0.75 0.92 -28.03 30.53 WB 178 3.50 0.40 -26.74 30.83 WB 179 3.33 0.30 -27.65 31.13 WB 180 3.71 0.49 -27.34 31.43 WB 181 3.61 0.30 -24.82 31.73 WB 182 3.67 0.33 -28.12 33.03 WB 183 3.57 0.44 -24.33 33.33 WB 184 4.50 0.45 -22.39 33.63 WB 185 3.28 0.30 -28.08 33.93 WB 186 3.14 0.46 -24.76 34.23 270 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 187 3.05 1.00 -20.33 34.53 WB 188 2.98 0.35 -24.91 34.83 WB 189 3.69 0.64 -21.70 35.13 WB 190 3.47 0.62 -25.84 35.43 WB 191 3.46 0.80 -24.07 35.73 WB 192 3.67 0.62 -26.66 36.03 WB 193 3.13 0.61 -27.50 36.33 WB 194 3.96 0.51 -26.92 36.63 WB 195 3.20 0.68 -26.78 36.93 WB 196 3.53 0.84 -27.61 37.23 WB 197 3.70 0.56 -28.24 37.53 WB 198 2.95 0.83 -23.20 37.83 WB 199 3.53 0.27 -26.00 38.13 WB 200 3.42 0.49 -26.29 38.43 WB 201 3.82 0.27 -25.73 38.73 WB 202 3.73 0.46 -25.93 39.03 WB 203 3.28 0.45 -26.61 39.33 WB 204 3.53 0.40 -25.21 39.63 WB 205 6.03 0.34 -25.86 39.93 WB 206 3.54 0.51 -21.32 40.23 WB 207 3.39 0.58 -21.56 40.53 WB 208 4.08 0.44 -27.35 40.83 WB 209 4.15 0.52 -27.70 41.13 WB 210 4.04 0.54 -27.00 41.43 WB 211 6.13 0.34 -26.23 41.73 WB 212 3.32 0.79 -18.91 42.03 WB 213 3.63 0.82 -22.03 42.33 WB 214 3.19 0.61 -28.60 42.63 WB 215 5.25 0.94 -25.96 42.93 WB 216 3.20 0.37 -28.03 43.25 WB 217 3.27 0.58 -26.76 43.55 WB 218 3.44 0.29 -27.77 43.85 WB 219 5.94 0.25 -27.41 44.15 WB 220 3.03 0.46 -27.26 44.45 WB 221 6.55 0.33 -23.18 44.75 WB 222 3.98 0.18 -25.44 45.05 271 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 223 3.23 0.33 -27.52 45.35 WB 224 3.00 0.54 -27.45 45.65 WB 225 3.65 0.61 -27.28 45.95 WB 226 6.51 0.23 -27.86 46.25 WB 227 3.71 0.45 -28.24 46.55 WB 228 3.60 0.36 -25.46 46.85 WB 229 6.70 0.76 -18.29 47.15 WB 230 3.59 0.62 -25.54 47.45 WB 231 5.84 0.25 -26.70 47.75 WB 232 5.96 0.84 -19.86 48.05 WB 233 6.18 0.37 -27.72 48.35 WB 234 3.65 0.39 -24.18 48.65 WB 235 3.47 0.29 -26.88 48.95 WB 236 6.32 0.08 -27.25 49.25 WB 237 3.12 0.64 -27.69 49.55 WB 238 3.43 0.74 -26.50 49.85 WB 239 3.41 0.57 -26.17 50.15 WB 240 3.15 0.56 -28.08 50.45 WB 241 3.84 0.65 -28.24 50.75 WB 242 3.45 0.78 -25.84 51.05 WB 243 4.07 0.55 -28.03 51.35 WB 244 5.97 0.34 -22.14 51.65 WB 245 6.36 0.36 -22.72 51.95 WB 246 6.72 0.18 -27.57 52.25 WB 247 3.19 0.50 -22.61 52.55 WB 248 6.62 0.14 -26.71 52.85 WB 249 3.19 0.29 -26.42 53.15 WB 250 3.07 0.31 -27.86 53.45 WB 251 6.60 0.28 -28.94 53.75 WB 252 4.23 0.43 -21.73 54.05 WB 253 3.80 0.50 -27.35 54.35 WB 254 1.69 0.86 -28.43 54.65 WB 255 2.30 1.29 -27.92 54.95 WB 256 3.25 0.74 -28.62 55.25 WB 257 3.24 0.73 -27.22 55.55 WB 258 3.42 0.66 -28.30 55.85 272 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 259 3.86 0.42 -24.72 56.15 WB 260 3.16 0.60 -28.59 56.45 WB 261 3.41 0.86 -25.31 56.75 WB 262 5.92 0.49 -27.54 57.05 WB 263 3.66 0.37 -26.57 57.35 WB 264 4.15 0.27 -25.94 57.65 WB 265 3.97 0.74 -26.29 57.95 WB 266 3.54 0.57 -28.02 58.25 WB 267 3.73 0.70 -27.34 58.55 WB 268 3.59 0.54 -28.49 58.85 WB 269 3.38 0.41 -27.47 59.15 WB 270 6.70 0.23 -26.59 59.45 WB 271 3.66 0.41 -27.05 59.75 WB 272 6.51 0.27 -26.87 60.05 WB 273 3.85 0.53 -28.16 60.35 WB 274 4.80 0.53 -28.00 60.65 WB 275 6.34 0.16 -27.95 60.95 WB 276 3.64 0.46 -27.35 61.25 WB 277 3.73 0.37 -28.66 61.55 WB 278 3.11 0.70 -28.62 61.85 WB 279 3.61 0.38 -27.01 62.15 WB 280 3.25 0.79 -28.48 62.45 WB 281 6.34 0.35 -26.64 62.75 WB 282 3.46 0.35 -27.43 63.05 WB 283 3.82 0.61 -28.20 63.35 WB 284 3.32 0.14 -27.55 63.65 WB 285 4.21 0.66 -28.56 63.95 WB 286 3.52 0.31 -28.46 64.25 WB 287 3.28 0.90 -28.91 64.55 WB 288 3.73 0.82 -28.50 64.85 WB 289 3.60 0.76 -28.50 65.15 WB 290 6.73 0.28 -24.07 65.45 WB 291 3.08 0.84 -25.22 65.75 WB 292 5.09 0.29 -25.98 66.05 WB 293 3.03 0.81 -28.51 66.35 WB 294 3.19 0.70 -27.83 66.65 273 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 295 3.53 0.50 -24.96 66.95 WB 02 4.68 0.53 -28.14 67.05 WB 296 3.15 0.15 -28.72 67.25 WB 03 4.61 0.80 -26.61 67.35 WB 04 4.41 0.62 -27.58 67.50 WB 297 4.31 0.18 -27.89 67.55 WB 05 4.79 0.70 -28.46 67.60 WB 06 4.50 0.86 -27.62 67.80 WB 07 4.27 0.87 -23.52 67.84 WB 298 3.08 0.14 -27.51 67.85 WB 08 4.59 0.68 -28.34 68.04 WB 299 3.50 0.23 -28.56 68.15 WB 09 4.53 0.71 -27.24 68.24 WB 10 4.54 0.58 -28.15 68.44 WB 300 2.48 0.39 -28.21 68.45 WB 11 4.30 0.67 -25.40 68.64 WB 301 2.95 0.37 -27.65 68.75 WB 12 4.38 0.46 -28.48 68.84 WB 13 4.71 0.34 -28.73 69.04 WB 302 2.83 0.74 -28.93 69.05 WB 14 4.54 0.39 -27.76 69.24 WB 303 3.28 0.29 -28.19 69.35 WB 15 4.72 0.40 -27.05 69.54 WB 16 4.59 0.46 -28.77 69.74 WB 17 4.66 0.42 -27.76 69.86 WB 18 4.43 0.67 -28.96 70.06 WB 19 4.36 0.55 -28.41 70.33 WB 20 4.46 0.81 -26.37 70.61 WB 21 4.73 0.66 -27.53 70.81 WB 22 4.32 0.59 -28.60 71.01 WB 23 4.62 0.43 -28.86 71.21 WB 24 4.51 0.64 -28.84 71.41 WB 25 4.61 0.45 -28.09 71.61 WB 26 4.73 0.81 -26.48 71.81 WB 27 4.68 0.45 -27.87 72.01 WB 28 4.50 0.72 -26.79 72.21 274 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 29 4.77 0.68 -28.51 72.41 WB 30 4.65 0.45 -27.32 72.61 WB 31 4.54 0.39 -26.54 72.91 WB 32 4.63 0.56 -28.12 73.11 WB 33 4.72 0.43 -27.07 73.31 WB 34 4.77 0.47 -27.07 73.51 WB 35 4.66 0.45 -27.39 73.61 WB 36 4.69 0.40 -27.13 73.71 WB 37 4.51 0.54 -28.28 73.93 WB 38 4.70 0.59 -26.76 74.03 WB 39 4.39 0.51 -27.68 74.16 WB 40 4.48 0.70 -27.17 74.26 WB 41 4.45 0.52 -27.41 74.36 WB 42 4.04 0.57 -25.38 74.46 WB 43 4.43 0.52 -29.16 74.56 WB 44 4.33 0.47 -27.87 74.66 WB 45 4.56 0.50 -27.51 74.76 WB 46 4.61 0.69 -27.07 74.86 WB 47 4.44 0.36 -26.14 74.96 WB 48 4.80 0.52 -27.75 75.06 WB 49 4.58 0.61 -28.08 75.16 WB 50 4.61 0.50 -27.44 75.26 WB 51 4.74 0.46 -27.39 75.36 WB 52 4.44 0.65 -27.08 75.46 WB 53 4.57 0.42 -27.40 75.56 WB 54 4.61 0.46 -26.97 75.66 WB 55 4.41 0.61 -27.30 75.76 WB 56 4.69 0.51 -26.20 75.96 WB 57 4.79 0.64 -26.31 76.16 WB 304 3.03 0.53 -26.18 76.23 WB 58 4.63 0.44 -26.53 76.36 WB 305 3.88 0.53 -26.16 76.53 WB 59 4.66 0.47 -26.60 76.56 WB 306 3.71 0.41 -26.15 76.73 WB 60 4.46 0.51 -27.09 76.76 WB 307 3.72 0.51 -26.47 76.88 275 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 61 4.32 0.57 -27.09 76.96 WB 308 0.42 -26.95 77.03 WB 62 4.94 0.22 -26.82 77.16 WB 309 3.54 0.40 -28.43 77.18 WB 310 3.05 0.53 -28.38 77.28 WB 63 4.75 0.36 -27.02 77.36 WB 311 3.10 0.55 -28.35 77.38 WB 312 3.74 0.61 -28.14 77.48 WB 64 4.89 0.52 -27.24 77.56 WB 313 3.05 0.61 -27.94 77.58 WB 314 3.54 0.61 -28.17 77.68 WB 65 4.76 0.47 -27.25 77.76 WB 315 3.15 0.61 -28.49 77.78 WB 316 3.63 0.57 -28.45 77.88 WB 66 4.55 0.55 -26.76 77.96 WB 317 3.42 1.50 -28.43 77.98 WB 318 3.07 0.47 -27.97 78.08 WB 67 4.74 0.45 -27.00 78.16 WB 319 3.24 0.62 -28.49 78.28 WB 68 4.41 0.39 -29.02 78.36 WB 320 3.09 0.65 -28.23 78.48 WB 69 4.58 0.36 -28.59 78.56 WB 321 3.07 0.68 -28.05 78.68 WB 70 4.90 0.43 -28.52 78.76 WB 322 3.34 0.54 -28.21 78.88 WB 71 4.34 0.50 -28.48 78.96 WB 72 4.36 0.48 -28.59 79.16 WB 323 3.51 0.51 -28.19 79.18 WB 73 4.45 0.60 -28.51 79.36 WB 324 2.95 0.75 -28.28 79.48 WB 74 4.85 0.54 -28.40 79.56 WB 75 4.70 0.60 -28.52 79.76 WB 325 3.08 0.73 -28.10 79.78 WB 76 4.51 0.59 -28.70 79.96 WB 326 3.11 0.70 -28.33 80.08 WB 77 4.68 0.56 -28.56 80.16 276 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 78 4.35 0.62 -28.67 80.36 WB 327 2.89 0.80 -28.38 80.38 WB 328 3.21 0.67 -27.60 80.68 WB 329 3.57 0.63 -28.02 80.98 WB 330 3.27 0.82 -27.90 81.28 WB 331 3.42 0.71 -28.31 81.58 WB 332 3.22 0.65 -28.36 81.88 WB 333 5.17 0.63 -28.09 82.18 WB 334 3.84 0.68 -28.36 82.48 WB 335 3.62 0.65 -28.12 82.78 WB 336 3.26 0.79 -28.69 83.08 WB 337 3.48 0.56 -27.85 83.38 WB 338 3.07 0.50 -25.55 83.78 WB 339 3.01 0.33 -25.46 84.18 WB 340 3.56 0.42 -25.32 84.48 WB 341 3.49 0.40 -25.13 84.78 WB 342 3.65 0.34 -25.52 85.18 WB 343 3.27 0.34 -25.87 85.58 WB 344 3.51 0.33 -25.44 85.88 WB 345 3.58 0.41 -24.98 86.28 WB 346 3.64 0.22 -25.91 86.68 WB 347 3.49 0.36 -25.66 87.08 WB 348 3.80 0.50 -24.99 87.48 WB 349 3.20 0.55 -24.71 87.88 WB 350 6.23 0.47 -24.72 88.28 WB 351 3.99 0.44 -24.99 88.68 WB 352 5.15 0.44 -25.16 89.08 WB 353 3.21 0.73 -25.90 89.48 WB 354 3.44 0.39 -24.98 89.88 WB 355 3.86 0.42 -25.66 90.28 WB 356 4.21 0.35 -25.02 90.68 WB 357 3.29 0.34 -25.65 91.08 WB 358 3.27 0.35 -25.80 91.48 WB 359 5.06 0.21 -25.86 91.88 WB 360 3.39 0.30 -25.66 92.28 WB 361 3.74 0.35 -26.11 92.68 277 Haida Gwaii Whiteaves Bay data Identifier Amt. (Mg) TOC% \u00CE\u00B413Cadj TN% \u00CE\u00B415N Ht (m) WB 362 3.67 0.25 -26.08 93.08 WB 363 7.18 0.18 -26.89 93.48 WB 364 3.77 0.32 -26.04 93.88 WB 365 3.54 0.31 -25.95 94.28 WB 366 3.67 0.36 -26.19 94.68 WB 367 3.61 0.35 -25.15 95.08 WB 368 3.29 0.37 -25.58 95.48 WB 369 3.37 0.30 -25.42 95.88 WB 370 3.56 0.32 -25.15 96.28 WB 371 3.42 0.34 -25.06 99.67 WB 372 3.12 0.34 -25.44 100.07 WB 373 5.74 0.23 -26.03 101.97 WB 374 5.98 0.25 -26.06 102.37 WB 375 3.73 0.35 -25.40 102.77 278 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 001 4.35 1.17 -27.08 676.04 SB 002 4.33 1.02 -27.35 675.84 SB 003 4.25 0.92 -27.33 675.64 SB 004 4.29 1.11 -27.23 675.44 SB 005 4.14 0.95 -25.92 675.24 SB 006 4.24 1.37 -27.64 675.04 SB 007 4.17 0.83 -27.14 674.84 SB 008 4.28 1.21 -27.68 674.64 SB 009 4.19 0.90 -27.33 674.44 SB 010 4.08 1.19 -27.70 674.24 SB 011 4.15 1.12 -27.60 674.04 SB 012 4.21 1.07 -27.58 673.84 SB 013 4.23 0.80 -27.52 673.64 SB 014 4.35 1.03 -27.62 673.44 SB 015 4.32 1.04 -27.92 673.24 SB 016 4.31 0.90 -27.36 673.04 SB 017 4.23 0.80 -27.50 672.84 SB 018 4.01 0.99 -27.59 672.64 SB 019 4.26 0.95 -27.83 672.44 SB 020 4.24 0.87 -27.38 672.24 SB 021 4.01 1.13 -27.58 672.04 SB 022 4.37 0.84 -27.04 671.84 SB 023 4.03 0.86 -27.11 671.64 SB 024 4.04 0.90 -27.39 671.44 SB 025 4.40 1.41 -27.36 671.24 SB 026 4.33 1.22 -27.49 671.04 SB 027 4.11 1.45 -27.70 670.84 SB 028 4.15 2.49 -27.87 670.64 SB 029 4.08 1.77 -27.67 670.44 SB 030 4.34 1.94 -28.11 670.24 SB 031 4.17 0.94 -27.71 670.04 SB 032 4.35 1.07 -27.72 669.84 SB 033 4.14 1.63 -28.06 669.64 SB 034 4.30 1.59 -28.19 669.44 SB 035 4.34 1.06 -27.79 669.24 SB 036 4.01 1.29 -27.25 669.04 SB 037 4.19 0.97 -27.52 668.84 SB 038 4.27 1.93 -27.75 668.64 279 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 039 4.24 1.78 -27.69 668.44 SB 040 4.23 1.27 -27.99 668.24 SB 041 4.16 1.88 -28.24 668.04 SB 042 4.30 1.26 -28.00 667.84 SB 043 4.15 1.01 -27.97 667.64 SB 044 4.38 1.70 -28.19 667.44 SB 045 4.01 2.84 -27.90 667.24 SB 046 4.18 1.46 -28.22 667.04 SB 047 4.17 1.84 -28.47 666.84 SB 048 4.20 1.21 -28.01 666.64 SB 049 4.30 2.27 -27.86 666.44 SB 050 4.13 2.63 -28.28 666.24 SB 051 4.11 0.89 -27.88 666.04 SB 052 4.13 1.01 -27.62 665.84 SB 053 4.40 1.59 -27.94 665.64 SB 054 4.19 1.13 -27.56 665.44 SB 055 4.07 1.47 -27.99 665.24 SB 056 4.08 1.57 -27.94 665.04 SB 057 4.18 1.31 -28.10 664.84 SB 058 4.23 1.57 -27.93 664.64 SB 059 4.27 1.90 -28.33 664.44 SB 060 4.24 1.01 -27.65 664.24 SB 061 4.32 0.98 -27.50 664.04 SB 062 4.23 0.91 -27.57 663.84 SB 063 4.04 0.85 -27.23 663.64 SB 064 4.20 0.90 -27.36 663.44 SB 065 4.21 0.90 -26.98 663.24 SB 066 4.38 0.89 -27.01 663.04 SB 067 4.22 0.81 -27.26 662.84 SB 068 4.00 0.88 -26.62 662.64 SB 069 4.29 0.98 -26.73 662.44 SB 070 4.02 0.94 -26.51 662.24 SB 071 4.05 0.81 -26.66 662.04 SB 072 4.31 1.11 -26.96 661.84 SB 073 4.21 0.82 -27.08 661.64 SB 074 4.07 0.82 -26.82 661.44 SB 075 4.01 1.06 -27.16 661.24 SB 076 4.32 1.46 -27.41 661.04 280 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 077 4.21 1.02 -26.89 660.84 SB 078 4.06 1.30 -27.22 660.64 SB 079 4.21 0.93 -26.98 660.44 SB 080 4.20 1.49 -26.73 660.24 SB 081 4.36 0.79 -26.91 660.04 SB 082 4.31 0.93 -27.30 659.84 SB 083 4.01 0.81 -26.98 659.64 SB 084 4.12 0.74 -26.82 659.44 SB 085 4.20 0.76 -26.72 659.24 SB 086 4.24 0.73 -26.27 659.04 SB 087 4.08 0.65 -26.38 658.84 SB 088 4.18 0.59 -26.58 658.64 SB 089 4.12 0.73 -26.81 658.44 SB 090 4.25 0.79 -26.62 658.24 SB 091 4.18 0.70 -26.85 658.04 SB 092 4.29 2.40 -24.91 657.84 SB 093 4.29 0.70 -26.45 657.64 SB 094 4.23 0.78 -26.87 657.44 SB 095 4.14 0.79 -26.96 657.24 SB 096 4.05 0.57 -26.66 657.04 SB 097 4.23 1.02 -26.41 656.84 SB 098 4.37 1.11 -26.57 656.64 SB 099 4.35 0.96 -26.44 656.44 SB 100 4.06 0.93 -25.67 656.24 SB 101 4.16 0.78 -25.44 656.04 SB 102 4.33 0.77 -25.80 655.84 SB 103 4.33 0.94 -25.85 655.64 SB 104 4.21 0.82 -26.20 655.44 SB 105 4.26 0.98 -25.96 655.24 SB 106 4.23 1.01 -26.29 655.04 SB 107 4.25 1.04 -26.06 654.84 SB 108 4.21 1.01 -26.00 654.64 SB 109 4.17 0.74 -26.21 654.44 SB 110 4.23 0.84 -26.20 654.24 SB 111 4.17 0.93 -26.10 654.04 SB 112 4.16 0.87 -26.24 653.84 SB 113 4.16 0.86 -25.99 653.64 SB 114 4.09 0.69 -27.21 653.44 281 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 115 4.16 0.88 -26.12 653.24 SB 116 4.29 0.86 -25.61 653.04 SB 117 4.25 0.88 -25.65 652.84 SB 118 4.22 0.99 -25.12 652.64 SB 119 4.11 0.90 -25.28 652.44 GAP 1 SB 120 4.40 0.84 -26.48 645.87 SB 121 4.38 0.79 -26.78 645.72 SB 122 4.03 0.79 -26.59 645.57 SB 123 4.16 0.70 -26.51 645.42 SB 124 4.09 0.82 -27.01 645.27 SB 125 4.01 0.77 -26.59 645.12 SB 126 4.13 0.81 -26.83 644.97 SB 127 4.16 0.74 -26.96 644.82 SB 128 4.29 0.83 -26.55 644.67 SB 129 4.03 0.86 -26.35 644.52 SB 130 4.05 0.81 -26.11 644.37 SB 131 4.15 0.87 -26.54 644.22 SB 132 4.15 0.98 -25.77 644.07 SB 133 4.03 0.91 -25.97 643.92 SB 134 4.35 0.91 -25.54 643.77 SB 135 4.19 0.91 -25.62 643.62 SB 136 4.36 0.90 -25.82 643.47 SB 137 4.23 0.99 -25.87 643.32 SB 138 4.33 0.98 -26.10 643.17 SB 139 4.29 0.89 -26.42 643.02 SB 140 4.02 1.47 -26.97 642.87 SB 141 4.14 0.87 -26.19 642.72 SB 142 4.04 0.81 -26.69 642.57 SB 143 4.27 0.81 -26.72 642.42 SB 144 4.14 0.79 -26.51 642.27 SB 145 4.26 0.86 -26.69 642.12 SB 146 4.15 0.79 -26.74 641.97 SB 147 4.18 0.70 -26.75 641.82 SB 148 4.04 0.64 -26.63 641.67 SB 149 4.36 0.72 -26.92 641.52 SB 150 4.13 0.70 -27.14 641.37 SB 151 4.12 0.68 -26.30 641.22 282 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 152 4.17 0.66 -27.24 641.07 SB 153 4.34 0.69 -27.16 640.92 SB 154 4.23 0.74 -27.45 640.77 SB 155 4.12 0.77 -27.28 640.62 SB 156 4.12 0.88 -27.26 640.47 SB 157 4.15 0.76 -27.23 640.32 SB 158 4.30 0.78 -27.32 640.17 SB 159 4.16 0.72 -27.49 640.02 SB 160 4.07 0.83 -25.85 639.87 SB 161 4.33 0.70 -27.27 639.72 SB 162 4.19 0.75 -27.54 639.57 SB 163 4.34 0.74 -27.56 639.42 SB 164 4.20 0.88 -26.80 639.27 SB 165 4.23 0.87 -27.38 639.12 SB 166 4.40 0.66 -27.13 638.97 SB 167 4.06 0.65 -26.98 638.82 SB 168 4.17 0.67 -27.04 636.72 SB 169 4.11 0.74 -27.68 636.57 SB 170 4.22 0.65 -27.50 636.42 SB 171 4.20 0.60 -27.19 636.27 SB 172 4.21 0.70 -27.89 636.12 SB 173 4.31 0.69 -27.69 635.97 SB 174 4.33 0.78 -27.71 635.82 SB 175 4.34 0.75 -27.95 635.67 SB 176 4.31 0.70 -27.72 635.52 SB 177 4.33 0.86 -27.94 635.37 SB 178 4.31 0.75 -27.83 635.22 SB 179 4.01 0.72 -27.93 635.07 SB 180 4.08 0.61 -27.37 634.92 SB 181 4.32 0.65 -27.37 634.77 SB 182 4.11 0.69 -27.39 634.62 SB 183 4.05 0.73 -27.46 634.47 SB 184 4.23 0.66 -27.41 634.32 SB 185 4.14 0.65 -27.56 634.17 SB 186 4.27 0.56 -27.51 634.02 SB 187 4.05 0.61 -27.67 633.87 SB 188 4.19 0.59 -27.35 633.72 SB 189 4.15 0.68 -27.94 633.57 283 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 190 4.06 0.74 -27.99 633.42 SB 191 4.29 0.55 -27.82 633.27 SB 192 4.08 0.62 -27.98 633.12 SB 193 4.26 0.59 -27.69 632.97 SB 194 4.36 0.69 -27.92 632.82 SB 195 4.28 0.66 -27.80 632.67 SB 196 4.18 0.64 -27.91 632.52 SB 197 4.12 0.65 -27.84 632.37 SB 198 4.17 0.55 -27.52 632.22 SB 199 4.27 0.55 -27.73 632.07 SB 200 4.22 0.58 -27.67 631.92 SB 201 4.16 0.59 -27.75 631.77 SB 202 4.16 0.57 -27.79 631.67 SB 203 4.18 0.52 -27.55 631.57 SB 204 4.33 0.52 -27.83 631.47 SB 205 4.19 0.48 -27.21 631.37 SB 206 4.17 0.55 -27.55 631.27 SB 207 4.19 0.55 -27.45 631.17 SB 208 4.22 0.63 -27.59 631.07 SB 209 4.30 0.58 -27.58 630.97 SB 210 4.23 0.55 -27.29 630.87 SB 211 4.24 0.55 -27.16 630.77 SB 212 4.03 0.51 -27.56 630.67 SB 213 4.34 0.68 -28.53 630.57 SB 214 4.34 1.26 -27.93 630.47 SB 215 4.28 0.48 -27.60 630.37 SB 216 4.22 0.61 -27.95 630.27 SB 217 4.08 0.87 -28.87 630.17 SB 218 4.01 0.97 -28.84 630.07 SB 219 4.27 0.61 -28.13 629.97 SB 220 4.18 0.61 -27.64 629.87 SB 221 4.33 0.59 -28.10 629.77 SB 222 4.32 0.54 -27.77 629.67 SB 223 4.21 0.52 -27.83 629.57 SB 224 4.22 0.72 -27.92 629.47 SB 225 4.31 0.71 -28.18 629.37 SB 226 4.19 0.71 -28.17 627.27 SB 227 4.20 0.72 -27.82 627.17 284 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 228 4.26 0.77 -27.87 627.07 SB 229 4.35 1.67 -29.51 626.97 SB 230 4.07 0.84 -28.47 626.87 SB 231 4.15 0.80 -28.60 626.77 SB 232 4.32 1.27 -28.12 626.67 SB 233 4.38 0.94 -28.16 626.57 SB 234 4.06 0.57 -27.92 626.47 SB 235 4.02 0.88 -27.86 626.37 SB 236 4.16 0.69 -28.05 626.27 SB 237 4.11 0.98 -27.50 624.23 SB 238 4.35 0.94 -27.20 624.13 SB 239 4.36 0.87 -27.28 624.03 SB 240 4.27 0.87 -26.79 623.93 SB 241 4.07 0.95 -26.84 623.83 SB 242 4.24 0.94 -26.99 623.73 SB 243 4.04 0.99 -27.09 623.63 SB 244 4.20 0.81 -26.74 623.53 SB 245 4.17 0.92 -27.03 623.43 SB 246 4.16 1.04 -27.11 623.33 SB 247 4.13 0.97 -27.34 623.23 SB 248 4.13 1.04 -27.35 623.13 SB 249 4.28 1.12 -27.72 623.03 SB 250 4.35 1.19 -27.46 622.93 SB 251 4.19 1.16 -27.35 622.83 SB 252 4.18 1.23 -27.72 622.73 SB 253 4.30 1.26 -27.68 622.63 SB 254 4.25 1.21 -27.54 622.53 SB 255 4.35 1.19 -27.60 622.43 SB 256 4.17 1.15 -27.63 622.33 SB 257 4.32 1.08 -27.53 622.23 SB 258 4.15 1.23 -27.77 622.13 SB 259 4.36 1.10 -27.38 622.03 SB 260 4.31 1.07 -27.29 621.93 SB 261 4.30 1.02 -27.34 621.83 SB 262 4.14 1.03 -27.39 621.73 SB 263 4.07 1.00 -27.37 621.63 SB 264 4.05 0.90 -27.28 621.53 SB 265 4.35 0.98 -27.42 621.43 285 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 266 4.26 0.97 -27.42 621.33 SB 267 4.23 0.93 -27.35 621.23 SB 268 4.20 0.83 -27.28 621.13 SB 269 4.17 1.14 -27.36 621.03 SB 270 4.11 0.93 -27.60 620.93 SB 271 4.12 1.40 -26.95 620.83 SB 272 4.12 1.40 -27.59 620.73 SB 273 4.29 1.07 -27.60 620.63 SB 274 4.20 0.98 -27.57 620.53 SB 275 4.31 1.03 -27.50 620.43 SB 276 4.16 1.07 -27.44 620.33 SB 277 4.22 1.30 -27.33 620.23 SB 278 4.31 1.12 -27.16 618.13 SB 279 4.12 1.06 -27.20 618.03 SB 280 4.29 1.09 -27.36 617.93 SB 281 4.29 1.14 -27.48 617.83 SB 282 4.26 1.05 -27.45 617.73 SB 283 4.10 1.11 -27.59 617.63 SB 284 4.10 1.12 -27.37 617.53 SB 285 4.01 1.04 -27.34 617.43 SB 286 4.13 1.10 -27.34 617.33 SB 287 4.26 1.10 -27.20 617.23 SB 288 4.12 0.98 -27.68 617.13 SB 289 4.07 1.83 -27.22 617.03 SB 290 4.29 1.05 -27.96 616.93 SB 291 4.12 1.30 -28.26 616.83 SB 292 4.03 1.07 -28.24 616.73 SB 293 4.06 1.14 -28.39 616.63 SB 294 4.22 1.33 -28.49 616.53 SB 295 4.29 1.27 -28.28 616.43 SB 296 4.15 1.24 -28.25 616.33 SB 297 4.03 1.23 -28.21 616.23 SB 298 4.17 1.11 -28.02 616.13 SB 299 4.23 1.15 -27.97 616.03 SB 300 4.26 1.05 -27.80 615.93 SB 301 4.05 1.33 -27.79 615.83 SB 302 4.29 1.10 -27.55 615.73 SB 303 4.13 1.14 -27.52 615.63 286 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 304 4.17 1.03 -27.56 615.53 SB 305 4.12 0.95 -27.19 615.43 SB 306 4.06 1.26 -27.71 615.33 SB 307 4.16 1.47 -28.10 615.23 SB 308 4.34 1.35 -28.02 615.13 SB 309 4.01 1.44 -28.00 615.03 SB 310 4.26 1.45 -28.06 614.93 SB 311 4.29 1.28 -27.93 614.83 SB 312 4.16 1.32 -27.92 614.73 SB 313 4.31 1.36 -27.90 614.63 SB 314 4.20 1.34 -27.90 614.53 SB 315 4.35 1.10 -27.93 614.43 SB 316 4.33 1.02 -27.95 614.33 SB 317 4.33 0.77 -28.13 614.23 SB 318 4.63 0.88 -28.12 614.13 SB 319 4.36 0.96 -28.32 614.03 SB 320 4.24 1.00 -28.26 613.93 SB 321 4.78 0.85 -28.29 613.83 SB 322 4.87 0.90 -28.38 613.73 SB 323 4.68 1.15 -28.43 613.63 SB 324 4.50 1.10 -28.27 613.53 SB 325 4.15 1.07 -27.99 613.43 SB 326 4.21 1.04 -28.05 613.33 SB 327 4.19 2.27 -28.24 613.23 SB 328 4.24 1.18 -27.86 613.13 SB 329 4.26 1.04 -27.53 613.03 SB 330 4.67 1.06 -27.80 612.93 SB 331 4.02 1.07 -27.74 612.83 SB 332 4.25 0.91 -27.67 612.73 SB 333 4.76 0.98 -27.49 612.63 SB 334 4.21 0.97 -27.13 612.53 SB 335 4.85 0.95 -27.43 612.43 SB 336 4.44 1.01 -27.32 612.33 SB 337 4.85 0.86 -27.36 612.23 SB 338 4.81 0.86 -26.94 612.13 SB 339 4.50 0.92 -27.48 612.03 SB 340 4.69 0.92 -27.55 611.93 SB 341 4.17 0.85 -27.19 611.83 287 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 342 4.91 0.83 -27.12 611.73 SB 343 4.36 0.84 -27.01 611.63 SB 344 4.17 0.89 -27.34 611.53 SB 345 4.81 0.79 -26.64 611.43 SB 346 4.26 0.77 -26.76 611.33 SB 347 4.83 0.79 -27.23 611.23 SB 348 4.85 0.78 -26.87 611.13 SB 349 4.29 0.75 -26.90 608.68 SB 350 4.05 0.76 -27.06 608.58 SB 351 4.60 0.72 -26.44 608.48 SB 352 4.77 0.64 -27.03 608.38 SB 353 4.78 0.81 -27.28 608.28 SB 354 4.60 0.67 -27.37 608.18 SB 355 4.27 1.10 -27.59 608.08 SB 356 4.32 0.96 -27.41 607.98 SB 357 4.24 0.92 -27.48 607.88 SB 358 4.61 0.96 -27.39 607.78 SB 359 4.58 0.92 -27.18 607.68 SB 360 4.27 0.96 -27.51 607.58 SB 361 4.28 0.96 -27.36 607.48 SB 362 4.19 0.82 -27.47 607.38 SB 363 4.76 0.89 -27.97 607.28 SB 364 4.46 0.95 -27.59 607.18 SB 365 4.12 0.86 -27.24 607.08 SB 366 4.91 0.96 -27.46 606.98 SB 367 4.77 0.83 -27.53 606.88 SB 368 4.51 0.84 -27.57 606.78 SB 369 4.19 0.86 -27.19 606.68 SB 370 4.33 0.81 -27.42 606.58 SB 371 4.23 0.82 -27.45 606.48 SB 372 4.39 0.82 -27.45 606.38 SB 373 4.92 0.85 -27.38 606.28 SB 374 4.71 0.81 -27.26 606.18 SB 375 4.44 0.84 -27.14 606.08 SB 376 4.29 0.74 -27.37 605.98 SB 377 4.89 0.80 -27.23 605.88 SB 378 4.35 0.97 -27.25 605.78 SB 379 4.18 0.93 -27.23 605.68 288 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 380 4.66 0.85 -27.07 605.58 SB 381 4.37 0.93 -27.34 605.48 SB 382 4.90 0.87 -27.29 605.38 SB 383 4.64 0.93 -27.36 605.28 SB 384 4.52 1.01 -27.20 605.18 SB 385 4.59 0.86 -26.90 605.08 SB 386 4.48 0.81 -27.10 604.98 SB 387 4.48 0.87 -27.08 604.88 SB 388 4.50 0.95 -27.30 604.78 SB 389 4.20 0.92 -27.28 604.68 SB 390 4.08 0.87 -27.17 604.58 SB 391 4.56 0.68 -27.11 604.48 SB 392 4.43 0.87 -27.39 604.38 SB 393 4.69 0.78 -27.46 604.28 SB 394 4.17 0.91 -27.19 604.18 SB 395 4.04 1.02 -27.20 604.08 SB 396 4.76 0.95 -27.55 603.98 SB 397 4.62 0.97 -27.44 603.88 GAP 2 SB 398 4.38 0.64 -26.63 548.03 SB 399 4.76 0.63 -26.98 547.98 SB 400 4.72 0.61 -27.29 547.93 SB 401 4.79 0.72 -27.63 547.88 SB 402 4.06 0.67 -26.89 547.83 SB 403 4.66 0.56 -26.72 547.78 SB 404 4.07 0.71 -26.27 547.73 SB 405 4.17 0.75 -26.84 547.68 SB 406 4.28 0.67 -26.93 547.63 SB 407 4.26 0.73 -26.93 547.58 SB 408 4.27 0.72 -27.06 547.53 SB 409 4.18 0.65 -26.86 547.48 SB 410 4.31 0.80 -26.96 547.43 SB 411 4.19 0.61 -26.09 547.38 SB 412 4.16 0.72 -26.65 547.33 SB 413 4.31 0.79 -27.06 547.28 SB 414 4.10 0.93 -27.53 547.23 SB 415 4.25 0.99 -28.04 547.18 SB 416 4.36 0.69 -26.39 547.13 289 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 417 4.03 0.80 -27.04 547.08 SB 418 4.31 0.72 -27.02 547.03 SB 419 4.34 0.81 -27.66 546.98 SB 420 4.40 0.77 -26.83 546.93 SB 421 4.32 0.73 -27.64 546.88 SB 422 4.29 0.73 -27.00 546.83 SB 423 4.16 0.69 -26.51 546.78 SB 424 4.35 0.68 -26.65 546.73 SB 425 4.23 0.80 -27.07 546.68 SB 426 4.33 0.70 -26.60 546.63 SB 427 4.09 0.80 -26.82 546.58 SB 428 4.36 0.63 -26.75 546.53 SB 429 4.23 0.69 -26.44 546.48 SB 430 4.37 0.75 -27.05 546.43 SB 431 4.24 0.51 -26.64 546.38 SB 432 4.15 0.74 -27.00 546.33 SB 433 4.18 0.68 -26.29 546.28 SB 434 4.38 0.64 -27.04 546.23 SB 435 4.10 0.66 -26.77 546.18 SB 436 4.17 0.70 -26.98 546.13 SB 437 4.27 0.79 -27.45 546.08 SB 438 4.38 0.93 -27.36 546.03 SB 439 4.36 0.61 -26.39 545.98 SB 440 4.38 0.68 -26.87 545.93 SB 441 4.31 0.63 -26.86 545.88 SB 442 4.04 0.73 -27.26 545.83 SB 443 4.25 0.66 -27.08 545.78 SB 444 4.35 0.74 -27.03 545.73 SB 445 4.35 0.59 -27.20 545.68 SB 446 4.36 0.85 -27.11 545.63 SB 447 4.22 0.68 -26.61 545.58 SB 448 4.07 0.70 -27.04 545.53 SB 449 4.17 0.71 -26.86 545.48 SB 450 4.18 0.68 -27.59 545.43 SB 451 4.06 0.74 -27.41 545.38 SB 452 4.34 0.74 -27.11 545.33 SB 453 4.03 0.75 -27.30 545.28 SB 454 4.11 0.82 -27.75 545.23 290 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 455 4.23 0.75 -27.50 545.18 SB 456 4.15 0.77 -27.20 545.13 SB 457 4.10 0.68 -27.64 545.08 SB 458 4.25 0.75 -27.14 545.03 SB 459 4.18 0.43 -28.29 544.98 SB 460 4.25 0.67 -27.90 544.93 SB 461 4.08 0.80 -27.91 544.88 SB 462 4.37 0.76 -27.76 544.83 SB 463 4.25 0.76 -27.75 544.78 SB 464 4.34 0.89 -27.80 544.73 SB 465 4.31 0.67 -27.99 544.68 SB 466 4.14 0.74 -27.19 544.63 SB 467 4.29 0.75 -27.22 544.58 SB 468 4.25 0.72 -27.76 544.53 SB 469 4.30 0.77 -26.86 544.48 SB 470 4.01 0.62 -27.22 544.43 SB 471 4.19 0.72 -27.33 544.38 SB 472 4.07 0.73 -27.11 544.33 SB 473 4.06 0.68 -26.46 544.28 SB 474 4.32 0.73 -26.69 544.23 SB 475 4.34 0.75 -26.92 544.18 SB 476 4.11 0.85 -26.70 544.13 SB 477 4.36 0.85 -27.12 544.08 SB 478 4.07 0.74 -27.05 544.03 SB 479 4.28 0.87 -27.46 543.98 SB 480 4.28 0.80 -27.46 543.93 SB 481 4.30 0.83 -27.20 543.88 SB 482 4.38 0.64 -27.69 543.83 SB 483 4.03 0.77 -27.35 543.78 SB 484 4.18 0.76 -27.47 543.73 SB 485 4.26 0.71 -27.17 543.68 SB 486 4.12 0.75 -27.58 543.63 SB 487 4.23 0.70 -27.93 543.58 SB 488 4.37 0.72 -27.48 543.53 SB 489 4.04 0.70 -27.29 543.48 SB 490 4.39 0.65 -27.57 541.93 SB 491 4.18 0.76 -27.52 541.88 SB 492 4.06 0.69 -28.28 541.83 291 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 493 4.35 0.87 -27.71 541.78 SB 494 4.29 0.80 -27.75 541.73 SB 495 4.09 0.84 -28.10 541.68 SB 496 4.40 0.68 -27.53 541.63 SB 497 4.28 0.59 -28.11 541.58 SB 498 4.30 0.68 -28.13 541.53 SB 499 4.25 0.82 -28.08 541.48 SB 500 4.23 0.71 -27.82 541.43 SB 501 4.32 0.75 -27.87 541.38 SB 502 4.19 0.68 -27.64 541.33 SB 503 4.23 0.85 -28.05 541.28 SB 504 4.10 0.68 -27.93 541.23 SB 505 4.22 0.65 -27.62 541.18 SB 506 4.32 1.11 -28.03 541.13 SB 507 4.04 0.72 -28.04 541.08 SB 508 4.07 0.69 -27.49 541.03 SB 509 4.28 0.76 -27.89 540.98 SB 510 4.23 0.81 -27.81 540.93 SB 511 4.35 0.71 -27.14 540.88 SB 512 4.23 0.67 -27.37 540.83 SB 513 4.36 0.70 -27.62 540.78 SB 514 4.05 0.65 -27.53 540.73 SB 515 4.30 0.61 -27.32 540.68 SB 516 4.07 0.66 -27.24 540.63 SB 517 4.05 0.69 -27.50 540.58 SB 518 4.28 0.68 -27.22 540.53 SB 519 4.09 0.82 -27.91 540.48 SB 520 4.25 0.74 -27.47 540.43 SB 521 4.05 0.85 -27.76 540.38 SB 522 4.35 0.83 -27.73 540.33 SB 523 4.28 0.99 -27.98 540.28 SB 524 4.36 0.88 -27.46 540.23 SB 525 4.21 1.00 -27.58 540.18 SB 526 4.17 1.05 -28.00 540.13 SB 527 4.27 0.92 -27.95 540.08 SB 528 4.25 0.85 -27.87 540.03 SB 529 4.40 0.90 -27.31 539.98 SB 530 4.38 0.75 -27.61 539.93 292 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 531 4.38 0.94 -27.73 539.88 SB 532 4.26 0.87 -27.73 539.83 SB 533 4.15 0.83 -27.65 539.78 SB 534 4.20 0.74 -27.83 539.73 SB 535 4.34 1.01 -27.76 539.68 SB 536 4.31 0.91 -27.61 539.63 SB 537 4.22 0.79 -27.84 539.58 SB 538 4.35 0.75 -27.08 539.53 SB 539 4.14 0.94 -27.48 539.48 SB 540 4.04 0.85 -27.76 539.43 SB 541 4.26 0.86 -28.02 539.38 SB 542 4.15 0.64 -27.76 539.33 SB 543 4.23 1.02 -27.45 539.28 SB 544 4.17 0.80 -27.58 539.23 SB 545 4.21 0.80 -27.22 539.18 SB 546 4.11 0.68 -27.35 539.13 SB 547 4.05 0.93 -27.90 539.08 SB 548 4.37 0.95 -27.68 539.03 SB 549 4.08 0.68 -27.62 538.98 SB 550 4.22 0.75 -27.40 538.93 SB 551 4.21 0.89 -27.31 538.88 SB 552 4.09 0.67 -27.11 538.83 SB 553 4.28 0.78 -27.81 538.78 SB 554 4.37 0.76 -27.64 538.73 SB 555 4.30 0.79 -27.42 538.68 SB 556 4.36 1.01 -27.90 538.63 SB 557 4.25 1.07 -27.77 538.58 SB 558 4.26 1.12 -27.94 538.53 SB 559 4.22 1.09 -27.91 538.48 SB 560 4.06 0.88 -27.77 538.43 SB 561 4.27 0.95 -27.77 538.38 SB 562 4.31 0.94 -27.38 538.33 SB 563 4.27 1.04 -27.54 538.28 SB 564 4.12 0.94 -27.80 538.23 SB 565 4.17 0.99 -27.52 538.18 SB 566 4.20 0.85 -28.05 538.13 SB 567 4.27 0.98 -27.89 535.83 SB 568 4.31 1.25 -27.94 535.78 293 Northern Alaska South Barrow #3 Core data core ID Amt. (Mg) TOC% \u00CE\u00B413Cadj depth (m) SB 569 4.40 0.71 -28.14 535.73 SB 570 4.30 0.90 -27.93 535.68 SB 571 4.17 1.03 -27.79 535.63 SB 572 4.16 1.02 -27.87 535.58 SB 573 4.20 1.06 -27.82 535.53 SB 574 4.04 1.13 -27.87 535.48 SB 575 4.24 0.79 -27.70 535.43 SB 576 4.08 0.38 -28.75 535.38 SB 577 4.22 1.01 -27.67 535.33 SB 578 4.11 0.90 -27.79 535.28 SB 579 4.19 0.93 -27.45 535.23 SB 580 4.25 0.90 -27.60 535.18 SB 581 4.31 0.77 -28.04 535.13 SB 582 4.18 0.82 -27.74 535.08 SB 583 4.28 0.76 -27.59 535.03 SB 584 4.39 0.81 -27.73 534.98 SB 585 4.32 0.77 -27.35 534.93 SB 586 4.11 0.76 -27.48 534.88 SB 587 4.33 0.79 -27.96 534.83 SB 588 4.13 0.99 -27.61 534.78 SB 589 4.38 0.84 -27.46 534.73 SB 590 4.20 0.82 -28.02 534.68 SB 591 4.34 0.77 -28.18 534.63 SB 592 4.04 0.82 -27.80 534.58 SB 593 4.36 0.83 -27.59 534.53 SB 594 4.25 0.88 -28.06 534.48 SB 595 4.26 0.89 -27.75 534.43 SB 596 4.11 0.97 -27.91 534.38 SB 597 4.28 0.76 -27.97 534.33 SB 598 4.35 0.67 -27.46 534.28 SB 599 4.01 0.82 -27.52 534.23 SB 600 4.25 0.87 -27.94 534.18 SB 601 4.15 0.81 -27.91 534.13 SB 602 4.12 0.69 -27.75 534.08 SB 603 4.31 0.77 -27.44 534.03 "@en . "Thesis/Dissertation"@en . "2013-05"@en . "10.14288/1.0073839"@en . "eng"@en . "Geological Sciences"@en . "Vancouver : University of British Columbia Library"@en . "University of British Columbia"@en . "Attribution-NonCommercial-NoDerivatives 4.0 International"@en . "http://creativecommons.org/licenses/by-nc-nd/4.0/"@en . "Graduate"@en . "Pliensbachian\u00E2\u0080\u0093Toarcian (Early Jurassic) extinction in western North America"@en . "Text"@en . "http://hdl.handle.net/2429/44233"@en .