"Arts, Faculty of"@en . "Anthropology, Department of"@en . "DSpace"@en . "UBCV"@en . "McKechnie, Iain Mitchell Patrick"@en . "2014-06-30T00:00:00Z"@en . "2013"@en . "Doctor of Philosophy - PhD"@en . "University of British Columbia"@en . "This dissertation examines multiple scales of Indigenous history on the Northwest Coast from the disciplinary perspective of archaeology. I focus on cultural lifeways archaeologically represented in two key domains of human existence: food and settlement. The dissertation consists of six individual case studies that demonstrate the utility of applying multiple spatial and temporal scales to refine archaeological understanding of cultural and historical variability on the Northwest Coast over the Mid-to-Late Holocene (ca. 5,000-200 BP). The first of three regionally scaled analyses presents a coast-wide examination of fisheries data indicating that Pacific herring (Clupea pallasii) exhibit a pervasive and previously under-recognized importance in Northwest Coast Indigenous subsistence practices. Next, I use zooarchaeological data from the southern British Columbia coast to identify a pattern of regional coherence in Coast Salish and Nuu-chah-nulth hunting traditions reflecting the scale of intergenerational cultural practice. The third study re-calibrates the settlement history of a small and historically significant locality in Coast Tsimshian territory (Prince Rupert Harbour) to clarify the temporal resolution of existing radiocarbon datasets and test inferences about social and political change. Following this regional exploration of scale, I document site-specific temporal variability in archaeological fisheries data from a Nuu-chah-nulth \u00E2\u0080\u0098big-house\u00E2\u0080\u0099 reflecting climatic and socio-economic change. I examine Indigenous oral histories and archaeological datasets to evaluate these parallel records of settlement in the neighbouring territory of an autonomous Nuu-chah-nulth polity before and during the occupation of a large defensive fortress. Finally, I demonstrate how everyday foodways are archaeologically expressed and reflect ecological differences and active management strategies within several spatially associated sites over millennial timescales. These linked case studies offer new clarity into long-standing debates concerning archaeologically relevant scales of cultural-historical variability on the NWC. They collectively demonstrate an enduring regional and temporal coherence for key aspects of indigenous resource use and settlement and a historical dynamism at finer scales. I argue this has cultural, historical, and archaeological significance as well as relevance for contemporary understandings of the Northwest Coast environment. I conclude that a focus on the pervasive aspects of the everyday over millennia offers insight into individual actions across broader patterns of history."@en . "https://circle.library.ubc.ca/rest/handle/2429/45670?expand=metadata"@en . "AN ARCHAEOLOGY OF FOOD AND SETTLEMENT ON THE NORTHWEST COAST by Iain Mitchell Patrick McKechnie B.A., University of California, Santa Cruz, 1999 M.A., Simon Fraser University, 2005 A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in THE FACULTY OF GRADUATE AND POSTDOCTORAL STUDIES (Anthropology) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) December 2013 \u00C2\u00A9 Iain Mitchell Patrick McKechnie, 2013 \u00E2\u0080\u0093 ii \u00E2\u0080\u0093 Abstract This dissertation examines multiple scales of Indigenous history on the Northwest Coast from the disciplinary perspective of archaeology. I focus on cultural lifeways archaeologically represented in two key domains of human existence: food and settlement. The dissertation consists of six individual case studies that demonstrate the utility of applying multiple spatial and temporal scales to refine archaeological understanding of cultural and historical variability on the Northwest Coast over the Mid-to-Late Holocene (ca. 5,000-200 BP). The first of three regionally scaled analyses presents a coast-wide examination of fisheries data indicating that Pacific herring (Clupea pallasii) exhibit a pervasive and previously under-recognized importance in Northwest Coast Indigenous subsistence practices. Next, I use zooarchaeological data from the southern British Columbia coast to identify a pattern of regional coherence in Coast Salish and Nuu-chah-nulth hunting traditions reflecting the scale of intergenerational cultural practice. The third study re-calibrates the settlement history of a small and historically significant locality in Coast Tsimshian territory (Prince Rupert Harbour) to clarify the temporal resolution of existing radiocarbon datasets and test inferences about social and political change. Following this regional exploration of scale, I document site-specific temporal variability in archaeological fisheries data from a Nuu-chah-nulth \u00E2\u0080\u0098big-house\u00E2\u0080\u0099 reflecting climatic and socio-economic change. I examine Indigenous oral histories and archaeological datasets to evaluate these parallel records of settlement in the neighbouring territory of an autonomous Nuu-chah-nulth polity before and during the occupation of a large defensive fortress. Finally, I demonstrate how everyday foodways are archaeologically expressed and reflect ecological differences and active management strategies within several spatially associated sites over millennial \u00E2\u0080\u0093 iii \u00E2\u0080\u0093 timescales. These linked case studies offer new clarity into long-standing debates concerning archaeologically relevant scales of cultural-historical variability on the NWC. They collectively demonstrate an enduring regional and temporal coherence for key aspects of indigenous resource use and settlement and a historical dynamism at finer scales. I argue this has cultural, historical, and archaeological significance as well as relevance for contemporary understandings of the Northwest Coast environment. I conclude that a focus on the pervasive aspects of the everyday over millennia offers insight into individual actions across broader patterns of history. \u00E2\u0080\u0093 iv \u00E2\u0080\u0093 Preface Chapter 2 is a multi-authored manuscript that was submitted for publication on September 16, 2013 and accepted for publication on December 12, 2013. I am the principal author and analyst on this manuscript, which has been co-written primarily with Dana Lepofsky. Other authors include Madonna L. Moss, Virginia L. Butler, Trevor J. Orchard, Gary Coupland, Fredrick Foster, Megan Caldwell, and Ken Lertzman. Aubrey Cannon has generously provided unpublished data and commented on multiple versions of this manuscript. Andrew Martindale provided column and auger sample fauna for osteological identification by the author. Chapter 3 has been previously published as: McKechnie, Iain and Rebecca J. Wigen (2011) In Human Impacts on Seals, Sea Lions, and Sea Otters: Integrating Archaeology and Ecology in the Northeast Pacific, edited by Todd J. Braje and Torben C. Rick, pp. 129\u00E2\u0080\u0093166. University of California Press, Berkeley. \u00C2\u00A9 2011 by the Regents of the University of California. I designed the study, compiled data, completed quantitative and cartographic analyses, wrote much of the chapter and worked with co-author Rebecca Wigen on developing concepts, revisions and edits. Chapter 4 is a manuscript co-authored with Morley Eldridge. I conducted background research and designed and conducted calibration analyses, and co-wrote the paper and conducted the bulk of the revisions on multiple drafts over a period of four years. An initial version of this chapter was originally presented in 2008 at the Society for American Archaeology meetings in Vancouver and subsequent drafts have been heavily revised in correspondence with Andrew Martindale and Susan Marsden. Chapter 5 has been previously published as: McKechnie, Iain (2012) Zooarchaeological Analysis of the Indigenous Fishery at the Huu7ii Big House and Back Terrace, Huu-ay-aht Territory, Southwestern Vancouver Island. In Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound, by Alan D. McMillan and Denis E. St. Claire, pp. 154\u00E2\u0080\u0093186. \u00C2\u00A9 2012 Archaeology Press, Simon Fraser University, Burnaby, BC. This manuscript was conducted in the service of the Huu7ii Archaeological Project with the support of the Huu-ay-aht First Nation and the 2006 University of Victoria archaeological field school. I took part in the primary fieldwork, field based sampling, coordinated sorting of samples and conducted osteological identification and am the sole author of this manuscript, which was written in 2011 and completed in the Spring of 2012. \u00E2\u0080\u0093 v \u00E2\u0080\u0093 Table of Contents Abstract .................................................................................................................................... ii\t \u00C2\u00A0Preface ..................................................................................................................................... iv\t \u00C2\u00A0Table of Contents .....................................................................................................................v\t \u00C2\u00A0List of Tables ........................................................................................................................ viii\t \u00C2\u00A0List of Figures ......................................................................................................................... ix\t \u00C2\u00A0List of Abbreviations ........................................................................................................... xiv\t \u00C2\u00A0Glossary of Technical Terms ................................................................................................ xv\t \u00C2\u00A0Acknowledgements .............................................................................................................. xvi\t \u00C2\u00A0Dedication ........................................................................................................................... xviii\t \u00C2\u00A0Chapter 1.\t \u00C2\u00A0 Archaeological Scales of Analysis on the Northwest Coast ........................ 1\t \u00C2\u00A0Introduction ....................................................................................................................... 1\t \u00C2\u00A0Theoretical Framework ..................................................................................................... 2\t \u00C2\u00A0Food and Place as Multi-scalar ......................................................................................... 6\t \u00C2\u00A0Dissertation Structure ...................................................................................................... 11\t \u00C2\u00A0Conclusion ...................................................................................................................... 18\t \u00C2\u00A0Chapter 2.\t \u00C2\u00A0 Archaeological Data Provide Alternative Hypotheses on Pacific Herring (Clupea pallasii) Distribution, Abundance, and Variability on the Northwest Coast 19\t \u00C2\u00A0Introduction ..................................................................................................................... 19\t \u00C2\u00A0Material and Methods ..................................................................................................... 26\t \u00C2\u00A0Results ............................................................................................................................. 31\t \u00C2\u00A0Discussion ....................................................................................................................... 41\t \u00C2\u00A0Conclusion ...................................................................................................................... 49\t \u00C2\u00A0Chapter 3.\t \u00C2\u00A0 Towards a Historical Ecology of Pinniped and Sea Otter Hunting Traditions on the Coast of Southern British Columbia ................................................ 51\t \u00C2\u00A0\u00E2\u0080\u0093 vi \u00E2\u0080\u0093 Introduction ..................................................................................................................... 51\t \u00C2\u00A0Context ............................................................................................................................ 52\t \u00C2\u00A0Methods........................................................................................................................... 59\t \u00C2\u00A0Results ............................................................................................................................. 64\t \u00C2\u00A0Discussion ....................................................................................................................... 80\t \u00C2\u00A0Conclusions ..................................................................................................................... 87\t \u00C2\u00A0Chapter 4.\t \u00C2\u00A0 Re-Calibrating Archaeological Chronologies on the Northern Northwest Coast: Radiocarbon Data from Prince Rupert Harbour .............................................. 91\t \u00C2\u00A0The Importance of Calibration in Radiocarbon Dating .................................................. 92\t \u00C2\u00A0Calibrating Radiocarbon Age Estimates for Marine Samples ........................................ 99\t \u00C2\u00A0Variable Reporting of Radiocarbon Dates and Calibration Results ............................. 106\t \u00C2\u00A0Summed Probability as an Analytical Method ............................................................. 110\t \u00C2\u00A0Recalibrating the Radiocarbon Record for Prince Rupert Harbour .............................. 113\t \u00C2\u00A0Calibration Issues in Village Abandonment Dates from Shell ..................................... 122\t \u00C2\u00A0Recalibrating Radiocarbon Dates on Human Remains from Prince Rupert Harbour .. 134\t \u00C2\u00A0Interpretive Implications of Recalibrations on Individual Dates .................................. 138\t \u00C2\u00A0Conclusions ................................................................................................................... 141\t \u00C2\u00A0Chapter 5.\t \u00C2\u00A0 Zooarchaeological Analysis of the Indigenous Fishery at the Huu7ii Big House & Back Terrace, Huu-ay-aht Territory, Southwestern Vancouver Island .... 144\t \u00C2\u00A0Methods......................................................................................................................... 145\t \u00C2\u00A0Results ........................................................................................................................... 157\t \u00C2\u00A0Discussion and Interpretation ....................................................................................... 192\t \u00C2\u00A0Conclusions ................................................................................................................... 201\t \u00C2\u00A0Chapter 6.\t \u00C2\u00A0 Indigenous Oral History & Settlement Archaeology in the Broken Group Islands, Western Vancouver Island .............................................................................. 203\t \u00C2\u00A0Indigenous Oral History and the Archaeological Record ............................................. 203\t \u00C2\u00A0Translating Between Archaeological and Oral Historical Information ........................ 205\t \u00C2\u00A0Objectives ..................................................................................................................... 206\t \u00C2\u00A0\u00E2\u0080\u0093 vii \u00E2\u0080\u0093 Analysis of Archaeological and Oral Historical Sequences ......................................... 207\t \u00C2\u00A0Historical Background .................................................................................................. 208\t \u00C2\u00A0The Broken Group Study Area ..................................................................................... 211\t \u00C2\u00A0Settlement History of Ts\u00E2\u0080\u0099ishaa, The Ts\u00E2\u0080\u0099ishaa7ath Origin Village ................................ 221\t \u00C2\u00A0Oral Historical Accounts of the Maktl7ii7ath Local Group Territory .......................... 224\t \u00C2\u00A0Archaeological Case Studies in Maktl7ii7ath Local Group Territory .......................... 228\t \u00C2\u00A0Discussion ..................................................................................................................... 248\t \u00C2\u00A0Conclusion .................................................................................................................... 253\t \u00C2\u00A0Chapter 7.\t \u00C2\u00A0 Zooarchaeological Scales of Variability on the Northwest Coast: A Barkley Sound Case Study ............................................................................................. 257\t \u00C2\u00A0Introduction ................................................................................................................... 257\t \u00C2\u00A0Theoretical and Methodological Context ...................................................................... 258\t \u00C2\u00A0Analysis of Scale and Variability ................................................................................. 262\t \u00C2\u00A0Methods......................................................................................................................... 266\t \u00C2\u00A0Results ........................................................................................................................... 267\t \u00C2\u00A0Discussion ..................................................................................................................... 278\t \u00C2\u00A0Conclusion .................................................................................................................... 284\t \u00C2\u00A0Chapter 8.\t \u00C2\u00A0 Conclusion: Significance, Limitations, and Future Directions .............. 287\t \u00C2\u00A0Limitations .................................................................................................................... 298\t \u00C2\u00A0Variation and Scale into the Future .............................................................................. 300\t \u00C2\u00A0References .............................................................................................................................303\t \u00C2\u00A0Appendices For Chapter 2 ..................................................................................................365\t \u00C2\u00A0 \u00E2\u0080\u0093 viii \u00E2\u0080\u0093 List of Tables Table 2.1 Archaeological data for Pacific herring for all sites (N = 171)\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6..366 Table 2.2 Individual assemblage data for the 50 sites with within site\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6......376 Table 2.3 Distance between archaeological sites and herring spawning locations\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6\u00E2\u0080\u00A6379 Table 3.1 Ecological characteristics for pinnipeds and sea otters in British Columbia. ......... 61\t \u00C2\u00A0Table 3.2 Site Names, Locations, References, and Number of Identified Specimens. .......... 73\t \u00C2\u00A0Table 3.3 Sex Distribution of Steller Sea Lion Remains from Selected Sites. ....................... 79\t \u00C2\u00A0Table 4.1 Paired marine and terrestrial radiocarbon dates from Prince Rupert Harbour. .... 119\t \u00C2\u00A0Table 4.2 Recalibrated radiocarbon dates on shell from uppermost shell midden layers ..... 119\t \u00C2\u00A0Table 4.3 Carbon isotope values. .......................................................................................... 131\t \u00C2\u00A0Table 4.4 Recalibrated radiocarbon dates on human bone collagen. .................................... 132\t \u00C2\u00A0Table 5.1 Column samples containing identified fish remains screened .............................. 155\t \u00C2\u00A0Table 5.2. List of identified taxa (NISP) in the column sample assemblage. ....................... 155\t \u00C2\u00A0Table 6.1. The 15 largest shell midden sites in the Broken Group Islands ........................... 219\t \u00C2\u00A0Table 6.2. Lineage household names (ushtakimilh) names at the village of Maktl7ii .......... 226\t \u00C2\u00A0Table 6.3 Radiocarbon dates from archaeological sites in the Maktl7ii7ath study area. ...... 255\t \u00C2\u00A0Table 7.1 Examined vertebrate remains from the seven study sites ..................................... 269\t \u00C2\u00A0Table 7.2 Examined invertebrate remains from the seven study sites .................................. 270\t \u00C2\u00A0Table 7.3. Mean lengths and standard deviations in cm for all measured fish ..................... 273\t \u00C2\u00A0 \u00E2\u0080\u0093 ix \u00E2\u0080\u0093 List of Figures Figure 1.1 Map of the Northwest Coast .................................................................................. 12\t \u00C2\u00A0Figure 1.2 Temporal scales of analysis. .................................................................................. 13\t \u00C2\u00A0Figure 2.1. Percent abundance of herring bones in archaeological sites ................................ 30 Figure 2.3. Boxplots showing percent abundance of herring bones ....................................... 32\t \u00C2\u00A0Figure 2.4. Percent abundance of herring bones in archaeological sites by sub-regions. ....... 32\t \u00C2\u00A0Figure 2.5. Percent abundance of herring bones over time in sites ........................................ 35\t \u00C2\u00A0Figure 2.6. Mean rank order and abundance of archaeological herring bones ....................... 38\t \u00C2\u00A0Figure 2.7. Boxplot showing variability in the archaeological abundance of herring bones. . 39\t \u00C2\u00A0Figure 2.8. Schematic figure representing three alternative hypotheses regarding the relationship between modern and archaeological abundance and variability of herring populations. ............................................................................................................................. 45\t \u00C2\u00A0Figure 3.1. Map of the study area showing. ............................................................................ 64\t \u00C2\u00A0Figure 3.2. Relative composition of marine and terrestrial mammals (%NISP) for all sites. 69 Figure 3.4. Relative abundance of marine versus terrestrial mammals for sites dating to within the four examined periods. .......................................................................................... 95 Figure 4.2. Example of calibration effects on a radiocarbon date with a large error range. ... 97\t \u00C2\u00A0Figure 4.3. Map of the northern Northwest Coast showing .................................................. 102\t \u00C2\u00A0Figure 4.4. Schematic depicting the steps required for radiocarbon calibration. ................. 105\t \u00C2\u00A0Figure 4.5. An example of three typical radiocarbon dates on terrestrial charcoal .............. 111\t \u00C2\u00A0Figure 4.6. Local marine reservoir (Delta-R) estimates over the Holocene ......................... 118\t \u00C2\u00A0Figure 4.7. Village abandonment data obtained from Archer 2001 and Archer 1992 .......... 125\t \u00C2\u00A0\u00E2\u0080\u0093 x \u00E2\u0080\u0093 Figure 4.8 Comparison of different calibration criteria applied to radiocarbon dates on human remains from Prince Rupert Harbour .................................................................................... 135\t \u00C2\u00A0Figure 5.1 Excavating column samples from the sidewall of excavation units .................... 146\t \u00C2\u00A0Figure 5.2. Plan view showing column samples locations recovered from House 1. ........... 147\t \u00C2\u00A0Figure 5.3. Perspective view of the Huu7ii village looking west. ........................................ 148\t \u00C2\u00A0Figure 5.4. Photo of column sample taken from the north wall of the back terrace unit . .... 149 Figure 5.6. Comparison between the relative abundance of fish remains in the column sample assemblage ............................................................................................................................ 159\t \u00C2\u00A0Figure 5.7. Taxonomic composition of the entire column sample fish assemblage. ............ 161\t \u00C2\u00A0Figure 5.8. Relative abundance over time for the ten most numerous fish taxa. .................. 168\t \u00C2\u00A0Figure 5.9. The number of specimens (NISP) and individuals (MNI) per cubic metre ....... 168\t \u00C2\u00A0Figure 5.10. Temporal trends (left to right) in the abundance of four important fish taxa i . 173\t \u00C2\u00A0Figure 5.11. Herring and anchovy abundance in individual column sample levels ............. 176\t \u00C2\u00A0Figure 5.12. NISP and MNI data for the entire column sample assemblage. ....................... 177\t \u00C2\u00A0Figure 5.13. Histograms showing fish length estimates. ...................................................... 180\t \u00C2\u00A0Figure 5.14. Measured salmon vertebrae from the column sample assemblage .................. 181\t \u00C2\u00A0Figure 5.15. The identification rate for fish, mammals and birds in the column sample assemblage. ........................................................................................................................... 183\t \u00C2\u00A0Figure 5.16. Identification rate for fish remains in the 12 examined column sample assemblages. .......................................................................................................................... 183\t \u00C2\u00A0Figure 5.17. Estimated number of individual fish per litre in three deposits at Huu7ii.. ...... 184\t \u00C2\u00A0Figure 5.18. Collectors curve for taxonomic richness in the column sample assemblage. .. 185\t \u00C2\u00A0Figure 5.19. Taxonomic richness of the House 1 versus the back terrace assemblages. ...... 186\t \u00C2\u00A0\u00E2\u0080\u0093 xi \u00E2\u0080\u0093 Figure 5.20. Scatterplot showing the non-linear relationship between the weight of bone and the weight of shell in individual column sample levels. ....................................................... 187\t \u00C2\u00A0Figure 5.21. Burning and other modifications to bone specimens in the column sample assemblage by fish, mammal, and bird categories. ............................................................... 190\t \u00C2\u00A0Figure 5.22. Relative percent of cranial vertebral and caudal elements for the 10-most numerous fish in the column sample assemblage. ................................................................ 191\t \u00C2\u00A0Figure 6.1 Territorially bounded local group areas in the Broken Group Islands during the 18th century as described in oral historical accounts synthesized by St. Claire (1991). ...... 213\t \u00C2\u00A0Figure 6.2. Alex Thomas (left), Frank Williams and his daughter (centre) in Port Alberni in 1914. Edward Sapir (right) pictured in New Haven in 1934 ................................................ 215\t \u00C2\u00A0Figure 6.3. Tseshaht place names as detailed in oral historical. ........................................... 217\t \u00C2\u00A0Figure 6.4 Recorded archaeological sites in the Broken Group and Sechart Channel. ........ 218\t \u00C2\u00A0Figure 6.5. The size and number of shell midden sites in different quadrants of the Broken Group Islands.. ...................................................................................................................... 220\t \u00C2\u00A0Figure 6.6. Site map of the village of Ts'ishaa (204 and 205T) showing oral historically named locations of household lineages (ushtakimilh) .......................................................... 224\t \u00C2\u00A0Figure 6.7. Overview of the Maktl7ii7ath study area with individual study sites ................ 225\t \u00C2\u00A0Figure 6.8. Oblique perspective view of Maktl7ii looking north. ......................................... 230\t \u00C2\u00A0Figure 6.9. Calibrated Radiocarbon dates from Maktl7ii (site 206T) at Wouwer Island. .... 231\t \u00C2\u00A0Figure 6.10. Terraced midden with house platforms at site 206T, Outer Wouwer Island. ... 232\t \u00C2\u00A0Figure 6.11. Oblique perspective view of Huumuuwaa. ...................................................... 234\t \u00C2\u00A0Figure 6.12. Calibrated radiocarbon dates from Huumuuwaa (site 304T). .......................... 234\t \u00C2\u00A0Figure 6.13. Oblique view of the archaeological deposits at the site of Shiwitis (82T).. ..... 237\t \u00C2\u00A0\u00E2\u0080\u0093 xii \u00E2\u0080\u0093 Figure 6.14. Calibrated radiocarbon dates for Shiwitis, 82T. ............................................... 237\t \u00C2\u00A0Figure 6.15. Southwest facing view of Ch'ituukwachisht. .................................................... 238\t \u00C2\u00A0Figure 6.16. Calibrated radiocarbon dates on charcoal from deposits at upper Huts\u00E2\u0080\u0099atswilh (site 129T, top), N. Cree Island (site 132T, middle), and S. Cree (site 131T, bottom).. ...... 239\t \u00C2\u00A0Figure 6.17. Perspective view of a topographic model of Huts\u00E2\u0080\u0099atswilh s ............................ 241\t \u00C2\u00A0Figure 6.18. View from upper Huts\u00E2\u0080\u0099atswilh (site 129T) looking west. ................................ 242\t \u00C2\u00A0Figure 6.19. Dates from lower Huts\u00E2\u0080\u0099atswilh showing calibrated ages from core samples. .. 244\t \u00C2\u00A0Figure 6.20 Map of the lower village (83T) and upper defensive site (129T). ..................... 247\t \u00C2\u00A0Figure 6.21 Dates on cultural deposits on upper Dicebox .................................................... 248\t \u00C2\u00A0Figure 6.22. Settlement history in Maktl7ii7ath Local Group territory ................................ 250\t \u00C2\u00A0Figure 7.1. Map of the Maktl7ii7ath study area in the Broken Group Islands. .................... 261\t \u00C2\u00A0Figure 7.2. Arial overview of the Maktl7ii7ath study area with individual study sites. ....... 261\t \u00C2\u00A0Figure 7.3 Proportion of fish, birds and mammals in the examined assemblages. ............... 269\t \u00C2\u00A0Figure 7.4 Proportion of fish taxa by number of identified speciemens (NISP) .................. 271\t \u00C2\u00A0Figure 7.5. Proportion of shellfish by weight from the five study sites ................................ 271\t \u00C2\u00A0Figure 7.6. Distribution of fish lengths for three ubiquitous and abundant fish taxa ........... 273\t \u00C2\u00A0Figure 7.7. Identified fish remains from the five study sites. ............................................... 275\t \u00C2\u00A0Figure 7.8. Proportion of shellfish from the five study sites. ............................................... 276\t \u00C2\u00A0Figure 7.9. Four measures of abundance applied across the five study sites for anchovy (left) and clam (right).. ................................................................................................................... 277\t \u00C2\u00A0Figure 7.10. Ancient DNA identifications of rockfish in the study area (from Rodrigues, McKechnie, and Yang 2013). Size of pie chart is scaled to sample size. ............................. 280\t \u00C2\u00A0Figure 7.11. Anthropogenically altered beach at Shiwitis (Site 82T).. ................................. 282\t \u00C2\u00A0\u00E2\u0080\u0093 xiii \u00E2\u0080\u0093 Figure 8.1 Illustration by John Webber depicting the inside a portion of the house of Mowachaht Muchalaht Chief Maquinna. ............................................................................. 291\t \u00C2\u00A0Figure 8.2 Recorded shell midden sites on the British Columbia Coast. ............................. 302\t \u00C2\u00A0 \u00E2\u0080\u0093 xiv \u00E2\u0080\u0093 List of Abbreviations NWC \u00E2\u0080\u0093 Northwest Coast NISP \u00E2\u0080\u0093 Number of Identified Specimens MNI \u00E2\u0080\u0093 Minimum Number of Individuals DFO \u00E2\u0080\u0093 Department of Fisheries and Oceans Canada TEK \u00E2\u0080\u0093 Traditional Ecological Knowledge LEK \u00E2\u0080\u0093 Local Ecological Knowledge \u00E2\u0080\u0093 xv \u00E2\u0080\u0093 Glossary of Technical Terms Number of Identified Specimens \u00E2\u0080\u0093 Used to describe the total number of skeletal specimens which have been confidently identified to at least family, genus or species level (i.e., not including \u00E2\u0080\u0098unidentified\u00E2\u0080\u0099 specimens). This is not the same as miscellaneous unidentified bone specimens which are sometimes referred to as simply Number of Specimens (NSP). Specimen \u00E2\u0080\u0093 skeletal fragment. The term \u00E2\u0080\u0098specimen\u00E2\u0080\u0099 is used rather than \u00E2\u0080\u0098bone\u00E2\u0080\u0099 because skeletal elements are often broken into multiple pieces and can therefore be counted more than once in a zooarchaeological assemblage. Taxa - a term used to describe different animals (or plants) that are not necessarily all of the same taxonomic category. For instance \u00E2\u0080\u0098Northern anchovy\u00E2\u0080\u0099 is a species while \u00E2\u0080\u0098rockfish\u00E2\u0080\u0099 (Sebastes spp.) is a genus designation (following Linnaean taxonomy). Relative abundance (relative frequency, %NISP or % weight) \u00E2\u0080\u0093 simply the percentage of a particular item (e.g., herring, clam weight) relative to all other specimens within the same category (e.g., identified fish remains, identified shell fragments). Rank order abundance (%NISP) \u00E2\u0080\u0093 the relative rank order of a particular item (e.g., herring) in comparison to all other specimens within the same category (e.g., identified fish remains, identified shell fragments). Ubiquity \u00E2\u0080\u0093 Frequency of occurrence or the percentage of contexts in which a certain item is present versus absent. For example, herring would be \u00E2\u0080\u0098ubiquitous\u00E2\u0080\u0099 if it was present in every level of every excavation unit at the site. Column sample \u00E2\u0080\u0093 Small (typically 10x10 cm) vertical \u00E2\u0080\u0098column\u00E2\u0080\u0099 of archaeological sediment excavated from the sidewalls of an excavation area. Archaeological \u00E2\u0080\u0098site\u00E2\u0080\u0099 \u00E2\u0080\u0093 An archaeological term describing a place with material and human altered sediments indicative of past human use and settlement. In this dissertation, I use the terms site, shell midden, and settlement interchangeably. Shell midden \u00E2\u0080\u0093 A widely used archaeological term to describe places in coastal, near-coastal, and aquatic areas containing anthropogenic (human created) sediments. On the Northwest Coast, shell middens are typically composed of varying quantities of shell, bone, fire altered rocks (e.g., boiling stones), artifacts, charcoal, greasy silt, decomposed plant and other organic remains, architectural features, and a host of other cultural materials, including isolated and multiple human interments. Shell midden sediments can range widely in volume from small instantaneously deposited lenses to massive deposits enveloping entire landforms and spanning broad periods of time (ca. 100-10,000 years). Midden \u00E2\u0080\u0093 a term that \u00E2\u0080\u009Chas its roots in the Scandinavian languages, meaning material that accumulates about a dwelling place\u00E2\u0080\u009D (Stein 1996:638). \u00E2\u0080\u0093 xvi \u00E2\u0080\u0093 Acknowledgements This dissertation was completed only because a number of supportive, inquisitive, positive, patient, and hard working people entrusted me with their time, equipment, and funding. I am *very* grateful to have had this opportunity to think about food with good people and at several amazing places. I cannot thank my supervisors Andrew Martindale and Michael Blake enough for their support and mentoring over the years, for enabling and expanding my research horizons, and making time at massively important moments. Thanks to Andrew for finding ways to support me in the field, in the lab, and in energizing discussions with scrap paper and coffee. Thanks to Mike for his sagely wisdom, timely advice, and for sharing his expertise in archaeological cartography and visualization which has been useful in each chapter. Thanks also to committee members Bruce Miller and Aubrey Cannon for inspiration, insight and support in numerous timely ways over many years, and to other examiners for thoughtful critique and constructive feedback (Robert Losey, David Pokotylo, and Greg Crutsinger). I have been especially fortunate to have had fantastic feedback from a number of co-authors on Chapters 2, 3, and 4. I owe Dana Lepofsky and Madonna Moss an inordinate amount of gratitude for their support and efforts in navigating the process of writing Chapter 2. I thank Dana in particular for her boundless energy, for bringing me into the Herring School at SFU, and timely encouragement during a re-think of the structure of this dissertation. I am grateful to Ken Lertzman, Virginia Butler, and Trevor Orchard for productive correspondence, and to Gary Coupland, Megan Caldwell and Fred Foster for contributing and organizing data. Rebecca Wigen at UVic and Morley Eldridge at Millennia were amazing collaborators during numerous revisions in Chapters 3 and 4 respectively. Thank to editors Andrew Martindale, Susan Marsden, Duncan McLaren, Torben Rick & Todd Braje. Alan McMillan and Denis St. Claire are owed large debts of my gratitude for their support of this dissertation throughout its several stages and for being great mentors (and editors). Denis St. Claire\u00E2\u0080\u0099s boundless energy and ability to coordinate boats and equipment has been particularly vital to the Hiikwis Archaeological Project, to which this dissertation contributes. Alan and Denis were also the directors of the Huu7ii Archaeological Project \u00E2\u0080\u0093 generously supported by Huu-ay-aht First Nations. I am especially grateful to Tseshaht and Huu-ay-aht First Nations who provided me with the privilege of conducting fieldwork in their territories. Particular thanks to Les Sam, Boyd Gallic, Darrell Ross, and Chris Robinson at Tseshaht for their support of the Hiikwis Archaeological Project and to Barry Watts, Jordan Dick, Hank Gus, Jeff Gallic, and Cody Robinson for their help in the field. I especially thank Wanda Robinson for her amazing organizational and culinary ability at the Nettle Island field cabin. I thank Robert Dennis at Huu-ay-aht for enabling fieldwork at Huu7ii and Judy Johnson, Henry Williams, Arthur Peters, Gabe (Hip) Williams, and Alec Frank and numerous others for help in the field, and many volunteer UVic \u00E2\u0080\u0098Rockwashers\u00E2\u0080\u0099 for help in processing column samples in the lab. Ian Sellers, Ted Knowles, Pete Dady, and Ian Sumpter are most thoroughly thanked for their patient persistence for volunteering long days of coring, augering, and mapping in the Broken Group in 2008, 2009, & 2011. Thanks also to Stephanie Sketchley for her organizational perseverance, to Bryn Letham, Alex Clark, Kelsey McLean, Butch Neilson, Mike Blake, Andrew Martindale, Ben Nelson, Audrey Dallimore, Quentin Mackie, Daryl Fedje, Adam Love, Rodney Steadman, Jacob Earnshaw, Byron Malloy, Byron Fry, Tim Sumpter, Russ Markel, and Nicole Smith for help in the field. At Parks, I thank Karen Haugen, Jen Yakimishin, Arlene Armstrong, Darren Salisbury, Mike Collyer, and Kim St. Claire (nee Seward-Hannam) and to Gwyn Langemann in Calgary. Thanks to Karey and Henk at Sechart Lodge for use of their showers and grocery delivery. Thanks to John Southon for radiocarbon dates and to Paula Reimer and Maarten Blaauw for answering 14C queries. I am further thankful to Dongya Yang, Camilla Speller, and Antonia Rodrigues for their Ancient DNA analyses. \u00E2\u0080\u0093 xvii \u00E2\u0080\u0093 Patricia Ormerod, Wayne Point, Mike Blake and Andrew Martindale are thanked for facilitating the use of key UBC Lab of Archaeology mapping/coring/augering equipment and enabling boat and van transportation which was absolutely critical to this project. I owe massive thanks to Steve Weisman for hundreds of hours of sorting and conversation in the LOA lab at MOA as well as Mary Vickers, Alexandra Tucker, Kelsey Timler, Kav Golka, Sara Wilson, Kathryn Youngs, Abby Manuel, Brett Patychuk, Cora denHartigh, Chenise McClarty, and other LOA Lab night volunteers for all their help. Tremendous thanks to Susan Crockford and Becky Wigen for vital assistance with fish/bird/ mammal bone identification at the University of Victoria Zooarchaeology Lab and for their many years of mentorship. Thanks to Ann Stahl, Quentin Mackie, and Becky for enabling department access and to Yin Lam, Dorothy Wong, Daryl Fedje, and Quentin for hosting me (often w Nicole/Flora) in Victoria. The opportunity to conduct research in the Broken Group Islands was made possible by the support of the Tseshaht First Nation and Pacific Rim National Park Reserve of Canada as well as Parks Canada\u00E2\u0080\u0099s Western and Northern Service Centre. Further support was provided by the Nuu-chah-nulth Tribal Council, BC Hydro, the Laboratory of Archaeology at UBC, Denis St. Claire, and Audrey Dallimore at Royal Roads University. The Broken Group research was conducted under Parks Canada Agency Research and Collection Permits #PRN-2008-1579 and 2009-2737. Research at Huu7ii was supported by the Huu-ay-aht First Nation with in-kind support from SFU and the University of Victoria and was conducted under BC Archaeology Branch Permit 2004-162. I have been extremely fortunate to have the support of a SSHRC CGS scholarship, a Charles and Alice Borden Fellowship, and awards from the UBC Faculty of Arts to support this dissertation research. I thank Audrey Dallimore for geological enthusiasm and capability, for enabling fieldwork in the Broken Group in 2010 and 2011, and for coordinating Malcolm Nicol and Randy Enkin\u00E2\u0080\u0099s scanning of percussion cores at the Geological Survey. Thanks to Brad Anholt and Beth Rogers at the Bamfield Marine Sciences Centre and Larry Johnson, Johnson Ginger, and Rita Johnson at Huu-ay-aht First Nations and as well as numerous BMSC staff for the opportunity to teach Coastal Field Archaeology. Phenomenal thanks to Jane and Ewen McKechnie and to Steve and Kay Weisman who enabled this. I am thankful to the many other UBC anthropology people such as Sue Rowley, Charles Menzies, Pat Moore, Brian Chisholm, Leslie Robertson, Julie Cruikshank, RG Matson, Mike Richards, Eleanore Asuncion, Kyla Hicks, Joyce Ma, Vinay Kamat, Peter Johansen, and William McKellin for various important conversations, opportunities, and acts of kindness. I particularly thank Irena at the MOA Caf\u00C3\u00A9 and Rima Wilkes, Jennifer Kramer and Pat for the many lunches as well as Charles and the Anth Running Club. Further thanks to Hisham Zerriffi, Cornelia Scheffler, Chris Arnett, Nigel Haggan, Jasmine Johnstone, Molly Malone, Duncan McLaren, Anne Salomon, Bill Angelbeck, Gwaliga Hart, Lee Brown, Russ Markel, Jessie Morin, Marina La Salle, Rich Hutchings, Gerald Singh, Margot-Hessing Lewis, Ale Diaz, Sue Formosa, Steve Daniel, Ed Gregr, Jerry Maedel, Tony Pitcher, and Daniel Pauly. Thanks to the department for hosting inspiring visiting speakers and for sushi & coffee. Blake Edgar and Greta Lindquist at UC Press and Roy Carlson at SFU Archaeology Press kindly provided permission to include Chapters 2 and 5 in this dissertation. Grant Keddie at the Royal BC Museum, Tanya Anderson at the Canadian Museum of Civilization, and Jessica Desany Ganong at the Peabody Museum of Archaeology and Ethnology kindly provided image permissions. Finally, I thank my parents Jane and Ewen McKechnie for priceless grandparenting, Nicole\u00E2\u0080\u0099s parents, Ron and Pat Smith for sharing many meals, editorial advice, and help with English as a first language, and to Linda and Bill Murdoch for hosting me from the island my first term in Vancouver. Nicole is an award-deserving contributor to this dissertation. her positive editing and countless moments of support have made this dissertation entirely possible. I thank you and Flora immensely for enduring far too much archaeology talk and for completing my contemporary world in our mutuality of being. \u00E2\u0080\u0093 xviii \u00E2\u0080\u0093 Dedication This dissertation is dedicated to Flora Frances McKechnie and Nicole Fenwick Smith. \u00E2\u0080\u0093 1 \u00E2\u0080\u0093 Chapter 1. Archaeological Scales of Analysis on the Northwest Coast \u00E2\u0080\u009CArchaeological synthesizing on the Northwest Coast and elsewhere is always a process in which we move back and forth between scales of analysis. The ecological foundation for patterns found through analysis of a single site or sites occupied by a local group is often lost when \u00E2\u0080\u0098scaling up.\u00E2\u0080\u009D Madonna Moss (2012:16). Introduction This dissertation examines multiple scales of Indigenous Northwest Coast history from the disciplinary perspective of archaeology. I investigate social life materially represented in subsistence and settlement data to examine temporal and geographic scales of historical variability over the mid-to-late-Holocene (ca. 5,000-200 BP). Scholars have long argued that the Northwest Coast is not a homogenous place but exhibits cultural and historical variation that is, in part, influenced by environmental and ecological differences (Boas 1887; Suttles 1987). These factors have long been central to archaeological investigations of human history on the coast but the temporal and spatial scale of this variation remains unknown, yet must have major consequences for archaeological interpretation (Ames 1991; Cannon 2001; Moss 2012). Confronting this interpretive challenge requires moving beyond the archaeological tendency to generalize from a single case or a small set of cases to examine many examples at a variety of scales and evaluate how robustly aspects of human and environmental history are archaeologically supported and conversely, which aspects of archaeological interpretation are in need of revision. As stated by Madonna Moss in this chapter\u00E2\u0080\u0099s epigram, a central interpretive challenge is the issue of scale. As a conceptual heuristic, scale is both a virtue and a constraint in the archaeological \u00E2\u0080\u0093 2 \u00E2\u0080\u0093 interpretation of Indigenous histories on the Northwest Coast. Scale is virtuous because archaeological data are effective at illuminating the enduring presence of Indigenous People occupying vast landscapes over hundreds of generations throughout the coast. However, a spatially and temporally coarse scale also severely constrains the analytical and interpretive resolution of the conventional archaeological record (Bailey 2007) and correspondingly, can contribute to a lack of richness and dynamism in Indigenous history when this is often a result of inadequate archaeological methods and a lack of preservation. Applying a large-scale archaeological perspective also has the particularly negative consequence of homogenizing a host of potentially significant aspects of cultural, historical, and environmental variation. This analytically erases the possibility of identifying archaeological patterning and social practices as expressed at smaller scales. This dissertation confronts the issue of scale across six chapters within the unifying theme of \u00E2\u0080\u0093 the everyday over millennia. This theme focuses on two fundamental domains of human existence that have a pervasive archaeological expression: food and settlement. My methodology focuses on the everyday over millennia through six case studies that variously draw on zooarchaeological, settlement, and radiocarbon data from multiple localities and time periods on the Northwest Coast. These linked case studies query the strength of archaeological patterning at multiple spatial and temporal scales and demonstrate the interpretive utility of contextualizing archaeological history alongside the intergenerational history represented in Indigenous oral historical narratives and place names. Theoretical Framework The theoretical framework I employ to examine the everyday over millennia is inspired by the historical approach of Braudel (1970; 1980, 1981), particularly the Annaliste \u00E2\u0080\u0093 3 \u00E2\u0080\u0093 concept of the longue dur\u00C3\u00A9e, which is the broad centennial and millennial scale of history beyond \u00E2\u0080\u0098the history of events.\u00E2\u0080\u0099 This broad scale and inclusive conception of historical analysis is applicable to the narration of archaeological history in part because: 1) Archaeological data are exceptional for their encompassing representation of time as well as a range of temporal scales. This \u00E2\u0080\u0098depth\u00E2\u0080\u0099 ensures that a broad assemblage of items and human circumstances are encountered in archaeological sequences. 2) Archaeological data often represent the material legacy of a multitude of people and this collective representation (at a community scale over long periods of time) constitutes a rich representation of human history. 3) Archaeologists intuitively recognize that many aspects of the archaeological record produce strong patterning at millennial scales but often lack a nontrivial explanation to account for such patterning. 4) Adopting a broad spatial and temporal scale facilitates the exploration and recognition of smaller-scale variability and encourages the comparison of phenomena across multiple scales to account for conjunctures and/or disjunctures in historical phenomena. Archaeologists in western North America periodically invoke Braudel and the Annaliste concept of the longue dur\u00C3\u00A9e (Ames 1991; Cannon 1996; Hull 2005) but also recognize that emphasizing patterning at a millennial-scale can be simultaneously amorphous, homogenizing, and under-specific (Silliman 2012). Archaeologists have also commonly evoked a similarly amorphous term of \u00E2\u0080\u0098long-term history\u00E2\u0080\u0099 (Lightfoot 1993; Mitchell and Scheiber 2010), which can also underspecify why or how certain aspects of the archaeological record persist over millennial-scale time frames. Such concepts remain useful descriptive heuristics under which archaeological data exhibit patterning but often lack detailed exploration and explanation. On the Northwest Coast, those who focus on millennial-scale history do not feel compelled to adopt Braudel\u00E2\u0080\u0099s historical scales, but rather emphasize the Indigenous robusticity and stability of cultural institutions (Cannon 2002, \u00E2\u0080\u0093 4 \u00E2\u0080\u0093 2011a; Martindale and Letham 2011). This approach considers the agency of Indigenous peoples and their ritual cosmology as an explanatory factor in the persistence of certain archaeological patterns rather than the more common characterization of Northwest Coast societies as shaped and \u00E2\u0080\u009Csubject to biological and environmental constraint[s]\u00E2\u0080\u009D (Cannon 2011a:57). This dissertation engages with millennial-scale and centennial-scale patterning and variability in Northwest Coast archaeology. I draw on Braudel\u00E2\u0080\u0099s (1970; 1981) concepts of the longue dur\u00C3\u00A9e and conjunctures (medium-term change over human lifetimes) but do so in combination with Bourdieu\u00E2\u0080\u0099s concept of practice (1977). I combine these concepts to investigate how the habitus (i.e. culturally situated orientations or dispositions) of the \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 is expressed as a persistent practice unfolding over human generations and over millennia of human history. As stated by Ortner (2005:33), habitus emerges from practice and represents \u00E2\u0080\u009Ca system of dispositions that incline actors to act, think, and feel in ways consistent with the limits of the structure.\u00E2\u0080\u009D This structure, rooted in history, is inherently cultural and: \u00E2\u0080\u009CCulture is a product of acting social beings trying to make sense of the world in which they find themselves, and if we are to make sense of a culture, we must situate ourselves in the position from which it was constructed\u00E2\u0080\u009D (Ortner 1984:130, emphasis in the original). Similarly, in this dissertation I attempt to identify the practice-based outcome of everyday life as represented in the fine-scale archaeological record of food and settlement. Like Ortner\u00E2\u0080\u0099s subjectivity-oriented view of culture, I seek to adopt the perspective \u00E2\u0080\u009Cfrom which it was constructed\u00E2\u0080\u009D by applying it to interpreting the structured depositional practices of the \u00E2\u0080\u0093 5 \u00E2\u0080\u0093 archaeological record of food and settlement (cf. Garrow 2012; Thomas 2012). I argue that regular and individually practiced elements of everyday life over millennia can be recognized in archaeological deposits when examined at a fine depositional scale and collectively across millennia and broad geographic regions. Aspects of the longue dur\u00C3\u00A9e are archaeologically and anthropologically significant as they are reflective of actions expressed on the highly regular timescale of the \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 which nevertheless durably persist over the \u00E2\u0080\u0098long-term.\u00E2\u0080\u0099 Everyday life is an important but understudied domain of history which \u00E2\u0080\u009Cconsists of the little things one hardly notices in time and space\u00E2\u0080\u009D such as \u00E2\u0080\u009Cthe ways people eat, dress, or lodge\u00E2\u0080\u009D (Braudel 1981:29). Such durable practices are often under-recognized or remain un-questioned (Giddens 1984:41) but importantly transcend the \u00E2\u0080\u009Cevent-centered\u00E2\u0080\u009D focus of conventional historical discourse (Fogelson 1989). This under-recognition extends to archaeological interpretation wherein everyday quotidian practices are often considered historically inconsequential relative to ritual and symbolic practices. In the words of Garrow (2012:85, emphasis added), such \u00E2\u0080\u009Cmaterial signatures of \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 practice have been under-theorized and all too often ignored\u00E2\u0080\u009D and yet it is \u00E2\u0080\u009Cvital to consider the \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 as well as the \u00E2\u0080\u0098ritual\u00E2\u0080\u0099 processes which lie behind the patterns we uncover in the ground.\u00E2\u0080\u009D Thus, certain pervasive forms of archaeological data are under-considered given their interpretive relevance to understanding the \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 and such a lack of focus inhibits exploration of patterning in such data at a finer scale as well as the integration of such data across larger scales of history. Bourdieu\u00E2\u0080\u0099s (1977) concept of practice, as an iterative unfolding of culturally constructed subjectivities and dispositions (Ortner 2005), provides a direct causal linkage between the everyday and the longue dur\u00C3\u00A9e. That is, regularized everyday practices exhibit \u00E2\u0080\u0093 6 \u00E2\u0080\u0093 remarkable persistence across the vast timescales of the longue dur\u00C3\u00A9e, even amidst change in their form and expression, because macro-scale patterns are the collective product of everyday actions and not vice versa. This is particularly applicable to those intergenerational secular rituals of food and settlement that curate and recreate \u00E2\u0080\u0098traditional\u00E2\u0080\u0099 practices, the material signatures of which are richly represented in archaeology. Thus, an archaeological analysis of the longue dur\u00C3\u00A9e has the potential to highlight the power of individuals as causal agents in the construction of history. At the same time, given the volatility of individual agency, evidence of millennia-long stability in everyday practices speaks to the power of cultural traditions and the conventions that foster such continuity. Food and Place as Multi-scalar This dissertation examines two forms of archaeological data that are fundamentally regular aspects of everyday life, that stretch across human generations and are pervasive aspects of the archaeological record on the Northwest Coast: 1) food and 2) the places people live. The animals people eat and the places their remains are deposited offer particular insight into the quotidian daily practices of ancient communities as we can be confident that animal and/or animal products are eaten on a daily basis and that consumption most often occurs in settlements where people spend a great deal of time (homes). Evidence of food consumption and deposition at specific places has enduring interpretive value in archaeology precisely because such data are the result of everyday practice (Atalay and Hastorf 2006; Gifford-Gonzalez 2008). Moreover, because consumption likely occurs on such a fine temporal scale, archaeological evidence of food can be examined at a large variety of scales: from a handful of sediment deposited in a day or a week to an entire midden deposited over centuries or millennia. While there is an unknown degree of temporal averaging and taphonomic bias \u00E2\u0080\u0093 7 \u00E2\u0080\u0093 built into such a multi-scalar comparison (Bailey 2007; Kowalewski, et al. 1998; Lyman 2003a), observations of the repetitive occurrence and stable proportionality for certain foods in archaeological deposits (aka. structured deposition) represents a strong archaeological pattern and creates a basis for inferring the consistent importance of certain foods and thus a continuity in practice over generations. Food Deposition as Social Practice This dissertation focuses on the identification of such collective depositional patterns and argues that a focus on food at particular places provides insight into how \u00E2\u0080\u009Cone is disposed to do and to be in certain ways because of one\u00E2\u0080\u0099s experiences in social settings\u00E2\u0080\u009D (Pauketat 2001:4). I argue that such a simple insight \u00E2\u0080\u0093 the repetitive continuity of a cultural practice \u00E2\u0080\u0093 is often lost amid the more common tendency to emphasize dynamic historical change, the symbolic dimensions of practices, and especially to avoid the equally hazardous narrative trope of an unchanging Indigenous history. A contrary perspective on food is that it can be an abstraction that may have little or no consequence or relationship to major historical and social transformations. In other words, a tsunami of social change may have just occurred but breakfast may remain the same for the next hundred years. However, it is also the case that breakfast wouldn\u00E2\u0080\u0099t be possible without that skilled but often mundane activity of collecting and preparing food on which those \u00E2\u0080\u0098higher-level\u00E2\u0080\u0099 historical dynamics unfold. While this is not to suggest, for instance, that a history of medieval Europe can be deduced from an analysis of farming practices alone, but rather, that such a history cannot be holistically achieved without attentiveness to everyday activities on a broad spatial and temporal scale (Braudel 1981). It is additionally noteworthy \u00E2\u0080\u0093 8 \u00E2\u0080\u0093 that large-scale social transformations often begin from everyday circumstances that become magnified through \u00E2\u0080\u009Cthe amplification of minor differences\u00E2\u0080\u009D consequentially turning \u00E2\u0080\u009Csmall issues into big events,\u00E2\u0080\u009D and thus \u00E2\u0080\u009Chow microhistories become macrohistories and vice versa\u00E2\u0080\u009D (Sahlins 2013:161). Such disruptions can alter the ways people and communities relate to one another and this correspondingly has the potential to be reflected in the proportion and intensity of food being collected, processed, consumed and deposited into the archaeological record. Food pervades most social relations due to its regularity of consumption and its often fundamental centrality in social and ceremonial gatherings (Dietler and Hayden 2001). Food has human and historical agency as it both represents a culturally unique transformation of various animals and plants into a category of \u00E2\u0080\u0098food\u00E2\u0080\u0099 and in so doing, both reflects and facilitates the perpetuation of a \u00E2\u0080\u009Cspecifically cultural and historical consciousness\u00E2\u0080\u009D (Ortner 2005:34). As in Bourdieu\u00E2\u0080\u0099s habitus, such a consciousness does not structurally determine a given future, but it likely informs, influences, and guides practices into the future. Put more simply, such a regular aspect of everyday life also defines what is considered \u00E2\u0080\u0098acceptable\u00E2\u0080\u0099 in a given setting, or conversely, what is not considered acceptable1. Such a formulation is equally applicable to how a person \u00E2\u0080\u0098goes about living\u00E2\u0080\u0099 in a place and what and how one prepares and consumes food and disposes of waste. Animals consumed as human food are often themselves considered to have levels of agency in that they require skill and attentiveness to catch or harvest, prepare and/or care for 1 For instance, in Nuu-chah-nulth territory on western Vancouver Island, an ethnographic account states that a \u00E2\u0080\u009Ccommon punishment for eating alone was to have one\u00E2\u0080\u0099s mouth \u00E2\u0080\u0098stretched\u00E2\u0080\u0099 by two husky individuals, one of whom hooked a finger in either corner of the mouth and gave a yank. Food boxes and dishes of the offender would be smashed up, contents and all\u00E2\u0080\u009D (Drucker 1951:388) \u00E2\u0080\u0093 9 \u00E2\u0080\u0093 (e.g., Ingold 1980; Russell 2012). On the Northwest Coast, animals are often associated with a human-like animism which emerges from specific social and cosmological histories with people (e.g., Drucker 1951:151; Hill 2012; Losey 2010). Animal products and products made from animals also obtain a spiritual significance and serve a host of other practical and secondary uses (e.g., Hodgetts and Rahemtulla 2001). The disproportionate abundance of certain foods combined with exceptional preservation conditions further structures archaeological perspectives on food and food waste. The presence of these foods represents actions of processing and provisioning and involves connections between people and landscapes and of people to each other. It is through such quotidian, under-recognized routines, that communities are commonly defined and persist through time. Place and the Practice of Settlement An even more elemental aspect to the human history of everyday life concerns place. As stated by Casey (1993:21) \u00E2\u0080\u009Cthere is no (grasping of) time without place\u00E2\u0080\u009D and thus, place is a primary category and a fundamentally \u00E2\u0080\u0098prior\u00E2\u0080\u0099 condition. Places emerge over the long-term through routinized practices in a \u00E2\u0080\u009Clandscape of habit\u00E2\u0080\u009D that is both consciously and non-discursively built upon the \u00E2\u0080\u009Cwell-trodden (or well-paddled) paths\u00E2\u0080\u009D laid out by one\u00E2\u0080\u0099s predecessors (Mackie 2001:25). Places are central to the expansive history of the longue dur\u00C3\u00A9e as articulated by Braudel and the Annaliste school of historians. Places are not simply a backdrop to the arc of human history but act as \u00E2\u0080\u009Ca major continuous constraint on human activity\u00E2\u0080\u009D (Thornton 2008:20). Archaeological sites, such as shell midden settlements on the Northwest Coast, are particularly significant as human places, as they are constructed localities where groups of \u00E2\u0080\u0093 10 \u00E2\u0080\u0093 people have \u00E2\u0080\u009Cinvest[ed] themselves into their landscape\u00E2\u0080\u009D (Momaday 1974:80) and literally have mixed \u00E2\u0080\u009Ctheir labour with the earth\u00E2\u0080\u009D (Williams 2005:76). Such physically altered places acquire a broader social and historical significance that can come to signify \u00E2\u0080\u009Cpoints in the geography of a community where time and space intersect and fuse\u00E2\u0080\u009D (Bakhtin 1981:84). Similarly, the surrounding landscape in which archaeological sites develop and in which human histories unfold, both shape people\u00E2\u0080\u0099s history and are correspondingly shaped by people over time (Ingold 1993). In this fashion, human places and landscapes can be said to be ongoing co-productions, reflecting the regular actions of people and natural processes alongside more periodic but larger-scale transformations. Under these conditions, an attempt \u00E2\u0080\u009Cto separate natural history from social history becomes extremely problematic\u00E2\u0080\u009D as \u00E2\u0080\u009Cthe idea of nature contains, though often unnoticed, an extraordinary amount of human history\" (Williams 2005:67-76). For the purposes of this dissertation, shell midden deposits are considered human places or settlements. I consider shell middens to be physically created places where people spent a great deal of time, and whose everyday practices have cumulatively produced a physical assemblage of material spanning generations of persistent human activity (Blukis Onat 1985; Cannon 2003). I recognize that shell middens are only one of a large variety of other \u00E2\u0080\u0098types\u00E2\u0080\u0099 of sites on the Northwest Coast and comprise a relatively narrow aspect in what was undoubtedly a much more widely used coastal and inland landscape (e.g., Oliver 2007). Moreover, materials preserved in these deposits do not encompass the full spectrum of human activities that took place there (e.g., Croes 2003) and neither do the materials represented reflect purely incidental utilitarian deposition or an exclusively ritualized deposition (cf. Klokler 2008; Luby and Gruber 1999; Luby, et al. 2006; McNiven 2013; \u00E2\u0080\u0093 11 \u00E2\u0080\u0093 Villagran, et al. 2011). However, shell middens deposits do contain an immense amount of material evidence of past human activity (Moss 2011a) and persist as settled places because these built environments act to structure future deposition (Mackie 2001:62). Moreover, because shell middens are observed to be localities where people spend a great deal of time, it can be reliably assumed that many of the materials deposited in shell midden sites are related to the everyday human activity that occurred there. As physical deposits, shell middens are places where \u00E2\u0080\u009Ctime takes on flesh and becomes visible for human contemplation\u00E2\u0080\u009D and thus may serve as \u00E2\u0080\u009Cmonuments to the community itself\u00E2\u0080\u009D (Bakhtin 1981:84). Dissertation Structure This dissertation is divided into two parts consisting of six individual chapter contributions as well as an introduction and conclusion. The first half of the dissertation examines Indigenous subsistence and settlement histories as archaeologically expressed on large, medium, and small geographic scales on the Northwest Coast (Chapters 2, 3, and 4). The second half of the dissertation consists of three chapters that focus on archaeological data specific to Barkley Sound on the southwest coast of Vancouver Island (Chapters 5, 6, and 7). These multiple temporal and spatial scales and mixture of regional and local case studies (Figure 1.1 and Figure 1.2) focus on foodways and settlement practices and indicate that Indigenous history on the Northwest Coast exhibits remarkable continuity at the scale of the everyday. \u00E2\u0080\u0093 12 \u00E2\u0080\u0093 Figure 1.1 Map of the Northwest Coast showing geographic scales of analysis that are the focus for chapters in this dissertation. \u00E2\u0080\u0093 13 \u00E2\u0080\u0093 Figure 1.2 Temporal scales of analysis that are the focus for chapters in this dissertation. Vertical scale approximates north-south gradient. Regional Scales of Human and Ecological Histories Chapter 2 adopts a historical ecological approach (e.g., Jackson, et al. 2001) to investigate the archaeology of Indigenous herring fisheries and ecological baselines within much of the Northwest Coast as a singular region over the Holocene. I, along with several co-authors, demonstrate that Pacific herring (Clupea pallasii), have an under-recognized importance in the history of Indigenous fisheries on the Northwest Coast. We compile and integrate a wealth of formerly scattered black and grey-literature data on fine screen fish remains (171 sites) and identify a pervasive coast-wide pattern with a high archaeological importance of herring relative to all other fish species. We observe herring to have a more significant role in Indigenous subsistence economies than has been previously appreciated and further demonstrate that herring exhibit strong spatial variability in abundance across this large region. We explore variability at the finest archaeologically reported scale within 50 individual sites across the coast and argue this indicates herring use and herring habitat was \u00E2\u0080\u0093 14 \u00E2\u0080\u0093 broadly consistent over time and within regions and micro-regions. We further hypothesize that pre-industrial herring population variability differs from modern variability that characterizes industrial fisheries and advocate that research at an archaeologically finer temporal scale is needed to qualify the strength of these observations. Chapter 3 narrows in geographic scope and shifts from Indigenous fisheries to Indigenous hunting traditions on the southern British Columbia coast, from the Strait of Georgia in the Salish Sea around to western Vancouver Island (Figure 1.1). Along with co-author Rebecca Wigen, I examine the spatial and temporal coherence of mammalian hunting practices as represented in a compilation of existing zooarchaeological data within the linguistically distinct Nuu-chah-nulth and Coast Salish culture areas. These regions have long been recognized as distinct by Indigenous Peoples and anthropological researchers (e.g., Boas 1887), but this study compiles a wealth of existing zooarchaeological data to further demonstrate how this distinctiveness is archaeologically represented in hunting practice. Thus, within each of these two regions, we observe an unexpected degree of geographical and temporal coherence in the proportionality of zooarchaeological data. These results yield further insight into the ecological contribution that human hunting had on the marine and terrestrial environments over the Holocene and offer an improved basis for interpreting how these \u00E2\u0080\u0098traditions\u00E2\u0080\u0099 and/or \u00E2\u0080\u0098cultural adaptations\u00E2\u0080\u0099 contributed to and/or complemented other domains of human subsistence and ritual practice (e.g., burning, plant harvesting, spiritual training and preparation). Although these hunting practices have an obvious environmental distribution (outer coast versus an inland sea), this by no means explains the cultural persistence of these practices which are direct products of repetitive, culturally embedded, \u00E2\u0080\u0093 15 \u00E2\u0080\u0093 intentional social action (e.g., Ingold 2000) operating over an intergenerational time scales and at the scale of the everyday or at least every year. Chapter 4 refines the temporal and spatial scale of analysis to examine settlement history in the heavily investigated area of Prince Rupert Harbour. I explore the issue of radiocarbon calibration and the chronological resolution of archaeological phenomena in marine settings with co-author Morley Eldridge. We employ available marine reservoir data from the Dundas Islands and Prince Rupert Harbour to re-evaluate the archaeological settlement dynamics of two long-standing radiocarbon chronologies that have played a significant interpretive role in Northwest Coast archaeology, particularly concerning the antiquity of Indigenous oral history and regional conflict (Cybulski 1999; Martindale and Marsden 2003). This re-analysis narrows and affirms the previous identification of this medium-scale temporal patterning in Coast Tsimshian settlement dynamics and burial practices. Our re-analysis advocates for continued improvements in the use of calibration techniques for marine and marine influenced dates in order to clarify and strengthen previously published chronologies by re-engaging with the cumulative legacy of existing data. We conclude by emphasizing the value of acknowledging ambiguities and contingencies in archaeological claims to knowledge (Gero 2007; Martindale n.d.), which can limit archaeologists\u00E2\u0080\u0099 ability to resolve the temporal scale in key periods in human history. Case Studies of Indigenous History in Barkley Sound The second half of the dissertation examines the archaeological history of subsistence and settlement in two different First Nation territories in Barkley Sound on Southwest \u00E2\u0080\u0093 16 \u00E2\u0080\u0093 Vancouver Island and demonstrate how these reflect both an enduring cultural continuity at a broad scale and provide vital insight into intermediate timescales including periods of dramatic historical change in settlement and more subtle change in foodways. Chapter 5 examines fine-scale coherence and variability in archaeological fisheries data from a single Nuu-chah-nulth \u00E2\u0080\u0098big\u00E2\u0080\u0099 house. I explore the depositional patterning in these everyday foods at multiple depositional and spatial scales within this very large house (35 x 17 m, or 115 x 50 ft) in the centre of a village in the named origin place of the Huu-ay-aht First Nation (McMillan and St. Claire 2012). Two additional column sample time series dating to the mid-Holocene (ca. 5000 to 3000 years BP) and associated with a period of higher sea levels, further contextualize these later household associated data. This chapter also evaluates the interpretive utility of converting the number of fish bones (NISP) into numbers of fish (MNI) and extrapolates these in terms of fish per cubic metre. I also examine temporal trends in the use of key species (hake, salmon, rockfish) that indicate coherent shifts in the marine climate during a period of when there is a regional increase in dated defensive sites on western Vancouver Island (ca. 800-500 years BP). This study demonstrates the relevance of exploring subsistence and climatic variability in a single large house over a 1000 year period. Zooming out to a multi-site scale but remaining in Barkley Sound, Chapter 6 examines the settlement history in an archipelago identified in numerous Indigenous oral historical accounts as the territory of an autonomous Nuu-chah-nulth polity (the Maktl7ii7ath local group) in a portion of the Broken Group Islands (Sapir and Swadesh 1955; St. Claire 1991). This group of islands has a series of archaeologically anchored Indigenous place names and an oral historical record describing it as the defended year-round territory of the \u00E2\u0080\u0093 17 \u00E2\u0080\u0093 Maktl7ii7ath prior to the contact era (St. Claire 1991). I examine the archaeological evidence of settlement in this portion of the archipelago and evaluate how this corresponds to oral historically documented social history. I employ multiple measures (e.g., midden surface areas, radiocarbon dates) to archaeologically characterize this human settlement history. Collectively, these observations demonstrate these islands were enduringly occupied over multiple generations, consistent with the Nuu-chah-nulth concept of \u00E2\u0080\u0098home\u00E2\u0080\u0099 or \u00E2\u0080\u0098chiefly territory\u00E2\u0080\u0099 (hahuulhi). This dynamic settlement chronology, reveals considerable shifts exemplified in the archaeologically \u00E2\u0080\u0098sudden\u00E2\u0080\u0099 occupation of the elevated fortress of Huts\u00E2\u0080\u0099atswilh. This micro-regional settlement history is an archaeological example of the intermediate scale of intergenerational history akin to Braudel\u00E2\u0080\u0099s conjunctures and that transcends \u00E2\u0080\u0098event-centred\u00E2\u0080\u0099 history (cf. Braudel 1970; Fogelson 1989). Chapter 7 builds from the previous chapter but focuses on the fundamentally regular activity of collecting and consuming food as a frame for contextualizing the everyday over millennia. Centering on zooarchaeological data from numerous auger samples collected from two settlements on Dicebox Island as well as associated villages on adjacent islands, this chapter details how archaeological evidence of ubiquitous foods are represented at multiple temporal and spatial scales over the last 2,500 years. This includes data from the territory as a whole between sites, and within individual levels at a single site. I demonstrate that certain foods constitute a consistent majority of the repetitively deposited items. I then evaluate if the relative proportion and/or density (bones per litre) of these foods persist over time, and explore the ecological factors by which they may differ spatially and/or temporally. I relate these to community processes identified in the settlement organization explored in the previous chapter. \u00E2\u0080\u0093 18 \u00E2\u0080\u0093 Conclusion These linked multi-scalar studies examine the most regularly practiced aspects of daily life at a variety of scales on the Northwest Coast. I identify the outcome of traditional foodways as represented in the archaeological record both over long temporal scales but also at small depositional scales (e.g., handfuls of dirt). The results provide evidence of both historical continuity and considerable stability of practice. I interpret this as evidence of both a thoughtful cultural adaptation to the circumstance of people\u00E2\u0080\u0099s lives in these environments and as evidence of an extraordinary effort at maintaining historical continuity via a construct of tradition. Thus, the stability of the longue dur\u00C3\u00A9e is achieved not as a fixed cultural entity but is iteratively reproduced within a cultured continuity of everyday practice. This approach seeks to confront the analytical tendency to consider certain aspects of the archaeological record to be static, essentialized and normative. Rather, I seek to demonstrate how archaeological attentiveness to identifying regular and structured depositional practices in a community setting and its variation holds potential for enriching archaeological understandings of the everyday practices of Indigenous peoples. Finally, I examine how both the Indigenous oral historical record and archaeological chronology of settlement enrich archaeological interpretations of continuity and change in settlement practices on the Northwest Coast. \u00E2\u0080\u0093 19 \u00E2\u0080\u0093 Chapter 2. Archaeological Data Provide Alternative Hypotheses on Pacific Herring (Clupea pallasii) Distribution, Abundance, and Variability on the Northwest Coast2 Introduction Low trophic-level fish (\u00E2\u0080\u009Cforage fish\u00E2\u0080\u009D) are experiencing global declines with increasing recognition of widespread and cumulative ecological, cultural and economic impacts (Pikitch, et al. 2012; Pinsky, et al. 2011; Smith, ADM, et al. 2011). Both Indigenous and non-Indigenous peoples on the Northwest Coast of North America recognize Pacific herring (Clupea pallasii) as a foundation species in coastal food webs (Menge, et al. 2013) that serve an essential role in maintaining social\u00E2\u0080\u0093ecological systems (e.g., Garibaldi and Turner 2004; Platten and Henfrey 2009). Herring and its roe are critical prey for a host of fish (e.g., hake, Pacific cod, dogfish, salmon), bird and marine mammal predators (Robinson and Ware 1999; Therriault, et al. 2009; Ware and Schweigert 2002). Herring is also central to the social, cultural, and economic relations of coastal indigenous communities, many of which seek to continue their traditional fisheries for herring and herring roe on kelp or other substrates (Harris 2000; Thornton, et al. 2010a; Thornton, et al. 2010b). Since 1882 and continuing into recent decades, industrial fishing of herring has helped support many communities across the Northwest Coast (Funk 2010; Tester 1935). Populations of this once highly abundant forage fish have been dramatically reduced across much of its North Pacific range relative to levels seen in the mid 20th century (Funk 2 This multi-authored chapter was submitted for consideration for publication on September 16, 2013 and accepted for publication on December 12, 2013 (see preface for more information). Due to length, the tables associated this chapter are presented in the appendix to this dissertation. \u00E2\u0080\u0093 20 \u00E2\u0080\u0093 2010; Schweigert and Linekin 1990; Schweigert, et al. 2010). Fisheries scientists have proposed various factors accounting for these declines and sustained low abundances even after reductions of fishing pressure. These include climate-induced ecological changes in distribution of predators and prey (Ware and Schweigert 2002), disease (Marty, et al. 2010), over-fishing (Pauly, et al. 2001), and the rebound of marine mammal populations that prey on herring (Schweigert, et al. 2010). Assessing these potential drivers and moving forward with conservation requires baseline information on herring abundance and distribution prior to its depletion. Yet current knowledge of Pacific herring distribution and abundance is based on biomass estimates that date back only to the mid 20th century \u00E2\u0080\u0093 well after the onset of industrial fishing. Most modern ecological data lack sufficient time depth for establishing baselines for marine ecosystem management and/or recovery (McClenachan, et al. 2012); this has the potential to dramatically underestimate the degree of population loss (the \u00E2\u0080\u0098shifting baseline\u00E2\u0080\u0099 problem [Pauly 1995]) and inhibit recovery efforts. As a result, traditional and local ecological knowledge (TEK/LEK) as well as paleoecological and archaeological data are increasingly important for informing these baselines (Rick and Lockwood 2013). Such data provide both the long-term perspective needed to assess pre-industrial ecological states, and ecologically and culturally salient baselines for conservation (Jackson, et al. 2011; Turner, et al. 2008; Wolverton and Lyman 2012). Here, we compile the archaeological record of fisheries in the Northeast Pacific from Alaska, British Columbia, and Washington (Figure 2.2) to assess the spatial and temporal distribution of pre-industrial herring abundance. The abundance and spatial distribution of archaeological fish bones reveals the widespread importance of herring to indigenous peoples \u00E2\u0080\u0093 21 \u00E2\u0080\u0093 throughout the region, and indicates the abundance of herring in coastal ecosystems over the past several thousand years. Herring were highly abundant along southwest Vancouver Island and in the Salish Sea. The archaeological record indicates that places with abundant herring were consistently harvested over time, and suggests that the areas where herring massed or spawned were more extensive and less variable in the past than today. Although archaeological data are expressed over different spatial and temporal scales than modern fisheries population estimates, archaeological data highlight both a disjunction between pre-industrial and contemporary herring abundances and distribution, and the need to revise and expand the data on which current fisheries management and conservation are based. History of Herring and Herring Roe Fishing For many First Nations and Native American groups from Alaska to Washington, the nutritionally valuable and readily harvested herring and its roe were integral to daily lives and worldviews (e.g., Boas 1932; Bouchard and Kennedy 1989; Brown and Brown 2009). This is reflected in photographs, interviews, oral histories, and indigenous place names (e.g., Teeshoshum, \u00E2\u0080\u009Cmilky waters from herring spawn;\u00E2\u0080\u009D K\u00E2\u0080\u0099i:na?a, \u00E2\u0080\u009Cherring guts on rocks\u00E2\u0080\u009D (Sapir and Swadesh 1955:35); Yaaw Teiyi \u00E2\u0080\u009CHerring Rock,\u00E2\u0080\u009D the sacred place where the first herring arrived; Shaan Daa \u00E2\u0080\u009CWhite Island,\u00E2\u0080\u009D also known as \u00E2\u0080\u009CFish Egg Island\u00E2\u0080\u009D, named for the whiteness created by the spawning activity each spring; and \u00E2\u0080\u009CSilver Bay\u00E2\u0080\u009D, because in the winter, there were so many herring \u00E2\u0080\u009Cif you looked at it in the moonlight you\u00E2\u0080\u0099d see the backs of the herring... and it would look silver\u00E2\u0080\u009D (Thornton, et al. 2010a:60, 334, 489). Many coastal groups maintained family-owned locations for harvesting herring and herring roe from anchored kelp fronds, eelgrass, or boughs of hemlock or cedar trees (Drucker 1951; Powell \u00E2\u0080\u0093 22 \u00E2\u0080\u0093 2012). Herring was harvested at other times of the year than the spawning period when massing in local waters (Grant 1857:300; McKechnie 2005b:103; Moss, et al. 2011; Sapir and Swadesh 1939:223) but most ethnohistorical observations identify late winter and springtime spawning as a key period of harvest for both roe and fish. Processed herring and roe were consumed in large quantities and traded widely among coastal First Nations (Bouchard and Kennedy 1989:8; Jewitt 1807:6). Sustainable harvests were encouraged by building kelp gardens (Brown and Brown 2009:xviii), wherein some roe-covered fronds were not collected, by minimizing noise and movement during spawning events, and by elaborate systems of kin-based rights and responsibilities that regulated herring use and distribution (Bouchard and Kennedy 1989; Powell 2012; Sproat 1868:224; Thornton, et al. 2010a; Thornton, et al. 2010b). Industrial commercial fishing of herring began in earnest on the Northwest Coast in the late 19th century primarily using beach and drag seines to catch fish and then render (\u00E2\u0080\u009Creduce\u00E2\u0080\u009D) oil and meal (Stick and Lindquist 2009; Taylor 1955; Tester 1935; Thornton, et al. 2010a). In British Columbia, officials prohibited the reduction of herring into oil and fertilizer in 1910 (Carrothers 1941:111) and noted that the larger bays were \u00E2\u0080\u009Cbeing gradually deserted by the larger schools where they were formerly easily obtained\u00E2\u0080\u009D (Tester 1935:7). In 1927 (23 years prior to the current management baseline for estimating historical biomass), the fishery on eastern Vancouver Island, British Columbia, processed 31,103 metric tons of herring (Munro and Clemens 1931:5), which is roughly two times the annual harvest rate in 2012 and approximately 38% of the current biomass estimate for the entire Strait of Georgia (Department of Fisheries and Oceans Canada 2012b; 2013:43). In Alaska, the herring reduction industry began in 1882, with state-wide harvests reaching a peak of 75,000 metric \u00E2\u0080\u0093 23 \u00E2\u0080\u0093 tons in 1929 (Funk 2010:257). By the 1930s, the effects on population numbers and/or behaviour caused fisheries scientists to express deep concern about the effects of over-fishing on herring (Carrothers 1941:111; Taylor 1955:110; Tester 1933:289), as the fishery shifted to deeper water harvest technology (Tester 1935:7) and underwent a considerable spatial expansion in search effort (Taylor 1955:110). By the late 1960s, as a result of increased fleet efficiency, reduction fisheries, and poor recruitment, the herring populations of British Columbia and Washington collapsed (Outram and Humphreys 1974). This led Canada to permanently close its reduction fishery in 1968 (Pearse 1982:15), followed by the closure in Washington state in the early 1980s (Stick and Lindquist 2009). In the early 1970s, the herring industry in Japan collapsed (Morita 1985), which increased East Asian demand for herring eggs taken from North American waters. As a result, a Sac Roe Fishery (SRF) began throughout the Northeast Pacific, targeting female herring just prior to spawning. In the last decade, coast-wide declines in herring numbers (Cleary, et al. 2010; Stick and Lindquist 2009) have led to a greatly reduced SRF, now limited to a few regions such as the Strait of Georgia (Salish Sea) and Sitka and Togiak, Alaska (Funk 2010). In addition to these larger ventures, a relatively small food and bait fishery has persisted since the early 20th century. Since the 1970s licenses and legal judgments have been issued to First Nations in Canada (Newell 1999; Powell 2012), Native Americans in Washington (Boxberger 1999), and Alaskan Natives (Thornton, et al. 2010a) that support food, social, and ceremonial fisheries and in some cases commercial fishing. Government fishery managers, scientists, and local and indigenous peoples lack consensus on the cumulative consequences of ongoing commercial fisheries on herring populations. Many First Nations, Native Americans, Alaska Natives and other local fishers, \u00E2\u0080\u0093 24 \u00E2\u0080\u0093 based on personal observations and traditional knowledge, hypothesize that local herring stocks, on which they consistently relied for generations, have been dramatically reduced and made more difficult to access following 20th century industrial fishing (Haida Marine Traditional Knowledge Study Participants, et al. 2011; Hebert and Thornton 2010; Thornton, et al. 2010a; Thornton, et al. 2010b). In contrast, fisheries managers (Hay, et al. 2008; Schweigert, et al. 2010) identify commercial fishing as only one of several potential causes for the coast-wide decline in herring and/or persistent lack of recovery since implementing conservation measures (Hay, et al. 2009). In the Strait of Georgia (BC), managers hypothesize that populations have in fact not been depleted, but rather have shifted spatially due to climatic factors and predator abundance (Hay, et al. 2009; Ware and Schweigert 2002). The unresolved nature of these alternative hypotheses regarding the primary factors responsible for temporal and spatial shifts in herring populations represent a barrier to achieving consensus on the need and strategies for improving herring conservation and management. Zooarchaeological data offer a record of the pre-industrial abundances and distribution of herring, providing a longer-term perspective that can illuminate and contextualize these debates. Herring and the Archaeological Record On the Northwest Coast, stratified, shell-bearing archaeological sites (shell middens) provide long-term records of human-animal interactions. Although animal bones can enter archaeological sites through a range of cultural and non-cultural processes (Erlandson and Moss 2001), the majority of fish bones in coastal middens are associated with human processing or consumption of fish. Even though animal products were widely traded in the \u00E2\u0080\u0093 25 \u00E2\u0080\u0093 past, the bulk of zooarchaeological remains in shell midden deposits tend to be composed primarily of resources harvested nearby (Cannon 2000a; Lepofsky, et al. 2007; Szpak, et al. 2013). Thus, in most cases, zooarchaeological remains, including marine fish bones, can be used as proxies for local distribution and abundance. This synthesis of zooarchaeological fisheries records builds on the increasing number of recent analyses employing rigorous methods of fish bone recovery and quantification, particularly the use of column sampling and fine-screen mesh (\u00E2\u0089\u00A4 3.2 mm) that is critical to ensuring adequate proportional representation of small-bodied fish such as herring (Casteel 1972). Our estimates of herring relative abundance and rank order are based on a standard zooarchaeological measure of \u00E2\u0080\u009Cnumber of identified specimens\u00E2\u0080\u009D (NISP), which is not equivalent to biomass or meat weight. Rather, NISP is correlated with numbers of individual animals (Grayson 1984) and thus can be used as a culturally and taphonomically filtered proxy of past, local fish populations (Lyman 2008). Most archaeological deposits, including the zooarchaeological data presented here, have relatively low chronological resolution as calibrated radiocarbon age-range estimates often span more than a century; thus, specific deposits incorporate a degree of time averaging (Kowalewski, et al. 1998; Lyman 2003a). Despite interpretive challenges, ancient fish bone records provide long-term and spatially explicit data on past use and abundance of herring and other fish. This is particularly so in the case of our dataset, which compiles rigorously screened data from numerous sites across the Pacific Northwest Coast that represent multiple temporal and regional scales. This is the largest available dataset of fine-screened archaeological fish assemblages from the Northwest Coast and offers new insight into the taxonomic composition of indigenous fisheries. \u00E2\u0080\u0093 26 \u00E2\u0080\u0093 Figure 2.1. Percent abundance of herring bones in archaeological sites with >50 identified fish bones (N = 171 sites). Only two sites lack herring bones. Herring is abundant (>60% of total fish NISP) in sites throughout the Strait of Georgia in southern British Columbia. In 71% of sites, herring makes up at least 20% of the site\u00E2\u0080\u0099s total assemblage of fish bones. Material and Methods3 Data Sources To examine the distribution and abundance of herring represented in archaeological sites of the Northwest Coast (SE AK, BC, WA), we compiled a database of all well-sampled sites with adequately recovered and identified fish bone assemblages located within 1 km of 3 Due to length, the tables for this chapter (Chapter 2) are presented in an appendix to this dissertation. \u00E2\u0080\u0093 27 \u00E2\u0080\u0093 the current marine shoreline (Table 2.1). This entailed an extensive literature review of published and grey literature on zooarchaeological analyses completed over the past 40 years. Given the small size of herring bones (e.g., vertebral centra are 4 mm or less in diameter), we only included sites where the zooarchaeological remains were systematically recovered using a fine-screen mesh (equal to or less than 3.2 mm). All zooarchaeological remains were identified by established analysts or students working under the analysts\u00E2\u0080\u0099 direct supervision using comparative collections (Table 2.1). Only sites containing a minimum of 50 fish bone specimens identified to at least family level were included. Over 91% of the 171 sites have more than 100 fish bones. This is a reasonable threshold for assessing the relative abundance of the three-most ubiquitous and abundant taxa at a site (Butler and Campbell 2004:340). For the site-based analyses presented in Figure 2.1 and the regional analyses presented in Figures 2.3-2.6, we combined zooarchaeological data from sites with multiple excavated sub-assemblages to ensure adequate sample sizes (Tables 2.1 and 2.2). We used the relative abundance of herring bones among all identified fish (number of identified specimens, NISP) as a proxy for the relative abundance of herring in archaeological sites. We did not convert herring bone NISPs to estimates of \u00E2\u0080\u0098meat-weight\u00E2\u0080\u0099 or biomass as suitable taphonomic, allometric, and stratigraphic data are lacking for the majority of sites. We judge our assessment of the abundance of identified herring bones as a proportion of all fish bones to be conservative given that herring have fewer vertebrae than some larger fish taxa that can fracture into numerous identifiable fragments (e.g., Oncorhynchus spp., Squalus acanthias). While quantitative data on herring bone density measurements have not been conducted, their bones are notably smaller and more delicate \u00E2\u0080\u0093 28 \u00E2\u0080\u0093 relative to most other measured fish species (Smith, RE, et al. 2011a). Moreover, the cellular structure of herring bone likely makes them more susceptible to microbial degeneration relative to other fish (Szpak 2011:3367) indicating this species would likely be less well-preserved in the archaeological record. Chronological Assessments To assign age-ranges to archaeological assemblages, we used stratigraphic and chronological information provided in the original research reports and in subsequent published archaeological research (Table 2.1). The majority of sites contain at least one radiocarbon date (88%) while the remainder contain temporally diagnostic artifacts and are on landforms consistent with late Holocene sea levels (N = 21 sites). Radiocarbon age estimates were assigned age categories based on dates re-calibrated using the Intcal09 curve (Reimer, et al. 2009) and represent the 2-sigma calibrated range. Most radiocarbon samples were taken from the basal layers of individual site deposits. Time Series Analysis To examine within-site temporal patterns at the finest possible depositional scale, we identified a sub-set of 50 sites that report chronologically distinct vertical levels or stratigraphic layers (Figures 2.6 and 2.7, Table 2.2). We included only those sites that have been dated in their basal and terminal levels either by radiocarbon dating or the presence of historic artifacts of European manufacture. Assemblages had to have 50 or more identified fish bones per vertical level or stratigraphic layer and three or more levels in sequence. To ensure each datapoint had 50 or more specimens, some assemblages from contiguous strata or levels with less than 50 identifiable specimens were combined (Table 2.2). \u00E2\u0080\u0093 29 \u00E2\u0080\u0093 Due to differences in how assemblages were collected in situ, and the level of detail in which they are reported, we present these data in four analytically separate categories in Figure 2.5. These are; 1) assemblages collected from multiple areas within a site according to temporally distinct strata (ca. 500-2000 years per stratigraphic layer, n=14 sites); 2) sites with a single column sample of contiguous levels (N = 19 sites); 3) sites with multiple column samples with multiple contiguous levels spanning broad temporal intervals (N = 8 sites), and 4) Sites with multiple column samples with multiple contiguous levels spanning broad temporal intervals in Barkley Sound (N = 9 sites). Comparison to Modern Spawning Records To evaluate archaeological data in relation to historic and contemporary herring spawning records from British Columbia (Figure 2.7.), we calculated the distance between our archaeological study sites and historically documented herring spawning localities (Table 2.3). We restricted this analysis to sites within areas with modern herring spawning observations as these represent a long-term ecological record of cumulative spawning activity and are derived from annual surveys of spawning length, thickness, timing and other variables collected by the Department of Fisheries and Oceans Canada (DFO) (Department of Fisheries and Oceans Canada 2012a) since 1937 (McCarter, et al. 2005). Where available, we also documented the proximity to spawning areas identified in Traditional and Local Ecological Knowledge studies in BC (Haida Marine Traditional Knowledge Study Participants, et al. 2011; Hessing-Lewis, et al. 2011). These are based on interviews with knowledgeable indigenous and non-indigenous local residents with lifetimes of perspective on the fisheries in their traditional territories. Distance to closest documented spawn was \u00E2\u0080\u0093 30 \u00E2\u0080\u0093 determined by placing geo-rectified maps of cumulative spawning localities into Google Earth and measuring the shortest route by water (\u00C2\u00B1 500 m) to an archaeological site. To assess the potential disjuncture between archaeological data and the cumulative record of historical spawning activity, we compared the six quantitative categories used by DFO (ranging from \u00E2\u0080\u0098vital\u00E2\u0080\u0099 to \u00E2\u0080\u0098minor\u00E2\u0080\u0099(McCarter, et al. 2005)) to correspondingly ranked archaeological abundance categories (e.g., 100-80% NISP = vital, 1-20% = minor). If there was more than one record of cumulative spawning within a distance of 2 km, we chose the most abundant ranking for the comparison. If the ordinal grouping in the two records differed by more than one category, we considered the two records to represent a disjuncture in abundance (Table 2.3). Figure 2.2. Rank order of herring bones in 171 adequately sampled archaeological sites. Herring is among the top three most abundant fish taxa in 88% of the sites and is absent in only two sites. \u00E2\u0080\u0093 31 \u00E2\u0080\u0093 Results In the study area, 171 coastal archaeological sites have been sufficiently sampled to evaluate the past distribution and archaeological abundance of herring (Figure 2.1, Table 2.1). These sites span the early Holocene (10,700 cal BP) to the contact-era (AD 1740-1860) with the bulk of sites dating to within the past 2,500 years (76%). The dataset contains 435,777 fish bones confidently identified to family, genus, and/or species. These specimens represent a wide range of taxa, with each site containing a minimum of 50 identified bone specimens. Coast-wide similarities in the taxonomic richness of assemblages and relative abundance of numerically dominant fish taxa indicate that the archaeofaunal sample sizes are adequate to assess the distribution and relative abundance of herring (Table 2.1). Herring Ubiquity, Abundance, and Rank Order Within this dataset, herring is the single-most ubiquitous fish taxon in archaeological sites throughout the Northwest Coast, occurring in 169 of the 171 assemblages (99%) representing various physiographic settings. It is also the single-most numerically abundant taxon in the dataset representing 49% of all identified fish bones, with a site average of 47% (\u00C2\u00B133% NISP). Herring is similarly the 1st ranked taxon by NISP in over half (56%) of the 171 assemblages and is among the two most numerous taxa in 80% of assemblages (Figure 2.2). Herring comprises more than 20% of fish bones in 71% of the assemblages (NISP). The high ubiquity, rank order, and relative abundance of herring in most sites is remarkable given the many other taxa present in each site (mean N of fish taxa = 10.2, SD = 3.2) and that some ubiquitous genera, such as Oncorhynchus spp. (salmon and trout) and Sebastes spp. (rockfish) comprise many species (7 and 36 respectively). The occurrence and relative abundance of herring in the majority of sites surpasses any other fish taxa, and demonstrates \u00E2\u0080\u0093 32 \u00E2\u0080\u0093 a pervasive and previously under-recognized role for this species in indigenous economies spanning the Holocene. Figure 2.3. Boxplots showing percent abundance of herring bones in adequately sampled archaeological sites (N = 171) by major region (arranged north to south). Boxes encompass the middle 50% of cases, with the median represented as a vertical line. Whiskers encompass the 1.5 of the interquartile range, excluding outliers (represented as circles) and extreme outliers (represented as stars). Figure 2.4. Boxplots showing percent abundance of herring bones in archaeological sites by sub-regions (arranged north to south) within the Salish Sea region (N = 78). Note the lower abundance of herring in the Fraser Delta and Puget Sound. Note: Gulf Islands includes the Canadian Gulf Islands and the US San Juan Islands. \u00E2\u0080\u0093 33 \u00E2\u0080\u0093 Spatial Distribution of Herring Geographically, nearly half the sites in the database (46%) are concentrated around the Salish Sea, reflecting more intensive archaeological investigation in this region. Only two assemblages, both in the northern portion of the study area, do not contain herring bones (Table 2.1). The occurrence of sites with significant proportions of herring throughout the study area (Figure 2.3) indicate that the spatial gaps in the distribution of analyzed sites reflect gaps in archaeological sampling, rather than gaps in the ancient distribution of herring. Thus, as a proxy for the number of fish harvested in coastal waters, these data reflect a widespread abundance of herring in marine ecosystems in the deep past. Archaeological sites numerically dominated by herring occur throughout the study area. Sites in the Central Coast, West Coast Vancouver Island, and Salish Sea areas are consistently dominated by herring bones with median abundance values representing over 40% of the assemblage (Figure 2.3). Most sites where herring represents less than 20% of NISP (n=50) are in the northernmost portion of the study area (Haida Gwaii, northern BC Coast, and SE Alaska) or in the southernmost portion of the study area (Puget Sound and the lower mainland of British Columbia) (Figures 2.1, 2.3, and 2.4). Even in the northern portion of the study area, herring still ranks among the three most ubiquitous and abundant fish in 75% of sites (Table 2.1). The proportion of herring in Haida Gwaii exhibits high variability amongst nearby sites with salmon, rockfish, and dogfish (Squalus acanthias) representing the other dominant fish taxa. On the northern BC Coast, sites around Prince Rupert and on the Dundas Islands tend to be dominated by salmon and oil-rich smelts (Osmeridae). In southeast Alaska, salmon tends to dominate followed by herring, Pacific cod (Gadus macrocephalus), and sculpins (Cottidae). \u00E2\u0080\u0093 34 \u00E2\u0080\u0093 The Salish Sea region exhibits exceptionally high archaeological abundance of herring, but with substantial variation among sub-regions (Figures 2.3 and 2.4). Sites along the northwestern and northeastern Strait of Georgia (eastern Vancouver Island, E. Strait of Georgia, respectively), the southwestern Strait of Georgia (southern Vancouver Island), and the Gulf Islands, have the highest mean relative abundances for herring (Figure 2.4). These high numbers contrast with sites in Puget Sound (N = 10, x\u00CC\u0085 = 20.6%) and along the Fraser Delta (N = 10, x\u00CC\u0085 = 33.7%), which tend to have abundance values less than 40%. This spatially variable distribution of herring across broad regions of the Northwest Coast and around the Salish Sea suggests that although herring was a ubiquitous component of coastal ecosystems, herring biomass (vis-\u00C3\u00A0-vis other fish taxa) was relatively greater in some regions than others. \u00E2\u0080\u0093 35 \u00E2\u0080\u0093 Figure 2.5. Percent abundance of herring bones over time in sites with dated samples (N = 50, Table 2.2). Sites are arranged into three groups a) Multiple samples within a site aggregated into broad temporal groupings (ca. 500-2000 years per stratigraphic layer, N = 14 sites), b) Sites with a single vertical \u00E2\u0080\u0098column\u00E2\u0080\u0099 of contiguously sampled levels (N = 19), and c) Sites with multiple column samples from multiple site areas including Barkley Sound (N = 17). Each site is associated with one or more radiocarbon dates and/or contain historic-era artifacts in the upper levels indicating they date ca. AD 1778-1850 (Table 2.1). \u00E2\u0080\u0093 36 \u00E2\u0080\u0093 Temporal Patterns The sites in the dataset span the past 10,700 calendar years. The majority of the sites (78%) have temporal components that fall within the past 2,500 years whereas much smaller numbers of sites have components that date prior to this period and only a handful that date before 5,000 years ago (Table 2.1). This temporal distribution broadly corresponds to the proportion of dated archaeological sites in the region (Ames and Maschner 1999:54), and is a product of site taphonomy, relative search effort, and geographically variable Holocene sea levels and shoreline locations (Mackie, et al. 2011). Of the 171 sites, 50 (29%) meet our criteria for examining detailed intra-site temporal patterning (Figures 2.5 and 2.6, see Materials and Methods). Similar to the overall dataset, over half of these sites are from the Salish Sea region (N = 27, 54%) and the majority date to within the past 2,500 years. Collectively, the 50 sites are comprised of 571 vertically distinct levels, each with sample sizes of >50 NISP (Table 2.2). Herring occurs in the overwhelming majority of the 571 individual levels (99.3%) and the mean relative abundance of herring exceeds 20% in the majority of the 50 sites (88%). Variance in the relative abundance of herring is less than \u00C2\u00B110% within 96% of individual sites (Figure 2.6, Table 2.2), further supporting an interpretation of broad consistency over time. Variance is correlated with sample size, however (Spearman\u00E2\u0080\u0099s rho=0.303, p=0.032, n=50), which reflects the influence of several thoroughly sampled sites from a single region with moderate variability (Barkley Sound, see Figure 2.5D). Herring ranks as the first or second-most numerous fish taxon in the majority (92.1%) of the 571 stratigraphic levels. Seven sites exhibit mean abundances lower than 20% and low rank orders (less than 3) for herring over time (Figure 2.6). Herring is only absent from four \u00E2\u0080\u0093 37 \u00E2\u0080\u0093 individual levels in four sites which also have consistently low mean abundance values (<4% NISP; Figure 2.5). The broad consistency in abundance through time is particularly striking considering these archaeological deposits likely represent a variety of cultural and ecological contexts, seasons of occupation, and depositional processes over multiple temporal scales (Cannon 2000a; Lyman 2003a). Of the two sites in our sample that exhibit highest intra-site variability (more than \u00C2\u00B110% variance), one (Burnaby Narrows, Figure 2.5A5) spans a key period of subsistence change (Orchard 2011b) and sea-level change (Mackie, et al. 2011). In the other (Spring Cove, Figure 2.5B10), dramatic increases in anchovy in two of five occupational levels reduce the relative abundance of herring, which correspondingly increases variance (cf., Lyman 2003a). As with modern fisheries landing data, social, economic, and environmental circumstances as well as analytical resolution and ecological drivers, needs to be considered when interpreting fine-scale variability within a specific site. Sites in close proximity and which are broadly contemporaneous display similar patterns of herring abundance over time (e.g., Tsawwassen, Beach Grove, and Crescent Beach around Boundary Bay [Figure 2.5A8-10]; Cape Lazo, Q\u00E2\u0080\u0099umu?xs, and DkSf-4 in Comox Harbour [Figure 2.5B6-7, 2.5C2a and 2.5C2b ]; and the Nanaimo Harbour / Departure Bay area [Figure 2.5C4a-b]). This similarity suggests sub-regional-scale coherence in herring abundance over time and indicates that the formation processes associated with each site assemblage (either human harvest of herring or natural taphonomic processes within a site) do not obscure the overall ability of the zooarchaeological data to accurately estimate abundance of herring in these settings. Furthermore, the observations of similar herring abundance values from multiple columns from spatially separate but temporally overlapping \u00E2\u0080\u0093 38 \u00E2\u0080\u0093 deposits within a single site (Figure 2.5C and Table 2.2) strengthens the inference that single column samples can document site-wide patterning for such abundant and ubiquitous taxa (McKechnie 2005b). These overall patterns thus support the value of zooarchaeological data as a proxy for past herring abundance. Figure 2.6. Mean abundance and rank order of archaeological herring bones from levels in adequately sampled and dated sites (N = 50; arranged north to south). Error bars represent \u00C2\u00B11 standard deviation (see Table 2.2 for supporting information). Collectively, these multiple analytical scales (Figures 2.5 and 2.6) and abundance measures indicate a pattern of relatively low temporal variability and broad consistency in \u00E2\u0080\u0093 39 \u00E2\u0080\u0093 herring use as represented among and within archaeological sites distributed across a large coastal region. In areas where herring was archaeologically abundant, it appears to be consistently available and abundant throughout the mid-to-late Holocene. Given the limits of radiocarbon dating and calibration, we cannot determine the extent to which the temporal resolution of our data may inhibit our ability to detect finer-scale (e.g. annual, decadal, multi-decadal) variation in herring abundance through time. If high variation in the local availability of herring over time was characteristic for this region, however, we hypothesize this would be apparent in more of our site sequences. Overall, this large-scale archaeological dataset indicates that herring was a desired and reliable source of food in coastal waters in most locations across the Northwest Coast. Figure 2.7. Boxplot showing variability in the archaeological abundance of herring bones in relation to proximity by water to historically surveyed spawning locations (\u00C2\u00B1500 m). Distance data calculated only for sites in BC within monitored areas (N = 79) and with components dating to within the last 2,500 years (Table 2.3). Archaeological data are grouped into relative abundance categories as per Figure 2.1. Comparison to Modern Spawn Location Data To evaluate the ecological longevity of herring spawning habitat and the extent that herring spawning areas have changed historically, we determined the proximity of \u00E2\u0080\u0093 40 \u00E2\u0080\u0093 archaeological sites to herring spawning areas documented as a result of federal fisheries monitoring efforts that support stock assessment in British Columbia (Department of Fisheries and Oceans Canada 2012a). We limited our analysis to archaeological deposits dating within the last 2,500 years (\u00E2\u0080\u0098late period sites\u00E2\u0080\u0099) and located within areas annually monitored for spawning activity since the mid 20th century (N = 79 sites). We observe that archaeological sites with greater than 80% herring are significantly closer (Mann-Whitney U = 525, Z=-2.936, p=0.003) to documented spawning locations than archaeological sites with less than 80% herring (Figure 2.7, Table 2.3). The significance of this difference is strengthened when we include data from nine sites with traditional and local ecological knowledge [TEK/LEK] (Haida Marine Traditional Knowledge Study Participants, et al. 2011; Hessing-Lewis, et al. 2011) on the BC Coast (Mann-Whitney U = 478, Z = -3.490, p=0.000). The strong correspondence between high archaeological abundance of herring bones and the multi-decade monitoring record of modern spawning sites is a compelling argument for the long-term site fidelity of herring spawning. In British Columbia, over 30% (N = 35) of the 114 late period archaeological study sites are in areas that have not been monitored for herring spawn by Department of Fisheries and Oceans (DFO) (mean distance from closest monitored shoreline is 7.5 km, SD = 7.9 km). Of these, 14% (N = 5 sites) have herring abundance values over 80%. Based on the correspondence between sites with >80% herring bone abundance and documented spawning sites, we hypothesize that these five sites were also close (i.e., within \u00C2\u00B1 2 km) to former herring spawn locations. Thus, archaeological sites can be used to expand the ecological baseline for spawning populations and identify localities that may no longer support spawning or are missed by modern monitoring efforts. \u00E2\u0080\u0093 41 \u00E2\u0080\u0093 We also note a general correspondence between six abundance categories used by federal fisheries monitoring in British Columbia (see Materials and Methods and Table 2.3) to characterize the cumulative location and intensity of herring spawning, and six ordinal categories of archaeological abundance of herring. That is, the majority of late period archaeological assemblages (48 of 79 sites; 60%) have similar categorical abundance values to the fisheries monitoring dataset (\u00C2\u00B1 2 km). However, 25% of these archaeological sites (N = 20 sites) contain a categorically greater abundance of herring than fisheries monitoring records whereas, only 10% have substantially lower abundance values than cumulative spawning records (N = 8 sites). Collectively, these comparisons both provide information on the longevity of spawning habitats as well as the potential depletion of spawning habitats in historic times. The predominant correspondence between ordinal categories in the ancient and modern data indicates correspondence between ancient productivity of herring and the productivity of contemporary spawning habitats in a given location. This in turn indicates consistency in the geography of spawning habitats from the distant to the more recent past. Furthermore, the quarter of archaeological sites that have higher abundances of herring than is predicted by the federal cumulative spawn records provides a basis for expanding our estimate of potentially productive spawning habitat beyond what is indicated by historic census records. Discussion This broad compilation of zooarchaeological data provides a previously unconsidered record of the past abundance and distribution of fisheries along the Northwest coast of North America. It establishes herring as among the most important marine fish on which coastal indigenous people relied in the past. These data expand greatly upon the extant \u00E2\u0080\u0093 42 \u00E2\u0080\u0093 paleoecological fisheries record (O\u00E2\u0080\u0099Connell and Tunnicliffe 2001; Tunnicliffe, et al. 2001; Wright, et al. 2005) and indicate that herring was both widespread across the coast, and a mainstay of ecological and socio-economic systems over the Holocene. Herring occurred in abundance (>40% NISP) in numerous sites across all areas of the coast but exhibited super-abundance (i.e., > 60% NISP) in a few regions such as the central and northern Strait of Georgia and west coast of Vancouver Island. Spatially clustered sites reflect similarly consistent abundance values and a few sites exhibit modest trends over time that may relate to physically driven habitat change, such as the geomorphological evolution of the Fraser Delta (e.g., Beach Grove and Tsawwassen). Only a small number of sites exhibit high variability whereas the more common pattern is of broadly sustained abundance over space and time (Figures 2.5 and 2.6). Accounting for Differences Between Archaeological and Modern Herring Dynamics Could the contrast between the consistency seen in the archaeological record and patterns in recent fisheries showing overall decline and increasing fluctuation in spawning distribution arise from insufficient temporal resolution in the archaeological data? Even though modern population fluctuations are observed at interannual to decadal timescales and the likely resolution of the archaeological record is at the centennial scale, it is unlikely that the temporal consistency in the archaeological record is an artifact of insufficient chronological resolution for several reasons. First, if extreme fluctuations in abundance and periodic absences were common in the past (for instance, such that sites with great abundance in some periods commonly declined to low levels) we would expect to see this reflected in our data, even if we couldn\u00E2\u0080\u0099t resolve the actual temporal trajectory in an \u00E2\u0080\u0093 43 \u00E2\u0080\u0093 individual population. Alternatively, our data may also represent subannual temporal variability, which is masked by the coarseness of existing calibrated radiocarbon records. Both scenarios indicate that in our sample of 571 levels, across 50 sites and representing ca. 7,000 years of harvesting data, we should observe some indication of absences and wide fluctuations in herring abundance if they were present. Yet these data show a remarkably consistent presence of herring with 99.3% of the 571 levels containing herring and a mean relative abundance exceeding 20% in 86% of the sites, with within site variance of less than \u00C2\u00B110% in 48 of 50 sites (Figure 2.7, Table 2.3). Second, because of the regularity with which ancient peoples \u00E2\u0080\u0098sampled\u00E2\u0080\u0099 their marine environments, the archeological record should be a sensitive indicator of marked declines in local abundance of herring as harvesters would conceivably switch to alternative prey when a target species declines to low levels. Such resource switching is observed in the regional zooarchaeological record, but is expressed over centennial scales and among a limited range of animal species other than herring (e.g., Butler and Campbell 2004; Orchard 2011b). In contrast, even substantial variability around a high mean abundance might not be reflected in our data if herring populations never declined to a threshold where people could meet their needs for consumption, storage, and trade. Likewise, another factor potentially contributing to consistency in the archaeological abundance of herring may be multi-season herring harvests, where fish, fish oil, and roe were processed for longer-term storage and extended consumption within a year (e.g., McKechnie 2005b:103; Moss, et al. 2011:283). This scenario may further dampen fluctuations in abundance over decadal or centennial time scales but also depends on consistent interannual availability of herring. \u00E2\u0080\u0093 44 \u00E2\u0080\u0093 We envision three alternative hypotheses to account for the difference we observe between archaeological and modern data in the apparent variability and abundance of herring populations (Figure 2.8). The first hypothesis asserts that the modern pattern of interannual dynamics is an accurate reflection of the past and that the archaeological record overestimates abundance and underestimates variability in ancient herring populations (Figure 2.8, Hypothesis 1). For the reasons advanced above, we argue that this hypothesis can be rejected. Our data seem most consistent with hypotheses that ancient herring populations had a higher mean abundance than in the last century (Hypotheses 2 and 3, Figure 2.8). Variability could have been damped relative to the industrially exploited populations of the 20th century (Hypothesis 2, Figure 2.8), or it could have been similar to those harvested populations (Hypothesis 3, Figure 2.8). In either case, abundance in the more distant past was sufficiently high to meet the needs of harvesters, leading to consistency in the archaeological record. \u00E2\u0080\u0093 45 \u00E2\u0080\u0093 Figure 2.8. Schematic representing three alternative hypotheses regarding the relationship between modern and archaeological abundance and variability of herring populations. The left side of the figure combines the smooth archaeological signal (blue lines) with three hypothesized trajectories for actual populations (black lines), each shown with its range of variability (grey stippled rectangles). In each case, the end of the timeline following the onset of industrial fishing represents a caricature of the recent pattern of herring population dynamics, characterized by great variability, an overall decline in abundance, and some periods of very low abundance. While the archaeological record shows consistency through time, the paleooceanographic record for the Northwest Coast illustrates considerable variability (e.g., Ivanochko, et al. 2008). Both ocean temperature and productivity vary throughout the late Holocene at different spatial and temporal scales, with some periods of interregional convergence (Anderson, et al. 2005a; Ersek, et al. 2012; Wright, et al. 2005). If ancient herring populations fluctuated in response to high frequency oceanographic variability (e.g., \u00E2\u0080\u0093 46 \u00E2\u0080\u0093 Tanasichuk 1997), the fluctuations were not of sufficient amplitude to influence the overall catch of ancient fishers (cf. Reitz, et al. 2008); the substantial variability in the paleooceanographic record is not matched by the archaeological record of herring. Moreover, it seems unlikely that climatic change alone is sufficient to account for substantial declines in Pacific herring populations and spawning distribution over the last 100 years (cf. Hsieh, et al. 2006; Lindegren, et al. 2013). The most parsimonious explanation for the difference between the modern pattern of variability in herring abundance and the long-term archaeological record is the onset of industrial-scale commercial fishing. Dramatic reductions and spatial shifts in herring populations were observed by fisheries managers in British Columbia and Alaska in the 1930s (Carrothers 1941:111; Funk 2010; Munro and Clemens 1931; Taylor 1955:110; Tester 1933:289; 1935:7). A period of decline between AD 1910 and 1970 is evident in the well-dated sediment cores from southeast Vancouver Island (O\u00E2\u0080\u0099Connell and Tunnicliffe 2001:187). Since the 1970s, further population constrictions have been observed by non-Native bait fisherman in the Georgia and Johnstone Straits (BC), including the widespread absence of herring in over 170 locations that previously supported spawning (Schweigert and Linekin 1990). Modern declines and contractions of the spawning range of Pacific herring are supported by recent syntheses of the traditional ecological knowledge of indigenous peoples in Alaska and British Columbia (Haida Marine Traditional Knowledge Study Participants, et al. 2011; Hebert and Thornton 2010; Hessing-Lewis, et al. 2011). These syntheses document quantifiable reductions in the magnitude of spawning events and loss of spawning locations in living memory. Furthermore, this shift in abundance is reflected in indigenous place \u00E2\u0080\u0093 47 \u00E2\u0080\u0093 names, which highlight locations of formerly abundant herring, but where few herring are found today (e.g., Ch'axa'y or \u00E2\u0080\u009CSizzling [with herring] Water\u00E2\u0080\u009D]) (Reimer/Yumks 2011:237). Archaeological abundance of herring is additionally mirrored in indigenous place names and origin narratives. For instance, the 2,400 year-old site of Nulu on the Central BC Coast, where herring made up 85\u00C2\u00B111% of the fish assemblage (Cannon, et al. 2011), is the place where, according to Heiltsuk oral tradition, Raven first obtained herring (Boas 1932). Conversely, a culturally affiliated site at the Koeye River 25 km away from Nulu (Table 2.1), is not associated with herring or herring spawn either in ethnographic or modern ecological data and exhibits low archaeological herring abundance (10% NISP). In the northeastern Salish Sea in southern British Columbia, the place name of Teeshoshum (\u00E2\u0080\u009Cwaters white with herring spawn\u00E2\u0080\u009D) is associated with a ca. 800 year old assemblage comprised of 93% herring bones (Table 2.1). However, extensive herring spawning in this ecologically suitable location have not been documented since 1998 (Department of Fisheries and Oceans Canada 2012a). Collectively, these sources of historical and ethnobiological data reinforce the contrast between the archaeological record and the dynamics of the modern fishery. In response to locally reduced or absent spawning populations, the current herring fishery in BC has contracted from its mid-20th century coast-wide focus and is now concentrated along the east coast of Vancouver Island, overlapping an area in which herring were super-abundant archaeologically (Department of Fisheries and Oceans Canada 2013; Therriault, et al. 2009). While other factors are also responsible for range contraction of harvested fish populations, range contractions are recognized as one of the effects of over-fishing (MacCall 1990; Worm and Tittensor 2011). \u00E2\u0080\u0093 48 \u00E2\u0080\u0093 Historical Baselines As Context for Modern Dynamics Historical baselines in natural resource management serve as reference conditions for understanding the context for modern population or ecosystem dynamics, providing both assessments of abundance and ranges of natural variability (Landres, et al. 1999; Rosenberg, et al. 2005). Our archaeological data, in combination with traditional ecological knowledge and early historic observations (Carrothers 1941; Funk 2010; Munro and Clemens 1931), suggest that late 20th century census data alone do not provide a sufficient baseline for assessing the abundance, distribution, and dynamics of Pacific herring in relation to industrial fishing since the 1880s. This discrepancy between recent and past dynamics is expressed in both temporal and spatial domains. In the temporal domain, the archaeological data challenge the notion that large fluctuations in abundance, including extremely low levels of abundance, are a regularly occurring component of population variability. In the spatial domain, our data argue against the idea that spawning was spatially erratic in the past, with little site fidelity (Hay, et al. 2009:1662). The archaeological data, in combination with the oral historical knowledge and early historic observations (Carrothers 1941; Funk 2010; Munro and Clemens 1931), suggest that industrial fishing already had a significant ecological impact on herring abundance and spawning location well before the initiation of coast-wide spawning censuses in British Columbia in 1938. Moreover, the historical baseline currently used for annual stock assessments that underpin current harvest allocations in BC begin only in 1951 (Department of Fisheries and Oceans Canada 2012b:11). The recent history of erratic population fluctuations, declines, and shifting spawning distributions exhibited by Pacific herring populations are not unusual among industrially harvested populations of forage fish worldwide (Anderson, et al. 2008; Hsieh, et al. 2006; \u00E2\u0080\u0093 49 \u00E2\u0080\u0093 Pikitch, et al. 2012; Pinsky, et al. 2011; Smith, ADM, et al. 2011). What is unique is our ability to provide a long-term temporal context for these recent dynamics. This is particularly relevant because these data reflect sustained continuous harvesting of herring populations for millennia prior to more than a century of modern industrial exploitation. Similar to historical research on the impacts of early industrial-era fishing on Atlantic cod (Rosenberg, et al. 2005) and herring in northern Europe (Barrett, et al. 2004; Poulsen 2008), and the long-term effects of human use of coral reefs (Kittinger, et al. 2011), the archaeological analyses synthesized here critically extend the temporal depth of ecological baselines for contemporary fisheries management. Conclusion Spatially and temporally extensive archaeological data on the relative abundance of herring bones in coastal archaeological sites along the Northwest Coast provide insight into the past distribution and abundance of Pacific herring and long-term human-herring interactions. Herring bones exhibit a remarkable degree of dominance within the archaeofauna across space and time in the majority of these records. Over the period represented well by the archaeological record (ca. 2,500-200 years BP), Pacific herring populations also appear to have exhibited higher abundance and greater consistency in their distribution than is indicated by the dynamics of industrially harvested populations over the past 50-100 years. The archaeological data indicate that in most parts of the study area, and particularly in the Strait of Georgia, herring remained consistently available to harvesters over thousands of years. Of the hypotheses posed in Figure 2.8, we reject hypothesis 1, that the archaeological data misrepresent the actual abundance and variability of herring. At present, we cannot distinguish between hypotheses 2 and 3: herring abundance appears to \u00E2\u0080\u0093 50 \u00E2\u0080\u0093 have been consistently high, but we cannot resolve the magnitude of variability in abundance. The archaeological record, in combination with local and traditional knowledge, early historical records, and paleoecological records, suggest that spawning locations were formerly more abundant and geographically extensive than is recorded by modern surveys. These data, particularly in the context of high harvest levels during the early industrial fishery and the subsequent contraction of effective spawning range, indicate that the currently utilized ecological baseline of the mid 20th century is inadequate for modern management. Our data support the idea (Hypotheses 2 and 3, Figure 2.8) that, if past populations of Pacific herring exhibited substantial variability, this variability was expressed around a high enough mean abundance such that there was adequate herring available for indigenous fishers to sustain their harvests while avoiding the extirpation of local populations. These records thus demonstrate a fishery that was sustainable at local and regional scales over millennia, and a resilient relationship between harvesters, herring, and environmental change that has been absent in the modern era. Archaeological data have the potential to provide a deep-time perspective on the interaction between humans and the resources on which they depend. Furthermore, it can contribute significantly towards developing temporally meaningful ecological baselines that avoid the biases of shorter term records. \u00E2\u0080\u0093 51 \u00E2\u0080\u0093 Chapter 3. Towards a Historical Ecology of Pinniped and Sea Otter Hunting Traditions on the Coast of Southern British Columbia4 Introduction Marine mammals (pinnipeds, cetaceans, and sea otters, Enhydra lutris) have been important to First Nations people in coastal British Columbia for millennia but their archaeological distribution is poorly known. While archaeological evidence of marine mammal hunting is known for numerous locations over the past 10,000 years of human occupation on the British Columbia Coast (e.g., Cannon 1991; Carlson 2003; Fedje, et al. 2005a; Matson 1976), it is remarkable that few studies have examined archaeological evidence of mammalian hunting traditions on broad regional and/or temporal scales. Considering the importance of these animals to the modern marine ecosystem, understanding the long-term human use and past distribution of marine mammals can add considerable perspective to contemporary knowledge of these ecologically important species and this highly valued marine ecosystem (cf. Jackson, et al. 2001; Lotze and Worm 2009; Pitcher 2005). The occurrence of marine mammal bones in archaeological contexts reflects the direct use and long-term occupation of this region by coastal First Nations people. Such occurrences indicate that humans have been participants in this ecosystem for at least the past 10,000 years and as such, likely directly or indirectly affected the distribution, growth, behavior, and relative densities of marine mammals and their prey (cf. Crockford, et al. 2002; 4 This chapter has been published previously as: McKechnie, Iain and Rebecca J. Wigen (2011) In Human Impacts on Seals, Sea Lions, and Sea Otters: Integrating Archaeology and Ecology in the Northeast Pacific, edited by T. J. Braje and T. C. Rick, pp. 129\u00E2\u0080\u0093166. University of California Press, Berkeley. (c) 2011 by the Regents of the University of California. See preface for more information. \u00E2\u0080\u0093 52 \u00E2\u0080\u0093 Etnier 2002a; Gifford-Gonzalez, et al. 2005; Lyman 2003b). In this chapter, we compile archaeological data on the distribution of pinnipeds and sea otters from archaeological assemblages along the coast of southern British Columbia. We evaluate the spatial and temporal extent of human hunting and explore the possible influence humans may have had on this aspect of the marine environment and conversely, discuss the potential significance that hunting these animals had to First Nations cultures in the region. Drawing on archaeological and ethnographic information, we ask three questions regarding the use of marine mammals by First Nations People over the past 8,000 years: \u00E2\u0080\u00A2 What marine mammals did aboriginal people in Southern British Columbia most commonly utilize? \u00E2\u0080\u00A2 How similar or how different are species occurrences and proportions relative to today? \u00E2\u0080\u00A2 Is there evidence of specialized or regional hunting traditions and if so, what might have been the potential impacts of these activities on the ancient marine ecosystem? Context Recent assessments of Northwest Coast archaeology have emphasized the need to adopt a regional approach to characterizing patterns of aboriginal resource use (Butler and Campbell 2004; Cannon 2001). In discussing the role of salmon on the central Northwest Coast, Cannon (2001:185) has noted that: \u00E2\u0080\u009CMore important than improved recovery and analysis of fish remains from individual site locations is the pressing need to expand research strategies beyond the individual sites to encompass a variety of site locations within specific regions. This is the only way to gain a better appreciation of the extent of regional, seasonal, and longer-term temporal variability in the focus and intensity of local fisheries.\u00E2\u0080\u009D \u00E2\u0080\u0093 53 \u00E2\u0080\u0093 Although Cannon\u00E2\u0080\u0099s suggestion is directed towards the archaeological assessment of fisheries, it is broadly applicable to archaeological research on the Northwest Coast. To date, only a few regional studies of indigenous resource use have been conducted (Butler and Campbell 2004; Crockford, et al. 2002; Hanson 1991; Hanson and Kusmer 2001; Hebda and Frederick 1990; Hobson and Driver 1989; Moss 2008; Orchard and Clark 2005). These efforts have provided a host of insights but have been hindered by the small number of published zooarchaeological studies, a methodological concern on sampling adequacy and taxonomic identification, and a focus on characterizing the \u00E2\u0080\u0098full range\u00E2\u0080\u0099 of animal consumption including birds, fish, mammals, and shellfish (e.g., Croes and Hackenberger 1988; Driver 1993; Hanson 1991; Mitchell 1990a). Additionally, many of these studies have focused on demonstrating quantitative differences in tabular form making it difficult to assess the potential continuity and spatial associations between adjacent sites and regions. In this chapter, we adopt an explicitly spatial approach to examining the archaeological expression of aboriginal hunting practices in southern British Columbia, focusing specifically pinnipeds and sea otters. We explore compositional patterns using zooarchaeological assemblage data from 75 temporally distinct assemblages from 58 sites spanning the past 8,000 years. We group assemblages into four broad temporal categories variously corresponding to broadly defined archaeological cultural historical \u00E2\u0080\u0098periods\u00E2\u0080\u0099 (300\u00E2\u0080\u00931,200; 1,200\u00E2\u0080\u00932,400; 2,400\u00E2\u0080\u00935,000; 5,000\u00E2\u0080\u00938,000 calibrated years BP) following Lepofsky et al. (2005; 2007) and Ames and Maschner (1999). These data provide context for considering the regional extent and potential influence that First Nations hunting practices had on the marine and terrestrial environments of southern British Columbia. \u00E2\u0080\u0093 54 \u00E2\u0080\u0093 The Modern Marine Ecosystem Our study area includes the western and eastern coasts of southern Vancouver Island, the Gulf Islands, and the British Columbia mainland and adjacent areas on the Olympic Peninsula in Washington State (approximately 40,000 km2, Figure 3.1). This region is temperate with occasional winter snowfall at sea level, high annual rainfall on western Vancouver Island (260-340 cm), and much lower rainfall on southeastern Vancouver Island and the Gulf Islands (70-110 cm) (Environment Canada 2009). The terrestrial environment is heavily forested with the densest forest cover on western Vancouver Island. The least densely forested areas are on southeastern Vancouver Island and the Gulf Islands, which contains patchy oak savanna woodlands and prairies with evidence of anthropogenic fire regimes extending back at least 2,000 years (Brown and Hebda 2002). The waters of the Strait of Georgia are heavily influenced by the seasonal input of freshwater from the Fraser River and other major rivers (Thomson 1981). Marine primary productivity is further influenced by nutrient-rich deep-water drawn into the Strait of Juan de Fuca from the north flowing California counter-current. This hyper-saline and oxygenated water is vertically mixed by the strong tidal currents in the Gulf Islands at which point this mixed surface water then flows back out the strait and northwards along the outer coast of Vancouver Island (Thomson, et al. 2007). The greatest annual primary productivity in the study area occurs off the coast of southwestern Vancouver Island (Ware and Thomson 2005), but strongly increases during Summer when wind-driven upwelling and prolonged sunlight along the continental shelf break facilitate massive blooms of plankton attracting a host of marine predators (Allen, et al. 2001). \u00E2\u0080\u0093 55 \u00E2\u0080\u0093 The contemporary distribution of pinnipeds on the southern British Columbia coast is represented by six species occupying a range of habitats and exhibiting seasonal fluctuations in abundance (Table 3.1). Harbor seals (Phoca vitulina) are the most numerous and ubiquitous species which are present year-round (Department of Fisheries and Oceans Canada 2010). Northern fur seals (Callorhinus ursinus) are slightly more abundant but do not come ashore and are found almost exclusively on the continental shelf off western Vancouver Island, primarily between December and May (Olesiuk 2009b:57). Steller Sea lions (Eumetopias jubatus) while resident year-round, do not have rookeries in the study area and are only seasonally present in the Strait of Georgia during Winter and Spring (Bigg 1988; Olesiuk 2009a). California sea lions (Zalophus californianus) are present in lesser numbers, and elephant seals (Mirounga angustirostris) migrate to the southern BC coast seasonally (Table 3.1). Elephant seals have recently established a year-round haul-out on Race Rocks on the southern tip of Vancouver Island where at least two elephant seals pups have been born in 2009 and 2010 (Racerocks.com 2010). The contemporary British Columbia sea otter population consists of 4,182 animals (Nichol, et al. 2009), which have increased from a re-introduced population of 89 animals transported from the Aleutian Islands to northern Vancouver Island in between 1969-1972 (Bigg and MacAskie 1978). This population has been increasing annually at a rate of 15% but has yet to entirely re-colonize its former habitat (Nichol, et al. 2009). A recent population reconstruction (Gregr, et al. 2008), incorporating habitat characteristics and contact-era (ca. AD 1774-1825) fur trade records, suggests that the British Columbia coast may have once supported a population of between 34,000 and 74,000 otters. Interestingly, this estimate \u00E2\u0080\u0093 56 \u00E2\u0080\u0093 excludes the Strait of Georgia due to the absence of sea otters in published archaeological assemblages (i.e., Hanson and Kusmer 2001). Until 1970, the federal Department of Fisheries and Oceans supported \u00E2\u0080\u0098population control\u00E2\u0080\u0099 programs, which offered incentives to reduce pinniped populations (particularly Harbor seals and Steller sea lions) (Jeffries, et al. 2003; Olesiuk 2009a). These controls ended when the marine mammal protections were implemented in 1970 (Olesiuk 2009b). Since this time, Harbour seal populations have increased dramatically from 10,000 animals to 105,000 animals (Department of Fisheries and Oceans Canada 2010). Harbour seals in particular were viewed as competitors with human fisheries, especially salmon, despite the fact that the bulk of their prey is composed of non-salmonids (Olesiuk, et al. 1990). Historic Changes to the Marine Ecosystem Marine mammals, particularly the trade of their highly valued pelts, were central throughout the process of early historic contact between First Nations and Europeans on the Northwest Coast of North America (Fisher 1977; Lutz 2008). Sea otters, fur seals, and then whales sequentially became the focus of international commercial industries, which rapidly diminished these culturally and ecologically significant animals over several decades. The corresponding impact of the sudden removal of these ecologically important animals is poorly known (due in part to the paucity of archaeological data and lack of long-term ecological studies) and this issue is a contemporary research concern for ecologists and managers throughout northern Pacific (e.g., Misarti, et al. 2009; Springer, et al. 2003). The early \u00E2\u0080\u0098contact period\u00E2\u0080\u0099 on the Northwest Coast (ca. AD 1774-1812) focused on the trade of sea otter pelts obtained from the western coast of Vancouver Island as well as the \u00E2\u0080\u0093 57 \u00E2\u0080\u0093 northern and central British Columbia Coast (Busch and Gough 1997; Dick 2006). Sea otter pelts were recognized as wealth items and chiefly regalia within aboriginal communities prior to contact with European trading vessels. Aboriginal exchange of pelts for highly valued European trade goods was therefore immediately incorporated into the indigenous economy and caused a large expansion of this particular hunting activity (Dick 2006; Lutz 2008). Over a few decades, the burgeoning trade led to an increasing scarcity of sea otters and increasing conflict between European and American traders and First Nation communities. European trading ships ceased to seek sea otter pelts on the southern British Columbia coast after 1812 following a series of dramatic incidents involving the destruction of European ships and crew as well as reprisal attacks on aboriginal villages (e.g., McMillan 1999:188). After a period of several decades of infrequent interaction, the European colonial presence on the coast increased substantially during the mid 19th century. In the 1880\u00E2\u0080\u0099s through the early 1900\u00E2\u0080\u0099s, commercial demand for seal pelts in the European market fueled a commercial sealing industry, which targeted northern fur seals off British Columbia Coast and Washington State (Crockford 1996). Dozens of schooners operating out of Victoria and Seattle employed hunters from First Nations communities on western Vancouver Island and the Olympic peninsula to travel offshore where hunters would launch traditional dugout canoes and use harpoons to approach and capture fur seals \u00E2\u0080\u0098sleeping\u00E2\u0080\u0099 in the surface waters (Swan 1883). In 1880 alone, over 12,000 fur seals were harvested off the coast of Vancouver Island (Swan 1887:397). Growing international pressure concerning the declining stocks of fur seals at their breeding grounds in the Bearing Sea in the late 19th century (Elliott 1886) \u00E2\u0080\u0093 58 \u00E2\u0080\u0093 eventually resulted in one of the first international treaties protecting marine wildlife in 1911, and ended the commercial harvest of fur seals in Canadian waters (Gay 1987). Also during this time, a nascent whaling industry developed on the British Columbia coast during the mid 19th century but expanded significantly after 1905, regionally diminishing whale populations, particularly sperm whales (Physeter macrocephalus), fin whales (Balaenoptera physalus), and humpbacks (Megaptera novaeangliae) (Gregr, et al. 2000). Non-migratory populations of humpback whales residing in coastal inlets were among the first whales targeted (Merilees 1985; Webb 1988). Whaling continued on the British Columbia coast until it was halted in 1967. Historic Changes Among Coastal First Nations The dramatic changes described above occurred in the context of tumultuous and tragic social, economic, and demographic changes within aboriginal communities over the last two centuries. Particularly devastating were introduced diseases (particularly smallpox and measles), which swept through the coastal population in the 1790\u00E2\u0080\u0099s and later in the 1850\u00E2\u0080\u0099s and 1860\u00E2\u0080\u0099s (Boyd 1999). These epidemics severely reduced the aboriginal population and correspondingly destabilized the highly structured social and political networks and dramatically reduced the scale and extent of the aboriginal subsistence economy. A large influx of European and Asian colonists in the mid to late 19th century introduced a host of new opportunities and challenges for sustaining and generating wealth in aboriginal communities. Many aboriginal people began to participate in the developing wage economy in forestry, fisheries, and agriculture on a seasonal basis (Lutz 2008). Among the highest wages were those offered to aboriginal hunters employed in the fur seal industry who earned \u00E2\u0080\u0093 59 \u00E2\u0080\u0093 seasonal incomes well above European counterparts (Lutz 2008). This industry persisted until the 1911 international fur seal treaty between America, Japan, Russia and Canada mandated a halt to this once central economic activity. Also during the late 19th century, anthropological fieldworkers began to document aboriginal peoples of the coast (e.g., Boas 1887). Numerous researchers compiled painstakingly detailed description and oral historical accounts of aboriginal lifeways including marine mammal hunting practices and rituals associated with hunting (e.g., Sapir and Swadesh 1939; Waterman 1920). Whaling was a particularly honored tradition among Nuu-chah-nulth peoples on the west coast of Vancouver Island and Makah peoples on the Olympic Peninsula (McMillan 1999) but sealing was also a respected skill in coastal aboriginal communities on both the east and west coasts of Vancouver Island (Drucker 1951; Elmendorf 1960; Suttles 1952). Hunters were typically chiefs or men of high status who underwent elaborate ritual and physical preparation (Arima 1988; Drucker 1951; Elmendorf 1960:86; Suttles 1952; Waterman 1920). Hunter\u00E2\u0080\u0099s wives were obligated to observe ritual practices prior to and during the hunt and contributed to labor to butchering and cooking as well as preparing materials which comprised all aspects of hunting technology (e.g., preparing sinew, seal skin floats, cedar bark rope, etc.). Hunting took place from dugout canoes using harpoon technology with seal-skin floats or from land using clubs. A host of historic and ethnographic accounts are available for further investigating hunting practices and techniques, some of which are described further in the sections below. Methods To investigate the regional expression of marine mammal hunting practices, we compiled available zooarchaeological data from academic papers, graduate student theses, \u00E2\u0080\u0093 60 \u00E2\u0080\u0093 and provincially reviewed archaeological permit reports on file at the British Columbia Archaeology Branch in Victoria or federally reviewed reports on file with Cultural Resource Management Services Division of Parks Canada (Calgary, AB) (Table 3.2). All sites are within 500 m of the coastline or along the banks of lower Fraser River (a major tidally influenced river draining into the Strait of Georgia). Sites that are more than 500 m from the ocean (\u00E2\u0080\u0098inland\u00E2\u0080\u0099 sites) are not included in the present analysis because they have not been investigated to the same degree as shoreline shell midden sites and generally contain poorly preserved faunal assemblages. All assemblages included in the analysis are from pre-contact archaeological deposits (ca. > AD 1774) and date to with the past 8,000 years. However, the majority of the assemblages date to the last 2,400 years. Most sites in the study sample (>90%) are from deposits which have been radiocarbon dated. Assemblages without radiocarbon ages are either from the upper layers of ethnographically identified village locations (assumed to date within the past 1,200 years) or are associated with temporally diagnostic artifact assemblages. To enable comparability between regions, we collapse assemblages into one of four broad temporal categories (300\u00E2\u0080\u00931,200; 1,200\u00E2\u0080\u00932,400; 2,400\u00E2\u0080\u00935,000; 5,000\u00E2\u0080\u00938,000 calibrated years BP, Table 3.2). \u00E2\u0080\u0093 61 \u00E2\u0080\u0093 Table 3.1 Ecological characteristics for pinnipeds and sea otters in British Columbia. Common Name Northern Fur Seal Harbour Seal Steller Sea Lion California Sea Lion Elephant Seal Sea Otter Species Callorhinus ursinus Phoca vitulina richardsi Eumetopias jubatus Zalophus californianus Mirounga angustirostris Enhydra lutris Mean Female Wt. (kg) 35-45 58 200-300 94 650 44 Mean Male Wt. (kg) 100-200 72.5 400-800 375 1800 33 Modern Period of Pup Dependency 7-8 months One month 12-24 months 12-24 months One month 5-8 months Modern Birthing Season, Location July \u00E2\u0080\u0093 Nov., Bearing Sea - California May - June, Throughout BC May - June, N BC May \u00E2\u0080\u0093 July, California Oregon Dec -March, California & Mexico Year-round with peak in Spring Modern Seasonal Occurrence in March - June, with peak in May Resident in BC year round Resident in BC year round Year round with peak Oct \u00E2\u0080\u0093 May Recently present year-round Resident year-round Prevalence in Archaeological Sites in Southern BC Ubiquitous on western Vancouver Island Ubiquitous Common Rare Rare but increasing Present on Western VI, rare/absent on Eastern VI Estimated Contemporary Population in BC (2000-2010) 123,000 105,000 18,400-19,700 2,000-3,000 Fewer than 100 4,712 Management Status in Canada Threatened Not at Risk Special Concern Not at Risk Not at Risk Special Concern IUCN Red List Status (2009) Vulnerable Least Concern Endangered Least Concern Least Concern Endangered References (Banfield 1974; Department of Fisheries and Oceans Canada 2007:5) (Baird 2001: 668; Olesiuk 1999: 28) (COSEWIC 2003) (Bigg 1985:17; Hancock 1970; NOAA Fisheries Service 2009) (Stewart 1999: 217-218) (Estes 1999:182; Nichol, et al. 2009) Zooarchaeological assemblages included in the analysis were identified using comparative osteological collections based at large universities in the region (e.g., Simon Fraser University, University of British Columbia, University of Victoria, and Washington State University) and/or at the Royal BC Museum in Victoria (RBCM). Individual bone specimens were identified to the lowest taxonomic level possible on the basis of their morphological similarly to comparative specimens. A majority of analyses have been \u00E2\u0080\u0093 62 \u00E2\u0080\u0093 conducted by or with the consultation of a few individuals who have a long history of work in the region (e.g., Susan Crockford; Jon Driver; Gay Frederick [formerly Calvert & Boehm]; Gregory Monks; Rebecca Wigen). Researchers at the University of Victoria and the RBCM have assembled a comprehensive comparative collection of vertebrates in the North Pacific including marine and terrestrial mammals from multiple individuals of different age and sex classes. Other universities have less comprehensive collections, but include common mammals like deer and harbor seal. Identification and quantification criteria have been reviewed and standardized for consistency. For identifications that appear questionable, as in the case of northern fur seals identified from Burrard Inlet in North Vancouver (i.e., Galdikas-Brindamour 1972), revised assessments of earlier identifications are used (i.e., Trost 2005). Faunal remains from the majority of sites included in the analysis were recovered from controlled excavations using standard 1/4\u00E2\u0080\u009D (6.35 mm) mesh. However, in one instance of salvage data recovery, mammal bones were collected by hand from excavated sediments (Site 46, Table 3.2). Mammal remains recovered from fine mesh bulk samples were not included in the analysis due to the small sample sizes and their lack of comparability with fauna from larger volumes of excavated sediment. In most cases, mammal bones were not recovered or were in very low numbers in these fine mesh bulk samples. Faunal identification data are quantified by number of identified specimens (NISP), which includes individual skeletal elements or fragments that can be confidently attributed to taxonomic categories of family, genus, or species level designation (e.g., cervidae, deer, or mule deer). Highly fragmented remains that cannot be confidently assigned to family, genus, or species level are not included in the analysis (e.g., unidentified sea mammal). \u00E2\u0080\u0093 63 \u00E2\u0080\u0093 Quantification of relative abundance is evaluated by comparing the percentage of identified specimens (NISP) from individual taxonomic categories to all other positively identified specimens (e.g., %NISP of pinnipeds). Species or genus-specific taxonomic identifications are shown in Table 3.2 but for the purpose of this analysis, specimens have been lumped into larger categories of \u00E2\u0080\u0098marine\u00E2\u0080\u0099 and \u00E2\u0080\u0098terrestrial\u00E2\u0080\u0099 mammals. The marine category includes all pinnipeds (seals and sea lions), delphinids (porpoises, dolphins), and sea otters. River otters (Lontra canadensis) are also included in the marine mammal category, as they inhabit coastal waters throughout the study area (Banfield 1974). Whales have been excluded due to the highly fragmentary nature of these large elements, which make identification and quantification problematic (Huelsbeck 1988; Monks 2001). Canids and domestic dogs (the only domesticated animal in the region) have been excluded from this analysis since it is not clear they are used for dietary consumption (Crockford 1997). Age and sex classification (based on epiphyseal fusion, eruption and wear of teeth) is not consistently recorded for most of the assemblages in this study. However, sex information is available for a small selection of the examined sites for adult-sized sea lions and fur seals (summarized in Table 3.3). Ideally, future analyses will examine incremental skeletal structures, such as growth lines in non-deciduous teeth as well as measuring the dimensions of diagnostic elements, which has been published for a handful of sites (e.g., Crockford, et al. 2002; Etnier 2002b). The compiled assemblage data have been entered into to a geographic information system (GIS) and plotted over a series of geospatial layers depicting the major river drainages, the terrestrial topography and the 200 and 500 m bathymetric contour lines (Figure \u00E2\u0080\u0093 64 \u00E2\u0080\u0093 3.2). Mammalian fauna from individual sites are represented as pie charts showing the relative abundance of mammals for different taxonomic categories. Our analysis examines the differences in the relative composition between marine and terrestrial mammals for all sites and time periods and then examines species-specific patterns for marine mammals. Figure 3.1. Map of the study area showing the distribution of archaeological assemblages used in the analysis and their chronological age-range. Results Seventy-five temporally distinct assemblages from 58 archaeological sites are included in this analysis, representing a combined total of 67,500 mammalian skeletal specimens (Table 3.2). Fifty-one assemblages are from the Strait of Georgia region, 21 are from the west coast of Vancouver Island, and three are from the Olympic Peninsula in Washington State. Site distribution is relatively more concentrated in areas impacted by modern development and sparse in other areas, particularly north of the city of Vancouver and along on southwestern Vancouver Island (Figure 3.1). Slightly more than half of the sites \u00E2\u0080\u0093 65 \u00E2\u0080\u0093 (53%, N=31) were excavated as part of academically driven research projects while the remainder were cultural resource management projects mandated under heritage legislation in the province of British Columbia. The most recent temporal period (ca. 300-1200 cal BP) contains the greatest number of assemblages (N=36) and this progressively decreases with assemblage age (1200-2400 cal BP, N=26; 2400-5000, N=11; 5000-8000, N=2). This trend reflects a sampling bias towards younger sites associated with modern sea levels and a lack of research devoted to locating and sampling earlier Holocene archaeological assemblages (cf., Fedje, et al. 2004; 2011b). Nine sites span multiple temporal periods whereas the remaining sites fall into one of the four broadly defined temporal periods (Table 3.2). Assemblage size (NISP) ranges widely from a low of five mammalian elements (Site 38) to a high of 51,937 (Site 50) but over 70% of the assemblages have more than 50 identified specimens. The largest assemblage from Ozette (Site 50), was the product of a decades-long collaborative project between the Makah tribe and a team of archaeologists from Washington State University (Huelsbeck 1994; Samuels and Daugherty 1991). Sites with small assemblages (< 50 NISP) are included in the analysis as they represent additional locations with confidently identified taxonomic information. \u00E2\u0080\u0093 66 \u00E2\u0080\u0093 Figure 3.2. Relative composition of marine and terrestrial mammals (%NISP) for all sites in the study area (n=58). Chronologically distinct periods have been combined in this figure. Spatial and Temporal Patterning Figure 3.2 illustrates the relative abundance of marine and terrestrial mammals among all examined assemblages spanning the past 8,000 years and shows a sharply defined regional difference in the use of marine mammals and terrestrial mammals between the west and east coasts of Vancouver Island. Sites in the Strait of Georgia show remarkable consistency, with terrestrial mammals comprising the majority of all mammalian assemblages. Sites on western Vancouver Island in contrast, are composed, primarily, of assemblages with more than 75% marine mammals (Figure 3.2). Notable exceptions to this pattern occur at two sites on western Vancouver Island, which contain a higher percentage of terrestrial than marine mammals. One of these sites is a rockshelter in a sheltered area of Hesquiat Harbour (Site 35) and the other is located on a highly exposed linear coastline near a river that is navigable for 10 km upstream but where the surf makes it hazardous to launch canoes in most conditions (Site 46). In the Strait of Georgia, sites that show the greatest \u00E2\u0080\u0093 67 \u00E2\u0080\u0093 percentage of marine mammals are in the Gulf Islands and the Saanich and Tsawwassen peninsulas (Sites 15-18, 21-24). Sites situated on the east coast of Vancouver Island show a consistently low relative percentage of marine mammals relative to elsewhere in the Gulf of Georgia. Shoemaker Bay (Site 45), a site at the head of an inlet, which deeply incises the western coast of Vancouver Island, has an assemblage dominated by terrestrial mammals that appears more similar to eastern Vancouver Island assemblages. Saltery Bay (Site 51), a multi-component site in the north of the study area, shows a notably high percentage of marine mammals relative to other sites in the Strait of Georgia. Figure 3.3 depicts the chronological sequence of the relative composition of marine and terrestrial mammals over the four distinct time periods. Terrestrial mammals represent a majority of the two earliest assemblages in the Strait of Georgia and this pattern follows in subsequent time periods with the notable exception of Saltery Bay (Site 51). Similarly, sites on western Vancouver Island have a majority of marine mammals and this is consistently represented over the three temporal periods with available data (Figure 3.3). These persistent compositional patterns in regionally distributed mammal assemblages indicate remarkable continuity over long periods of time. However, temporal trends within particular sites are also apparent, although more subtle. Some individual sites have suggestive chronological changes in the relative composition of marine versus terrestrial mammals. For instance, Helen Point (Site 21) shows a relative decrease in marine mammal composition over three successive time periods which speculatively correlates with greater terrestrial mammal composition on southeastern Vancouver Island (Sites 8-18). In contrast, sites on southwestern Vancouver Island and Washington State appear to show a progressive increase in marine mammal percentages over the three time periods (sites 34-50). While these trends \u00E2\u0080\u0093 68 \u00E2\u0080\u0093 are suggestive of regionalized subsistence economies, it is not within the scope of this paper to evaluate each of these possible local patterns in detail. \u00E2\u0080\u0093 69 \u00E2\u0080\u0093 Figure 3.3. Relative composition of marine and terrestrial mammals (%NISP) for all sites in the study area according to temporal period. Figure 3.4 shows the same temporal sequence of analyses shown in Figure 3.3, but depicts only the relative composition of three most common pinniped species (harbour seal, Steller sea lion, northern fur seal) and sea otter. This analysis purposefully excludes terrestrial and cetacean (delphinid) specimens, which correspondingly reduces sample size \u00E2\u0080\u0093 70 \u00E2\u0080\u0093 and alters the representation of relative percentage (i.e., %NISP refers to just these species). Nevertheless, this analysis indicates that assemblages on western Vancouver Island are predominantly composed of northern fur seal, followed by Steller sea lion and harbor seal. Northern fur seals are strikingly absent from all sites in the Strait of Georgia and harbor seals predominate the marine mammal assemblages in the Strait of Georgia followed distantly by Steller sea lions. Sea otters are present at eight sites on western Vancouver Island and Washington and comprise substantial proportions of four of these assemblages (Sites 34, 36, 44, & 45). Sea otters are present at five sites in the Strait of Georgia dating to between 2400 and 300 yr BP (Sites 13, 23, 52, and 53), which contrasts with the findings of Hanson and Kusmer (2001) who observed a near absence of sea otters in this region. Where sea otters are present, they represent small proportion of the assemblages except for one small marine mammal assemblage (NISP=4), which is composed entirely of sea otter (Site 13). Steller sea lions also are present at sites on southern and western Vancouver Island but are found in greater densities on western Vancouver Island and in lower percentages along the Strait of Georgia (Figure 3.3). California sea lions are identified in only six sites in the study area, all on the exposed Pacific coast (Table 3.2). Elephant seals are similarly present in small numbers at four sites in the western portion of study area (Sites 41, 42, 48, and 49). Guadalupe fur seals, whose skeletal morphology closely resembles the northern fur seal (Etnier 2002b), are only identified at one site, which also has the largest identified assemblage (Ozette, Site 50). \u00E2\u0080\u0093 71 \u00E2\u0080\u0093 \u00E2\u0080\u0093 72 \u00E2\u0080\u0093 Figure 3.4. Relative abundance of marine versus terrestrial mammals for sites dating to within the four examined periods (5000-8000, 2500-5000, 1200-2500 and 300-1200 BP). \u00E2\u0080\u0093 73 \u00E2\u0080\u0093 Table 3.2 Site Names, Locations, References, and Number of Identified Specimens (NISP) used in the analysis. All Mammal totals exclude whales and domestic dogs and canids. No. Region* Site # Period Site Name Reference Harbor S. Steller SL N Fur Seal C. Sea Lion Guadelupe Fur Seal Elephant seal Sea Otter Terrestrial Mammals All Mammals 1 ecVI DjSf-13 1200-2400 Buckley Bay Wigen 1980 4 10 0 0 0 0 0 221 235 2 ecVI DjSf-14 1200-2400 Tsable River Wigen 1980 2 4 0 0 0 0 0 194 203 3 ecVI DkSg-2 300-1200 Sandwick midden Wigen 1980 citing Capes,1964 3 0 0 0 0 0 0 64 75 4 ecVI DiSc-1 300-1200 Little Qualicum River Bernick and Wigen 1990 18 1 0 0 0 0 0 243 263 5 ecVI DiSc-26 1200-2400 Qualicum Beach Golf Course Willows et al. 2008 0 0 0 0 0 0 0 10 11 6 ecVI DhSb-3 1200-2400 Dogwood St., Parksville Wilson et al. 2006 0 0 0 0 0 0 0 12 12 7 ecVI DhRx-16 1200-2400 Departure Bay Wilson and Crockford 1994:118 6 0 0 0 0 0 0 733 743 8 scVI DcRu-74 1200-2400 Esquimalt lagoon Wet site McKechnie, 2004:51, Table 19 3 0 0 0 0 0 0 63 67 9 scVI DcRu-2 300-1200 Esquimalt Lagoon Hanson 1991:Table 6 citing Stevenson 1978 9 3 0 0 0 0 0 254 266 10 scVI DcRu-2 300-1200 Esquimalt/ Ft Rodd Hill Crockford, 1997b 17 0 0 0 0 0 0 337 358 11 scVI DcRu-78 300-1200 Ft Rodd Hill Mitchell 1981 0 5 0 0 0 0 0 88 96 12 scVI DcRu-136 300-1200 Esquimalt Harbour Wigen 2002 5 0 0 0 0 0 0 53 58 13 scVI DcRu-4 300-1200 Kosapsom Stewart and Wigen 2003 0 0 0 0 0 0 4 19 23 14 scVI DcRu-75 300-1200 Dallas Road Wilson et al. 2003:Table 28 14 0 0 0 0 0 0 126 143 15 scVI DcRt-16 300-1200 King George Terrace Wilson et al. 2004:107, Table 38 3 4 0 0 0 0 0 111 119 16 scVI DcRu-71 1200-2400 Eagles Nest Eldridge 2000:40 11 2 0 0 0 0 0 54 77 17 scVI DdRu-18 300-1200 Eventide Rd. Weathers et al. 2007:29, Table 6 2 5 0 0 0 0 0 37 44 18 scVI DdRu-5 1200-2400 Patricia Bay Kanipe et al. 2006:264, Table 46 9 10 0 0 0 0 0 145 165 \u00E2\u0080\u0093 74 \u00E2\u0080\u0093 No. Region* Site # Period Site Name Reference Harbor S. Steller SL N Fur Seal C. Sea Lion Guadelupe Fur Seal Elephant seal Sea Otter Terrestrial Mammals All Mammals 19 Gulf Is DgRu-3 1200-2400 Dionisio Point Ewonus, 2006:39, Table 2 3 0 0 0 0 0 0 88 96 20 Gulf Is DfRv-106 300-1200 Galiano Island Mason et al. 1995:21, Table 4 0 1 0 0 0 0 0 20 21 21.1 Gulf Is DfRu-8 300-1200 Helen Point Boucher, 1976: Assemblages 12 ,13, & 14 12 1 0 0 0 0 0 180 194 21.2 Gulf Is DfRu-8 1200-2400 Helen Point Boucher, 1976: Assemblages 5 & 6 18 5 0 0 0 0 0 93 119 21.3 Gulf Is DfRu-8 2400-5000 Helen Point Boucher, 1976: Assemblages 1, 2, 3, & 4 187 54 0 0 0 0 1 360 664 22 Gulf Is DhRt-1 300-1200 Pender Canal Hanson, 1995:34 62 0 0 0 0 0 0 406 525 23 Gulf Is 45SJ169 1200-2400 Decatur Island-169 Lyman, 2003 24 0 0 0 0 0 1 154 179 24 Gulf Is 45SJ165 1200-2400 Decatur Island-165 Lyman, 2003 10 0 0 0 0 0 0 77 87 25.1 Mnlnd DgRr-2 300-1200 Tsawwassen Butler & Campbell 2004:353, citing Kusmer, 1994 4 0 0 0 0 0 0 16 20 25.2 Mnlnd DgRr-2 1200-2400 Tsawwassen Butler & Campbell 2004:353, citing Kusmer, 1994 1 0 0 0 0 0 0 5 6 26 Mnlnd DgRr-1 1200-2400 Crescent Beach Hanson 1991:Table 6, citing Ham 1982 2 1 0 0 0 0 0 805 820 27 Mnlnd DhRr 18 1200-2400 Point Grey Coupland 1991:89 0 0 0 0 0 0 0 16 16 28 Mnlnd DhRr-6 300-1200 Belcarra (Tum-tu-may-whueton) Trost 2005:116 citing Charlton 1977:226 &Galdikas-Brindamour 1972 10 0 0 0 0 0 0 227 239 29 Mnlnd DhRr-8 300-1200 Cates Park (Whey-Ah-Wichen) Trost 2005:116 citing Williams 1974 14 0 0 0 0 0 0 227 251 30 Mnlnd DhRr-18 300-1200 Cove Cliff (Say-Umiton) Trost 2005:112 7 0 0 0 0 0 0 123 133 31.1 Mnlnd DgRr-2 300-1200 St. Mungo - late Boehm 1973: Table VIII 4 0 0 0 0 0 0 15 19 31.2 Mnlnd DgRr-2 1200-2400 St. Mungo - mid Boehm 1973: Table VIII 3 0 0 0 0 0 0 6 9 31.3 Mnlnd DgRr-2 2400-5000 St. Mungo - early Boehm 1973: Table VIII 23 0 0 0 0 0 0 212 235 32.1 Mnlnd DgRr-6 1200-2400 Glenrose Cannery-late Imamoto 1976: Table 2.1 6 0 0 0 0 0 0 70 76 32.2 Mnlnd DgRr-6 2400- Glenrose Cannery- Imamoto 1976: Table 2.1 23 0 0 0 0 0 0 193 216 \u00E2\u0080\u0093 75 \u00E2\u0080\u0093 No. Region* Site # Period Site Name Reference Harbor S. Steller SL N Fur Seal C. Sea Lion Guadelupe Fur Seal Elephant seal Sea Otter Terrestrial Mammals All Mammals 5000 mid 32.3 Mnlnd DgRr-6 5000-8000 Glenrose Cannery-early Imamoto 1976: Table 2.1 3 0 0 0 0 0 0 32 35 33 Mnld DhRp-17 1200-2400 Port Hammond Rousseau et al. 2003:101 3 0 0 0 0 0 0 264 267 34 wcVI DiSo-9 1200-2400 Hesquiat DiSo9Combined Calvert 1980: Tables 43 & 47 38 1 100 2 0 0 68 88 328 35 wcVI DiSo-16 300-1200 Hesquiat rockshelter Calvert, 1980: Table 39 0 0 0 0 0 0 0 66 126 36 wcVI DiSo-1 300-1200 Hesquiat Village combined Calvert, 1980: Tables 51, 55, 59, 63, & 67 86 41 228 29 0 0 90 155 844 37 wcVI DgSl-67 300-1200 Chesterman Beach Wilson 1994: Table 5 0 0 25 12 0 0 0 6 52 38 wcVI DgSk-1 300-1200 Long Beach Wigen, 2003: Table 2 0 1 0 0 0 0 0 0 5 39 wcVI DfSj-100 2400-5000 Little Beach Weathers et al. 2008; Wigen, 2008: Table 5 0 1 19 0 0 0 0 3 25 40 wcVI DfSj-57 1200-2400 Spring Cove, Ucluelet Spady& Wigen 2008: Table 15 1 0 49 0 0 0 0 5 57 41 wcVI DfSi\u00E2\u0080\u00935 300-1200 Ma'acoah Monks 2006:224 32 2 84 14 0 1 6 71 224 42.1 wcVI DfSi-16+17 300-1200 Tsishaa Late McKechnie 2007a: Tables 17 & 18, including Frederick & Crockford 2005 28 13 214 1 0 0 3 29 344 42.2 wcVI DfSi-16+17 1200-2400 Tsishaa Mid McKechnie 2007a: Tables 17 & 18, including Frederick & Crockford 2005 14 0 49 0 0 1 1 25 109 42.3 wcVI DfSi 16+17 2400-5000 Tsishaa Early McKechnie 2007a: Tables 17 & 18, including Frederick & Crockford 2005 2 7 35 0 0 0 2 131 408 43 wcVI DfSi-26 300-1200 Clarke Island McKechnie 2007: Tables 17 & 18 6 7 56 0 0 0 0 9 80 44.1 wcVI DfSh-7 300-1200 Huu7ii late Frederick et al. 2006: Tables 6, 8 & 10 21 24 46 0 0 0 12 49 298 44.2 wcVI DfSh-7 1200-2400 Huu7ii mid Frederick et al. 2006: Tables 6, 8 & 10 5 1 1 0 0 0 0 3 17 44.3 wcVI DfSh 7 2400-5000 Huu7ii Early Frederick et al. 2006: Tables 6, 8 & 10 9 16 39 0 0 0 3 65 169 45.1 wcVI DhSe-2 300-1200 Shoemaker bay Late Calvert and Crockford 1982:186 75 0 11 0 0 0 0 781 889 45.2 wcVI DhSe-2 1200-2400 Shoemaker bay Mid Calvert and Crockford 1982:186 13 2 0 0 0 0 0 182 201 45.3 wcVI DhSe 2 2400- Shoemaker bay Calvert and Crockford 1982:186 7 0 1 0 0 0 0 91 101 \u00E2\u0080\u0093 76 \u00E2\u0080\u0093 No. Region* Site # Period Site Name Reference Harbor S. Steller SL N Fur Seal C. Sea Lion Guadelupe Fur Seal Elephant seal Sea Otter Terrestrial Mammals All Mammals 5000 Early 46 wcVI DeSf-6 300-1200 Klanawa McKechnie, 2007b:7 7 0 8 0 0 0 0 59 75 47 wcVI DeSf-2 300-1200 Tsuxwkwaada McKechnie 2005: Table 1 3 0 185 0 0 0 0 3 191 48 WA 45CA213 1200-2400 Hoko R. Wet/Dry site Croes 1995: 71 7 1 3 0 0 0 0 10 22 49 WA 45CA21 300-1200 Hoko R. Rockshelter Wigen 2005: Table 4.4 114 40 2425 0 0 1 40 423 3318 50 WA 45CA24 300-1200 Ozette Huelsbeck 1994:28 w additions from Etnier 2007:200 377 10 47296 1 34 2 501 708 51007 51.3 Strait of Georgia DkSb-30 5000-8000 Saltery Bay Pegg et al. 2007: Table 20 17 0 0 0 0 0 0 108 129 51.2 Strait of Georgia DkSb-30 2400-5000 Saltery Bay Pegg et al. 2007: Table 20 6 2 0 0 0 0 0 76 163 51.1 Strait of Georgia DkSb-30 300-1200 Saltery Bay Pegg et al. 2007: Table 20 3 0 0 0 0 0 0 14 27 52 Gulf Is 45-SJ-24 300-1200 English Camp Pegg 1999:69 20 0 0 0 0 0 1 310 372 53 Strait of Georgia DKSf-4 300-1200 Comox Harbour Site Simonsen 1990:35, Table 11 2 7 0 0 0 0 2 262 292 54 wcVI DfSg-2 300-1200 Aguilar House Engish 2006:28 2 4 19 0 0 0 2 6 36 55 Gulf Is DgRv-2 1200-2400 Shingle Point Matson et al. 1999:61-71 4 0 0 0 0 0 0 107 116 56 Mnlnd DhRt-6 2400-5000 Locarno Beach Brolly and Muir 1993:52, Table 12, excluding 2mm mesh 1 0 0 0 0 0 0 67 68 57 Gulf Is DfRu-3 1200-2400 Harbour House, Saltspring Brolly et al. 1993:72, Table 8 0 1 0 0 0 0 0 127 129 58.1 ecVI DiSe-10 300-1200 Denman - Southern Rckshltr Eldridge 1987:5-57, Table 5-9 0 0 0 0 0 0 0 7 7 58.2 ecVI DiSe-10 2400-5000 Denman - Northern Rckshltr Eldridge 1987:5-59, Table 5-10 2 0 0 0 0 0 0 57 59 58.3 ecVI DiSe-10 2400-5000 Denman - Blufftop Hunting Ritual Eldridge 1987:5-67, Table 5-16 1 0 0 0 0 0 0 26 27 * ecVI-East Coast Vancouver Island, scVI-South Coast Vancouver Island, Gulf Is-Gulf Islands, wcVI-West Coast Vancouver Island, MnInd-Mainland of British Columbia, WA-Washington State. \u00E2\u0080\u0093 77 \u00E2\u0080\u0093 Age and Sex Data Due to the striking disparity between the archaeological and modern distributions of northern fur seal, age and sex data for this once commercially significant species have been discussed elsewhere in great detail (Burton, et al. 2001; Calvert 1980; Crockford, et al. 2002; Etnier 2002a, 2007; Gifford-Gonzalez, et al. 2005; 2011; Gustafson 1968; Moss, et al. 2006; Newsome, et al. 2007). These studies have identified adult males and females as well as immature pups less than 6 months of age in a variety of sites in the study area. Combining archaeological data from Ts\u00E2\u0080\u0099ishaa (Site 42) with historic accounts from schooner captains, aboriginal hunters, and merchants involved in the sealing industry of the 1880s (Swan 1883, 1887), Crockford et al. (2002:170) argue for the presence of a non-migratory breeding population of fur seals capable of giving birth in kelp-beds or kelp rafts at sea. Subsequent studies have used stable isotopes to posit the existence a mid-latitude breeding distribution in the Pacific Northwest (Newsome, et al. 2007). However, definitive evidence of a terrestrially-based northern fur seal rookery remains to be identified. Age and sex data from few species other than fur seal have been compiled. For sites in the study sample, a small portion of Steller sea lion skeletal remains from a limited number of sites has been confidently sexed. Table 3.3 shows the breakdown of sexed elements for sites where the information was available from controlled excavations. Although the sample is small (NISP=19), 95% of the sexed elements are from adult males with only two elements out of 40 either unknown or female. This pattern is similar to data from the Oregon Coast, which indicates that Steller sea lions from archaeological contexts are primarily adult males (Lyman 2003b). The incidence of females and/or juveniles of either \u00E2\u0080\u0093 78 \u00E2\u0080\u0093 sex are very low with only one definitively female element found on southeast Vancouver Island (Table 3.3). Additional Steller sea lion sex and age information comes from fauna collected from archaeological contexts during monitoring of construction projects (Table 3.3). Monitoring samples are collected without screening and have only general provenience. However, males dominate four large assemblages from the southern Strait of Georgia region. For example, at Patricia Bay (Site 18), at least 74 male Steller sea lion elements have been identified in contrast to only one female and three juveniles of indeterminate sex (Kanipe, et al. 2007; Wigen 2007b:24). Explanation for this pattern of male dominance could relate to seasonal availability of Steller sea lions and the differential behavior of the females and males at rookeries and haulout sites and/or preference on the part of the human hunters. Steller sea lions breed at communal rookeries in northern British Columbia during the summer but spend the winter at haulout sites throughout British Columbia (Olesiuk 2009a). Since the first survey in British Columbia in 1913 (Newcombe and Newcombe 1914), the most southerly sea lion breeding rookery is located on the north end of Vancouver Island. Non-breeding animals of both sexes use a limited number of established year-around haulouts found mainly on the west coast of Vancouver Island. In the winter, individuals of both sexes use a wider range of haulouts including the Strait of Georgia and southern Vancouver Island. This pattern suggests both males and females should be available to hunters in the winter, so it appears seasonal distribution does not explain the hunting pattern. Recent studies of the behavioral responses to human disturbance of hauled out Steller sea lions note that females and pups are the first to leave followed by the sub-adult males and, finally, the bulls (COSEWIC 2003; Szaniszlo 2005). The dominance of males in the \u00E2\u0080\u0093 79 \u00E2\u0080\u0093 archaeological sites may be a result of this behavioral pattern, with the males being easier for the hunters to target. Additionally, haulouts of \u00E2\u0080\u0098bachelor males\u00E2\u0080\u0099 are commonly present near rookeries and may have also been targeted by hunters. The extremely large size of males and these behavioral characteristics suggest that males were hunted while onshore rather than offshore as documented for fur seals. Table 3.3 Sex Distribution of Steller Sea Lion Remains from Selected Sites. Region* Site Total NISP Male Female Unknown Reference ecVI DiSe-7, Deep Bay 1 1 Wilson et al. 2004b:44 scVI DcRt-16, King George Terrace 4 4 Wilson et al. 2004a:94 scVI DdRu-5, Pat Bay 10 10 Kanipe et al. 2006:269 scVI DcRu-136, Esquimalt Lagoon 2 2 Wigen 2002 Gulf Is DfRv-106, Galiano Island 1 1 Mason et al. 1995:22 wcVI DiSo9-II, Hesquiat 1 1 Calvert 1980:155 wcVI DiSo1-I, Hesquiat 13 2 11 Calvert 1980:156 wcVI DiSo1-II, Hesquiat 1 1 Calvert 1980:157 wcVI DiSo1-III, Hesquiat 12 3 9 Calvert 1980:158 wcVI DiSo1-IV, Hesquiat 14 2 1 11 Calvert 1980:159 wcVI DjSf100, Little Beach 1 1 Wigen 2008:11 wcVI DfSi-26, Clarke Island 7 4 3 McKechnie 2007a:30 wcVI DfSi-16, Ts\u00E2\u0080\u0099ishaa 20 7 13 Unpublished data in Frederick and Crockford 2005 and McKechnie 2007 wcVI DfSh-7, Huu7ii 9 7 2 Unpublished data in Frederick et al. 2006 Total 96 44 5 47 % Sexable 90% 10% * ecVI = East Coast Vancouver Island, scVI = South Coast Vancouver Island, Gulf Is = Canadian Gulf Islands, wcVI = West Coast Vancouver Island, \u00E2\u0080\u0093 80 \u00E2\u0080\u0093 Harbor seals and sea otters are more difficult to sex, as sexual dimorphism is less extreme. In most cases, no information has been recorded about the sex of individuals. Age for harbor seals has been recorded in a few site reports (e.g., Frederick and Crockford 2005; Frederick, et al. 2006a; Wigen 2007b), indicating that harbor seals of all ages were regularly hunted in all areas. No age or sex information is available for the few identified California sea lion or elephant seal elements. Steller sea lions are present in sites along the southwestern portion of the study area, suggesting they may have been moderately abundant in the ancient environment. Elephant seals, California sea lions, and Guadalupe fur seals rarely occur in the assemblages, suggesting they were not commonly present at this latitude. However, taxonomic identifications of male California sea lions may be confused with female Steller sea lions (personal observations) and similarly, Guadalupe fur seals appear similar to northern fur seals (Etnier 2002b). Discussion This analysis has identified distinct regional patterns in marine and terrestrial mammal assemblages from archaeological sites on western Vancouver Island and the Strait of Georgia. These differences provide a basis for inferring that aboriginal hunting practices dramatically differed in focus between these two regions. In particular, aboriginal peoples in the Strait of Georgia region did not hunt marine mammals to the same degree as was practiced by aboriginal peoples along the Pacific coast of Vancouver Island and Washington state. Rather, terrestrial mammals appear to be relatively more important in the Strait of Georgia region and relatively less so on the exposed Pacific coast. \u00E2\u0080\u0093 81 \u00E2\u0080\u0093 These geographic patterns are robust at least over the past 5,000 years and have implications for a number of perspectives in anthropology and ecology. Of particular anthropological relevance is the apparent coherence and continuity of aboriginal marine and terrestrial hunting patterns within each of the two areas. This difference is most strongly apparent in the relative composition of marine versus terrestrial mammals (Figure 3.2 and Figure 3.3) and among the individual pinniped species (Figure 3.4). For specific localities within these two regions, variation in marine mammal use also appears to be spatially patterned, such as the slightly higher abundance of marine mammals in assemblages along southeastern Vancouver Island and in the Gulf Islands (Figure 3.2). These archaeological patterns are an outcome of the direct participation of generations of aboriginal people in this marine ecosystem. Such patterns reflect a continuity of skillful cultural knowledge and practical engagement with these marine environments. These regional differences are consistent with a host of ethnographic accounts and observations (e.g., Boas 1887; McMillan 1999; Suttles 1952), which identify the aboriginal peoples who occupy these two regions as belonging to two culturally and linguistically distinct groups (Wakashan speaking peoples [Nuu-chah-nulth, Ditidaht, Makah] and Salish speaking peoples). During the post-contact era (ca. AD 1774-1860) and continuing among First Nations today, these cultural differences include distinct patterns of residence, conventions of artistic representation, ritual practice, and a host of other regionally unique forms of cultural expression (ibid.). Our analysis provides insight into an aspect of these rich cultural histories, suggesting that cultural distinctions between regions and cultural similarities within regions have great antiquity in regards to hunting practice. Indeed, since our data relate to an important subsistence activity (hunting), such regional expressions can \u00E2\u0080\u0093 82 \u00E2\u0080\u0093 be hypothesized to have contributed structure to cultural history and community identity. This is illustrated in a 1922 quote discussing the biography of Tom Sayach\u00E2\u0080\u0099apis (ca. 1835\u00E2\u0080\u00931927), a respected cultural historian of the Tseshaht, a Nuu-chah-nulth Nation on southwestern Vancouver Island: Tom ate very little meat of land mammals in his early days. Indeed, like most of the [west] Coast people, he had a prejudice against deer meat and it was not until as a middle-aged man, he had come into contact with some of the deer hunting tribes of the interior of the island, that he learned to prize it, though to this day, venison has not for him the toothsome appeal of a chunk of whale meat (Sapir 1922:304). From an ecological perspective, such strong geographic patterning as seen through a lens of human hunting activity is relevant for considering the long-term ecological role of marine mammal species in the study area. Our data represent a unique and otherwise unattainable record of marine mammal abundance and distribution over the Holocene. This information helps contextualize contemporary marine mammal ecology by greatly extending temporal and geographic knowledge of these important species. In this regard, it is notable that the three most abundant pinnipeds in the archaeological assemblages (northern fur seals, harbor seals, and Steller sea lions) are also the three most numerically abundant in the contemporary environment (Table 3.1). The similarity in rank-ordered abundance between our archaeological data and contemporary pinniped population estimates indicates these three species have exhibited remarkable demographic resilience to human harvesting. The stable long-term trends observed in the archaeological data stand in contrast to the dynamic historical events documented over the past 200 years which resulted in the rapid extirpation of the sea otter (ca. AD 1774-1811) and culling of pinniped populations (ca. AD 1868-1970). In this context, it seems important to note that the sudden cessation of marine \u00E2\u0080\u0093 83 \u00E2\u0080\u0093 mammal hunting over the past 40 years appears to be unprecedented in a record spanning at least the past 5,000 years. The closest possible parallels to this sudden reduction in hunting intensity and corresponding population increase likely occurred after European diseases dramatically reduced aboriginal populations in the early contact era (ca. AD 1774-1811) and again in the mid-19th century, just prior to the beginning of the commercial fur seal industry (Boyd 1999). While it is striking to note the ongoing demographic consequences of historic-era commercial hunting, equally important are possible behavioral and ecological changes that occurred among targeted taxa over the Holocene. For instance, our compiled archaeological results indicate that aboriginal peoples regularly hunted fur seals for a minimum of 5,000 years on the exposed Pacific coasts of Vancouver Island and northwestern Washington. These observations correspond with the locations of fur seal hunting mentioned in historic-era harvest log-books (Olesiuk 2009b:42) indicating that the continental shelf off southwestern Vancouver Island has long been a major focus of fur seal foraging activity. However, many of these same historic-era harvest records (ca. AD 1891-1911) also indicate a strongly seasonal presence of fur seals at this latitude between January and May (Olesiuk 2009b:54), which contrasts with zooarchaeological data including osteometric evidence for localized breeding (Crockford, et al. 2002; Etnier 2002a; Gustafson 1968), and geochemical evidence for non-migratory mid-latitude foraging and prolonged pup weaning dating from the mid-Holocene (Newsome, et al. 2007). Tantalizingly suggestive historic accounts compiled from interviews with commercial fur sealers conducted a decade prior to the historic log-book-based records mentioned above, describe non-migratory behavior involving giving birth at sea in kelp beds and observing fur seals in the Strait of Georgia during late-\u00E2\u0080\u0093 84 \u00E2\u0080\u0093 summer when the migratory populations are in the Bering Sea (Crockford et al. 2002 citing Swan 1883). These data suggest historic-era commercial harvests may have eliminated or altered a behaviorally unique sub-population of fur seals and archaeological data yield important insight into the previously more diverse ecology of this currently strictly pelagic pinniped. Although the absolute number of animals harvested cannot be quantified with our zooarchaeological data alone, our analysis has several implications concerning potential human impacts on pinniped and sea otter populations. Firstly, our data indicate a regional focus on specific pinniped species, which implies that aboriginal peoples in these regions targeted specific taxa with regularity (e.g., annually or seasonally) as suggested by ethnographic accounts (e.g., Arima 1988; Suttles 1952). Secondly, our data represent only a small percentage of the approximately 4,000 currently documented coastal shell midden sites present in the study area (British Columbia Archaeology Branch n.d.). Thus, the collective impact of aboriginal hunting may be considerably greater than previously recognized if these regional patterns apply to several thousand unexamined sites. Finally, the temporal and spatial consistency of hunted marine mammals in these two regions indicates a degree of stability and continuity to aboriginal resource harvesting practices, which demonstrates the capacity for sustained harvesting in antiquity (Etnier 2007). Repetitive and consistent human participation in an ecosystem provides structure to ecological interactions (Bal\u00C3\u00A9e 2006; Hobbs and Fowler 2008; Liu, et al. 2007). However, determining and/or identifying \u00E2\u0080\u0098human impacts\u00E2\u0080\u0099 in such an archaeological context is challenging as these ecosystems may represent the end-point of centuries or millennia of human hunting. Future analyses that can incorporate more fine-grained osteometric \u00E2\u0080\u0093 85 \u00E2\u0080\u0093 information on age and sex distribution will yield much greater insight into these ecological impacts. Considering the ecological dimensions of human subsistence activity in coastal environments, Fedje et al. (2004) offer a distinction that seems to encompass the regional differences we observe in our analysis of mammalian hunting practice; the concept of \u00E2\u0080\u009Cmaritime\u00E2\u0080\u009D and \u00E2\u0080\u009Ccoastal adaptations.\u00E2\u0080\u009D These authors define a \u00E2\u0080\u009Cmaritime adaptation\u00E2\u0080\u009D as on which is: \u00E2\u0080\u009Cheavily reliant upon marine and coastal resources for the majority of subsistence needs. In contrast, a \u00E2\u0080\u0098coastal adaptation\u00E2\u0080\u0099 involves the fluent use of coastal and marine resources for at least some, and probably a significant portion, of the subsistence needs of a group of people. Coastal adaptations should therefore include the possibility of considerable use of terrestrial resources and inland areas\u00E2\u0080\u009D (Fedje, et al. 2004:112). This synthetic statement helps characterize the regional differences observed in our analysis suggesting that aboriginal peoples in the Strait of Georgia \u00E2\u0080\u0098adapted\u00E2\u0080\u0099 themselves to terrestrial mammal hunting more extensively than aboriginal peoples on western Vancouver Island. This observation is consistent with ethnographic and archaeological information from Coast Salish peoples in the Strait of Georgia region where widespread aboriginal burning practices aimed to enhance habitat for economically important plants such as camas and berries as well as entice foraging ungulates into recently burned areas with young re-growth (Brown and Hebda 2002; Suttles 1987; Weiser and Lepofsky 2009). Conversely, on western Vancouver Island, the logistics of hunting pelagic pinnipeds likely facilitated other forms of offshore subsistence activities (e.g., deep water fishing or birding) as well as voyaging and trading or vice versa (Arima 1988; McMillan 1999). \u00E2\u0080\u0093 86 \u00E2\u0080\u0093 However, while the concept of \u00E2\u0080\u009Cadaptation\u00E2\u0080\u009D (defined as \u00E2\u0080\u0098the suiting of one thing to another\u00E2\u0080\u0099) may help characterize human hunting activity expressed over thousands of years, a term that is more appropriately scaled to human lifetimes is \u00E2\u0080\u0098tradition\u00E2\u0080\u0099 (defined as \u00E2\u0080\u0098the act of handing down\u00E2\u0080\u0099). Considering long-term hunting practices as \u00E2\u0080\u0098traditions\u00E2\u0080\u0099 accommodates the intergenerational transference of complex hunting knowledge in a way that anthropological uses of \u00E2\u0080\u0098adaptation\u00E2\u0080\u0099 often under-specify or assume outright (cf., Moss 2008). The regionally consistent patterns of terrestrial and marine mammal use were additionally constrained or enabled by environmental differences between the two regions. For instance, the open prairie oak savanna woodlands of the Strait of Georgia versus the very moist dense coastal rainforest on western Vancouver Island (Suttles 1987) or the higher primary productivity on the exposed Pacific coast versus the Strait of Georgia (Ware and Thomson 2005). However, while it is essential to consider the environmental factors that may contribute how aboriginal peoples may have focused on marine or terrestrial mammals, these factors cannot adequately explain how or why generations of people choose to participate in an ecosystem in a particular manner. Rather, explanations that posit an environmental imperative for human action risk imposing a deterministic assumption about human actions when that is not the case. Thus, although the relative contribution of hunting was undoubtedly constrained or enabled to some degree by the presence or absence of animals in different regions within the study area, human agency cannot be ignored. Rather than a simple consequence of available resources, social factors critically underpin the hunting practice in ways that may be much more meaningful than the simple availability of resources. Hunting was a conscious decision first and then a reflection the environmental availability. \u00E2\u0080\u0093 87 \u00E2\u0080\u0093 In contrast to an environmentally determined model of human subsistence, aboriginal peoples the Northwest Coast are increasingly recognized as highly active in \u00E2\u0080\u009Cmanaging\u00E2\u0080\u009D subsistence resources through explicit cultural definitions of harvesting practice and etiquette including selective harvesting, seasonal restrictions on use and/or consumption, and proprietorship over resources which was contingent on sustained productivity (Berkes and Turner 2006; Hunn, et al. 2003; Moss 2011a; Trosper 2009; Weiser and Lepofsky 2009). The data presented here suggest that generations of people actively chose to pursue an elaborate, highly structured, and specialized activity \u00E2\u0080\u0093 primarily pelagic marine mammal hunting or primarily terrestrial mammal hunting in different regions. These hunting traditions embody the skill, training, and cooperation, which develop over years, and decades among groups of individual hunters and which are passed on to subsequent generations in a community of continuous social practice (cf., Bourdieu 1977). The consistency and continuity with which certain species were utilized in the archaeological assemblages over 5,000 years suggests that similar cultural practices informed aboriginal peoples\u00E2\u0080\u0099 hunting behavior in the archaeologically documented past. Conclusions We posed three questions at the beginning of this chapter: \u00E2\u0080\u00A2 What marine mammals did aboriginal people in Southern British Columbia most commonly utilize? \u00E2\u0080\u00A2 How similar or how different are species occurrences and proportions relative to today? \u00E2\u0080\u00A2 Is there evidence of specialized or regional hunting traditions and if so, what might have been the potential impacts of these activities on the ancient marine ecosystem? \u00E2\u0080\u0093 88 \u00E2\u0080\u0093 In answer to the first question, fur seals, harbor seals, and Steller sea lions respectively appear to be the most heavily utilized pinnipeds on the southern BC coast over the last 8,000 years. Sea otters appear to be utilized by peoples on western Vancouver Island and the Olympic Peninsula, and are present only in low numbers at several sites in the Strait of Georgia. The least utilized pinnipeds appear to be California sea lions, elephant seals, and Guadalupe fur seals, which are rare but are found in sites along the southwestern margin of the study area. These are relative differences however, and thus even small relative percentages may represent large numbers of harvested animals. In regards to the second question, our data expand knowledge of the range and relative composition of marine mammal species that are currently threatened or endangered (Table 3.1). Examining the archaeological distribution of these species provides context for understanding their recovery since historical over-exploitation and the extent to which aboriginal peoples may have influenced the marine ecosystem before the historic-era fur trade. Considering the fact that the 58 examined sites represent less than 1% of the approximately 4,000 currently documented coastal shell midden sites present in the study area (British Columbia Archaeology Branch n.d.), the perspective gained here hints at the potential cumulative influence of large indigenous human populations in the region. The extent of this influence is unknown at this point, but did not appear to have reached the stage where people dramatically change their hunting patterns. Such basic insights help refine our understanding of the influence of human activity on the environment and, conversely, the influence of animals on human activity. And finally, this study indicates that aboriginal people on the outer coast of Vancouver Island and Washington state developed a specialized hunting tradition targeting \u00E2\u0080\u0093 89 \u00E2\u0080\u0093 northern fur seals. This conclusion is further supported by several ethnographic and historic sources (e.g., Arima 1988; Sapir and Swadesh 1955; Swan 1887; Waterman 1920) and by artifactual data from pre-contact archaeological sites in the study area (cf., McMillan 1999; Samuels and Daugherty 1991). Ethnographically, such hunting activity involved paddling considerable distances offshore, sometimes beyond sight of land, to stealthily approach \u00E2\u0080\u0098sleeping\u00E2\u0080\u0099 fur seals to within range for launching a harpoon. Such specialized journeys entailed considerable physical risks and logistical preparation. Ethnographic accounts emphasize how participants ritually engaged in elaborate physical and spiritual training and how hunters utilized specifically designed canoes, sharpened paddles, unique harpoons, yew-wood throwing sticks, seal skin floats to tire harpooned animals, as well as long lengths of cedar-bark and elk-hide rope (Arima 1988; Sapir and Swadesh 1955; Waterman 1920). In contrast to western Vancouver Island and northwestern Washington, aboriginal peoples in the Strait of Georgia appear to have developed marine mammal hunting traditions that targeted harbor seals and to a lesser extent Steller sea lions. As discussed previously, harbor seals are non-migratory, nearshore foragers who regularly \u00E2\u0080\u0098haul-out\u00E2\u0080\u0099 on land and are the most common seal present on the coast today (Department of Fisheries and Oceans Canada 2010; Jeffries, et al. 2003). Ethnographic accounts describe hunting \u00E2\u0080\u0098hauled-out\u00E2\u0080\u0099 harbor seals from boats, by hand with clubs, or by stealthily placing nets in the water below haul-out areas, particularly outside sea caves (Suttles 1952; Waterman 1920). Steller sea lions, on the other hand, occur seasonally in the Strait of Georgia during winter and spring when herring aggregate prior to spawning (COSEWIC 2003). Ethnographically, sea lions were hunted by numerous Coast Salish peoples although a few groups were recognized as particularly skilled specialists (Suttles 1952). Ethnographic descriptions of hunting both \u00E2\u0080\u0093 90 \u00E2\u0080\u0093 Steller sea lions and harbor seals were associated with elaborate forms of physical and spiritual preparation, which occurred prior to and during the hunt and were observed by both hunters and their wives (Elmendorf 1960; Suttles 1952). While hunting practices documented in archaeological contexts represent only a small portion of the rich ancestral pasts of coastal First Nation communities, zooarchaeological data represent a readily available source of information that broadens our contemporary understanding of ancient coastal lifeways and cultural environments. Such ecologically significant marine data are essentially non-existent outside of archaeological contexts and future research will benefit from additional analyses of new sites, regions, and time periods as well as more detailed osteometric observations. The compilation of zooarchaeological information explored in this paper suggest coherent patterns indicating the persistence of hunting traditions practiced continuously over 5,000 years. These patterns help unravel the complex interwoven cultural and ecological histories and environments on the Northwest Coast. \u00E2\u0080\u0093 91 \u00E2\u0080\u0093 Chapter 4. Re-Calibrating Archaeological Chronologies on the Northern Northwest Coast: Radiocarbon Data from Prince Rupert Harbour5 This chapter investigates the methods by which radiocarbon dating and radiocarbon calibration has been practiced and interpreted by archaeological researchers on the northern Northwest Coast. We note that past efforts to report and interpret archaeological chronologies have used a variety of approaches and we suggest this affects the current understanding of the region\u00E2\u0080\u0099s human past. This paper attempts to add clarity to the process of radiocarbon dating and radiocarbon calibration with a focus on interpreting radiocarbon dates obtained on marine or marine-influenced organisms (particularly marine shell and bone collagen). We first discuss the process of radiocarbon dating and radiocarbon calibration, highlighting some of the complexities and nuances of the method. We then discuss three important radiocarbon datasets that underpin the archaeological chronology of social history and culture change in the region; 1) terminal village occupation dates on shellfish compiled by Archer (2001), and 2) human burial dates compiled by Cybulski (1975), and 3) this same dataset reanalyzed by Ames (2005). We then compile a database of currently existing paired marine and terrestrial dates to derive a regionally specific marine reservoir estimate (Delta-R or \u00E2\u0088\u0086R) that allows us to integrate these different datasets into the same chronological timescale. Our resulting re-calibrations refine previously identified chronological and settlement patterns and in some cases shift the understood age of past events by up to 500 years. These chronological shifts 5 This chapter is co-authored with Morley Eldridge. See preface for more information. \u00E2\u0080\u0093 92 \u00E2\u0080\u0093 and patterns have implications for archaeological understandings of the antiquity of social complexity, warfare, and settlement patterns within the region. The Importance of Calibration in Radiocarbon Dating Radiocarbon dating is the most widespread method for evaluating the chronology of archaeological phenomena within the past 40,000 years. However, many levels of uncertainty underlie this chronological method and these are not always well understood by archaeologists (Bronk Ramsey 2008). Most archaeologists are familiar to a degree with how radiocarbon dating works \u00E2\u0080\u0093 it is a method that measures the rate of decay of radioactive carbon isotopes (denoted as 14C) relative to stable carbon isotopes (13/12C), a measurable difference that begins when organic tissue stops growing \u00E2\u0080\u0093 hence the cessation of carbon uptake. The half-life of radiocarbon is approximately 5,730\u00C2\u00B140 years, meaning that half the radioactive carbon (14C) present in a living organism will have decayed in this time (Godwin 1962). Radiocarbon dates provide a result in \u00E2\u0080\u0098years before present\u00E2\u0080\u0099 and an instrumental measurement error (e.g., 4,500\u00C2\u00B150 14C yr bp)6. However, a crucial aspect of radiocarbon dating is that radiocarbon time (i.e., radiocarbon years) is not equivalent to conventional time (years measured by the astronomical 365 day-long year). Thus, calibration is necessary to convert radiocarbon years into historically equivalent and comparable measures of time. This is because radiocarbon \u00E2\u0080\u0098years\u00E2\u0080\u0099 (14C yr bp or RYBP) are variably affected by changes in the 6 We encourage researchers to follow international convention when reporting normalized radiocarbon dates (\u00E2\u0080\u009Cradiocarbon years bp\u00E2\u0080\u009D \u00C2\u00B1 error), and calibrated dates (\u00E2\u0080\u009Ccal yr BP\u00E2\u0080\u009D or \u00E2\u0080\u009Ccal yr BC/AD\u00E2\u0080\u009D) along with a one or two sigma age range. We also encourage researchers to submit their dates to both the Canadian Radiocarbon Database (CARD) and any shell wood pair data to the CHRONO Marine Database. The greater the available information, the greater the chance our collective archaeological knowledge will appropriately incorporate both the ambiguity and the qualified certitude in our regional archaeological chronologies (cf. Gero 2007). \u00E2\u0080\u0093 93 \u00E2\u0080\u0093 global carbon cycle, which is a product of global climate change and solar radiation that influences the concentration of 14C in the atmosphere and ocean at different points in the earth\u00E2\u0080\u0099s history (Guilderson, et al. 2005). This fluctuation is particularly influenced by changes in the global carbon cycle such as forests expanding following deglaciation, major changes in ocean circulation, increased carbon emissions during the industrial revolution, atomic bomb testing, and ongoing land clearance and burning of fossil fuels. Radiocarbon calibration (often known simply as calibration) is a probabilistic method for reconciling the variation in 14C with conventional time, thereby allowing for a measurement of radiocarbon years (14C yr bp or RYBP) to be converted into an age estimate in chronological years before present or calendrical (BC/AD) years (Bowman 1990). Due to the uncertainties inherent in the dating method (measurement error) and the uncertainties of past global carbon (calibration error) the resulting value of a given calibrated radiocarbon age is more appropriately considered an age-range within which a dated sample has a probability of occurring (rather than a single number with a standard error). Since radiocarbon varies depending on global climatic change, the same period in time can be represented by different amounts of 14C variation and vice versa. Calibration of radiocarbon years is further complicated for samples taken from marine or marine influenced contexts; these require a second form of calibration for the marine reservoir effect discussed at length in a following section. Calibrating Radiocarbon Age Estimates While the radiocarbon method was invented in the mid-20th century (Libby, et al. 1949), it was not until the 1980s that radiocarbon calibration \u00E2\u0080\u0098curves\u00E2\u0080\u0099 became commonly \u00E2\u0080\u0093 94 \u00E2\u0080\u0093 available for use in archaeology (Bowman 1990). The most widely used calibration curve developed from an increasingly refined series of dendrochronologically dated tree-ring sequences constructed from the Pacific Northwest (Stuiver 1982) and preserved in peat bogs in Germany and Ireland (Pearson and Stuiver 1986; Stuiver and Pearson 1986; Stuiver, et al. 1998). Researchers radiocarbon dated individual tree rings, each of which had a known-age and compared the difference between the measured radiocarbon age and the actual age to produce a graphical representation of this relationship (Figure 4.1). This \u00E2\u0080\u0098curve\u00E2\u0080\u0099 essentially documents the concentration of atmospheric 14C as preserved in the individual tree growth rings. The most crucial feature of the curve is the overall rate of 14C decline over time but also the undulations that occur over shorter-time periods (due to variations in atmospheric 14C which is influenced by variations in solar activity and other processes of global carbon circulation) and the varying thickness of the curve itself that reflects the relative confidence that a date of a certain age falls within a calendar age-range (due to measurement limitations). These features are discussed in two short examples. \u00E2\u0080\u0093 95 \u00E2\u0080\u0093 Figure 4.1 A 2,400 year-long section of the atmospheric radiocarbon curve covering the period between 1,400 and 3,000 years before present (Reimer et al. 2009). Figure modified from a graphical output of CALIB 6.1 (Stuiver and Reimer 1993). Variability in Radiocarbon Years Consider a radiocarbon date obtained from the leaf of a terrestrial plant (one that derived its carbon from the atmosphere), yielding a measure of 4000 \u00C2\u00B1 25 radiocarbon years before present (14C yr bp). When calibrated using the 2009 radiocarbon curve7 at a 95.5% 7 The internationally recognized radiocarbon curve is refined approximately every four years at an international meeting. The latest curve (INTCAL13) was published September 22, 2013 (Reimer, et al. 2013) and has not yet been incorporated into the dates reported in this dissertation. \u00E2\u0080\u0093 96 \u00E2\u0080\u0093 confidence interval (i.e., the 2-sigma calibrated range8), the resulting age estimate falls between 4,520 and 4,080 years ago. Thus, the process of calibration adds between 80-520 calendar years to this particular radiocarbon age and also significantly increases the apparent uncertainty (from \u00C2\u00B125 to up to \u00C2\u00B1220 years). In addition, differences between radiocarbon and calibrated radiocarbon ages vary over time as a result of changes in the shape of the curve. Variable Uncertainty in the Atmospheric Calibration Curve Consider another date from a terrestrial plant leaf, 9,500 \u00C2\u00B1 25 14C yr bp. When calibrated at a 95.5% confidence interval, this date has a calibrated age-range estimate of between 10,670 and 11,070 years ago. Thus, in contrast to the example above, this calibration increases the numeric difference between the radiocarbon date and calibrated age range by 1,170 to 1,570 years. The statistical uncertainty of this calibrated sample at the a 95.5% confidence interval is slightly less, approximately \u00C2\u00B1 200 years. It could be reduced further by examining the fluctuations in the curve for the period of time in question as shown in Figure 4.2. These transformations and increased uncertainties are integral to translating radiocarbon time into conventional (calendrical/astronomical) time and illustrate how some periods in time have increased uncertainty regardless of the precision of individual radiocarbon measurements (Guilderson, et al. 2005). 8 Probability is simply an expression of the likeliness of something occurring. The convention in analytical statistics is to calculate uncertainty in terms of standard deviations (which can be approximated as the average difference of a set of values from its average) often called \u00E2\u0080\u0098sigmas\u00E2\u0080\u0099 or \u00E2\u0080\u0098standard deviations\u00E2\u0080\u0099. In a normal distribution (where there is a central tendency to the values) 1-sigma encompasses about 67% of the probability and 2-sigmas encompasses about 95%. Sigmas are thus notations for quantifying confidence and are sometimes displayed using the symbol \u00CF\u0083. In describing the 2-sigma calibrated range, this is specific to the calibration probability and does not refer to the uncertainty for the individual date. \u00E2\u0080\u0093 97 \u00E2\u0080\u0093 Figure 4.2. Example of calibration effects on a radiocarbon date with a large error range. The vertical height of the \u00E2\u0080\u0098calibrated age-range\u00E2\u0080\u0099 shown on the x-axis represents the relative probability of the date occurring during a particular period. \u00E2\u0080\u0098Correcting\u00E2\u0080\u0099 for Isotopic Fractionation In addition to the fluctuations in the global carbon cycle, radiocarbon measurements are further influenced by molecular transformations as carbon is differentially incorporated into organic tissues and \u00E2\u0080\u0098hardparts\u00E2\u0080\u0099 such as bones and shells. These alterations have a smaller magnitude than the differences between marine and terrestrial carbon, but can still result in an age estimate being shifted by hundreds of years. Since the carbon-14 isotopes are slightly heavier (because of their additional neutrons), they have a greater inertia and react slower \u00E2\u0080\u0093 98 \u00E2\u0080\u0093 than carbon 12 (12C) and carbon 13 (13C) isotopes (Stuiver and Polach 1977). In other words, a plant that absorbs atmospheric carbon will have a particular carbon 14 value; but that same carbon 14 will be slightly depleted when absorbed by a herbivore consuming plant tissues. Likewise this difference would be slightly larger in a carnivore that regularly consumes herbivores that in turn regularly consumes plant tissues and so on. Thus, even though these different organisms may have lived at exactly the same time, the 14C levels in their tissues will exhibit slight differences, thus requiring further adjustment based on their tropic level. Fortunately, these differences in normalization have been determined by measuring the ratio of the stable carbon isotopes (13C to 12C) in a variety of organisms (Stuiver and Polach 1977). Since 13C and 12C stable isotopes do not change over time, the measured difference is exactly half that of the difference between the 12C and the 14C atoms, allowing for an accurate estimate of 14C in the original organism from the measured levels of 13C to 12C in the sample. Although direct measurements are often not available for samples analysed prior to the 1970s values can be retroactively estimated if the type of dated organism is known (e.g., 0.0 for shellfish, -25.0% for terrestrial charcoal)9 (Stuiver and Polach 1977) but see Southon ( 2011) for further discussion. 9 Most modern radiocarbon laboratories tend to provide a direct measurement of the C12/13 ratio prior to processing a sample so that the value can be factored into the normalization formula but exceptions still exist. In those cases, without directly measuring the isotopic fractionation, radiocarbon samples are \u00E2\u0080\u0098normalized\u00E2\u0080\u0099 according to assumed values (Stuiver and Polach 1977) which for C3 plants is typically -25.0 and for C4 plants ranges between -9 and -12 . Assumed values are commonly indicated by a lack of a decimal place (e.g., -25 rather than -24.5). \u00E2\u0080\u0093 99 \u00E2\u0080\u0093 Calibrating Radiocarbon Age Estimates for Marine Samples The global carbon cycle has two primary domains\u00E2\u0080\u0094 the atmosphere and the ocean \u00E2\u0080\u0094within which carbon circulates at dramatically different rates. For instance, atmospheric carbon circulates virtually instantaneously throughout the globe but marine carbon circulates at much slower rates as much of the carbon is \u00E2\u0080\u0098trapped\u00E2\u0080\u0099 in deep-ocean currents for hundreds of years before it is released once again into the biosphere through upwelling. This temporal lag in marine radiocarbon is known as the marine \u00E2\u0080\u0098reservoir\u00E2\u0080\u0099 (Stuiver and Braziunas 1993). On a global scale, marine carbon circulates an average of 405 years behind the rate of terrestrial carbon (Hughen, et al. 2004) but there is considerable variation regionally and over time. The difference between the global average and the local average of marine reservoir time is known as the \u00E2\u0088\u0086R (hereafter Delta-R). Thus, marine or marine influenced samples require additional calibration compared with terrestrial/atmospheric samples, as shown by the following example. Example 3 - Calibrating a Marine Date Consider a date on a marine clamshell from the same archaeological deposit as the 4000\u00C2\u00B125 14C yr bp plant leaf discussed in Example 1. Because this clam absorbed its carbon from the marine environment (via filter feeding on marine plankton), the radiocarbon age will likely appear significantly \u00E2\u0080\u0098older\u00E2\u0080\u0099 in radiocarbon time than a contemporaneously deposited terrestrial sample (e.g., wood charcoal). Thus, estimating the age of this \u00E2\u0080\u0098marine date\u00E2\u0080\u0099 requires additional calibration for the local marine reservoir, incorporating both the global reservoir value (R) and the Delta-R value, the values for which are derived from two methods discussed below. \u00E2\u0080\u0093 100 \u00E2\u0080\u0093 Method 1: Dating Modern Known-Age Shells Collected in the Historic Era The first of the two principal methods for estimating the age of the marine reservoir is to radiocarbon date living shells that were collected from a known place on a known date prior to the commencement of atomic bomb testing in the 1940s (which significantly disrupted the global carbon cycle). Because oceanic currents are relatively stable, this method provides a historical estimate of the local marine reservoir (Delta-R) that can be applied to radiocarbon calibration of marine samples. However, marine reservoir can be particularly variable in large estuary environments where there is considerable influx of freshwater from rivers, which dilutes (and thus lowers) the marine carbon absorbed by organisms living in those environments (Rick, et al. 2012). Conversely, deepwater fjords in proximity to glacial-fed rivers and tidewater glaciers may contain significant quantities of \u00E2\u0080\u0098old\u00E2\u0080\u0099 meltwater \u00E2\u0080\u0098trapped\u00E2\u0080\u0099 in glacial ice for thousands of years (hence older carbon) (Hutchinson, et al. 2004). On the Northern Northwest Coast a recent study (McNeely, et al. 2006) has dated a series of historic marine shells collected from Haida Gwaii and Southeast Alaska prior to atomic testing in the mid-20th century (Figure 4.3). For example, they obtained a radiocarbon date on a giant razor clam (Ensis siliqua) collected in AD 1937 from Masset on northern Haida Gwaii that returned a radiocarbon age of 850\u00C2\u00B140 14C yr bp. Thus, even though this clam lived during the early 20th century (and died in 1937), it absorbed \u00E2\u0080\u0098older\u00E2\u0080\u0099 marine carbon that had been circulating in the ocean for approximately 690 years. This 690-year difference includes the global average of marine carbon circulation of 405 years plus 285 years for the local marine reservoir effect. This local difference is due to \u00E2\u0080\u0098upwelling\u00E2\u0080\u0099 of \u00E2\u0080\u0098older\u00E2\u0080\u0099 carbon as deep oceanic currents collide with the continental shelf immediately west and south of Haida Gwaii and then circulate throughout Hecate Strait (Thomson 1981). \u00E2\u0080\u0093 101 \u00E2\u0080\u0093 McNeely et al.\u00E2\u0080\u0099s (2006) data are relevant to the interpretation of archaeological chronologies based on marine shell dates but unfortunately lack observations from Prince Rupert Harbour. Nevertheless, results from this study reveal regional differences and sub-regional variability in the marine reservoir age for both Southeast Alaska and Haida Gwaii (Figure 4.3). Of particular significance are the comparatively low Delta-R values from Haida Gwaii relative to deep-water fjords and channels of Southeast Alaska, which have Delta-R values consistently above 400 (Figure 4.3). This is likely a reflection of several large active glaciers and ice fields in coastal Southeast Alaska releasing particularly old freshwater carbon into the deep and potentially slowly circulating coastal fjords. While the sample remains small, the Southeast Alaska data also indicate a west-to-east trend of increasing Delta-R values closer to the mainland coast (Figure 4.3). Currently, there are no available data for Prince Rupert Harbour, thereby limiting the direct utility for interpreting marine influenced radiocarbon ages for sites in this area. A further complicating factor in the use of this method is that oceanic circulation patterns during the early 20th century or historic era may not be very representative of marine circulation during previous millennia. On the Northwest Coast, the early 20th century is the \u00E2\u0080\u0098end\u00E2\u0080\u0099 of the \u00E2\u0080\u0098Little Ice Age\u00E2\u0080\u0099 and is associated with significant receding of glaciers and an increase in associated meltwater (Menounos, et al. 2009). \u00E2\u0080\u0093 102 \u00E2\u0080\u0093 Figure 4.3. Map of the northern Northwest Coast showing the location and local marine reservoir age (Delta-R value) for radiocarbon dated shellfish of known age collected from the northern Northwest Coast prior to atomic testing in the mid 20th century (data from McNeely et al. 2006). Also shown are the locations of late 20th century glaciers (data from www.geogratis.cgdi.gc.ca). Method 2: Dating \u00E2\u0080\u0098Shell-Wood\u00E2\u0080\u0099 Pairs from Archaeological and Paleoecological Contexts The second method for obtaining a local estimate of marine reservoir age is to date contemporaneously deposited marine and terrestrial carbon samples to identify the marine reservoir age from a pre-modern context by the difference in age between the two samples. Thus, in contrast to dating early 20th century marine shell discussed above, this method can assess the magnitude of the marine reservoir prior to the modern era and therefore, is more appropriate for assessing the marine reservoir during the archaeological period of interest. \u00E2\u0080\u0093 103 \u00E2\u0080\u0093 This method consists of estimating the total marine reservoir age at particular points in the past (expressed as R[t]) by dating depositionally \u00E2\u0080\u0098paired\u00E2\u0080\u0099 terrestrial and marine samples from coastal archaeological or paleoecological contexts and subtracting the global reservoir estimate from the combined age difference. This method is appropriate only if marine and terrestrial organic remains are short-lived and were contemporaneously deposited in archaeological contexts such as rapidly accumulating shell midden or a paleoecological context such as wave cut terrace or delta (Southon, et al. 1990; Southon and Fedje 2003). Because archaeological and paleoecological deposits contain long temporal records, this method allows for the assessment of changes in reservoir age at a variety of temporal and spatial scales (Ingram and Southon 1996). For instance, Kennett et al. (1997) dated multiple paired archaeological marine and terrestrial samples from southern California and estimated temporal fluctuations in the reservoir age over the past 10,000 years that could be linked to changes in paleoclimate and upwelling intensity. This in turn allowed for a more refined chronology of the palaeoceanography and the pre-contact cultural dynamics of pre-colonial Southern California (Kennett and Kennett 2000). The drawback to this method is that geological and archaeological researchers must be very attentive to the strength of the stratigraphic association for each sample and recognize the potential interpretive hazards of each data point. In addition, the age difference between samples is less precise than Method 1 because both must rely on age estimates rather than known dates. To recap, the total marine reservoir is a combination of the global reservoir and the local correction (Delta-R) and can be estimated over time and regionally. The local component of reservoir correction is expressed as Delta-R. Considering our 4600\u00C2\u00B125 14C yr \u00E2\u0080\u0093 104 \u00E2\u0080\u0093 bp shell date example once more, if we know that the Delta-R for this period is 350\u00C2\u00B150 radiocarbon years for this point in time, it is possible to more accurately estimate the calibrated age of this shell using online calibration software (e.g., CALIB). This sample produces an age range estimate between 4180-4510 calibrated years BP (at 95.5% probability) and therefore appears more recent than the original measured age and is much more consistent with the original 4,000 \u00C2\u00B1 25 14C yr bp estimate on the (presumably) contemporaneous leaf in the example that started this discussion. Animals who consume both marine and terrestrial carbon Humans, bears, and dogs are among many organisms that ingest substantial amounts of foods (and hence carbon) \u00E2\u0080\u0098mixed\u00E2\u0080\u0099 from both the marine and terrestrial environment. Radiocarbon ages obtained on their bones require a modified form of marine calibration involving estimating the proportion of marine carbon in their diets. Fortunately, the proportion of marine carbon in the diet can be estimated using the commonly applied 13/12C stable isotope ratio measurement that documents the differential uptake of marine and atmospheric carbon by comparison to a regionally specific food web model (e.g., Chisholm, et al. 1982, 1983; Chisholm 1986; Szpak et al. 2009). Calibration programs such as CALIB and OXCAL offer a way to incorporate the percentage of marine carbon in the diet (if known) and calibrate the result using on marine curve. Figure 4.4 provides a summary of the interpretive steps needed in calibration a \u00E2\u0080\u0098mixed\u00E2\u0080\u0099 marine and terrestrial sample as well as for strictly terrestrial and strictly marine samples. \u00E2\u0080\u0093 105 \u00E2\u0080\u0093 Figure 4.4. Schematic depicting the steps required for radiocarbon calibration. Note the high number of steps required for calibrating mixed marine and terrestrial samples. Assuming an isotopic value of -25 for charcoal is really only appropriate for northerly latitudes that lack C4 plants (Teeri and Stowe 1976). Estimating Marine Reservoir in Highly Migratory Animals A further complicating factor is the difficulty of interpreting radiocarbon dates on migratory marine animals. For instance, many marine mammal species are highly migratory and forage far offshore throughout the Pacific Ocean meaning their marine carbon is derived from a variety of sources. Japanese researchers have extensively dated archaeological northern fur seal (Callorhinus ursinus) bones and associated human remains from Jomon period sites in Japan (Yoneda, et al. 2004). The highly migratory fur seals returned radiocarbon ages consistently older than contemporaneously deposited human remains. This same study further demonstrated that both fur seal and human remains exhibited respectively older radiocarbon ages than contemporaneously deposited terrestrial mammal bones (Yoneda, et al. 2004). On the Northwest Coast, researchers seeking to directly date whalebone or salmon or any number of highly migratory species should consider this additional uncertainty of the Delta-R estimate. \u00E2\u0080\u0093 106 \u00E2\u0080\u0093 Variable Reporting of Radiocarbon Dates and Calibration Results Despite the long history of radiocarbon dating, the convention of reporting calibrated radiocarbon ages remains somewhat of a marginal practice in some regions and time periods even through some journals offer very specific recommendations (e.g., Canadian Journal of Archaeology10). The perceived complexity of radiocarbon calibration combined with the diversity of approaches taken by different researchers and research teams, and the quickening pace of research has contributed to considerable variation in the reported radiocarbon record along the northern Northwest Coast. Given the fact that observations have been compiled over 40 years and with different research goals and standards in mind, such variation is entirely understandable. However, this has led to a circumstance where different sets of dates have not been directly compared because they differ in their marine or terrestrial content and have not been placed on a singular calibrated timescale. More importantly, published and repetitively cited accounts of the age of certain cultural events and epochs are understood to have occurred during particular time periods when these ages are more appropriately unanchored points in radiocarbon years lacking appropriate temporal uncertainty. In contrast, all archaeologists report and regularly discuss un-calibrated radiocarbon dates (e.g., 285\u00C2\u00B150 14C yr bp) despite the fact that the actual age-range of a given set of dates may not be directly comparable or change significantly when calibrated (especially if marine influenced). While this may be seen to be necessary for disciplinary discourse, it neglects the fact that a given radiocarbon result may not be the same time scale. For instance, two books concerning archaeology on the Northwest Coast (Ames and Maschner 1999; Fedje and 10 http://canadianarchaeology.com/caa/publications/cja/guidelines-cjajca-authors \u00E2\u0080\u0093 107 \u00E2\u0080\u0093 Mathewes 2005) do not attempt to calibrate radiocarbon dates but rather, offer readers a conversion table that lists a limited series of atmospheric radiocarbon dates in relation to their \u00E2\u0080\u0098true\u00E2\u0080\u0099 age in astronomical years. While this convention is welcome (it reduces translation error when discussing dates), it perpetuates a \u00E2\u0080\u0098floating\u00E2\u0080\u0099 chronology that is not tied to the astronomical (calendrical) timescale which becomes confusing when dealing with more than a handful of calibrated ages whose differences and similarities may be non-linearly different when calibrated. As a result, many researchers continue to report and discuss radiocarbon dates in un-calibrated radiocarbon years and do not use the full potential of current calibration methods. This is unfortunate because it decreases the accuracy and utility of radiocarbon dates used in archaeological interpretations. Unfortunately for the archaeologist, achieving sufficient and clear knowledge of the process of radiocarbon calibration is hindered in part by different commercial or research laboratories, whose calibration estimates may not always be the most up-to-date or appropriate for a given region, time period, or context, particularly for marine or marine influenced samples. Moreover, different laboratories often have slightly differing reporting conventions and terminologies that make it difficult to interpret and compare radiocarbon results11. To complicate matters further, academic journals in archaeology have a range of radiocarbon reporting standards and some major journals appear to lack consistently applied standards. These challenges add to the general level of confusion about the process and utility of calibration. 11 For instance, Beta Analytic, a major North American Lab, reports both a \u00E2\u0080\u0098measured\u00E2\u0080\u0099 radiocarbon age and a \u00E2\u0080\u0098conventional\u00E2\u0080\u0099 (aka. \u00E2\u0080\u0098normalized\u00E2\u0080\u0099) radiocarbon age but does not clearly indicate that the difference is a correction for isotopic fractionation and may create confusion in deciding on which age estimates to subject to calibration. \u00E2\u0080\u0093 108 \u00E2\u0080\u0093 Fortunately, researchers can retroactively re-calibrate adequately reported radiocarbon dates12. This can be quickly accomplished using free online software such as CALIB or OxCAL13. These programs enable researchers to calibrate large sets of dates and produce tabular and graphical representations of the calibrated age-ranges. Indeed, many academic radiocarbon laboratories (e.g., UC Irvine) no longer conduct the additional interpretive step of calibration. Rather, they leave such decisions up to the individual researchers in recognition that critical contextual details that underlie calibration are best achieved by researchers who selected and submitted the dated material. This is a positive step and we hope will encourage the wider adoption of calibration on the Northwest Coast. Following Bronk Ramsey (2008:249) we contend that \u00E2\u0080\u009Cmore than ever, it is necessary for users of radiocarbon to understand and engage with the science that underlies the method.\u00E2\u0080\u009D Problems in Interpreting the \u00E2\u0080\u0098Calibrated Intercept\u00E2\u0080\u0099 One formerly widespread format for archaeological reporting of calibrated radiocarbon dates is to report the \u00E2\u0080\u0098calibrated intercept\u00E2\u0080\u0099 (Bowman 1990). This intercept is the point at which a radiocarbon age (the number before the \u00E2\u0080\u009C\u00C2\u00B1\u00E2\u0080\u009D) \u00E2\u0080\u0098intersects\u00E2\u0080\u0099 the calibration curve. However, due to fluctuations in the width and thickness of the curve (Figure 4.1) and the change from a bell-curve from the radiocarbon age to an irregular curve in the calibrated age-range, there is rarely a \u00E2\u0080\u0098single\u00E2\u0080\u0099 point or number at which this intersection occurs and 12 One issue that can be extremely confusing and could lead to duplicated data when examining a date, is that multiple laboratory names are occasionally published for the same samples. For instance, TO-2352 is the laboratory number in CARD and MacDonald and Inglis 1981, whereas Coupland, et al. (2003:154) references the same date as ISOTRACE-2352. Both are correct, as radiocarbon laboratories often subcontracted out specialized work (such as early accelerator dates), and in these cases both laboratories issued their own numbers. Identification of specific radiocarbon samples is necessary to evaluate and, where appropriate, modify calibration methods. 13 Calibration programs are available for free download or online calculation Calib: http://calib.qub.ac.uk/calib/ or Oxcal: http://c14.arch.ox.ac.uk \u00E2\u0080\u0093 109 \u00E2\u0080\u0093 often the curve is intersected in two or more places and may even have relatively equal probabilities (Figure 4.2). Thus, while the intercept number is intuitively appealing, it is no longer considered an acceptable reporting convention (Telford, et al. 2004). Rather than a single numerical age bounded by an error estimate, a more appropriate representation of a calibrated age is the probability of a date occurring within an estimated time frame. Free calibration programs readily calculate both the 1-sigma and 2-sigma calibrated age-ranges, representing the 68.3% probability and 94.5% probability respectively. While some may find that reporting and discussing an age-range is awkward, it remains the most justifiable way to succinctly characterize the uncertainty of a calibrated age estimate. Fortunately, calibration programmes also provide researchers with an effective and efficient method for representing the probability visually through graphical imagery (e.g., Figure 4.7). As several Northwest Coast researchers have shown, these graphical features readily accommodate the detailed analysis of individual dates (e.g., Ames 2012:180; Grier 2006:102) or show the levels of relative probability for a series of dates (e.g., Lepofsky, et al. 2005:274; Morin, et al. 2008/09:24). Visual tools illustrate the relative temporal distribution of a suite of dates as well as show the undulations and features of the calibration curve itself. Through such graphical analyses, it is possible to identify chronological similarities and differences that may not be readily apparent in tabular form. Problems in Summarizing Multiple Dates A persistent problem related to the intercept concept are summary figures such as histograms commonly employed to illustrate trends from multiple dates and sites over time both on the un-calibrated (Fladmark 1975:298) and calibrated timescales (Maschner \u00E2\u0080\u0093 110 \u00E2\u0080\u0093 1991:930). For instance, a recent study from the Great Basin (Louderback, et al. 2011) presents a series of un-calibrated radiocarbon dates in histograms with 150-250-year bin categories to evaluate paleo-demographic sequences. Curiously, this study does not attempt to calibrate these compiled dates but if this had been conducted, the individual dates would likely span more than one histogram bin category and therefore blur the sharp peaks and troughs inferred to represent demographic trends. An alternative and more appropriate method of displaying the cumulative results of multiple radiocarbon dates is by calibrating a set of dates on a single linear timescale using the \u00E2\u0080\u0098summed probability\u00E2\u0080\u0099 method discussed below. Summed Probability as an Analytical Method Summed probability is a Bayesian method for pooling or combining the statistical probabilities from multiple calibrated radiocarbon dates into a singular age-range on the calendrical (linear) time scale. The method provides a statistically robust way to combine the calibrated age-range for a series of dates and hence is suitable for assessing the radiocarbon chronology from multiple archaeological contexts (Figure 4.5). It is particularly useful for combining the probabilities from sets of radiocarbon dates to evaluate whether they cluster in time, are spread over a long interval, or exhibit modes and/or toughs potentially indicative of an archaeological phase or event. Archaeologists have used summed probabilities of radiocarbon datasets to refine archaeological chronologies for some time (Eighmy and LaBelle 1996; Gregory 2001) but as the use of freeware calibration tools such as Calib and OxCal have proliferated, the method has been increasingly applied to address a wide range of hypotheses such as the assessment of cultural historical phases (Ames 2012; Lepofsky, et al. 2005; Morin, et al. 2008/09), proxies for paleo-demography (Peros, et al. 2010; Williams \u00E2\u0080\u0093 111 \u00E2\u0080\u0093 2013), and indicative of population migrations and colonization events (Buchanan, et al. 2008; Collard, et al. 2010; Hamilton and Buchanan 2007). Figure 4.5. An example of three typical radiocarbon dates on terrestrial charcoal calibrated using the 2009 atmospheric curve and underlain by the summed probability of these three dates. Model produced using Calib 6.10 (Reimer et al. 2009). A flurry of recent research has identified several interpretive and methodological concerns with summarizing radiocarbon data using summed probabilities (Bamforth and Grund 2012; Guilderson, et al. 2005; Steele 2010; Williams 2012). A recent review by Williams (2012) identifies that summed probability profiles are not direct paleo-demographic proxies but are subject to several potential biases, including small sample sizes, taphonomic loss of older sites (Surovell and Brantingham 2007), and calibration effects such as plateaus in the radiocarbon curve that constrain or exaggerate the resolution of some age-ranges. In this latter instance, a given set of dates that concentrates along a relatively \u00E2\u0080\u0098flat\u00E2\u0080\u0099 portion of \u00E2\u0080\u0093 112 \u00E2\u0080\u0093 the curve may cause a spuriously high summed probability that would not be as prominent if the same a cluster of dates were taken from a \u00E2\u0080\u0098steep\u00E2\u0080\u0099 part of the calibration curve (see Guilderson et al. 2005 for some examples). Other researchers have identified that the inclusion or exclusion of even a single dataset of dates may influence the shape of the summed probability curve and hence a demographic trends resulting from it (Steele 2010). Bamforth and Grund (2012) have demonstrated that with simulations wherein the periodicity (chronological spacing) of dates directly impacts on the shape of a summed probability profile and that this is particularly strong during the dramatic fluctuations in the radiocarbon curve during the Pleistocene-Holocene transition. While these researchers caution against uncritical and overly detailed interpretation of the undulations within summed probability profiles, they by no means dismiss the analytical potential of the approach. Rather, the bulk of caution is directed at using summed probabilities as proxies for trends in human demography without examining the many complex factors underlying this method. For instance, the shape and modality of the summed probability can be influenced by but also evaluated against major perturbations in the radiocarbon curve (Bamforth and Grund 2012; Buchanan, et al. 2011; Guilderson, et al. 2005; Williams 2012). The taphonomic fact that older sites are less likely to survive to the present (Surovell and Brantingham 2007) can be considered and even modelled for a given area (Williams 2012). Sample sizes and sample contexts can be evaluated and the strength of the claim can be adjusted accordingly. These studies critically highlight the continued importance of archaeological context as well as an understanding of the constraints of the radiocarbon method, particularly when interpreting what is often a partial and imperfectly sampled archaeological record. The summed probability method remains the most suitable \u00E2\u0080\u0093 113 \u00E2\u0080\u0093 method for examining the relative distribution of dates from well-sampled archaeological landscapes over broad periods of time. Unlike the studies described above, our use of summed probability is not to evaluate demographic trends using a the largest number of available dates, but to apply this method to systematically collected dates from the terminal deposits of individual village sites and individual dates on individual human burials. These targeted dating criteria are therefore much less subject to the vagaries of conventional archaeological chronologies (e.g., date the deepest deposit, etc.) and offer a chance to evaluate these interpretively signficant radiocarbon datasets on a single calibrated time scale. Recalibrating the Radiocarbon Record for Prince Rupert Harbour In light of the insights and challenges presented above, the remainder of the paper focuses on the radiocarbon record for the northern Northwest Coast. We use marine reservoir estimates and the summed probability method to re-evaluate the regional archaeological chronology with the aim of integrating interpretively significant datasets that derive from marine, marine-influenced, and terrestrial radiocarbon dates. In particular, we examine 1) dates on village abandonment in Prince Rupert Harbour and 2) dates on human remains from Prince Rupert Harbour. We focus on this area as it has been subject to intensive and sustained archaeological research for over 40 years and analyse these particular datasets as they represent an important set of radiocarbon observations that underpin the cultural chronology of the northern Northwest Coast during the past 3,000 years. \u00E2\u0080\u0093 114 \u00E2\u0080\u0093 Methods for Estimating Marine Reservoir on the Northern Northwest Coast Despite the prevalence of radiocarbon dates on marine and marine influenced samples in archaeological and paleoclimate records, the northern Northwest Coast lacks a time series estimate of Delta-R. Previous research by Southon et al. (1990) and Southon and Fedje (2003) presented an extensive, data-rich study of the total [R(t)] marine reservoir effects for three areas of the British Columbia Coast, but do not provide a specific estimate of Delta-R. Rather they provide only a total marine reservoir correction value, which is not useful for calibration because particular Delta-R values are required for each dated sample when using calibration programs14. Another paper by Deo et al. (2004) constructed a Delta-R curve for the southern Northwest Coast, but this only covers the past 3,000 years and is applicable to a very specific region of the Salish Sea. The lack of a published Delta-R estimate or a time series of estimates as well as the perceived complexity of calculating Delta-R values has led to several differing conventions for researchers calibrating marine influenced samples. For instance, many researchers simply subtract an average of the total reservoir age offset (often noted as 600 or 650 years) from normalized marine shell dates and then calibrate the date on the atmospheric curve as if it was terrestrial source material (e.g., Fedje and Mathewes 2005; McLaren 2008; Orchard 2009). Others, including ourselves (Eldridge, et al. 2008), have previously used various estimates for Delta-R based on generalized estimates for the whole Northwest Coast (following Ames 2005 citing Stuiver and Braziunas 1993:156) or specific unpublished estimates recommended by specialists (Moss 1989: citing a personal communication from 14 Delta-R values should not be calculated simply by subtracting the average global reservoir (405 years) from the total reservoir as the global reservoir value also fluctuates through time (Reimer et al. 2009). \u00E2\u0080\u0093 115 \u00E2\u0080\u0093 Stuiver). Some researchers (e.g., Coupland et al. 2006; 2010) have avoided the issue of marine reservoir by exclusively dating charcoal, which is subject to its own possible error from the \u00E2\u0080\u0098old wood\u00E2\u0080\u0099 effect15 (Moss 2011a:6; Schiffer 1986). Estimating Delta-R for Prince Rupert Harbour and the Dundas Islands Given the considerable numbers of marine and marine influenced radiocarbon dates obtained from sites in and around Prince Rupert Harbour and the lack of a regionally specific reservoir correction (Delta-R), we created a local record of marine reservoir by compiling all available information into a regional time series (Table 1, Figure 4.6). This record contains 10 dates from 5 shell-wood pairs spanning the past 7,000 years, and which were reported in McLaren (2008) and Southon and Fedje (2003) from the Dundas Islands and Prince Rupert respectively. Each of these marine \u00E2\u0080\u0093 terrestrial pairs has been appropriately normalized for isotopic fractionation and dated using accelerator mass spectrometry (AMS) at either Lawrence Livermore National Laboratories or the Keck Carbon Cycle AMS facility at the University of California Irvine. Terrestrial charcoal from these samples was specially obtained using a microscope and selecting identifiably deciduous or recent growth plant remains while the outermost growth margin from shells was submitted for dating (personal communications with Fedje and McLaren, 2008). The Dundas Islands data were obtained from archaeological deposits at Far West Point (GcTr-6) and the Connell Island Site (GcTr-7) and date within three broad intervals in the late, middle, and early Holocene (McLaren 2008; McLaren, et al. 2011). The Prince Rupert data were obtained from archaeological deposits on Tugwell Island just off the northern entrance to the Harbour and date to within 15 This phenomena represents the potential for wood or charcoal to come from old-growth trees or driftwood, the inner parts of which have the potential to be centuries older than the age at which it was deposited. \u00E2\u0080\u0093 116 \u00E2\u0080\u0093 the late Holocene (Southon and Fedje 2003:107). There are additional shell wood-pair data from Kitimat, 120 km southeast of Prince Rupert (ibid.), but these data were excluded as they date exclusively to the late Pleistocene and were obtained on burrowing shellfish from freshwater-influenced estuary subject to freshwater reservoir effects (e.g., Ingram and Southon 1996; Rick et al. 2012). For each terrestrial date in this modest database, we determined the global marine reservoir for that particular point in radiocarbon time [R(t)] using the raw observations that comprise the Intcal04/09 calibration curve, kindly provided to us by Paula Reimer (Personal Communication, 2008, 201316). These data show the finest existing temporal record for the terrestrial ages that collective make up the published Intcal04/09 calibration curve (Reimer, et al. 2004; Reimer, et al. 2009). The difference in age between coeval Marine04 and IntCal04 dates reveals the known global difference between ocean circulation and atmospheric circulation and thus represents the most specific estimate of global R(t) values through time. After assigning a total reservoir estimate [R(t)] and its error based on the terrestrial radiocarbon age of each of the five shell-wood pairs in the sample, we then subtracted the local shell-wood pair difference from the total (global) reservoir estimate to derive a single Delta-R estimate for each datapoint (\u00C2\u00B1 1 standard deviation). The resulting Delta-R values are then plotted over time (Figure 4.6) and overlaid by polynomial curve to infer temporal trends. The resulting figure represents the current temporal record of marine reservoir for Prince Rupert and the Dundas Islands. 16 These data are also available from the Radiocarbon journal website: www.radiocarbon.org \u00E2\u0080\u0093 117 \u00E2\u0080\u0093 This marine reservoir record spans 7,000 years and indicates only moderate fluctuation in regionally specific marine reservoir (Delta-R) during the mid-to-late Holocene (Figure 4.6). Notably, the two similar elevated estimates for the Dundas Islands show virtually no difference between the early and mid-Holocene (Delta-R values of 215, 225 respectively). The two late-Holocene marine\u00E2\u0080\u0093terrestrial pairs from Tugwell Island just outside Prince Rupert Harbour are slightly elevated relative to the two mid and early-Holocene pairs from Dundas (Figure 4.6). This may be due to the proximity of Prince Rupert to the estuarine Skeena River mouth (~25km southeast). However, this may also reflect a climatic influence during this period which overlaps with the \u00E2\u0080\u0098neoglacial\u00E2\u0080\u0099 expansion between 3,500-1,000 calendar years ago, when regional climate was colder relative to earlier periods (Menounos, et al. 2009:2063). Similarly, the most recent shell-wood pair (from late period deposits in the Dundas Islands) is slightly elevated relative to the middle and early Holocene shell-wood pairs which occurs during the onset of the \u00E2\u0080\u0098little ice age\u00E2\u0080\u0099 (ibid). Overall, the observed fluctuation in reservoir age over time is low to moderate but the temporal trends appears broadly similar to the more variable pattern identified for Haida Gwaii by Southon and Fedje (2003:100). \u00E2\u0080\u0093 118 \u00E2\u0080\u0093 Figure 4.6. Local marine reservoir (Delta-R) estimates over the Holocene for marine\u00E2\u0080\u0093terrestrial pairs from Prince Rupert Harbour and the Dundas Islands (error bars represent \u00C2\u00B11 standard deviation). Source data obtained from Southon and Fedje (2003) and McLaren (2008). The Delta-R estimate during the last 3,000 years is of the greatest relevance to the datasets presented in this paper. \u00E2\u0080\u0093 119 \u00E2\u0080\u0093 Table 4.1 Paired marine and terrestrial radiocarbon dates from Prince Rupert Harbour and the Dundas Islands and estimates of the local marine reservoir value depicted in Figure 4.6. Lab # Marine Material 14C Date \u00C2\u00B1 Lab # Terrestrial Material 14C Date \u00C2\u00B1 Total Age Offset \u00C2\u00B1 R(t) Poly Local Delta-R poly \u00C2\u00B1 Reference CAMS-49626 Saxidomus 2370 50 CAMS-49625 charcoal 1560 40 810 60 355 455 60 Southon and Fedje 2003:107 CAMS-49624 Protothaca 2780 50 CAMS-49623 charcoal 2040 50 740 70 340 400 70 Southon and Fedje 2003:107 UCIAMS-21882 Shell 4200 15 UCIAMS-21985 charcoal 3645 25 555 30 330 225 30 McLaren 2008:222 UCIAMS-21881 Mytilus 7510 20 UCIAMS-21984 charcoal 6925 50 585 55 370 215 55 McLaren 2008:224 UCIAMS-21980 Mytilus 1395 15 UCIAMS-21983 charcoal 640 60 755 60 385 370 60 McLaren 2008:240 Table 4.2 Recalibrated radiocarbon dates on shell from the uppermost shell midden layers in village sites in Prince Rupert Harbour reported by Archer (1992, 2001). Lab Number Site Number 14C Age 14C Age \u00C2\u00B1SD Delta-R Value Delta-R \u00C2\u00B1SD 2-sigma Cal BP Dated Material Reference WSU-4366 GbTo-2 1705 90 390 60 650-1090 marine shell Archer 1992:4 WSU-4367 GbTo-4 1980 90 360 60 940-1370 marine shell Archer 1992:4 WSU-4368 GbTo-4 2980 70 300 60 2150-2670 marine shell Archer 1992:4 WSU-4369 GbTo-7 2720 100 320 60 1770-2320 marine shell Archer 1992:4 WSU-4370 GbTo-7 2360 95 350 60 1320-1830 marine shell Archer 1992:4 WSU-4371 GbTo-8 2130 90 360 60 1083-1550 marine shell Archer 1992:4 WSU-4372 GbTo-8 2550 70 340 60 1570-2030 marine shell Archer 1992:4 WSU-4373 GbTo-9 2490 90 340 60 1480-2000 marine shell Archer 1992:4; Archer 2001:212 WSU-4374 GbTo-9 2360 90 350 60 1330-1820 marine shell Archer 1992:4; Archer 2001:212 WSU-4375 GbTo-28 1900 90 360 60 880-1300 marine shell Archer 1992:4 \u00E2\u0080\u0093 120 \u00E2\u0080\u0093 Lab Number Site Number 14C Age 14C Age \u00C2\u00B1SD Delta-R Value Delta-R \u00C2\u00B1SD 2-sigma Cal BP Dated Material Reference WSU-4376 GbTo-28 2600 70 330 60 1650-2120 marine shell Archer 1992:4 WSU-4377 GbTo-32 2435 100 350 60 1380-1930 marine shell Archer 1992:4; Archer 2001:212 WSU-4378 GbTo-32 2220 70 350 60 1250-1640 marine shell Archer 1992:4; Archer 2001:212 WSU-4379 GbTo-46 2530 95 340 60 1510-2060 marine shell Archer 1992:5; Archer 2001:212 WSU-4380 GbTo-46 2590 95 340 60 1570-2130 marine shell Archer 1992:5; Archer 2001:212 WSU-4381 GbTo-57 2470 90 340 60 1460-1980 marine shell Archer 1992:5; Archer 2001:212 WSU-4382 GbTo-57 2220 65 350 60 1250-1620 marine shell Archer 1992:5; Archer 2001:212 WSU-4383 GbTo-59 2870 70 310 60 1980-2470 marine shell Archer 1992:5 WSU-4384 GbTo-59 3200 80 290 60 2360-2880 marine shell Archer 1992:5 WSU-4385 GbTo-64 2370 60 350 60 1380-1800 marine shell Archer 1992:5 WSU-4386 GbTo-64 3840 60 250 60 3290-3700 marine shell Archer 1992:5 WSU-4387 GbTo-66 2500 60 340 60 1530-1940 marine shell Archer 1992:5; Archer 2001:212 WSU-4388 GbTo-66 2590 90 340 60 1590-2120 marine shell Archer 1992:5; Archer 2001:212 WSU-4389 GbTo-70 2445 90 340 60 1410-1930 marine shell Archer 1992:5; Archer 2001:212 WSU-4390 GbTo-70 2510 100 340 60 1490-2050 marine shell Archer 1992:5; Archer 2001:212 WSU-4391 GbTo-77 3210 100 290 60 2350-2940 marine shell Archer 1992:5; Archer 2001:212 WSU-4392 GbTo-77 2925 100 300 60 2030-2670 marine shell Archer 1992:5; Archer 2001:212 \u00E2\u0080\u0093 121 \u00E2\u0080\u0093 Lab Number Site Number 14C Age 14C Age \u00C2\u00B1SD Delta-R Value Delta-R \u00C2\u00B1SD 2-sigma Cal BP Dated Material Reference WSU-4393 GbTo-78 2760 90 320 60 1830-2330 marine shell Archer 1992:6; Archer 2001:212 WSU-4394 GbTo-78 2425 80 350 60 1400-1880 marine shell Archer 1992:6; Archer 2001:212 WSU-4395 GbTo-89 2450 70 340 60 1470-1920 marine shell Archer 1992:6; Archer 2001:212 WSU-4396 GbTo-89 2490 90 340 60 1480-2000 marine shell Archer 1992:6; Archer 2001:212 WSU-4397 GbTn-9 2710 90 320 60 1780-2300 marine shell Archer 1992:6 WSU-4398 GbTn-9 4090 70 240 40 3600-4050 marine shell Archer 1992:6 WSU-4399 GcTo-6 2310 70 350 60 1300-1730 marine shell Archer 1992:6; Archer 2001:212 WSU-4400 GcTo-6 2140 90 360 60 1100-1570 marine shell Archer 1992:6; Archer 2001:212 WSU-4401 GcTo-27 2280 95 350 60 1270-1750 marine shell Archer 1992:6 WSU-4402 GcTo-27 2695 70 330 60 1790-2270 marine shell Archer 1992:6 WSU-4403 GcTo-28 2440 100 340 60 1400-1950 marine shell Archer 1992:6 WSU-4404 GcTo-28 2290 90 350 60 1280-1750 marine shell Archer 1992:6 WSU-4405 GcTo-39 2280 90 350 60 1270-1730 marine shell Archer 1992:6 WSU-4406 GcTo-39 2430 100 340 60 1390-1930 marine shell Archer 1992:6 WSU-4407 GcTo-51 2580 100 340 60 1550-2130 marine shell Archer 1992:7 WSU-4408 GcTo-51 2070 90 360 60 1030-1500 marine shell Archer 1992:7 WSU-4409 GcTo-52 2360 75 350 60 1350-1810 marine shell Archer 1992:7; Archer 2001:212 WSU-4410 GcTo-52 2530 90 340 60 1520-2050 marine shell Archer 1992:7; Archer 2001:212 \u00E2\u0080\u0093 122 \u00E2\u0080\u0093 Recalibrating Marine-Influenced Dates from Prince Rupert Harbour Using this marine reservoir time series, it is possible to re-calibrate marine and marine-influenced radiocarbon dates from Prince Rupert Harbour and the Dundas Islands on a single integrated timescale. Thus, we reviewed and compiled all available information on each archaeological date from the region, including the type of dated material (e.g., charcoal, shell), both the standard and normalized radiocarbon ages, and we compiled the measured estimates for the percentage of marine carbon in the diet of dated human remains discussed below. For each marine or marine-influenced date, we derived an age-specific Delta-R value, an uncertainty for Delta-R and then re-calibrated all dates using the CALIB calibration programme (Calib 6.1.1) using the appropriate calibration curve for each individual date (i.e., atmospheric, marine, or mixed marine-terrestrial). Outputs for these re-calibrations are presented in 2-sigma calibrated age-ranges (Table 2) as well as in summarized summed probability plots presented in Figure 4.7 and Figure 4.8. We re-evaluate the chronology of three series of radiocarbon dates from Prince Rupert Harbour: 1) dates on settlement history and village abandonment by David Archer (1992, 2001) and 2) dates obtained on human remains obtained during the North Coast Prehistory Project (Ames 2005; MacDonald and Inglis 1981). Calibration Issues in Village Abandonment Dates from Shell The first radiocarbon dataset we consider is from the extensive settlement history study of Prince Rupert conducted by David Archer (1992, 2001) who systematically dated shells collected from the upper layers from 23 village sites to evaluate the hypothesis that there was a period of widespread village abandonment in Prince Rupert Harbour. \u00E2\u0080\u0093 123 \u00E2\u0080\u0093 Archer\u00E2\u0080\u0099s initial study was based on 45 dates from 23 sites (1992) but he later refined this to a much smaller sample of 22 dates from the 11 village sites based on strict chronological criteria (2001). Each dated site had spatial and architectural characteristics that were hypothesized to be associated with either \u00E2\u0080\u0098egalitarian\u00E2\u0080\u0099 or \u00E2\u0080\u0098ranked\u00E2\u0080\u0099 village social organization based on the variation in house size and the configuration of house platforms. As observed by Archer, dates of \u00E2\u0080\u0098abandonment\u00E2\u0080\u0099 from these two types of sites sequentially clustered in time around AD 100, but the \u00E2\u0080\u0098egalitarian\u00E2\u0080\u0099 villages were abandoned first followed after by the \u00E2\u0080\u0098ranked\u00E2\u0080\u0099 villages. Archer interpreted the overall clustering in time of these dates as reflecting widespread abandonment of the harbour locality around AD 400 (Archer 2001:214). He interpreted further detail in the sequence of this abandonment as relating to the type of village that was abandoned first and argued this reflected evidence of a dramatic increase in social ranking throughout Prince Rupert Harbour immediately followed by widespread regional village abandonments. This innovative settlement study was pioneering in its effort to consistently stringent household measurements and sampling criteria to a series of sites within a small area in order to characterize settlement history. Archer\u00E2\u0080\u0099s radiocarbon methodology entailed dating relatively short-lived marine shells from the top of shell midden (terminal) deposits and calibrating these dates using an estimated marine reservoir age. However, as discussed below, there are complexities with his application of calibration techniques to this radiocarbon dataset that warrant further evaluation. Archer\u00E2\u0080\u0099s 2001 study employed stringent chronological criteria so that only half of his original dates presented in his 1992 report were retained as representative of \u00E2\u0080\u0098village abandonment\u00E2\u0080\u0099. A key aspect of his 2001 criteria specified that two dates on marine shell \u00E2\u0080\u0093 124 \u00E2\u0080\u0093 (obtained from the two separate areas of a village) needed to overlap at a 2-sigma calibrated age-range for inclusion. However, this range depends on the methods and calibration curve used to calibrate the dates. Moreover, the methodology for calibration differ between Archer\u00E2\u0080\u0099s 1992 and 2001 publications. In 1992, Archer subtracted 650 years from the un-calibrated radiocarbon ages and added 50 years to the standard deviation (1992:4). He then calibrated these \u00E2\u0080\u0098corrected\u00E2\u0080\u0099 dates using the then internationally recognized atmospheric calibration curve (Stuiver and Pearson 1986). This was a thoroughly appropriate treatment of these data at the time. However, in 2001, Archer presents only a subset of these original data (22 of 45 dates) and more importantly, rather than correcting and calibrating these dates, he appears to simply subtract 650 years from the un-calibrated age and adds 15 years to the standard deviation (2001:212). Thus, he does not technically calibrate these radiocarbon dates but only provides the original radiocarbon age (14C yr bp) minus 650 years. Moreover, he presents these same data in a figure that show these dates on the calendrical (calibrated) time scale, thus implying they are calibrated but without showing a calibrated age-range or describing the method by which these singular age estimates were determined. Given the known differences between calibrated and non-calibrated radiocarbon dates, and the uncertainty of estimating marine reservoir, we feel this important dataset can be usefully re-calibrated on the marine curve using a temporally specific Delta-R . The results of these recalibrations are presented below. \u00E2\u0080\u0093 125 \u00E2\u0080\u0093 Figure 4.7. Village abandonment data obtained from Archer 2001 (top) and Archer 1992 (bottom) but presented on calibrated time scale. Horizontal lines represent calibrated age-ranges for individual dates (at 2-sigma) and the dark black line represents the summed probability of all of the dates combined. Dashed vertical line indicates a peak in summed probability that is identical in both reported datasets. \u00E2\u0080\u0093 126 \u00E2\u0080\u0093 Recalibration of Prince Rupert Harbour Village Abandonment Dates Our recalibration of both Archer\u00E2\u0080\u0099s abridged (2001) and original (1992) data on shell dates from the upper layers of multiple village sites in Prince Rupert Harbour confirm that these two series of dates strongly cluster in time (Figure 4.7). Our results also indicate the probable peak of abandonment seems to have occurred around 1700 cal yr BP (ca. AD 300) which support Archer\u00E2\u0080\u0099s original contention for a widespread abandonment of villages occurring roughly the same period. Importantly, the large spike in summed probability is nearly identical between Archer\u00E2\u0080\u0099s much larger original (1992) sample which includes a variety of dates that were \u00E2\u0080\u0098rejected\u00E2\u0080\u0099 because of a lack of calibrated overlap with other dates from the same site or from potential site disturbance. Thus, the overall similarity of the two probability plots, particularly the strong and relatively sudden peak in summed probability around 1700 years ago, strengthens Archer\u00E2\u0080\u0099s original claim for inferring something most remarkable in the archaeological record: a broadly contemporaneous abandonment of more than a dozen villages in a small area over a relatively short archaeological timeframe. This overlap in time is additionally notable due to the fact that the previous dataset was inconsistently calibrated. The summed probability plot of the larger dataset does reflect a less narrowly defined temporal window than Archer\u00E2\u0080\u0099s refined sub-sample (2001) and further distinctions could be made based on the category of \u00E2\u0080\u0098egalitarian\u00E2\u0080\u0099 and \u00E2\u0080\u0098ranked\u00E2\u0080\u0099 village sites but we do not evaluate this distinction here. The inherent uncertainty in radiocarbon calibration and the error-ranges in the current dataset precludes us from determining if this process of village \u00E2\u0080\u0098abandonment\u00E2\u0080\u0099 occurred in the space of a few years or decades, but it is notable that bulk of abandonments appear to span the period between 1200 and 2000 cal yr BP. However, the fact that nearly two-thirds \u00E2\u0080\u0093 127 \u00E2\u0080\u0093 of the radiocarbon dates include the year AD 300 in their two-sigma ranges strongly suggests to use that a series of village sites stopped accumulating around that period in time. Archer\u00E2\u0080\u0099s study did not include all identified village sites in the harbour \u00E2\u0080\u0093 some villages were purposefully excluded as they were currently occupied (e.g., Metlakatla) and other untested sites may have been occupied throughout the late pre-contact period up through the early historic-era but await future investigation. Archer\u00E2\u0080\u0099s larger (1992) sample indicates a more prolonged period of occupation as it includes a few younger dates that may represent abandoned villages being subsequently reoccupied for brief periods or subsequent disturbance to the site deposits (e.g., GbTo-2, GbTo-4 and GbTo-28, see appendix). Ames (2005:110) also notes that some sites (e.g., Grassy Bay) seem to begin to be occupied after this period. Thus, despite the discrepancies between Archer\u00E2\u0080\u0099s original (1992) and later (2001) presentation of these radiocarbon data, as well as the complex issue of marine calibration, our data confirm and strengthen Archer\u00E2\u0080\u0099s hypothesis in regards to a relatively narrow period of village abandonment and place these results firmly on the calibrated time scale. This important dataset can now be compared to the dates on human remains discussed below. Marine Influenced Human Burial Dates in Prince Rupert Harbour The North Coast Prehistory Project (1966-1980), directed by George MacDonald of the Canadian Museum of Civilization, identified and excavated a variety of shell midden sites and obtained a large number of radiocarbon dates directly on human remains recovered from burial interments (n=40), several of which are associated with elaborate burial goods interpreted to reflect emergent social differentiation and the antiquity of warfare (Ames 2001, 2005; Cybulski 1975, 1990, 1999; Fladmark, et al. 1990; MacDonald and Inglis 1981; \u00E2\u0080\u0093 128 \u00E2\u0080\u0093 MacDonald and Cybulski 2001). Even though more than 40 years have passed since this multi-year excavation project, this particular set of radiocarbon dates on human interments presently comprises more than 20% of the currently reported dates in Prince Rupert Harbour and remains among the most intensively dated and studied burial population on the Northwest Coast (Cybulski 2006). The issue of calibrating radiocarbon dates on human remains in Prince Rupert Harbour has long been identified as a topic of interpretive concern. Cybulski (1992; 1996:65) noted that a date on a decapitated human burial from the Lachane site (GbTo-33) was \u00E2\u0080\u009Cuncorrected for possible fluctuations in atmospheric 14C or marine reservoir influences.\u00E2\u0080\u009D In his summary of the North Coast Prehistory Project, Ames (2005:50) endeavoured to compile and calibrate all available radiocarbon dates from the harbour locality and has a lengthy discussion regarding the marine reservoir and Delta-R values and whether to apply them in the calibration of human remains dates. He notes the lack of an estimate for Delta-R values and highlights the variation in values applied by different researchers. Ames also mentions that stable isotope measurements by Chisholm et al. (1982, 1983) indicate the diet of pre-contact peoples in Prince Rupert Harbour was predominantly marine-based (and thus subject to the marine reservoir). However, Ames does not apply such a correction when calibrating these dates. Rather, he states that while \u00E2\u0080\u009Csome correction seemed in order\u00E2\u0080\u00A6 I had no basis for making it, if, indeed it was necessary\u00E2\u0080\u009D (Ames 2005:50). Based on a review of existing and newly available data, this chapter offers a basis for revising this statement by compiling existing information about the age of the marine reservoir and marine diet estimations in Prince Rupert Harbour and the Dundas Islands. Indeed, the present study was prompted by our suspicion that the discrepancy between the dates on human remains and village \u00E2\u0080\u0093 129 \u00E2\u0080\u0093 abandonment dates (discussed below) might be reduced and/or reconciled if both radiocarbon datasets were calibrated appropriately, although at the time we did not realize the complexity of this task. Researchers working elsewhere on the northern Northwest Coast have addressed the challenge of interpreting radiocarbon ages on human remains by using isotopically derived estimates of marine diet and to determine if a date should be calibrated on a marine, terrestrial, or mixed marine-terrestrial curve. In Southeast Alaska, Dixon and colleagues used the isotopic composition of human bone obtained from an early-Holocene site (Dixon, et al. 1997:703) to demonstrate this individual had an overwhelming marine-based diet and subsequently calibrated the age of this individual using the marine curve (Fedje, et al. 2004:113; Kemp, et al. 2007:606). Researchers and First Nations descendants working with Kwaday dan ts\u00E2\u0080\u0099inchi, the \u00E2\u0080\u0098man long ago found,\u00E2\u0080\u0099 whose remains were recovered from a receding glacier in northwestern BC (700km NW of Prince Rupert), radiocarbon dated both his clothing and his bone collagen as well as measuring carbon and nitrogen isotopes to determine his long-term diet (Richards, et al. 2007). They then used these values along with dates on associated terrestrial artifacts to best estimate the calibrated age-range when the individual died (ibid:721-723). Unfortunately, the Little Ice Age climatic anomaly during this period, associated with a \u00E2\u0080\u0098flat\u00E2\u0080\u0099 part of the calibration curve, makes the calibrated age ranges particularly broad for this particular case. Correcting for marine reservoir can thus be greatly strengthened by using available isotopic data to ensure the appropriate curve is used. Within Prince Rupert Harbour and nearby Lucy Island, existing carbon isotopic data are available from at least 11 human burials (Chisholm, et al. 1983:397; Cui, et al. 2013:4; Cybulksi 2001) with carbon isotope values \u00E2\u0080\u0093 130 \u00E2\u0080\u0093 (delta C12/C13), and range from -16.9 to -11.6 (Table 4.3). According to the conversion formula applied to five individuals by (Chisholm, et al. 1983:397), this represents interpreted values of marine protein intake of 85%, 88%, 97%, 99%, and 100% respectively.17 From this sample, an estimated value of 90\u00C2\u00B110% marine protein can be considered conservative given the preponderance of higher marine dietary components noted elsewhere on the Northwest Coast (Chisholm 1986; Chisholm et al. 1982, 1983; Richards et al. 2007:721). While only three of these individuals have been directly radiocarbon dated (S-1729, GSC-2886, and GSC-2888), the remainder are from dated sites or associated with other dated individuals (Table 4.3). The Canadian Archaeological Radiocarbon Database (CARD) lists an additional two individuals from Garden Island (GbTo-23) with measured delta C12/C13 values of -14.3 (GSC-2888) and an outlier with a value of -16.9 (GSC-2886). Presumably, this latter individual could be from an interior or upriver location with poor salmon runs and no other marine protein intake. The presence of non-local marine reservoir values in marine foods further complicates the determination of an appropriate Delta-R for radiocarbon dated human remains. Indigenous populations on the Northern Northwest coast regularly consumed pelagic and highly migratory marine mammals, birds, and fish, especially salmon (Ames 1998; Brewster and Martindale 2011; Coupland, et al. 2010; Moss 2007a). Many of these subsistence resources, including shellfish (Burchell, et al. 2013), could be smoked, dried and/or stored as well as transported and extensive traded during seasonal excursions (Ames 17 These isotopic estimates lack corresponding nitrogen measurements (N15/14) which might help further refine estimates of marine derived protein intake. However, since the region lacks C4 plants and the range of C12/C13 values for prey species is relatively well constrained at this latitude (Szpak, et al. 2009:2738; Szpak, et al. 2012), these estimates are remain the best current estimate until future studies can be conducted. \u00E2\u0080\u0093 131 \u00E2\u0080\u0093 2002; Martindale and Jurakic 2004). Thus, estimates for Delta-R specific to local conditions remain potentially variable for both humans and their protein sources in the absence of specific data such as series of human-charcoal and salmon-charcoal paired dates from different localities. While trade and exchange as well as the consumption of migratory marine animals adds a level of uncertainty to the calculation of Delta-R values, our use of a local Delta-R for calibrating dates on human remains (Table 4.4) is certainly preferable to previous approaches that have not applied any correction for marine content. Table 4.3 Available carbon isotope values on directly dated and undated human remains from Prince Rupert Harbour and surrounding locations. Site # Burial Number 14C Lab# 14C Date \u00C2\u00B1 SD Estimated Marine Protein1 (Chisholm et al. 1983) Delta C13 Reference GbTo-23 Garden Island 201 GSC-2886 2660 260 -16.9 (Cybulksi 2001) GbTo-23 Garden Island 197 GSC-2888 2620 70 -14.3 (Cybulksi 2001) GbTp-1 Lucy Island Ancient 938 Beta-294715 5530 40 -12.9 (Cui, et al. 2013:4) GbTp-1 Lucy Island Ancient 939 Beta-317343 5930 40 -11.6 (Cui, et al. 2013:4) GbTo-33 Lachane 455 S-1729 2300 70 99% -13.0 (Chisholm, et al. 1983:397; Chisholm 1986:143) GbTo-31 Boardwalk 328 85% -14.1 (Chisholm, et al. 1983:397; Chisholm 1986:143) GbTo-31 Boardwalk 330 88% -13.9 (Chisholm, et al. 1983:397; Chisholm 1986:143) GbTo-36 Baldwin 509 97% -13.2 (Chisholm, et al. 1983:397; Chisholm 1986:143) GbTo-36 Baldwin 515 100% -12.7 (Chisholm, et al. 1983:397; Chisholm 1986:143) GgTj-6 Greenville 2D-48 93% -13.5 (Chisholm, et al. 1983:397; Chisholm 1986:143) GgTj-6 Greenville 3D-23 86% -14 (Chisholm, et al. 1983:397; Chisholm 1986:143) TOTALS N=10 N=5 93 \u00C2\u00B112% 13.6 \u00C2\u00B11.3 1 This estimate has an uncertainty of \u00C2\u00B110% (Chisholm et al. 1983:397). \u00E2\u0080\u0093 132 \u00E2\u0080\u0093 Table 4.4 Recalibrated radiocarbon dates on human bone collagen compiled from existing literature as synthesized by Cybulski (1992) and Ames (2005) and including moving average Delta-R estimates and estimated marine protein content based on Table 4.3. Lab # Site 14C Age 14C Age \u00C2\u00B1SD Delta-R Delta-R \u00C2\u00B1SD % Marine Carbon Material Cal yr BP (2-sigma) Reference S-2549 GbTo-30 Parizeau Pt 1640 90 450 60 90 collagen 640-1040 Cybulski 1992, Cited in Ames 2005 S-1430 GbTo-31 Boardwalk 1820 70 450 60 90 collagen 800-1210 Cybulski 1992, Cited in Ames 2005 S-1428 GbTo-23 Garden Island 1930 85 425 60 90 collagen 930-1310 Cybulski 1992, Cited in Ames 2005 S-1663 GbTo-33 Lachane 1930 60 425 60 90 collagen 960-1280 Cybulski 1992, Cited in Ames 2005 S-1731 GbTo-33 Lachane 1930 70 425 60 90 collagen 950-1290 Cybulski 1992, Cited in Ames 2005 S-1599 GbTo-36 Baldwin 1940 80 425 60 90 collagen 940-1310 Cybulski 1992, Cited in Ames 2005 S-1738 GbTo-36 Baldwin 2040 80 400 60 90 collagen 1060-1480 Cybulski 1992, Cited in Ames 2005 S-1667 GbTo-31 Boardwalk 2050 70 400 60 90 collagen 1080-1480 Cybulski 1992, Cited in Ames 2005 S-1732 GbTo-33 Lachane 2080 80 375 60 90 collagen 1130-1540 Cybulski 1992, Cited in Ames 2005 S-1733 GbTo-31 Boardwalk 2090 150 375 60 90 collagen 1000-1700 Cybulski 1992, Cited in Ames 2005 S-1740 GbTo-33 Lachane 2200 110 350 60 90 collagen 1250-1770 Cybulski 1992, Cited in Ames 2005 S-1736 GbTo-31 Boardwalk 2260 80 350 60 90 collagen 1320-1770 Cybulski 1992, Cited in Ames 2005 S-1734 GbTo-31 Boardwalk 2270 80 350 60 90 collagen 1330-1780 Cybulski 1992, Cited in Ames 2005 S-1729 GbTo-33 Lachane 2300 70 350 60 90 collagen, Delta C13 = -13.0 1380-1800 Cybulski 1992, Cited in Ames 2005 and Chisholm 1986:143 S-1728 GbTo-33 Lachane 2320 70 325 60 90 collagen 1410-1840 Cybulski 1992, Cited in Ames 2005 S-1730 GbTo-33 Lachane 2320 70 325 60 90 collagen 1410-1840 Cybulski 1992, Cited in Ames 2005 S-1283 GbTo-36 Baldwin 2320 80 325 60 90 collagen 1400-1860 Cybulski 1992, Cited in Ames 2005 S-1282 GbTo-33 Lachane 2390 80 325 60 90 collagen 1490-1950 Cybulski 1992, Cited in Ames 2005 S-2548 GbTo-30 Parizeau Pt 2430 90 325 60 90 collagen 1511-2015 Cybulski 1992, Cited in Ames 2005 S-1665 GbTo-31 Boardwalk 2510 70 300 60 90 collagen 1680-2120 Cybulski 1992, Cited in Ames 2005 S-1735 GbTo-31 Boardwalk 2510 100 300 60 90 collagen 1590-2160 Cybulski 1992, Cited in Ames 2005 S-1664 GbTo-31 Boardwalk 2530 70 300 60 90 collagen 1690-2130 Cybulski 1992, Cited in Ames 2005 S-1427 GbTo-18 Dodge Island 2565 80 300 60 90 collagen 1700-2200 MacDonald and Inglis 1981 S-1739 GbTo-33 Lachane 2610 100 300 60 90 collagen 1740-2300 Cybulski 1992, Cited in Ames 2005 GSC-2888 GbTo-23 Garden Island 2620 70 300 60 90 human bone, C12/13 = -14.3 1818-2274 MacDonald and Inglis 1981 and Cybulski in CARD \u00E2\u0080\u0093 133 \u00E2\u0080\u0093 Lab # Site 14C Age 14C Age \u00C2\u00B1SD Delta-R Delta-R \u00C2\u00B1SD % Marine Carbon Material Cal yr BP (2-sigma) Reference GSC-2886 GbTo-23 Garden Island 2660 260 300 60 90 collagen Delta C13 = -16.9 1470-2720 MacDonald and Inglis 1981 Cybulski in CARD 2001 S-1429 GbTo-23 Garden Island 2670 85 300 60 90 collagen 1850-2320 Cybulski 1992, Cited in Ames 2005 S-1425 GbTo-18 Dodge Island 2700 80 300 60 90 collagen 1890-2330 MacDonald and Inglis 1981 cited in CARD S-2723 GbTn-19- Ridley 2710 150 300 60 90 collagen 1720-2550 MacDonald and Inglis 1981 cited in CARD S-1666 GbTo-31 Boardwalk 2750 80 275 60 90 collagen assoc w copper 1940-2440 Cybulski 1992, Cited in Ames 2005 S-1432 GbTo-31 Boardwalk 2760 90 275 60 90 collagen 1930-2480 Cybulski 1992, Cited in Ames 2005 S-1285 GbTo-31 Boardwalk 2790 80 275 60 90 collagen 1980-2520 Cybulski 1992, Cited in Ames 2005 S-1426 GbTo-18 Dodge Island 2930 80 275 60 90 collagen 2190-2700 MacDonald and Inglis 1981 S-1595 GbTo-23 Garden Island 2980 70 275 60 90 collagen 2280-2740 Cybulski 1992, Cited in Ames 2005 S-1284 GbTo-31 Boardwalk 2980 85 275 60 90 collagen 2280-2740 Cybulski 1992, Cited in Ames 2005 S-1431 GbTo-31 Boardwalk 3040 130 275 60 90 collagen 2200-2880 Cybulski 1992, Cited in Ames 2005 S-1598 GbTo-36 Baldwin 3100 70 250 60 90 collagen 2390-2860 Cybulski 1992, Cited in Ames 2005 S-1433 GbTo-33 Lachane 3210 90 250 60 90 collagen 2500-3070 Cybulski 1992, Cited in Ames 2005 S-1741 GbTo-33 Lachane 3280 85 250 60 90 collagen 2680-3160 Cybulski 1992, Cited in Ames 2005 S-3361 GbTo-36 Baldwin 3370 90 250 60 90 collagen 2750-3250 Cybulski 1992, Cited in Ames 2005 \u00E2\u0080\u0093 134 \u00E2\u0080\u0093 Recalibrating Radiocarbon Dates on Human Remains from Prince Rupert Harbour The radiocarbon dates obtained on human remains from Prince Rupert Harbour were obtained from a multi-site burial population examined by Cybulski (1975). Dates were presented in summarized form in MacDonald and Inglis (1981:44) and re-presented in Ames (2005). These dates were also submitted to the Canadian Radiocarbon Database (CARD) database by Cybulski (www.canadianarchaeology.ca/). For our re-calibrations, we sought to compare our recalibrations using the \u00E2\u0080\u0098mixed\u00E2\u0080\u0099 marine-terrestrial curve and incorporating a predominantly marine diet estimate with Ames\u00E2\u0080\u0099 (2005) two sets of calibrations based on 1) the atmospheric (terrestrial) radiocarbon curve (without a marine reservoir correction) and 2) entirely on the marine curve. To accomplish this, we first reviewed reported dates and available isotopic data to ensure each was normalized appropriately and assumed a predominantly marine diet for the burial population (90% marine protein). This estimate seems conservative given the high frequency of higher values (higher marine dietary components) noted in the harbour and elsewhere on the Northwest Coast (e.g., Chisholm 1986; Chisholm et al. 1982, 1983; Richards et al. 2007:721). With these estimates, we recalibrate dates on human remains contrasting three different calibration methods using the CALIB programme to demonstrate the chronological consequences of calibration treatments (Figure 4.8). \u00E2\u0080\u0093 135 \u00E2\u0080\u0093 Figure 4.8 Comparison of different calibration criteria applied to radiocarbon dates on human remains from Prince Rupert Harbour. The top figure depicts Ames\u00E2\u0080\u0099 (2005) original dataset and analytical assumptions calibrated on the atmospheric (terrestrial) curve without a marine reservoir correction. The middle figure also presents Ames dataset but uses the marine calibration curve with a generic Delta-R estimate of a 390\u00C2\u00B125 discussed by Ames (2005:50). The bottom figure represents our reanalysis of all available data using an estimate of a 90% marine diet and a moving average of Delta-R (derived from Figure 4.5). Horizontal lines overlaid onto the summed probability profile represent calibrated age-ranges for individual dates (at 2-sigma). \u00E2\u0080\u0093 136 \u00E2\u0080\u0093 The top plot in Figure 4.8 shows the summed probability of Ames\u00E2\u0080\u0099 (2005:108) calibration method, reflecting the combined profile of individual dates on human remains calibrated on the atmospheric curve and assuming a 100% terrestrial diet. This summed probability plot reveals multiple dramatic spikes and a profound drop just after 1700 cal yr BP. These data appear broadly similar to Ames (2005:108), whose Figure 5.31 depicts a histogram of calibrated intercepts (cf. Telford et. al. 2004) indicating a large spike in frequency between 300 BC and 200 BC followed by another spike between AD 0 and AD 100. This calibration treatment indicates that the burial dates span a period of 3,000 years but abruptly ends after 1400 cal yr BP (AD 400). In contrast, the middle plot also shows Ames\u00E2\u0080\u0099 (2005) sample of dated human remains, but calibrated on the marine curve and incorporating a Delta-R value of 390\u00C2\u00B125, a generic value for the entire Northwest Coast (Stuiver and Braziunas 1993:156) and commonly recommended by Beta Analytic (e.g., Ames 2005:50). This plot is greatly smoothed relative to the top plot, reflecting the slow circulation of carbon in the marine domain relative to the atmospheric curve. Due the inclusion of a local marine reservoir effect, this plot shifts the age-range for this set of dates forward in time by nearly 1,000 years and also alters the shape and modularity of the summed probability profile. In this treatment, the greatest frequency of dated human remains post-dates the peak in village abandonment ca. 1700 cal yr BP. The bottom plot in Figure 4.8 depicts our re-calibration treatment on the marine calibration curve, incorporating a locally derived and temporally specific Delta-R (Figure 4.6), and assuming a 90% marine diet among the dated individuals. This plot also includes additional dated human burials not reported in Ames 2005 (e.g., GbTo-18, Dodge Island). Overall, this plot demonstrates the frequency of dated burials is highest just prior to and \u00E2\u0080\u0093 137 \u00E2\u0080\u0093 following the peak in village abandonments around 1700 cal yr BP and decreases after approximately 1500 cal yr BP, nearly a millennium after Ames\u00E2\u0080\u0099s previous analysis suggests (2005:111). These re-calibrations indicate this sample of human burials spans a 3,000-year time span, abruptly ending at 500 calibrated years BP; slightly later than the marine recalibration of Ames\u00E2\u0080\u0099 2005 data (Figure 4.8b), and dramatically later than his original calibrations (Figure 4.8c). Our revised calibrations of Ames\u00E2\u0080\u0099 compiled data and using a slightly larger sample considerably shifts the archaeological chronology of human interments in Prince Rupert Harbour, particularly in relation to Archer\u00E2\u0080\u0099s settlement data on village abandonment. This integrated chronology indicates dated human interments predate as well as significantly post-date the period of village abandonment. It is notable that the bulk of the dated interments appear are most frequent in the centuries immediately preceding and following the peak of village abandonments (ca. 1700 cal yr BP). This is followed by two moderate peaks of increased probability at approximately 1200 and 800 cal yr BP respectively. The sequential periodicity of these three peaks is intriguing but it may be due to plateaus in the radiocarbon curve (cf. Guilderson et al. 2005) and/or the sampling of multiple interments from a single contemporaneous deposit from a limited number of sites (cf. Geib and Hurst 2013). The relatively large error ranges for these dates and the relatively broad calibrated age-ranges also increases the chance these are artifacts of cumulative probabilities (Bamforth and Grund 2012). However, this periodicity may also relate to a temporal clustering of burial interments on a regional scale as observed in subterranean cairn and mound burials on the southern Northwest Coast (Brown 2003; Lepofsky, et al. 2000). \u00E2\u0080\u0093 138 \u00E2\u0080\u0093 Regardless of the factors that may influence these individual undulations, this integrated set of recalibrated dates demonstrates a robust and broader pattern of burial interments both preceding and succeeding the peak period of village abandonment. In particular, this dataset indicates the bulk of dated burials occur after the period of village abandonment. This suggests that the individuals interred at these no-longer accumulating village sites may have maintained affinities with the kin-groups or lineages who formerly occupied them. Support for such a pattern is observed in Tsimshian oral histories (adawx) which speak of a temporary abandonment of the Harbour after approximately 2000 years ago followed by a re-establishment of more aggregated seasonal Tsimshian settlements in Metlakatla Pass (Martindale and Marsden 2003:14). The broad temporal overlap also indicates a continuity of burial interments during the period when these villages remained intensively occupied (i.e., prior to ca. 1700 cal yr BP). Future research into Prince Rupert burial practices and village settlement histories might productively consider how human interments might be placed into previously deposited shell midden such as the presence of culturally sterile sediments overlaying shell midden deposits (e.g., Ames 2005:248-249). Interpretive Implications of Recalibrations on Individual Dates In addition to the considering the insights provided by modelling the summed probability profiles, our recalibrations revise the understood age-ranges of several individual burials that have been identified as important social and historical markers of social differentiation, large-scale migrations, warfare, and regional trade and interaction (Ames 2001; Cooper 2012; Cybulski 1994; Martindale and Marsden 2003; Matson and Coupland 1995). \u00E2\u0080\u0093 139 \u00E2\u0080\u0093 Of particular significance are the dates on human interments associated with copper artifacts which are interpreted to be socially valuable and widely traded prestige items. The copper artifacts associated with dated human interments from Prince Rupert Harbour are among the earliest dated evidence for copper (ca. 2000 cal yr BP) in northwestern North America (Ames 2005:234; Cooper 2007). This broadly traded material has not been sourced but the closest archaeologically identified source in is the Copper River in Alaska over 1200 km northwest of Prince Rupert (Cooper 2007). The earliest dated interment (Burial 325, S-1666) is a young male (12-16 yr) associated with copper (including disc earrings and cooper staining on several long bones) (MacDonald and Cybulski 2001:11) that has been previously interpreted by Ames (2005:214) to date to between 1020 \u00E2\u0080\u0093 800 BC (2970-2750 cal yr BP). In contrast, our recalibrations indicate this individual is approximately 800 years younger, dating to between 240 BC and AD 30 (2190-1980 cal yr BP). While this age-range is significantly more recent than previously postulated (Ames 2005; MacDonald 1983), this date remains considerably older than the earliest documented copper artifacts from the south-central Alaska and the Yukon, which post date AD 1000 (Cooper 2012:580) as well as older than dates associated with copper on the southern BC Coast (Blake 2004:110-111; Lepofsky, et al. 2000:400). Burial 322 (S-1667) also contained a \u00E2\u0080\u009Ccopper-wrapped wood bead,\u00E2\u0080\u009D several copper stained element an amber bead and over 250 shell disc beads including 25 dentalium beads (MacDonald and Cybulski 2001:11). Ames (2005:209) calibrates this individual to between 190 BC to AD 60 whereas our re-calibrations indicate a date of between AD 640 and AD 810. Similarly, a date on an elderly male with a carved raven pendant (Burial 410, S-1432) was originally calibrated by Ames (2005:215) to between 1100 BC and 800 BC whereas our \u00E2\u0080\u0093 140 \u00E2\u0080\u0093 revised calibration indicates an age range of between 300 BC and AD 10. Our recalibrations thus support earlier contentions for an alternate source for copper on the British Columbia Coast (Acheson 2003; Cooper, et al. 2008:1734; de Laguna, et al. 1964:204; Rapp, et al. 1990) and indicate these dates from Prince Rupert Harbour remain the earliest associated with cooper artifacts on the Northwest Coast. In sum, our re-calibrations and summed probability profiles shift and refine current understanding of archaeological chronologies in Prince Rupert Harbour. They highlight the interpretive relevance of re-engaging with the existing radiocarbon record with regard to settlement dynamics and burial practices as well as the age of prestige technologies such as copper associated with human burials. However, we must caution that our re-calibrations have effectively broadened the uncertainty associated with individual dates despite being placed on a calendrical timescale. Rather than a single point in time bracketed by an evenly distributed uncertainty (\u00C2\u00B1), these individual dates now incorporate an uncertainty for the global and local marine reservoir age, an assumed estimate and source of percentage of marine protein in the diet (in the case of human remains), as well as an instrumental uncertainty in the original radiocarbon measurement and finally, a calibration uncertainty. Despite this compounded uncertainty, these calibration methods are considerably improved relative to previous interpretations of radiocarbon dates commonly presented simply in radiocarbon years without calibration or lacking an appropriate incorporation of marine reservoir effects and influences. Moreover, this increase in uncertainty is greatly narrowed and countered by calibrating sets of radiocarbon dates from multiple sites and features into \u00E2\u0080\u0098summed probability profiles\u00E2\u0080\u0099 that combine and thereby reduce temporal uncertainty for a given group of dates. This renewed focus reveals peaks and troughs in the periodicity of \u00E2\u0080\u0093 141 \u00E2\u0080\u0093 historical phenomena such as the patterns of village abandonment and the frequency of human burials, but continues to require cautious interpretation and will be subject to revision as future research progresses. For example, there will be opportunities to further refine and re-evaluate archaeological chronologies and calibration techniques that highlight the interpretive limitations of the current dataset. However, our current effort has served to place previously disparate historical phenomena in and around Prince Rupert Harbour onto a singular calibrated (calendar) time-scale. Combined with the insights that will come from future research, such an approach has the potential to reconcile and refine the archaeological knowledge of the timing of culture historical events in and around Prince Rupert Harbour. Conclusions A central conclusion that became abundantly clear during the preparation and revision of this chapter is that archaeologists must increasingly confront, acknowledge, and contend with the uncertainty and complexity of radiocarbon calibration (Bronk Ramsey 2008; Geib and Hurst 2013). This is particularly critical in order to successfully overcome the collective legacy of previously published archaeological chronologies that include marine and marine influenced radiocarbon dates. Another key conclusion of this paper is that additional dating and estimation of marine reservoir values are needed to refine our understanding of archaeological chronologies on the Northwest Coast. The modest marine reservoir time series along the northern coast presented in this paper will benefit from additional shell-wood pairs in good association which will refine the regionally specific Delta-R estimate and better characterize its temporal and spatial variability. If community support for additional radiocarbon dating of \u00E2\u0080\u0093 142 \u00E2\u0080\u0093 human burials can be attained (e.g., Cui, et al. 2013), additional dating of these remains paired with charcoal and or salmon will help resolve the still unknown difference between mixed-marine and terrestrial carbon sources (cf. Cybulski 2010). In particular, paired dates from human collagen and associated elderberry seeds from box burials would help demonstrate the offset between marine influenced human collagen and terrestrial dates (cf. Cybulski 1992) as elderberry seeds represent only one year of growth and therefore are not subject to potential \u00E2\u0080\u009Cold wood\u00E2\u0080\u009D problems (e.g., Kennett, et al. 2002). As noted, highly migratory marine species (e.g., grey whales, salmon, and fur seals) can have very different reservoir values than contemporary local marine species. Rockfish-charcoal and salmon-charcoal pairs may also provide an estimate of the local and migratory Delta-R estimate respectively. Salmon bones (and their carbon) may therefore need a very different Delta-R than local shell or fish; and humans and other animals such as bears that derive a large part of their diet from salmon may also differ as a result. Another conclusion is that field archaeologists should be advised to hesitate before radiocarbon dating shell at sites close to large estuaries or melting glaciers as there is considerable potential for estuarine and glacially influenced environments to substantially affect Delta-R values (e.g., Hutchinson, et al. 2004; Rick, et al. 2012; Ulm 2002, 2006). This can vary significantly depending on the amount of glacial melt-water, limestone substrate, or freshwater precipitation. Thus, where samples or funds are limited, it may be best to avoid shell and marine mammal bones altogether and instead date terrestrial mammal bone or deciduous charcoal (e.g., rapidly growing botanical remains such as needles, twigs, and berries). More detailed macro and microscopic examination, documentation, and description of dating material is also warranted particularly for the most commonly dated material \u00E2\u0080\u0093 \u00E2\u0080\u0093 143 \u00E2\u0080\u0093 charcoal. Researchers working in other regions with higher resolution archaeological chronologies are increasingly identifying the species of all dated samples. While this remains uncommon on the Northwest Coast, increased attention to the identifying charcoal to species of dated samples as well as direct dating of relatively short-lived terrestrial animals such as deer bone is commonly recovered in coastal sites but deer on the outer coast often consume substantial amounts of marine plants requiring 13/12C isotopic and or mixed marine reservoir correction. We also recognize that further research into marine reservoir effects may not ultimately result in the refinement of existing archaeological chronologies but rather the opposite \u00E2\u0080\u0093 a broadening of the temporal uncertainty for marine or marine influenced radiocarbon dates (cf. Geib and Hurst 2013; Gero 2007). Our study has compiled existing information on the marine reservoir and Delta-R and shown it to be mildly variable through the Holocene. However, this record is still relatively sparse and it would be desirable to add many more data points as this will increase our level of confidence in the calibrated age-range of a particular sample or set of samples and may result in a broadening the calibrated uncertainty for a set of dates, reflecting an increase in accuracy but a reduction in precision. Following from Ames (2005), this study is only the second on the northern Northwest Coast to attempt to compile and recalibrate a series of previously published marine and marine influenced dates on a single calendrical \u00E2\u0080\u0093 calibrated time scale. However, it will undoubtedly not be the last, particularly as these interpretively significant and often-irreplaceable datasets become increasingly integrated with or re-evaluated by contemporary research. As in many areas of the coast, North Coast researchers must confront the implications and observations of earlier work in order to collectively contribute to a richer cultural history. \u00E2\u0080\u0093 144 \u00E2\u0080\u0093 Chapter 5. Zooarchaeological Analysis of the Indigenous Fishery at the Huu7ii Big House & Back Terrace, Huu-ay-aht Territory, Southwestern Vancouver Island18 This chapter describes the archaeology of the Indigenous fishery at the Huu7ii19, a large archaeological site (DfSh-7) on what is now called Diana Island across a channel from the contemporary community of Bamfield in southern Barkley Sound (Figure 1.1). The name \u00E2\u0080\u0098Huu-ay-aht\u00E2\u0080\u0099 translates as the \u00E2\u0080\u0098people of Huu7ii\u00E2\u0080\u0099 and is an Nuu-chah-nulth placename which identifies this locality as a place from which Huu-ay-aht came to be an autonomous community in the distant past (Huu-ay-aht First Nations 2000:36). This location has extensive shell midden deposits stretching over 300 meters of shoreline site and as much as 100 meters inland with at least 10 large and topographically well-defined house platforms arranged in an arc facing the modern shoreline (Mackie and Williamson 2003:110). Excavations by a team of non-Indigenous archaeologists and Huu-ay-aht community members with support of the Huu-ay-aht First Nations were conducted in 2004 and 2006 and focused on a large area within the single largest house platform (House 1) and two areas of an elevated terrace containing older mid-Holocene deposits (McMillan, et al. 2008; McMillan and St. Claire 2012). 18 This Chapter has been previously published as: McKechnie, Iain (2012) Zooarchaeological Analysis of the Indigenous Fishery at the Huu7ii Big House and Back Terrace, Huu-ay-aht Territory, Southwestern Vancouver Island. In Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound, by Alan D. McMillan and Denis E. St. Claire, pp. 154\u00E2\u0080\u0093186. (c) 2012 Archaeology Press, Simon Fraser University, Burnaby, BC. For more archaeological and ethnographic context for the site, readers are referred to the monograph in which this chapter appears as well as Alexander Mackie and Laurie Williamson\u00E2\u0080\u0099s 2003 chapter on settlement archaeology in this portion of Huu-ay-aht territory. 19 The name Huu7ii includes the numeral \u00E2\u0080\u00987\u00E2\u0080\u0099 to represent a glottal stop (where there is a purposeful pause). The name can be pronounced as \u00E2\u0080\u0098Hoo\u00E2\u0080\u00A6aye\u00E2\u0080\u0099 with a pause between the two syllables. \u00E2\u0080\u0093 145 \u00E2\u0080\u0093 This chapter examines fine-screened fish remains from 12 \u00E2\u0080\u0098column samples\u00E2\u0080\u0099 recovered during fieldwork conducted in 2004 and 2006. I coordinated the excavation, recovery, and analysis of these small \u00E2\u0080\u0098columns\u00E2\u0080\u0099 of precisely excavated archaeological sediment (Figure 5.1, Figure 5.4, and Figure 5.5). The goal of this research was to use this detailed recovery strategy to investigate how residents utilized fish over the past 5,000 years of human history represented at this large village settlement. Column sample excavation and analysis aimed to complement the analysis of larger vertebrate fauna recovered from excavation units that used larger \u00C2\u00BC\u00E2\u0080\u009D mesh sizes (Frederick 2012). The principal advantage of column sampling is that it provides a much more accurate assessment of the relative proportion and actual number of fish, mammal, and bird bones present in the site deposits (Casteel 1976a; McKechnie 2005a; Stewart and Wigen 2003). Methods Column samples were recovered as contiguous bulk samples from the sidewalls of excavation units (Figure 5.1). Column sample level dimensions were 20x10x5cm (1 litre of excavated matrix per individual level), with the exception of a column sample from the back terrace, which measured 20x20x5cm (2 litres per level, see Table 5.1). Column samples were excavated in 5cm arbitrary levels within which stratigraphically distinct layers were separated. Vertical elevations were referenced to an arbitrary datum elevation as well as recorded in depth increments below ground surface. Due to the considerable effort required to process each recovered column sample level, not all excavated columns or column sample levels could be subject to comprehensive faunal identification. Twelve column samples, six from the 2004 excavations and six from the 2006 excavations were selected for zooarchaeological identification and analysis (Figure \u00E2\u0080\u0093 146 \u00E2\u0080\u0093 5.3 and Figure 5.2). These samples represent the greatest horizontal and vertical extent of the excavated deposits that focused on the House 1 deposits which date to between approximately 1,500 and 400 years ago and two areas of the older back terrace deposits which date to between approximately 5,000 and 3,000 years ago. Eight column samples were examined from separate areas of House 1; four span the length of occupation and known depth-range while the remaining four were collected from the upper \u00E2\u0080\u0098house floor\u00E2\u0080\u0099 portions of the deposits exposed during block excavations in 2006 (Figure 5.2). Two column samples were examined from two separate areas of the back terrace deposits (Figure 5.3). An additional column sample was obtained near the mouth of the creek on the northern and western portion of the site in 2004. This small assemblage was obtained from shell midden deposits encountered during the construction of the field camp privy. Figure 5.1 Excavating column samples from the sidewall of excavation units (left). Bulk sediment samples were removed in 5cm levels, wet-screened through 1mm mesh, and material larger than 1mm saved. Vertebrate fauna larger than 2mm was picked from the samples in the laboratory by supervised volunteers (centre and right). \u00E2\u0080\u0093 147 \u00E2\u0080\u0093 Figure 5.2. Plan view showing column samples locations recovered from the House 1 excavations. Large squares are 2x2m excavation units and coordinates with arrows indicate the location of individual column samples. \u00E2\u0080\u0093 148 \u00E2\u0080\u0093 Figure 5.3. Perspective view of the Huu7ii village looking west showing the location and layout of the House 1 excavation units and the location of the back terrace units. \u00E2\u0080\u0093 149 \u00E2\u0080\u0093 Figure 5.4. Photo of column sample taken from the north wall of the back terrace unit (N4-6/E0-2) which has initial and terminal dates that span between 5,000 and 3,000 years ago. Figure 5.5. Photo of column sample taken from the South wall of unit N18-20/E2-4. This excavation unit reached a depth of 230 cm below surface and has an initial date of approximately 1,500 years BP. \u00E2\u0080\u0093 150 \u00E2\u0080\u0093 Processing Column samples were removed in 5-cm levels and wet-screened through 1-mm mesh. After each matrix sample had been wet-screened and had sufficient time to dry, sediments were passed through 2-mm mesh using nested geological brass sieves. Vertebrate fauna was systematically collected from 2-mm mesh in well-lit laboratory conditions by volunteer \u00E2\u0080\u0098rockwashers\u00E2\u0080\u0099 at the University of Victoria who graciously donated many evenings picking through these numerous samples (see Figure 5.1). Through their collective efforts, a large number of samples were processed, a task that would have taken an inordinate amount of time for a single person. After processing was completed, shell, bone, rock, and charcoal constituents from individual samples were weighed and these data were entered into an Excel spreadsheet. Vertebrate fauna from each individual column sample was retained while the remaining sample constituents were placed back in the original sample bag. Processed non-vertebrate faunal samples were then delivered to the repository at the Royal BC Museum. Identification Vertebrate fauna was morphologically identified with the aid of a binocular dissecting microscope (6.3-40x) and the use of the comparative skeletal collection at the University of Victoria Zooarchaeology Laboratory. Identification data were recorded by skeletal element in a row and column database, noting relevant osteological, taphonomic, and provenience information. This database was then converted to a working spreadsheet and finally imported into a stable relational database (FileMaker Pro). With the exception of fish spines, ribs, branchials, scales, and gill-rakers, identification was attempted for all skeletal elements recognizable to species, genus or family level. Confidence codes were assigned to each \u00E2\u0080\u0093 151 \u00E2\u0080\u0093 examined specimen to indicate the certainty of identification (for criteria, see Frederick and Crockford 2005). Briefly specimens were considered \u00E2\u0080\u0098identified\u00E2\u0080\u0099 (NISP) if they could be confidently assigned to a taxonomic level of family, genus, or species. The remaining specimens were classified as unidentified fish, bird, mammal, or unidentifiable bone (NSP). Considerable effort was taken to employ identical identification and quantification procedures for both column and unit sample fauna (i.e., Frederick 2012) including the use of the same comparative collection. However, some species level designations, such as distinguishing different species of greenling (Hexagrammos sp.) was attempted much less frequently in the column sample assemblage due to a lack of equivalent confidence between analysts. Quantification NISP \u00E2\u0080\u0093 Number of Identified Specimens The primary means of quantification used in this analysis as well as in the excavation unit assemblage (Frederick 2012) is the number of identified specimens (NISP). This measure represents the number of skeletal specimens that can be confidently identified to family, genus, or species. NISP is an indivisible quantitative measure fundamental to all zooarchaeological assemblages and is readily compared across archaeological contexts. NISP data are typically expressed in terms of the relative abundance (% frequency) of a particular item relative to all other identified specimens from a taxonomic class (e.g., herring is 81.3% of all identified fish remains). NISP does not include specimens that are only recognizable as \u00E2\u0080\u0098fish,\u00E2\u0080\u0099 which were designated as 'NSP\u00E2\u0080\u0099 (see Table 5.1). \u00E2\u0080\u0093 152 \u00E2\u0080\u0093 Ubiquity \u00E2\u0080\u0093 Frequency of Occurrence Ubiquity is an additional measure of abundance based on the presence or absence of items in a number of archaeological contexts. Ubiquity is calculated as the percentage of discrete contexts in which a certain taxon is found (frequency of occurrence). For example, herring can be considered \u00E2\u0080\u0098ubiquitous\u00E2\u0080\u0099 in the assemblage because this species is present in over 90% of the 168 examined column samples levels at the site. Ubiquity is used here to supplement the interpretations of abundance as it is not dependent on the proportion of other species. MNI \u00E2\u0080\u0093 Minimum Number of Individuals MNI is defined as the most commonly occurring, non-repeatable skeletal elements observed in a temporally distinct paleontological or archaeological context (Lyman 2008). There are several well-known methodological problems with MNI (Grayson 1984; Lyman 2008). MNI estimates are ultimately derived from NISP data and therefore cannot be used as an independent measure. The estimates produced by MNI calculations are particularly affected by how archaeological units of analysis (time periods and depositional events) are defined \u00E2\u0080\u0093 the smaller the number of categories, the fewer individuals. Conversely, the larger number of analytical categories, the higher the minimum estimates, which also increase the probability that single individuals might be counted more than once. Such uncertainty has makes the use of MNI problematic, particularly for large mammals such as whales and seals whose large skeletal remains may be widely distributed in an archaeological context due to food sharing, differential butchery, transport, and consumption as well as use of bone to make tools and or extract oil (e.g., Monks 2003). However, such factors are arguably much less likely for smaller-bodied fish that are much \u00E2\u0080\u0093 153 \u00E2\u0080\u0093 more likely to redistributed and discarded as individual animals than would be larger-bodied animals. The benefits to employing MNI estimates is the ability to translate numbers of bone elements to an estimate of the total number of animals represented in a given depositional context. Such a conversion allows for a more detailed comparative assessment of the relative contribution of individual animals, and may differ substantially from the %NISP estimate but is in no way a substitute for it. Given that the column sample faunal assemblage is comprised of small discrete volumes of spatially and temporally distinct deposits, I deemed it worthy of considering the use of MNI estimates for the column sample fish assemblage. A particular motivation was to address notion that Pacific herring, the most abundant fish in the assemblage, might comprise a relatively less important role in the assemblage if faunal counts are converted to MNI. Thus, I calculated fish MNI by using the most numerous non-repetitive elements present in an individual column sample level (1-2 litres of sediment). If many more repeatable elements such as vertebrae were present in a particular sample, I divided this count by the number of elements for particular fish taxa (e.g., 55 vertebrae per herring). NISP and MNI Per Litre and Cubic Metre In addition to NISP, ubiquity, and MNI, I also calculate the number of identified specimens per litre (NISP per litre) and the minimum number of individuals (MNI) per litre. These latter measures are derived from the combined excavated volume of individual column sample levels (e.g., 1 litre per 5cm level) and are then scaled up to cubic metres (i.e., 1,000 litres). These measures provide an \u00E2\u0080\u009Cabsolute\u00E2\u0080\u009D measure of abundance as opposed to relative \u00E2\u0080\u0093 154 \u00E2\u0080\u0093 percentage data (where a change in the abundance of a particular taxon may reflect a change in the abundance of another species). Importantly, these estimates do not account for variability within individual column level samples but rather, are generated by dividing the total number of specimens by the total examined volume. As such, these precise estimates should be considered tentative but nevertheless distinct from relative percentage data. They are used here to supplement and strengthen the overall abundance estimates by providing another level of scrutiny in assessing the taxonomic composition and temporal trends in the assemblage. Fish Size Estimations To measure fish size and trends in size over time, I used digital calipers (\u00C2\u00B10.1mm) to measure a select number of herring, greenling, rockfish, salmon, and hake skeletal elements. Estimating fish size is possible due to the predicable relationship between the dimension of individual bones and the length of individual fish (Casteel 1974). Here, I use published regression formulae for rockfish, greenling, and Irish lord (Orchard 2003) as well as new formulae I developed for hake and herring (McKechnie 2010a; McKechnie and Tollit n.d.). These regressions were based on comparative collections at the Zooarchaeology Lab at the University of Victoria and the National Marine Mammal Laboratory located in the NOAA Sand Point facility in Seattle, Washington. To estimate fish length for herring, I measured the greatest width of the 1st and 2nd vertebrae of herring (McKechnie and Tollit n.d.). For hake, I measured the width of the articular surface of the quadrate (McKechnie 2010a). For salmon, I measured the greatest transverse diameter of whole salmon vertebrae to attempt to distinguish salmon species from their relative size distribution (Cannon and Yang 2006). \u00E2\u0080\u0093 155 \u00E2\u0080\u0093 Table 5.1 Column samples containing identified fish remains screened through 2mm mesh. Column Sample Ex. Date Number of levels Excavated Volume (Litres)* Recovered Volume (Litres)* Orig. wt. (kilograms) Unid. Fish (NSP) NISP fish Total Fish NISP / Litre (ex. Vol.) N2-4/W18-20 Back terrace 2006 25 25.0 24.75 33.807 2,110 11,439 13,549 457.6 N4-6/E0-2 Back terrace 2004 17 34.0 42.25 51.821 2,061 5,920 7,981 174.1 N10-12/E2-4 W. wall 2004 19 19.0 24.3 25.063 3,855 2,928 6,783 154.1 N12-14/E6-8 S. wall 2004 12 12.0 16.25 16.802 1,320 1,086 2406 90.5 N18-20/E2-4 S. wall 2004 21 21.0 25.5 28.632 2,477 2,033 4,510 96.8 N18-20/E6-8 E. wall 2004 21 21.4 30.95 34.663 4,674 3,834 8,507 179.2 N10-12/E4-6 S. wall 2004 1 1.0 0.75 0.750 96 99 195 99.0 N14-16/E15.5-16 E. wall 2006 6 6.0 8.65 9.075 875 855 1,730 142.5 N18-20/E15.5-16 W. wall 2006 8 8.0 8.6 10.103 812 739 1,551 92.4 Privy Pit 2004 4 4 4.4 5.794 210 33 243 8.3 N18-20/E26-28 W. wall 2006 9 9.0 11.75 11.991 1073 717 1,790 79.7 N18-20/E34-36 E. wall 2006 25 25.0 32.5 36.338 4,185 2,795 6,980 111.8 Total N=168 185.4 230.65 264.839 23,748 32,492 56,225 175.2 * Excavated volume is based on the dimensions of the excavation whereas recovered volume is based on the volume of sediment recovered and measured using water displacement. Table 5.2. List of identified taxa (NISP) in the column sample assemblage. Fish Back House 1 Total Terrace Pacific herring Clupea pallasi 16,470 9,930 26,400 Anchovy Engraulis mordax 347 1,519 1,866 Salmon Oncorhynchus sp. 146 1,616 1,762 Greenling sp. Hexagrammos sp. 184 527 711 Hake Merluccius productus 3 637 640 Rockfish sp. Sebastes sp. 61 276 337 Dogfish shark Squalus acanthias 42 257 299 Perch sp. Embiotocidae 64 26 90 Sablefish Anoplopoma fimbria 1 64 65 Petrale sole Eopsetta jordani 43 43 Flatfish sp. Pleuronectiformes 2 25 27 Lingcod Ophiodon elongatus 1 26 27 Ratfish Hydrolagus colliei 5 18 23 Irish lord sp. Hemilepidotus sp. 1 18 19 *Prickleback sp. Stichaeidae 2 13 15 Pile perch Damalichthys vacca 1 12 13 Sculpin sp. Cottidae 1 11 12 Plainfin midshipman Porichthys notatus 12 12 White-spotted greenling Hexagrammos stelleri 7 2 9 Cabezon Scorpaenichthys marmoratus 8 8 *Eulachon Thaleichthys pacificus 6 6 *Clingfish sp. Gobiesocidae 6 6 Herring/sardine Clupeidae 3 3 6 Skate sp. Raja sp. (unident.) 5 5 *Tomcod Microgadus proximus 4 4 \u00E2\u0080\u0093 156 \u00E2\u0080\u0093 Fish Back House 1 Total Terrace Halibut Hippoglossus stenolepis 3 1 4 *Smelt sp. Osmeridae 3 3 Pacific cod Gadus macrocephalus 2 2 Red Irish lord Hemilepidotus hemilepidotus 1 1 2 Buffalo sculpin Enophrys bison 2 2 Gadid (not hake) Gadidae, not hake 1 1 *Sand lance Ammodytes hexapterus 1 1 *Capelin Mallotus villosus 1 1 *Shiner perch Cymatogaster gracilis 1 1 *Atka mackerel Pleurogrammus monopterygius 1 1 Dover sole Microstomus pacificus 1 1 Starry flounder Platichthys stellatus 1 1 *Gunnel sp. Pholididae 1 1 Total NISP Fish = 32,459 (Back terrace=17,539, House 1=15,067, Privy pit=33) Marine Mammals Porpoise/Dolphin Delphinidae/Phocoenidae 3 3 P. white-sided dolphin Lagenorhynchus obliquidens 3 3 Whale sp. Cetacea 1 1 2 Harbour porpoise Phocoena phocoena 1 1 Harbour seal Phoca vitulina 1 1 Fur seal Callorhinus ursinus 1 1 Pinnepedia, sm Pinnepedia 1 1 Dall's porpoise Phocoena dalli 1 1 Total NISP Marine Mammals = 13 Terrestrial Mammals Canid Canis sp. 4 4 Rodent (vsm) Rodentia (vsm) 2 2 *Vole sp. Microtus sp. 1 1 Mouse/vole sp. Rodentia (vsm) 1 1 Rodent (sm) Rodentia (sm) 1 1 Deer sp. Odocoileus sp. 1 1 *Shrew sp. Soricidae 1 1 Total NISP Terrestrial Mammals = 11 Domestic Mammals Domestic Dog Canis familiaris 7 7 Reptiles *Unid. frog Amphibian 1 1 Bird Duck (med) Anatidae (med) 1 1 Grand Total (NISP) 17,367 15,125 32,492 * Taxa found in the column samples but not in the excavation unit assemblage. \u00E2\u0080\u0093 157 \u00E2\u0080\u0093 Results The examined assemblage contains a total of 32,459 identified specimens (NISP) and a total of 57,890 skeletal specimens (including unidentified fish, birds, mammals). This examined assemblage comes from 168 discrete column sample levels representing a total excavated volume of 185.4 litres. Vertebrate remains are present in every examined sample. Fish are the overwhelmingly dominant taxonomic group in the column assemblage. Fish represent 99.9% of the total identified assemblage (NISP) and more than 99% of the taxa in each of the 12 examined column samples (Figure 5.15). Mammal and bird specimens are significantly less abundant although they are frequently encountered in the column sample assemblage. The majority of mammal and bird specimens are small, unidentifiable fragments of what were much larger once-complete elements. The extremely low proportion of identifiable mammal and birds in the column sample assemblage stands in contrast to the excavation unit assemblage, which has a much larger assemblage of identified (NISP) mammalian and bird remains (Frederick 2012). Taxonomic Richness Thirty-two unique fish taxa were recovered from the column sample assemblage (Table 5.2). Ten of these fish taxa as well as two small mammal taxa are not present in the excavation unit assemblage (Frederick 2012). These taxa are small-bodied and represent relatively minor proportions of the overall assemblage (denoted with asterisks in Table 5.2). It is notable that 12 of these taxa were not identified in the excavation unit assemblage, as their small bones are likely to have passed through the larger mesh sizes used during field-based faunal recovery in \u00C2\u00BC\u00E2\u0080\u009D mesh screens. \u00E2\u0080\u0093 158 \u00E2\u0080\u0093 By contrast, the excavation unit assemblage contains the same number of fish taxa (n=32) but includes 12 species that were not identified in the column sample assemblage20. These species also represent relatively minor components of the assemblage or were only identified to a genus level in the column sample assemblage (i.e., greenlings, perches, sculpins, and flatfish). Two of the largest species (sevengill shark and bluefin tuna) are rare in the site as a whole. Considering that the excavation unit assemblage is numerically larger and represents a dramatically larger examined volume (Frederick 2012), the absence these in the column sample assemblage is relatively unsurprising. However, it is important to consider how such small proportions may indeed represent significant and sizable contribution to the fishery, especially considering the un-sampled portions of the site and the time depth of human occupation. The adequacy of the sample is further discussed in Section 4. Contrasting the Column and Excavation Unit Assemblages The analysis of vertebrate fauna from 2-mm mesh identified a similar suite of fish species as the \u00C2\u00BC-inch excavation unit assemblage but resulted in a much greater recovery of small fish bones and thus a much greater number of bones per litre of examined volume. One of the interpretive consequences of such a shift is a dramatic change in the relative abundance of taxa present in the deposits. This shift has been widely observed by researchers working with fish assemblages throughout the Pacific (Casteel 1976a; McKechnie 2005a; Nagaoka 1994; Partlow 2006; Stewart, et al. 2003). 20 Sevengill shark, Bluefin tuna, Great sculpin, Spinyhead sculpin, Striped seaperch, Rock greenling, Kelp greenling, Rock sole, English sole, Sand sole, and Pacific sanddab. \u00E2\u0080\u0093 159 \u00E2\u0080\u0093 Figure 5.6 compares the relative percentage of the fine-screen column sample with the excavation unit assemblage, which illustrates the dramatic extent of the contrast. Notably, herring represent less than 4% of the excavation unit assemblage but they vastly outnumber all other fish in the column sample assemblage (81% NISP). This numerical dominance dramatically alters percentage data for all other species and has vital consequences for interpreting subsistence and resource harvesting practices in the site as a whole. Figure 5.6. Comparison between the relative abundance of fish remains in the column sample assemblage (2mm mesh) and the excavation unit assemblage (1/4-inch mesh) for the 10 most numerous taxa in the column assemblage. Numbered bars indicate rank order abundance. Within the excavation unit assemblage, hake numerically dominate followed by salmon, rockfish, greenling, and dogfish (Figure 5.6). However, there was a single deposit containing several thousand hake specimens (a specific unit in the sub-floor deposits of House 1). As this does not adequately represent the overall composition of the unit assemblage (across space and time), fish from this particular deposit were subtracted and percentage data recalculated (Frederick 2012). Figure 5.6 illustrates this \u00E2\u0080\u0098modified\u00E2\u0080\u0099 total, indicating that salmon are the most abundant fish taxa, followed by rockfish, hake, greenling and dogfish. Thus, while excavation unit assemblage provides invaluable perspective on the large-volume excavation, the collection strategy dramatically under-represents the taxonomic abundance of small fish and thus, fish in the site as a whole. \u00E2\u0080\u0093 160 \u00E2\u0080\u0093 Abundance and Ubiquity Figure 5.7 illustrates the overall composition of the column sample fish assemblage according to two analytically distinct measures of abundance (%NISP and % Ubiquity). It is interpretively significant that the relative abundance of fish specimens so closely corresponds to the sequence of ubiquity as this demonstrates that the most abundant taxa (%NISP) also occur very regularly in the deposit as a whole. Conversely, less numerous taxa occur very infrequently. There are some notable exceptions that indicate that some taxa are present in high numbers in only a few contexts (e.g., hake) while others are consistently present in low numbers (e.g., rockfish). These similarities and differences provide vital insight into the spatial and temporal variability of these taxa in the examined assemblage and are discussed in more detail below. The overall similarity between these two measures provides a level of confidence that the taxonomic composition of the total assemblage is broadly representative and that the numerically dominant species are also likely to be the most abundant in small portions of the assemblage. \u00E2\u0080\u0093 161 \u00E2\u0080\u0093 Figure 5.7. Taxonomic composition of the entire column sample fish assemblage according to two measures of abundance; 1) the relative percent of the number of identified specimens (%NISP) and 2) ubiquity (% occurrence in individual column levels). Taxa are shown in descending frequency according to %NISP with sample sizes indicated below bar charts. 90.5% 75.0% 75.0% 71.4% 25.0% 45.2% 42.3% 22.6% 14.3% 9.5% 11.9% 8.9% 7.7% 9.5% 6.0% 6.0% 5.4% 1.8% 4.2% 2.4% 0.6% 1.2% 1.2% 1.8% 2.4% 1.2% 2.4% 1.2% 0.6% 1.2% 0.6% 0.6% 0.6% 0.6% 0.6% 0.6% 0.6% 0.6% 0% 25% 50% 75% 100% Herring Anchovy Salmon Greenling Hake Rockfish sp. Dogfish Perch sp. Sablefish Petrale sole Flatfish sp. Lingcod Ratfish Irish lord sp. Prickleback sp. Pile perch Sculpin sp. Midshipman W-S greenling Cabezon Clingfish sp. Eulachon Herring/sardine Skate sp. Tomcod Halibut Smelt sp. Buffalo sculpin Pacific cod Red Irish lord Gunnel sp. Dover sole Starry flounder Sand lance Capelin Shiner perch Gadid (not hake) Atka mackerel !\"!\"#$%#&'(#$\"%&'\"()*+,-+.\"\"81.29% 5.76% 5.53% 2.21% 1.97% 1.04% 0.93% 0.28% 0.20% 0.13% 0.09% 0.08% 0.07% 0.06% 0.05% 0.04% 0.04% 0.04% 0.03% 0.02% 0.02% 0.02% 0.02% 0.02% 0.01% 0.01% 0.01% 0.01% 0.01% 0.01% <0.01% <0.01% <0.01% <0.01% <0.01% <0.01% <0.01% <0.01% 0% 25% 50% 75% 100% Herring Anchovy Salmon Greenling Hake Rockfish sp. Dogfish Perch sp. Sablefish Petrale sole Flatfish sp. Lingcod Ratfish Irish lord sp. Prickleback sp. Pile perch Sculpin sp. Midshipman W-S greenling Cabezon Clingfish sp. Eulachon Herring/sardine Skate sp. Tomcod Halibut Smelt sp. Buffalo sculpin Pacific cod Red Irish lord Gunnel sp. Dover sole Starry flounder Sand lance Capelin Shiner perch Gadid (not hake) Atka mackerel !\")*+,(#/01\"$\"234567\"\u00E2\u0080\u0093 162 \u00E2\u0080\u0093 Taxonomic Composition While a large number of fish taxa are present in the examined deposits (n=32), the ten most numerous taxa represent more than 99% of the identified specimens while the remaining 20 taxa represent less than 1% of the combined total. This indicates that the bulk of the fishing activity focused on a limited number of species. In the following section, I discuss the ten most abundant and ubiquitous fish as shown in Figure 5.7. To more fully document the temporal and spatial changes, I also employ MNI measures as well as density measures (e.g., NISP per m3) to further distinguish the relative differences in abundance (Figure 5.8 and Figure 5.9). Herring The column sample vertebrate faunal assemblage from Huu7ii is dominated by herring, which represents 81.3% (NISP) of the total column sample fish assemblage (Figure 5.7). Similarly, herring is the most ubiquitously occurring species in each of the 168 column sample levels (90.5%). Herring is also the most numerous taxon as indicated by MNI calculations (Figure 5.8). The numerical abundance and consistent ubiquity of herring suggest this species dominates the indigenous fishery at Huu7ii throughout the archaeologically examined period of human occupation. Herring is most dominant in the back terrace deposits (ca. 5,000-3,000 years ago), where herring represents 94.9% of NISP and 60% of MNI (Figure 5.8). Herring is less abundant but still dominates the House 1 assemblage in both the sub-floor and house-floor deposits (65% NISP and 25-28% MNI). Herring from the back terrace also exhibits a much higher NISP and MNI per m3 than in the later House 1 deposits (Figure 5.9), strongly indicating that herring use and as a consequence, fish utilization was more intense than in the \u00E2\u0080\u0093 163 \u00E2\u0080\u0093 House 1 deposits. Despite these differences, the consistency of herring utilization (as most strongly indicated by ubiquity) indicates continuity in cultural practice relating to herring throughout the examined history of occupation. Anchovy Anchovy are the second-most numerically abundant fish species recovered from the column sample assemblage (%NISP), but still only represent 5.8% of all identified specimens (Figure 5.7). Nevertheless, anchovy are tied with salmon as the second-most ubiquitous fish, occurring in 75% of all examined contexts (Figure 5.7). This indicates that even though anchovy (and salmon) are considerably less abundant than herring, they occur almost as regularly in the column sample assemblage. As estimated by MNI, anchovy represents over (10 %) of the assemblage but its rank-order abundance drops from second to third using this measure (Figure 5.8). Anchovy are sequentially more abundant over the three temporal periods rising from 2% of NISP in the in the back terrace assemblage to 6.8% in the sub-floor assemblage, to a high of 11.6% in house floor assemblage (Figure 5.8). Comparing the number of anchovy specimens (NISP) per cubic metre and individuals (MNI) per cubic metre indicates there is a progressive increase anchovy over time (Figure 5.9). As shown in Figure 5.8b, the ubiquity value of anchovy remains virtually the same over time, suggesting that this species was harvested consistently but was increasingly utilized relative to other fish. Salmon Salmon are the third-most numerically abundant taxon in the assemblage, representing 5.5% of the total NISP (Figure 5.7) and 8.8% of the total MNI (Figure 5.8). Salmon are also relatively ubiquitous in the examined assemblage indicating consistent use \u00E2\u0080\u0093 164 \u00E2\u0080\u0093 throughout the site deposits (70%). However, salmon are considerably more abundant in the House 1 deposits than in the back terrace based on NISP, ubiquity, and MNI, as well as NISP per m3 and MNI per m3. Within the House 1 deposits, salmon increase in abundance between the sub-floor and House floor deposits, reaching their highest abundance in the period just prior to site abandonment. Greenling The next most numerically abundant taxa is greenling which represents only 2.2% of the total NISP but has a ubiquity value of 70% indicating it is found in low quantities but is regularly utilized (consistently present in most examined contexts). This circumstance appears to impact the estimate of MNI for greenling, which shifts it to the second-most abundant fish in the assemblage (Figure 5.8). There is not a discernable difference in the relative abundance and ubiquity of greenling over the three temporal periods represented at the site, suggesting this taxon remained consistently important throughout the occupation of Huu7ii. Hake While hake is the fifth most numerically abundant fish in the column sample assemblage, it is significantly less ubiquitous than other abundant taxa. This indicates hake is not as consistently harvested over the 5,000-year period of occupation, but rather occurs much more frequently within a particular context or time frame (i.e., a portion of the House 1 deposits). This is strongly supported by the temporal trends in abundance where hake very rarely occur in the back terrace deposits but spike in abundance and ubiquity in the sub-floor deposits only to drop in abundance during the house floor deposits (Figure 5.8 and Figure 5.10). \u00E2\u0080\u0093 165 \u00E2\u0080\u0093 As discussed for the excavation unit assemblage (Frederick 2012), there is an extremely high density of hake remains present the lower portion of column N18-20/E6-8 from House 1 that disproportionately increases the calculated abundance for the total unit assemblage. In contrast to the very high abundance estimates for hake in the excavation unit assemblage, the column sample data indicate that this species represents not much more than 10% of the assemblage in the sub-floor deposits. Hake MNI estimates range from a low of 0.2% to a high of 9% in the sub-floor deposits (Figure 5.8). Rockfish Rockfish represent only slightly more than 1% of the total assemblage but are found in 45% of the examined column level samples (Figure 5.7) and comprise nearly 8% of the estimated MNI (Figure 5.12). Rockfish are considerably more abundant and ubiquitous in the sub-floor deposits than in either the back terrace or house-floor deposits indicating this taxon was used most intensively during this period (Figure 5.8). It is noteworthy that rockfish are slightly less abundant than greenling in the column assemblage (Figure 5.6) but strongly outnumber greenling in the excavation unit assemblage (Frederick 2012). This likely reflects a screen size bias favoring larger fish where greenlings have a smaller size-range than rockfish (McKechnie 2005a:217). Similarly, greenling has a slightly smaller size range than rockfish in this assemblage (Figure 5.13). Dogfish Dogfish are not particularly abundant in the column assemblage, representing less than 1% of total NISP (Figure 5.7). However, they do occur regularly throughout the examined occupational history as indicated by their consistent ubiquity (Figure 5.7b). \u00E2\u0080\u0093 166 \u00E2\u0080\u0093 Dogfish no not appear to change over time but do have slightly higher frequencies in the House 1 deposits relative to the back terrace (Figure 5.9). It is notable that dogfish are among the ten most abundant fish in the assemblage as these cartilaginous sharks have significantly fewer skeletal structures than bony fish, which would further diminish the potential importance of this taxon relative to others (Rick, et al. 2002). However, dogfish do have highly distinctive and numerous vertebrae (ca. 100-110) (Last, et al. 2007), which in comparison with many other fish, may help offset an otherwise sparse skeletal anatomy. Perch Perch is a taxonomic family level designation referring to several species (e.g., pile perch, surf perch, shiner perch) that occur moderately frequently in the overall assemblage but represent small abundance values (Figure 5.7 and Figure 5.12). Interestingly, perch progressively decrease in abundance over the three time periods among all relative abundance measures (%NISP, Ubiquity, %MNI) as well as absolute abundances as estimated by NISP and MNI per cubic metre (Figure 5.9). Thus, perch appear to be most regularly used in the back terrace deposits and progressively decrease over time. Sablefish Sablefish are long-lived fish that inhabit deep pelagic waters along the continental shelf edge (King, et al. 2000). Sablefish occur moderately regularly in the column sample assemblage (14% ubiquity) but represent only 0.2% of the total assemblage (NISP). Sablefish is nearly absent in the back terrace but increases in the later sub-floor and house-floor deposits (Figure 5.8). \u00E2\u0080\u0093 167 \u00E2\u0080\u0093 Based on visual comparison to mature fish in the UVic comparative collection, it appears the majority of the archaeological specimens are from small, juvenile-sized fish, which inhabit shallow inshore waters before maturation. The moderately frequent occurrence of juvenile-sized sablefish in the House 1 assemblage is consistent with the use of the nearshore water in the vicinity of site. However, additional measurements and metric comparison to known age specimens might improve understanding of where in the environment these fish were harvested. Petrale Sole In contrast to the noted ethnographic importance of halibut (Arima 1983; Sproat 1868:225), the most numerous \u00E2\u0080\u0098flatfish\u00E2\u0080\u0099 in the column assemblage is Petrale sole, a plate-sized flatfish that inhabits relatively deep-waters between 80 and 500 metres (Department of Fisheries and Oceans Canada 1999). Petrale sole represent a small portion of the overall assemblage and exhibit a moderate ubiquity in the House 1 deposits but are absent from the back terrace column deposits (Figure 5.8). Petrale sole are also the most numerous flatfish identified in the excavation unit assemblage (Frederick 2012), considerably out-numbering all other flatfish including halibut. \u00E2\u0080\u0093 168 \u00E2\u0080\u0093 Figure 5.8. Relative abundance over time for the ten most numerous fish taxa shown as percent of identified specimens (top), ubiquity (middle), and the estimated minimum number of individuals (bottom). Grouped bars represent fauna from the three temporally distinct deposits; the back terrace, sub-floor deposits, and the House 1 floor deposits. Figure 5.9. The number of specimens (NISP) and individuals (MNI) per cubic metre based on sampled volume from the back terrace (59 litres), the subfloor (37.4 litres) and house floor assemblages (82 litres). Note: estimates are overly precise (extrapolations) but nevertheless give a comparative basis for assessing change over time and differences between species and that are not limited to relative percentages. \u00E2\u0080\u0093 169 \u00E2\u0080\u0093 Other Fish As shown in Table 5.2 and Figure 5.7, a host of other fish taxa are present in the assemblage but occur in very low quantities relative to the ten most numerous taxa. It is important to consider that even though these taxa are not abundant in this assemblage, this may be not be warranted given the fact that this assemblage comes from 1) a small portion of what is a much larger site 2) may be present in greater quantities in unsampled areas; 3) represents several thousand years of human occupation, and 4) is only one of number of large shell midden sites whose occupants likely harvested similar species over extended time periods. Nevertheless, these small numbers of comparatively \u00E2\u0080\u0098rare\u00E2\u0080\u0099 taxa preclude a justifiable assessment of their abundance and history of use within this assemblage. Notably however, there is a relative paucity of halibut in the column and unit assemblages (Frederick 2012), a circumstance that is not uncommon in archaeological contexts on the Northwest Coast (Orchard and Wigen 2008). The relative lack of halibut appears to reflect differential processing of halibut (butchery on the beach and public distribution of meat) as well as the taphonomic effects of a lower bone density value for halibut relative to other fish (Smith, RE, et al. 2011a). Culinary processing such as the use of halibut for soup is a common practice that likely contributes further to the relative paucity of halibut in the assemblage (Blackman 1990; de Laguna 1972:392-400). There are some unique and relatively unexpected occurrences of taxa such as eulachon, a small oily smelt known to spawn in large rivers, as well as possible sardine in the back terrace and the House 1 deposits that indicates the potential presence of this southerly species that rarely occur at this latitude (see Wright et al. 2005) (Wright, et al. 2005). However, it is notable that that some of these \u00E2\u0080\u0098rare\u00E2\u0080\u0099 taxa occur much more frequently in the \u00E2\u0080\u0093 170 \u00E2\u0080\u0093 excavation assemblage than would be expected based on screen size alone. For example, rockfish, lingcod, cabezon, and Irish lords are recovered in considerably larger numbers in the unit assemblage (Frederick 2012). While this is likely due in part to a strategy that preferentially recovers large distinctive bones, it also potentially reflects the spatially restricted sample of fish remains in the column sample assemblage relative to the spatially and volumetrically larger sample from the excavation units. For instance, at the completely sampled house floor at Ozette, there are patterned concentrations of fish species in certain parts of the housefloor (Huelsbeck 1981) which might be less likely to be captured in column sample deposits (see discussion in Gray 2008) (Gray 2008). Thus, it remains possible that those taxa that are rare in the column assemblage might appear comparatively more abundant if a larger spatial area was sampled but cannot be resolved until more samples are examined at some future date. Temporal Trends House 1 (ca. 1,500-400 yr BP) To assess the temporal trends in fish use within the House 1 deposits, I plotted the relative abundance of three important fish taxa by individual column sample level (Figure 5.10). I compared trends across the four column samples spanning the full depositional sequence of House 1 and calculated both %NISP and NISP per litre in order to evaluate if an increase in the relative abundance might simply be a consequence of an increase in another taxa. This analysis also includes four additional column samples spanning the upper layers of the house floor deposits. \u00E2\u0080\u0093 171 \u00E2\u0080\u0093 Overall, there is strong similarity between %NISP and NISP per litre for the House 1 column samples providing greater confidence that the observed trends are not spuriously caused by fluctuations in other taxa, but reflect actual changes in abundance. In some cases however, the two sets of data (%NISP and NISP per litre) diverge which allows further clarification for specific temporal periods. For example, the increasing NISP per litre for salmon in column sample N10-12/W2-4 does not match the decrease in salmon %NISP due to a particularly large increase in herring per litre (left column of Figure 5.10). Secondly, there are distinct similarities in the temporal trends for specific taxa throughout the depositional sequence, suggesting that coherent change occurred in resource harvesting practices with the most striking trends noted for hake and salmon. In particular, hake do not occur in either the lowest levels or in the highest column sample levels from House 1 but exhibit a dramatic spike in abundance in the middle portion of each of the examined column samples (both %NISP and NISP per litre, Figure 5.10). In the upper layers of the house floor deposit (later in time), there are dramatic increases in salmon abundance (relative percent and bones per litre), in contrast to the consistently low abundance values for salmon in the lower, earlier levels. The increase in salmon also occurs in the upper levels of the four column samples that span only the house floor portions of House 1 deposits. The sudden and progressive increase in the relative abundance of salmon appears to follow the period of intense use of hake (Figure 5.10) suggesting a long-term and spatially coherent shift in the focus of resource harvesting practice at a household level. Thus, despite the potential for spatial variability in a house deposit, this aspect of resource use appears to have shifted throughout the house deposit sometime after approximately 800 years ago. \u00E2\u0080\u0093 172 \u00E2\u0080\u0093 In contrast to the dramatic shifts in the abundance of hake and salmon, the most numerous taxa in the assemblage, herring, show progressive long-term fluctuations in abundance that range widely over time. For instance, there are broad similarities in the trends in abundance between the four columns, with high abundances in both the lower levels and the upper levels contrasted with a period of comparatively low abundance in the middle portion of the depositional sequence. Interestingly, this corresponds to the peak in hake abundance suggesting that number of herring per litre drops when hake reach their highest frequencies (Figure 5.10). Similar to herring, anchovy appears to exhibit broad temporal trends in the House 1 deposits with somewhat consistent increases in abundance when herring decrease in abundance. \u00E2\u0080\u0093 173 \u00E2\u0080\u0093 Figure 5.10. Temporal trends (left to right) in the abundance of four important fish taxa in four different areas of the House 1 deposits (top row\u00E2\u0080\u0093hake, middle row\u00E2\u0080\u0093salmon, bottom row\u00E2\u0080\u0093herring). Dark lines indicate the relative percent (%NISP) while the grey lines indicate absolute abundance (NISP/litre) and are plotted on a secondary vertical axis. The individual data points on horizontal axis are 5-cm increments in depths below ground surface. Interpretations of patterning provided in the right column. Note the difference ranges for scales Interpretations Hake suddenly occur in the middle portion of House 1 deposit but are then consistently absent. Salmon exhibit dramatic increases in abundance in the upper levels of all column samples after Hake spike in abundance. Herring exhibits broad fluctuations in abundance but is generally less abundant when hake is present. Anchovy abundance broadly fluctuates but exhibits its highest abundances when hake spikes in abundance. \u00E2\u0080\u0093 174 \u00E2\u0080\u0093 Back Terrace (ca. 5,000-3,000 yr BP) Herring consistently dominates the fish assemblage in both of the back terrace column samples; representing more than 94% of total NISP and more than 70% of NISP in individual levels in all but three of the 38 examined levels from the two columns (Figure 5.11).This consistently high percentage is similarly reflected in the number of herring per litre, which vastly outnumbers all other taxa throughout the depositional sequence. These consistent trends occur in physically separate deposits (20+ metres apart) that have overlapping age ranges, and therefore likely represent a deposit-wide pattern over a broad 2,000-year period, between 5,000 and 3,000 years ago. Notably, the highest herring frequencies occur in the upper half of the deposits between 125 and 90 cm below the surface in both column samples (both %NISP and NISP per litre). This spike in abundance therefore appears to represent a deposit-wide period of particularly intense herring use. The most extreme case is a series of contiguous levels in column N2-4/W18-20, which culminates in a single 1-litre level containing nearly 5,000 individual herring bones, representing a minimum of 98 individual herring (approximately 15% of the entire fish assemblage). Observations by excavators recount a \u00E2\u0080\u009Ccrazy herring\u00E2\u0080\u009D layer in both excavation units where dense concentrations of herring were present across horizontal levels. Such high fish numbers suggests an intense collective effort focused on herring harvesting and processing presumably over a number of seasons or over the course of a few human generations based on the presence of multiple 5cm levels containing particularly high numbers of herring. Both preceding and following this period of particularly intense herring use, herring progressively rises and drops in abundance over successive levels (among both %NISP and \u00E2\u0080\u0093 175 \u00E2\u0080\u0093 NISP per litre). This indicates progressive fluctuations in fishing effort that likely relates to a combination of 1) the abundance of herring in the environment, 2) local conditions conducive to herring spawning habitat, and most vitally 3) the collective social capacity and incentive to collect and process that many fish. Anchovy are the second-most abundant fish species in the two back terrace column samples but are considerably less abundant than herring in all levels. Similar to herring, the back-terrace column sample N2-4/W18-20 contains many times more anchovy remains than in column N4-6/W0-2 reflecting the comparatively greater number of fish in this deposit. Both columns contain higher frequencies of anchovy in the levels preceding the dramatic spike in herring suggesting that inhabitants may increased their use of anchovy when herring harvests were lower. Combined with similar patterning observed in the House 1 deposits and considering that both fish are high in oil content and were likely caught using similar methods (e.g., rakes, nets). Collectively, the temporal trends in the abundance of the most numerous and ubiquitous taxa reflect active shifts in fishing practices between the deposition in the back terrace deposits and the creators of the House 1 deposits. The cultural and paleoenvironmental significance of these changes are discussed further in the discussion section. \u00E2\u0080\u0093 176 \u00E2\u0080\u0093 Figure 5.11. Herring and anchovy abundance in individual column sample levels from the two back terrace column samples. Dark lines indicate the relative percent (%NISP) while the grey lines indicate absolute abundance (NISP per litre) and are plotted on a secondary vertical axis. The individual data points on horizontal axis are 5-cm increments in depths below ground surface. NISP and MNI Figure 5.12 contrasts the NISP and MNI values for the ten most abundant taxa in the entire column assemblage. As previously discussed, MNI is a minimum estimate derived from the most numerous non-repetitive element in the individual column sample levels and is subject to numerous quantitative and thus interpretive uncertainties (Lyman 2008). Nevertheless, this derived measure of abundance further details the relative and rank order abundance of the top ten fish taxa. Interestingly, conversion to MNI for the entire assemblage roughly halves the relative abundance of herring (dropping from 81% to 39%). As a result, the percentages of all other taxa increase accordingly (Figure 5.12). This implies that these other fish have a consistently greater contribution relative to herring than indicated by NISP data alone. Conversion to MNI also shifts the rank order abundance for several of the 10 most abundant taxa. For example, greenling shifts from the fourth to the second most numerous fish (Figure 5.12). \u00E2\u0080\u0093 177 \u00E2\u0080\u0093 Conversely, the contribution of hake drops from the fifth to seventh rank despite an increase in %NISP. Overall, rank order abundance generally corresponds with NISP data and no other taxa occur in the top ten, confirming the numerical importance of these ten taxa. Figure 5.12. NISP and MNI data for the entire column sample assemblage. MNI represents the minimum number of individuals (%MNI) as derived from NISP data. Rank orders for NISP and MNI listed as numbers above bars. Fish Length Estimates Fish length estimations were conducted on hake, herring, rockfish, and greenling using linear regression introduced previously. I also measured the greatest transverse width of salmon vertebrae to estimate the range of salmon species present in the deposits following Cannon and Yang (2006). Fish length data are shown in Figure 5.13 and salmon vertebrae measurements in Figure 5.14. The sample was limited by the presence of measureable elements and therefore represents only a small percentage of all identified specimens. These data are combined from all time periods as small sample sizes preclude temporal comparisons. The majority of herring are estimated to be between 20 and 27 cm in length, indicating adult-sized (spawning-age) fish were the focus of harvesting based on comparison with studies conducted during the late 20th century (Hourston 1958; Tanasichuk 1997:2784; Taylor 1969:45) (e.g., Hourston 1958; Tanasichuk 1997:2784; Taylor 1969:45). Visual \u00E2\u0080\u0093 178 \u00E2\u0080\u0093 inspection of the histogram for herring length (indicates a normal distribution with a central tendency between 22.5-25.0 cm in length. This suggests that aboriginal harvesters targeted herring when they aggregated in large schools of mature adults. The absence of herring smaller than 18 cm (approximately 2+year old fish) suggests that juvenile schools were not targeted even though they are known to congregate in separate, smaller, more diffuse schools in bays and inlets (Hourston 1958). As shown in Figure 5.13, the estimated size-range for hake is between 30 and 55 cm, indicating the exclusive presence of adult-sized fish that are well past spawning size (Benson, et al. 2002). This suggests that the hake targeted by site occupants were not part of a year-round resident population that inhabit parts of southern BC (ibid.) but are likely part of the California/Oregon migratory population whereby the largest individuals seasonally migrate north into southern British Columbia during the height of summer. As noted by several fishery researchers (Agostini, et al. 2006; Benson, et al. 2002; McFarlane, et al. 2000), hake migration is strongly related to oceanographic shifts in climate whereby larger hake migrate further north during years with warm ocean temperatures. Thus the abundant presence of hake in the sub-floor deposits may indicate warmer ocean conditions when they occur in abundance (ca. 1,500-800 years BP). Rockfish and greenling length estimates suggest the majority of these two taxa are between 20 and 40 cm in length (Figure 5.13). The similar size distributions for both greenling and rockfish and the fact that they share rocky bottom kelp-bed associated habitats indicate they were harvested at the same time and likely using similar technologies. These size ranges are much smaller than fish caught in the modern sport fishery (McKechnie 2007a:218) but are identical to the lengths reported for the late-Holocene deposits at the \u00E2\u0080\u0093 179 \u00E2\u0080\u0093 village of Ts\u00E2\u0080\u0099ishaa in the Broken Group Islands (ibid.). Considering these taxa are non-migratory and vulnerable to over-harvesting, this small size is conducive to a more sustainable harvest strategy (Birkeland and Dayton 2005), wherein targeting younger, smaller animals that have less reproductive capacity is preferable to targeting older, larger fish that exert a disproportionate effect on the survivorship of larval offspring (Berkeley, et al. 2004). Salmon vertebrae measurements indicate a range of species present in the modest sample of whole vertebrae (Figure 5.14). The greatest proportion of vertebrae falls within the size-range of smaller species such as pink, sockeye, and coho indicating a possible concentration on these taxa. However, this contrasts slightly with measurements taken on a slightly a larger sample from the excavation units in 2004 (Frederick, et al. 2006b). A larger sample is needed to more fully document the salmon species represented and recent papers (Grier, et al. 2013; Huber, et al. 2011; Orchard and Szpak 2011) suggest that new metric and imaging approaches will yield useful insights. \u00E2\u0080\u0093 180 \u00E2\u0080\u0093 Figure 5.13. Histograms showing fish lengths measured elements of herring, hake, rockfish, and greenling. Herring is a fork length while the remaining are total lengths. Normal distribution curves shown over histograms. Note the large size of hake relative to other fish. 01234567890 10 20 30 40 50 60 70Greenling sp.n=400246810120 10 20 30 40 50 60 70Haken=290204060801001201401601800 10 20 30 40 50 60 70Pacific herringn=3940123456780 10 20 30 40 50 60 70Rockfish sp.n=29Fish Length in CentimetersMeasured ElementsMeasured ElementsMeasured ElementsMeasured Elements \u00E2\u0080\u0093 181 \u00E2\u0080\u0093 Figure 5.14. Measured salmon vertebrae from the column sample assemblage showing the size range for individual taxa as indicated using the criteria of Cannon and Yang (2006:132). Taphonomy, Formation Processes, and Sampling Taphonomy \u00E2\u0080\u009Cthe study of processes of preservation and how they affect information in the fossil record\u00E2\u0080\u009D (Behrensmeyer and Kidwell 1985:105) is a fundamental process that underlies the archaeological record (Schiffer 1987), and the archaeology of animal remains in particular (Gifford 1981; Lyman 1994). Researchers have noted that faunal assemblages may be altered, transformed, and/or contributed-to by non-human agents such as burrowing and scavenging animals and microbes (Erlandson and Moss 2001), as well as a host of physical processes such as soil chemistry and sea-level change (e.g., Linse 1992; Moss 1985; Stein 1984). Cultural practices such as butchery, transport, consumption, deposition, and re-deposition can also strongly condition the types and proportions of animal bones recovered in archaeological contexts (e.g., Monks 2003). In addition, it is critical to be aware that the results observed in a zooarchaeological analysis may reflect limitations of the sampling strategy rather than a purported observation of historical significance (Gray 2008; Thomas 1978). To assess how factors other than past human agency may have conditioned the skeletal assemblage and to consider how this may constrain the interpretive possibilities of 2468101214162 4 6 8 10 12 14 16 18Salmon Vertebral Diameter (mm)Salmon n=54CountAll SpeciesPinkSockeyeCohoChumChinook \u00E2\u0080\u0093 182 \u00E2\u0080\u0093 the assemblage, this section conducts analyses that explore how taphonomic and sampling factors may affect the assemblage and the strengths of the assemblage. Identification Rates The ratio between \u00E2\u0080\u0098identified\u00E2\u0080\u0099 (NISP) and \u00E2\u0080\u0098unidentified\u00E2\u0080\u0099 (NSP)21 specimens in a faunal assemblage reflects variability in the degree of identifiability and fragmentation as well as an analyst\u00E2\u0080\u0099s confidence in identification. To assess the potential relationship between identification and fragmentation, I examined the ratio of unidentifiable to identifiable remains for fish within the 12 column samples (Figure 5.16). Each of the nine column samples from the House 1 deposits has a similar rate of fish identification (ranging between 40-50% of all fish specimens). This consistency in identification suggests fragmentation is similar within the House 1 deposits despite the potential for variability in a house floor. In contrast, the two older back terrace column samples have much greater rates of identification (75\u00E2\u0080\u009385%). This is a counterintuitive result as one might expect a much higher degree of fragmentation and lower identifiability in such an older deposit (ca. 5,000\u00E2\u0080\u00933,000 yrs ago). However, in this case, the pattern of high identifiability is likely driven by the large numbers of identifiable herring remains in these two deposits (Figure 5.15 and Figure 5.16). Regardless, the fact that larger numbers of small herring are abundantly present in these older de- posits indicates excellent preservation conditions during that time. The single column from the privy pit has a notably low rate of identification (15%) but also has the smallest assemblage size and examined volume of the 12 column samples. 21 \u00E2\u0080\u0098Identified\u00E2\u0080\u0099 refers to specimens that can be assigned to taxonomic family, genus, or species whereas \u00E2\u0080\u0098unidentified\u00E2\u0080\u0099 refers to specimens that are only recognizable as fish (i.e., usually ribs, branchials, and/or fragmented bones). \u00E2\u0080\u0093 183 \u00E2\u0080\u0093 This undated sample is close to the modern shoreline, and in the context of regional sea level history and site formation processes, it likely dates to within the past 500 years. Figure 5.15. The identification rate for fish, mammals and birds in the column sample assemblage indicating the number of identified and unidentified specimens for each category. Figure 5.16. Identification rate for fish remains in the 12 examined column sample assemblages (site area is noted at bottom). Density of Recovered Fish Remains Based on the total number of fish remains present in the column sample assemblage, there is an estimated density of 175,000 identifiable (NISP) fish bones per cubic metre (from 2mm mesh). The identified assemblage from the temporally older and spatially distinct back terrace deposits (ca. 5,000-3,000 yr BP) has an even greater estimated density of 294 identifiable fish remains per litre while the younger House 1 deposits have an estimated 123 bones per litre. While these estimates do not incorporate the considerable variability among individual column sample levels, they nevertheless, indicate the considerable scale and intensity of the fishery as archaeologically represented in the examined deposits. The greater number of bones per cubic metre present in the older deposits demonstrates the excellent preservation conditions and indicates a comparatively more intensive fishery at the site during that time (Figure 5.17). Conversely, this may also indicate that house-floor deposits are not as conducive to the preservation of bone and/or are subject to differing depositional conditions (e.g., trampling and house cleaning). \u00E2\u0080\u0093 184 \u00E2\u0080\u0093 To further investigate the absolute abundance of fish remains in the examined assemblage, I developed estimates for the number of individual fish per litre in the three temporally distinct deposits. Figure 5.17 indicates that the older back terrace deposits have the highest number of fish per litre which is predominantly due to a greater number of herring overall as there are fewer other fish per litre in this deposit. Conversely, later in time in the sub-floor and house-floor deposits, there is a substantial reduction in the number of herring per litre but an increase in the number of other fish, which appears stable during both periods. Figure 5.17. Estimated number of individual fish per litre (numbers have been rounded up) in the three temporally distinct deposits at Huu7ii. Numbers of herring and all other fish are based on the total number of individual fish divided by the total examined volume for each deposit. Sampling Effort To assess the relationship between taxonomic richness and sampling effort within the column sample assemblage (cf., Lepofsky and Lertzman 2005; Lyman and Ames 2004, 2007; Monks 2000), I created \u00E2\u0080\u0098collectors curves\u00E2\u0080\u0099 depicting the stepwise relationship between taxonomic richness and sampling intensity (Figure 5.18 and Figure 5.19). These figures show \u00E2\u0080\u0093 185 \u00E2\u0080\u0093 the \u00E2\u0080\u0098rate\u00E2\u0080\u0099 at which new fish taxa are identified as new column levels are cumulatively added together (i.e., new taxa found in individual column levels). Unsurprisingly, this analysis reveals that the greater number of identified specimens, the more fish taxa were identified. Importantly however, the \u00E2\u0080\u0098rate\u00E2\u0080\u0099 of novel identifications slows dramatically as sample size increases. For instance, Figure 5.18 illustrates that 30 fish species were identified when the sample size reached 15,000 specimens but an additional 15,000 specimens needed to be examined before two additional fish species were identified. Overall, this \u00E2\u0080\u0098slowing\u00E2\u0080\u0099 in the rate of identification indicates that the analysis passed a threshold whereby a larger sample size does not dramatically increase the number of new taxa. Figure 5.18. Collectors curve for taxonomic richness in the column sample assemblage depicting the linear relationship between the cumulative number of identified specimens and the number of new taxa present. Figure 5.19 compares the collector\u00E2\u0080\u0099s curves for from House 1 and the back terrace. This comparison reveals the similar level of sampling intensity in the two temporally distinct deposits. Both assemblages contain similar sample sizes and have reached relatively \u00E2\u0080\u0098level\u00E2\u0080\u0099 portions on the \u00E2\u0080\u0098curve\u00E2\u0080\u0099. Interestingly, this comparison also indicates that a greater number of fish taxa are present in the House 1 assemblage (n=29) assemblage than in the back terrace (n=19), even though the back terrace has a moderately larger sample size. The differences in the shape of these curve is likely due to the high number of herring present in the back \u00E2\u0080\u0093 186 \u00E2\u0080\u0093 terrace but the overall differences in the number of taxa appears to indicate substantial differences in the use of fish in these different periods in time (an issue discussed elsewhere). Figure 5.19. Collector\u00E2\u0080\u0099s curve comparing the taxonomic richness of the House 1 and the back terrace assemblages. The observation that both assemblages appear to have passed the \u00E2\u0080\u0098steepest\u00E2\u0080\u0099 portion of the curve indicates that the level of sampling effort adequately encompasses the taxonomic richness of these deposits. That said, neither collector\u00E2\u0080\u0099s curve appears to \u00E2\u0080\u0098level out\u00E2\u0080\u0099 indicating that new fish species will likely be identified if additional samples are examined. Thus, the assemblage has by no means been sampled to complete \u00E2\u0080\u009Credundancy\u00E2\u0080\u009D (Lyman and Ames 2004) but appears adequate for evaluating differences between them due to similar sample sizes and similarly shaped collector\u00E2\u0080\u0099s curves. Shellfish and the Preservation of Bone There is a widely held observation in shell midden archaeology of an association between the presence of shellfish and the preservation of bones (Erlandson 2001; Linse 1992) whereby the deposition of shell creates alkaline conditions conducive to the preservation of bone. To assess whether this relationship has influenced the amount of bone present in the examined assemblage, I evaluated the strength of this relationship by using correlation between \u00C2\u00BC\u00E2\u0080\u009D shell and 2mm bone for 256 samples processed samples (weights in grams per litre). However, I found no correlation between these variables for the assemblage \u00E2\u0080\u0093 187 \u00E2\u0080\u0093 as a whole but rather observe a very insignificant relationship with wildly varying amounts of bone and shell in individual column samples (Figure 5.20). I further examined this relationship within the back terrace, as these older deposits (ca. 3,000-5,000 BP) presumably might be more affected by such a relationship but again found no correlation. These results suggest that the presence of shell does not have a direct influence on the amount of bone present in individual levels and provides support for the interpretation that the deposition of bone is a function of cultural practices rather than an artifact of bone diagenesis. However, this analysis does provide some support for a taphonomic distinction between the main village and the back terrace in that both shell and bone are found in a wider range of quantities per sample in the House 1 deposits relative to the back terrace (Figure 5.21). In particular, the maximum weight of \u00C2\u00BC-inch shell per litre and the bone weight in grams per litre have lower values in the back terrace than House 1. This suggests that either the amount of shell and bone is less abundant in these older deposits or is subject to greater rates of fragmentation. However, this does not appear to be positively correlated with the preservation of bone. Figure 5.20. Scatterplot showing the non-linear relationship between the weight of bone and the weight of shell in individual column sample levels within the back terrace and the House 1 deposits (n=256). Note the lack of a strong correlation overall but the comparatively higher range for shell and bone weight in the House 1 deposits. \u00E2\u0080\u0093 188 \u00E2\u0080\u0093 Condition of Bone Assemblage Numerous skeletal specimens from the examined column sample assemblage show evidence of burning, erosion, cutmarks, and/or digestion. The frequencies of these alterations have implications for the preservation of the bone assemblage and for interpreting food preparation and bone disposal practices in the site as a whole. Mammal bone specimens are the most frequently altered with a total of 45% of bone fragments from House 1 deposits and 8% of back terrace mammals being affected by either burning, erosion, cutmarks, and/or digestion (Figure 5.21). Bird bones are the second most frequently altered specimens with 8.1% from House 1 being affected while none of the bird from the back terrace appear altered. In contrast, very few fish remains show any evidence for burning, digestion, and no cutmarks were observed on fishbone. The majority of burned and mammal and bird bones consist primarily of unidentifiable fragments of what were much larger skeletal elements and so appear disproportionately frequent relative to the much more numerous fish remains. The low frequency of taphonomic alterations on fish remains may be masked by their vulnerability to burning and digestion, as such bones might be much less likely to survive the digestive process (Jordan 1997). However, the assemblage does not lack fish bones and thus, such processes do not account for the exponentially more abundant fish in the assemblage. Among the two examined areas of the site, the House 1 deposits contain considerably higher percentages of taphonomically altered bones as might be expected for deposition in a household context (e.g., cooking and consuming food). The large number of digested and eroded mammal bones indicate the influence of carnivore modification, most likely domestic dogs (Canis familiaris) that are abundant in the excavation units in both the back terrace and \u00E2\u0080\u0093 189 \u00E2\u0080\u0093 the House 1 assemblages (Frederick 2012). Gnawing and consumption of bone fragments (for grease and marrow), is a common canine activity and one that might have been conducive to temporarily removing some mammal bones from a floor surface. Comparatively few fish bones show evidence for digestion with the greatest frequencies for digestion occurring on salmon (0.6%) and then greenling (0.1%). Bird and fish bones also show higher percentages of burning and calcification in the House 1 deposits, which likely reflects culinary processing and/or bone disposal practices. Notably, the highest rate of calcification was observed for birds in the House 1 deposits suggesting high-temperature burning may have been a preferred culinary method or bone disposal method. Among the fish, Ratfish had the highest incidence of burning (34%) and this cartilaginous species lacks all skeletal structures except six teeth. Petrale sole was the second most frequently burnt and calcined fish (12%) followed by lingcod (10%), Irish lord (5%), greenling (3.5%), salmon (3.1%), and herring (0.8%). These frequencies suggest that roasting was a more common method of preparation among these taxa. Cut marks are noted on a few mammal bone shaft fragments, all of which are too fragmentary to identify to a specific skeletal element let alone to species. These nevertheless indicate the use of sharp-edged tools in the butchery process. \u00E2\u0080\u0098Chop marks\u00E2\u0080\u0099 are noted on several mammal bones indicating direct percussion by a blunt object. In addition, several thin and warped \u00E2\u0080\u0098chips of mammal bone may reflect bone artifact production or carving detritus. Collectively, this patterning suggests that the fish assemblage appears to be the least subject to destructive taphonomic factors and therefore most closely reflective of harvesting practices. In contrast, the highly fragmentary and rarely identifiable mammal remains from the column assemblage provide a much narrower view of the species utilized, but add \u00E2\u0080\u0093 190 \u00E2\u0080\u0093 considerable detail to the taphonomic dimension of mammal bones present in the fine screen column sample mesh. Figure 5.21. Burning and other modifications to bone specimens in the column sample assemblage by fish, mammal, and bird categories. Calcined refers to highly burnt \u00E2\u0080\u0098white\u00E2\u0080\u0099 bones. Element Representation The relative proportions of cranial, vertebral, and caudal elements for the ten most numerous fish taxa are depicted in Figure 5.22. There is broad consistency in that the greatest proportions of identified elements are vertebrae, followed by cranial and then caudal elements. Perch have the largest proportion of cranial elements but this is due in part to their numerous corn-kernel-like teeth that readily separate from the jaw structure, which over represents the proportion of cranial elements. However, this is not the case for hake, rockfish, and greenling, where over 25% of elements are from crania, which is disproportionate to their skeletal anatomy (see, Wigen 2005). These latter proportions may reflect the robusticity of these species\u00E2\u0080\u0099 cranial elements and/or high discard rate of head bones for these species as opposed to additional culinary processing. In contrast, salmon have one of the lowest proportions of cranial elements among the ten most numerous fish as has been observed elsewhere in Barkley Sound (Frederick and Crockford 2005:192) and on the Northwest Coast (Matson 1992; Orchard 2007; Wigen and !\"#$%&'& ()*+&,%--)*%& !\"#$%&'& ()*+&,%--)*%& !\"#$%&'& ()*+&,%--)*%&./0!& 121123& (/45&6#-7%8& 9:;<=& 9:9>=& ?:@=& @:<=& ':>=& 9=&*)A*B7%8& 9:?<=& 9:9<=& C:'=& 9=& >:@=& 9=&*#DE)-+$& 9=& 9=& 9:CF=& 9=& 9=& 9=&%-\"8%8& 9:9?=& 9:'@=& 'C:G=& C:9=& 9=& 9=&8BH%$D%8& 9:9@=& 9=& =& 9=& 9=& 9=&9=&@=&'9=&'@=&<9=&<@=&!\"#$%&%'(&\"\")*%+,-%.)\"\"/,.+&0\"1.,23\"6#-7%8& *)A*B7%8& *#DE)-+$& %-\"8%8& 8BH%$D%8& \u00E2\u0080\u0093 191 \u00E2\u0080\u0093 Stucki 1988; Wigen, et al. 1990; Wigen 2005:92). It remains unclear whether this pattern is due to the differential transport or preservation, the fragility of salmon head bones, or the processing techniques such as boiling, smoking, or some taphonomic combination thereof (Butler and Chatters 1994). However, the persistence of this pattern in the Huu7ii assemblage reaffirms this is as an intriguing and underexplored research question. Notably however, fragmentary salmon vertebrae make up over 80% of all identified salmon remains in the assemblage whereas complete vertebrae only represent 5% of identified salmon. Salmon vertebrae are visually distinct from other fish remains due to their unique skeletal shape and texture that produces higher identification rates (Cannon 2000a; Orchard 2007; Wigen and Stucki 1988). Thus, while the ubiquity of salmon throughout assemblage demonstrates it was regularly used, the high proportion of fragmentary vertebrae disproportionately contributes to an apparent lack of cranial elements. Therefore, Figure 5.22 also includes a second element ratio for salmon that excludes fragmentary vertebrae and correspondingly increases the proportion of cranial remains to nearly 13%. Figure 5.22. Relative percent of cranial vertebral and caudal elements for the 10-most numerous fish in the column sample assemblage. The cranial category includes the neurocranium, dermocranium, suspensorium, hyoid arch, and branchial arch. Vertebral category also includes the pectoral and pelvic girdle. Caudal includes tail elements and miscellaneous elements. \u00E2\u0080\u0093 192 \u00E2\u0080\u0093 Discussion and Interpretation Sampling Adequacy The column sample assemblage contains over 32,000 identified fish remains from 32 taxa collected from a number of temporally and spatially distinct contexts in House 1 and the back terrace. The most abundant taxa are also the most commonly occurring, indicating that the most numerous species are also the most widely utilized and thus, of particular significance in interpreting the collective social and economic practices of the people who created the House 1 and back terrace deposits. What is much more challenging to interpret in the column sample assemblage are those \u00E2\u0080\u0098rare\u00E2\u0080\u0099 taxa that represent food gathering and consumption activities that may be particularly socially valued but occur comparatively infrequently (e.g., whales and whaling). As noted by Sahlins (2010), the social construction of \u00E2\u0080\u009Cvalue\u00E2\u0080\u009D is often related to culturally defined notions of rarity or \u00E2\u0080\u009Calterity\u00E2\u0080\u009D (i.e., exoticness). He argues, \u00E2\u0080\u009Cscarcity is largely a function of exchange-value rather than the other way around\u00E2\u0080\u009D (Sahlins 2010:380). In an archaeological context, this implies that rarely occurring items in an assemblage can have particular cultural significance, but that scarcity will often frustrate attempts to adequately interpret their role in a given archaeological context (Gray 2008). Another interpretive hazard is that such archaeologically rare items may also reflect spontaneously random occurrences which may have less interpretive significance than they sometimes receive, especially in comparison to more common yet perhaps more mundane elements of everyday life that are nonetheless vital to social and economic relations. Whichever the case, the series of analyses presented in this paper indicate the column sample faunal assemblage is a robust sample, representing common and widespread food harvesting practices present in small volumes of closely examined cultural deposit. I focused \u00E2\u0080\u0093 193 \u00E2\u0080\u0093 on vertical \u00E2\u0080\u0098columns\u00E2\u0080\u0099 of sediment from multiple contexts to provide a strong basis for interpreting continuity and change over time, but recognize this approach is less conducive to understanding horizontal spatial patterning within a household context. However, in order to assess the more infrequent dimensions of social and cultural life an even larger sample size and a larger number of examined contexts is required. Fortunately, this is the case as the excavation unit assemblage (Frederick 2012) examined a much larger number of contexts and more rarely occurring taxa, revealing spatial patterning at a household scale. Continuity in Resource Use Although there is considerable dynamism in the composition and proportion of fish in individual column sample levels (Figure 5.10), there is also broad consistency and continuity across space and over time (Figure 5.7 and 5.8). The pervasiveness of fish remains in the site deposits reflects the importance of fish and fishing in the daily lives and social relations of the inhabitants of Huu7ii. Fishing targeted a wide range of species but intensively focused on a number of taxa. These are cultural patterns that indicate consistent and enduring connections between the site occupants and particular fish and the places from which they were harvested. Such information represents a vital and important aspect of how people created and sustained a community at this location for over five millennia. The numerical dominance of herring, as indicated by multiple measures of zooarchaeological abundance over a 5,000-year period, is particularly significant to the interpretation of social and economic practices at Huu7ii. Herring represent an excellent winter and spring food, one that is rich in oil and could be mass harvested and stored, then consumed over extended periods (Arima 1983; Hart, et al. 1939; Jewitt 1807; McKechnie 2005b; Smylie 2004). Herring and herring roe were a form of wealth that could be traded and \u00E2\u0080\u0093 194 \u00E2\u0080\u0093 distributed widely in a village setting, between households, within families, at feasts, and with other Nations in Barkley Sound and beyond (Drucker 1951:40; Jewitt 1807:37; Sproat 1868:224). Herring length estimates (Figure 5.13) indicate harvests concentrated on adult-sized fish that were likely caught in large schools prior to spawning. In summary, the dominance of herring in the column sample assemblage, both over time and consistently within small volumes of cultural deposit, reflects a particularly concentrated collective investment, focused on processing large numbers of fish for immediate consumption and long-term storage. Community members likely simultaneously harvested the other marine predators (birds and marine mammals) that also consume herring (Monks 1987) and thus the occurrence of herring likely was a highly anticipated and prepared for time of the year. Of course, numerous other fish species are also regularly present in the assemblage and these additionally provide a basis for interpreting the persistent and everyday aspects of social, economic, and ecological relations at Huu7ii. In particular, anchovy and salmon are two taxa that represent similarly abundant contributions to the overall assemblage and are ubiquitously present, indicating highly regular use. Anchovy, as a small schooling fish, may have served as a supplement the comparatively more intensive harvest of herring. Salmon similarly occur in consistent but relatively low frequencies throughout the 5,000-year record but increase dramatically during the last 500 years of occupation (see discussion below). Greenling and rockfish are two non-migratory taxa that also consistently occur in the assemblage and could be readily obtained within the vicinity of the village at all times of the year. Numerous other fish species additionally played important roles in the daily lives of site inhabitants and further examination of these individual taxa is needed. \u00E2\u0080\u0093 195 \u00E2\u0080\u0093 From a methodological standpoint, one of the intriguing consequences of calculating the minimum number of individual (MNI) fish in the fine-screened column sample assemblage is the considerable difference between the MNI and NISP values of herring. Herring are the most numerous fish according to both measures, but the NISP value for herring is more than twice as large as its MNI value (Figure 5.12). This suggests that even though herring represent more than 80% of total NISP, their nutritional contribution (according to MNI) may be equivalent to less numerous but individually larger fish such as salmon. However, this does not account for the uncertainty of MNI calculations (Lyman 2008) and the complexity of estimating \u00E2\u0080\u0098meat weight\u00E2\u0080\u0099, as well as how fish taxa (particularly salmon) change considerably in abundance over time. These estimates will remain under-resolved until additional analyses consider variability over time and space. However, the present analysis represents an important first step towards reconciling the difference between NISP and MNI data for small column sample assemblages. Change in Resource Use While there is a strongly expressed continuity in resource use within the Huu7ii fish assemblage, there are two particularly robust temporal changes in the abundance of fish in the House 1 deposits: 1) salmon increase dramatically during the last 500 years of occupation, reflecting a considerable shift from earlier periods and 2) hake occur in large numbers in the middle of the House 1 deposits (prior to ca. 700 BP) but are absent or have low frequencies during all other periods (Figure 5.10). The shift towards increasing use of salmon indicates a broad cultural change in the social economy of House 1. This change may reflect an expansion of the political territory of the village, such as securing access to a productive salmon river from, or in cooperation with, \u00E2\u0080\u0093 196 \u00E2\u0080\u0093 another polity. This also may be a product of a more intensive fishery in the immediate vicinity of the village, such as the large troll-based sport fishery for salmon that currently operates off Kirby Point on Diana Island (1 km from the village). Alternatively but not exclusively, the sharp increase in salmon may additionally reflect progressively more favorable oceanographic conditions conducive to the intensification of salmon fishing at a community scale. A similar and contemporaneous trend has been observed at other archaeological sites in Barkley Sound, such as at the Ts\u00E2\u0080\u0099ishaa village in Tseshaht territory (McKechnie 2005b, 2007a; McMillan, et al. 2008) and at Ma\u00E2\u0080\u0099acoah in Toquaht territory (Monks 2006), as well as elsewhere on the northern (Orchard and Clark 2005) and southern (Wigen 2005) Northwest Coast. These local and regional patterns may relate to large-scale climatic changes in the North Pacific that occurred after AD 1200 (Anderson, et al. 2005b; Finney, et al. 2002; McKechnie, et al. 2008). Future research will help identify the cultural, historical, and climatic circumstance that may be driving these trends. Of particular importance will be identifying the particular salmon species targeted during this period of increased salmon utilization. Further examination of salmon specimens from Huu7ii is warranted and new morphometric techniques (Grier, et al. 2013; Huber, et al. 2011), as well as ancient DNA (Cannon and Yang 2006; Cannon, et al. 2011), have the potential to address this question of a period of dynamic change. The second particularly notable change in the abundance of fish in the column sample assemblage is the sharp increase in the abundance of hake throughout the House 1 deposits (Figure 5.10). While this trend is moderately observed at other sites in Barkley Sound (McKechnie 2007a:214), it appears to be much more strongly expressed at Huu7ii. This increase may reflect a local specialization, such as community access to particularly \u00E2\u0080\u0093 197 \u00E2\u0080\u0093 productive hake fishing locale. However, since hake are strongly influenced by marine climate (Agostini, et al. 2006; Benson, et al. 2002; McFarlane, et al. 2000) and the length measurements suggest that primarily large migratory adults were harvested (rather than a local population in which a range of sizes would be expected), the occurrence of hake provides support to the interpretation of a period of warmer ocean conditions prior to AD 1200. It is also significant that the sharp increase in hake occurs prior to the increase in salmon, further suggesting that climatic factors may be influencing this cultural change. Within the back terrace, the most notable temporal change is the higher abundance and higher numbers of herring per litre relative to the House 1 deposits (Figure 5.8 and 5.9), indicating fisheries were comparatively more intensive during this mid-Holocene occupation (ca. 5000\u00E2\u0080\u00933000 BP). The progressive increase in the middle levels of both deposits suggest a particularly intensive peak in the utilization of herring followed by a progressive decline in the upper levels (Figure 5.11). A possible factor that might have contributed to the higher abundance of herring in the back terrace is the beach sand present beneath these cultural deposits that represent a former intertidal zone. This may have been an ideal habitat for herring spawning and a reason for intensive human use and settlement. These fine beach sands were likely deposited when sea levels were 3\u00E2\u0080\u00934 m higher during the mid-Holocene (Friele and Hutchinson 1993) and contrast with the steep rocky intertidal storm beach that dominates the shoreline today. It is therefore possible that during the back terrace occupation, the intertidal zone may have been a herring spawning location and may even have included a fishtrap. Additional paleo-topographic reconstruction of this raised beach landform will help add substance to this interpretation. \u00E2\u0080\u0093 198 \u00E2\u0080\u0093 Seasonality The dominance of herring in the back terrace deposits may indicate a comparatively more seasonal use of the site during the period between 5,000 and 3,000 years ago. However, a year-round use of the site is indicated by the continuous deposition of shell midden sediments (Figure 5.4) and the occurrence of mammalian, bird, and fish species that are summer and fall seasonal indicators abundantly present in the back terrace excavation unit assemblage (Frederick 2012). In addition, the back terrace column samples also consistently contain anchovy and salmon that may have been more readily obtained in summer and fall, as well as herring that may have been used more readily in winter and spring. Later in time, during the House 1 occupation, the column sample assemblage contains a comparatively more \u00E2\u0080\u0098even\u00E2\u0080\u0099 distribution of fish taxa (Figure 5.8) as well as a greater number of fish taxa (Figure 5.19), suggesting an even more substantial degree of year-round occupation. This is also supported by the excavation unit assemblage. An important consideration in determining the seasonality of archaeological faunal assemblages is that the assessment of seasonality is often based on the presence of \u00E2\u0080\u0098indicator species\u00E2\u0080\u0099 in a given depositional context (e.g., Ford 1989). However, a key question is how consistently these species occur: are they ubiquitous and abundant, or are occurrences rare and a thus a reflection of sampling intensity rather than site seasonality? This column sample assemblage provides important insight into this issue, as there are several fish taxa that are both abundant and ubiquitously occurring. Thus, their consistent presence across dozens of depositional contexts (i.e., multiple small temporal snapshots) may represent a more robust indication of seasonality than infrequently occurring but more seasonally diagnostic taxa. Another key issue in the determination of seasonality is that species may occur over a broader range of seasons than is conventionally understood, especially considering the \u00E2\u0080\u0093 199 \u00E2\u0080\u0093 impact of 20th century industrial commercial exploitation of the marine environment. For instance, the consistent occurrence of adult herring in archaeological deposits on the Northwest Coast is often interpreted to mean that herring were targeted exclusively during the spring spawning season. However, this does not often encompass the range of variability on the coast, particularly relating to the period prior to herring spawning. A series of historic observations suggest that herring were harvested during a much longer period of the year. For instance, John Jewitt\u00E2\u0080\u0099s (1807) journal recounting his two and a half years of captivity among the Mowachaht in Friendly Cove (120 km north of Barkley Sound) documents the consumption of herring and her- ring roe multiple times in all months of the year except July and August (McKechnie 2005b:103). Modern industrial herring fisheries convention- ally began fishing for herring in October, well prior to the winter and spring spawning periods (Mackinson 1999; Taylor 1955:111; Tester 1933:287). Department of Fishery and Oceans records going back to the 1940s (McCarter, et al. 2005) show herring spawn once regularly occurred as early as late January and early February on western Vancouver Island. These observations indicate that pre-industrial herring populations may be better characterized as a late fall and mid-winter food in addition to the spring spawning period. Moreover, the storability of herring would mean that it could be consumed for months afterwards and thus well into summer. Similarly, anchovy is a species said to occur more frequently in summer as it is at the northern edge of its latitudinal range (Department of Fisheries and Oceans Canada 2002). However, paleoecological analysis of fish scales recovered from a geological sediment core in nearby Effingham Inlet (~15 km north of Huu7ii) indicates that anchovy were the dominant fish in that particular inlet over the past 4,000 years (Wright, et al. 2005:376), \u00E2\u0080\u0093 200 \u00E2\u0080\u0093 implying that anchovy are a resident non-migratory population and were likely avail- able throughout much the year. The occurrence of salmon from multiple size ranges and species (Figure 5.14) suggests that this taxon could also have been harvested at multiple times of the year (spring through fall). Thus, rather than seeking to identify seasonal indicators, it is additionally important to consider the consistent utilization of species that are present during multiple seasons as a proxy for year-round site occupation. Comparisons to Other Assemblages The examined column sample assemblage can be compared with available precontact archaeological faunal assemblages in Huu-ay-aht territory and elsewhere in Barkley Sound. However, there are only three other sites within Huu-ay-aht territory where fine-screen fish remains (smaller than \u00C2\u00BC-inch mesh) have been used. The nearby Huu-ay-aht village of Kiix7in (DeSh-1) contains a modest vertebrate assemblage (NISP = 700), in which herring are the most abundant fish (48% NISP fish), followed by salmon (22%) and greenling (16%) (Wigen 2003b). Herring is considerably less abundant (18% NISP) but still the second most frequent fish in the small assemblage (NISP = 171) from the adjacent defensive site at Kiix7in (DeSh-2), which is dominated by greenling (49%) (Wigen 2003b). An additional fine-screen assemblage (NISP = 187) from the Klanawa Rivermouth (DeSf-6), 20 km south of Cape Beale (McKechnie 2007c:9), is dominated by salmon and greenling and only contains a small percentage of herring (6% NISP). Northwest of Huu7ii in the Broken Group Islands, herring are also dominant among the fish at the large Tseshaht village of Ts\u00E2\u0080\u0099ishaa (DfSi-16&17), followed by anchovy, rockfish, greenling, and salmon (McKechnie 2005b; McMillan, et al. 2008). Herring is similarly the most abundant fish (58%) in a small (NISP = 151) column sample assemblage \u00E2\u0080\u0093 201 \u00E2\u0080\u0093 from a defensive site on Clarke Island (DfSi-26) in close proximity to Ts\u00E2\u0080\u0099ishaa (McKechnie 2007b:29) and is overwhelmingly dominant (85%) in an assemblage from Dodd Island in a protected portion of the Broken Group Islands (Wigen 2009). Two fine-screened assemblages from Ucluelet Harbour in western Barkley Sound, Little Beach (DfSj-100) (Weathers, et al. 2008; Wigen 2008a) and Ittatsoo North (Brolly and Pegg 1998:167), have identically high abundance values for herring (79% of NISP). The fish identified from Ma\u00E2\u0080\u0099acoah (Monks 2006), Tukw\u00E2\u0080\u0099aa and Ch\u00E2\u0080\u0099uumat\u00E2\u0080\u0099a (Monks 2011) in Toquaht territory as well as Shoemaker Bay at the head of Alberni Inlet (Calvert and Crockford 1982) are not directly comparable to the Huu7ii column assemblage as these excavations did not use column sample recovery methods. However, it is notable that these \u00C2\u00BC-inch assemblages contain only a negligible number of hake specimens, which further indicates the uniqueness of the Huu7ii assemblage. However, all sites appear to have significant frequencies of rockfish, greenling, and salmon and a host of other taxa that speak to the common utilization of these fish in other archaeological contexts. Grasping the variability will require considerably more analysis to fully synthesize and assess the spatial and temporal variability. Conclusions This study has explored the archaeological expression of vertebrate faunal remains, particularly fish, from the ancestral village site of Huu7ii. I analyzed over 58,000 vertebrate specimens containing over 32,000 identified remains from 12 column samples representing 168 depositional contexts spanning 5,000 years of human occupation. Fish bones were the most numerous and commonly encountered vertebrate elements, followed distantly by mammals and birds. Herring was the most numerous and consistently present fish species, \u00E2\u0080\u0093 202 \u00E2\u0080\u0093 followed distantly by anchovy, salmon, greenling and a host of other taxa. I analyzed these frequencies using multiple measures of abundance and argue they reflect cultural, social, and economic relations within the village. Collectively, these site specific and regional patterns indicate the vital importance of herring in indigenous precontact fisheries in Barkley Sound and have broader significance for interpreting the archaeological history of fishing on the Northwest Coast. The column sample assemblage from Huu7ii further confirms that small fish are grossly under-represented using conventional recovery techniques, which contributes to the under-recognized role of herring relative to the more well-known and disproportionately emphasized taxa such as salmon (cf., Coupland, et al. 2010; Monks 1987). A long-standing gap in understanding is the lack of column sample analysis, which is widely recognized to offer the most precise determination of the relative abundance of fish. A full-scale comparison of the temporal and spatial variability is ongoing and will yield more detailed insights into the regional character and intensity of ancient Nuu-chah-nulth fishing practices. \u00E2\u0080\u0093 203 \u00E2\u0080\u0093 Chapter 6. Indigenous Oral History & Settlement Archaeology in the Broken Group Islands, Western Vancouver Island This chapter examines both archaeological and Indigenous oral historical information specific to a precontact Nuu-chah-nulth local group territory on western Vancouver Island and asks whether these data reveal parallel sequences of human settlement. I draw on the past 2,500 years of archaeological records in the Broken Group Islands and the corpus of oral historical information specific to this area to help evaluate and inform the dynamic human settlement history in this region. The archaeological data have been compiled from previous surveys and excavations as well as from new field investigations and radiocarbon dating I coordinated in 2008, 2009, and 2011 while the oral historical and ethnohistorical observations have been translated and compiled by previous researchers over the past century. Indigenous Oral History and the Archaeological Record The historical accuracy of Indigenous accounts of history and its specific applicability to archaeological interpretation has faced numerous empirical and interpretive challenges. Archaeologists in North America and elsewhere have a long history of ignoring and/or regularly questioning the validity, specificity and historical accuracy of oral historical accounts, due in part to the pervasive critique that orally transmitted histories are qualitatively less robust than textually or archaeologically based accounts and/or are regularly contrived to suit practical political purposes and thus, are fundamentally \u00E2\u0080\u0098biased\u00E2\u0080\u0099 (Mason 2000; McGhee 2008). Oral histories are also considered to be subject to inherent degradation over time (Vansina 1985) a viewpoint often associated with the settler-colonial mentality that aboriginal culture has been similarly \u00E2\u0080\u0098degraded\u00E2\u0080\u0099 due to cultural \u00E2\u0080\u0098assimilation\u00E2\u0080\u0099. \u00E2\u0080\u0093 204 \u00E2\u0080\u0093 Such pervasive skepticism(s) helps explain the continued privileging of colonial historical accounts over Indigenous historical experiences, exemplified by the treatment of Indigenous oral history in courts of law (Miller 1992, 2011). Yet, as much as orally performed histories may indeed reflect a strategically deployed emphasis crafted to suit particular circumstances and audiences (Cruikshank 1997), the narrative content in these histories retains an enduring historical structure that fundamentally transcends, collapses, and contextualizes the contemporary moment and can extend far beyond the colonial experience. Archaeologists who seek to include Indigenous oral accounts in their interpretive practice are frequently charged with perpetuating a teleological (self-referenced) account of history (McGhee 2008). This skepticism similarly relates to a perception that oral historical accounts lack analytical rigor (Mason 2006) a critique somewhat akin to the \u00E2\u0080\u0098cultural historical\u00E2\u0080\u0099 approach dismissed by a generation of \u00E2\u0080\u0098processual\u00E2\u0080\u0099 archaeologists (Trigger 1991). However, one fundamental problem with this critique is that it effectively minimizes consideration of oral history as a legitimate source for archaeological insight and therefore further displaces the narration of Indigenous history from Indigenous voices (Atalay 2008). It also fails to acknowledge the pervasive influence that oral historical accounts have had on archaeological observations and in guiding interpretive processes and ethnographic and ethnoarchaeological analogues, particularly in North American anthropology and archaeology, which has been unmistakably shaped and influenced by the Indigenous societies they study (Darnell 2000; Valentine and Darnell 1999). Recent commentators contend that a productive aim in evaluating Indigenous oral narrative and historiography in the context of archaeological interpretation is not to seek to empirically \u00E2\u0080\u0098verify\u00E2\u0080\u0099 the culturally embedded forms of place-based oral historical knowledge \u00E2\u0080\u0093 205 \u00E2\u0080\u0093 (Crowell and Howell 2013; Cruikshank 2001; Green 2008; Martindale n.d.). Rather, it is to recognize that both Indigenous oral history and archaeological interpretation represent attempts to narrate and assign causality to human history, but that each perspective encompasses rich but incomplete sources of knowledge about the past, which can be made richer and stronger when examined in parallel (Martindale 2006). Such an approach avoids a dichotomous questioning of \u00E2\u0080\u0098Indigenous knowledge\u00E2\u0080\u0099 or its decontextualized incorporation into scientific and bureaucratic enterprises (Cruikshank 2005:256-257). Rather, it refocuses on how particular translations between oral historical and archaeological knowledge can advance and thus strengthen understanding of the human and environmental past. Translating Between Archaeological and Oral Historical Information A number of disciplines recognize Indigenous oral histories have a richness and sophistication that require detailed consideration of how they are performed, reflected upon, and persist over multiple generations (Basso 1996; Cruikshank 1997; Marsden 2001; Martindale 2006; Miller 2011). These factors do not necessarily detract from their utility in comparison to archaeological data. Rather, they require an anthropological engagement with historical and cultural setting and linguistic dimensions that strengthens understanding of how aspects of oral narratives can persist across generations in the absence of written texts. Such an engagement helps further refine and identify the embedded cross-linked event chronology in oral narrative and its links to geographically grounded place names and named actors. As stated by Fogelson (1989:141), oral historical events \u00E2\u0080\u009Cdo not generate the narrative, but events are selected to cohere to story lines, frameworks, or plots that result in intelligible narratives.\u00E2\u0080\u009D \u00E2\u0080\u0093 206 \u00E2\u0080\u0093 Archaeologists and Indigenous scholars concerned with understanding human history on the Northwest Coast have increasingly identified the considerable potential for examining parallels between Indigenous oral histories and archaeological histories. A number of recent studies identify parallel alignments between oral historical accounts of places and archaeological evidence that stretch into the precontact past (Atleo 2004; Crowell and Howell 2013; de Laguna 1960; Ludwin, et al. 2005; Marsden 2001; Martindale and Marsden 2003; Martindale 2006; McLaren 2003; McMillan 1999; McMillan and Hutchinson 2002; McMillan 2009). Archaeological researchers have further observed a widespread correspondence between geological events, particularly earthquakes and tsunamis and Indigenous oral histories (Bornhold, et al. 2007; Budhwa 2002; Losey 2000, 2002, 2005; Ludwin, et al. 2005; Ludwin, et al. 2007). Others have identified intergroup conflict and political alliances and amalgamation as a significant event-scale temporal marker regularly recounted in Indigenous oral histories (Angelbeck and McLay 2011; Huu-ay-aht First Nations 2000; Martindale and Marsden 2003; Swadesh 1948). Objectives Archaeological research that is attentive to Indigenous oral narrative addresses two important aspects of human history that are often poorly investigated in conventional archaeological research: 1) an Indigenous perspective and narration of the historical phenomena represented in archaeological data and 2) a focus on the intergenerational timescales spanning the lifetimes of individuals that fall between the broad millennial-scale of the longue dur\u00C3\u00A9e (Ames 1991; Smith 1992) and the \u00E2\u0080\u009Cevent based\u00E2\u0080\u009D scale of conventional history (Fogelson 1989). \u00E2\u0080\u0093 207 \u00E2\u0080\u0093 To date, Indigenous oral histories have not been broadly integrated with or evaluated alongside the conventional archaeological chronologies on the Northwest Coast (but see Martindale and Marsden 2003; McLaren 2003). Following the principle of relative sequential ordering (Martindale 2006), this chapter examines Indigenous oral histories, as well as the place names embedded within them, to evaluate Indigenous timelines of sequential and overlapping historical events alongside archaeological sequences of settlement. I specifically compare these distinct datasets to evaluate the ages of occupation between spatially and ethnographically affiliated settlements in close proximity to each other and within these large shell middens. I observe the oral historical sequences and the archaeological settlement chronology to show overlapping and complementary patterns that document the growth, expansion, and dynamically shifting residence patterns at multiple village sites over the past two and half millennia. I argue this comparison adds vital detail and an Indigenous perspective to archaeological settlement histories at an intergenerational scale and anthropologically enriches interpretation of the relationships between spatially associated archaeological sites. This research also addresses sources of archaeological settlement variability that have not been particularly well documented on the Northwest Coast (Cannon 2001; Moss 2012). Analysis of Archaeological and Oral Historical Sequences The analyses undertaken in this chapter compare and evaluate oral historical information alongside archaeological data at multiple spatial and temporal scales specific to the Broken Group Island archipelago. In the spatial domain, I use contact-era Indigenous territories and place names identified in oral historical narratives compiled by St. Claire (1991) to compare with archaeological site locations and site sizes compiled through \u00E2\u0080\u0093 208 \u00E2\u0080\u0093 previous survey efforts (Haggarty and Inglis 1985). I compare these datasets by creating separate geospatial layers in ARCGIS with place names, territorial boundaries and archaeological settlement locations. To calculate archaeological site size, I geo-rectified images from the original hand-drawn survey maps onto a high resolution DEM and basemap22 to trace the outline of shell bearing site deposits indicated by the appropriately scaled site map to generate a polygon estimate of shell-bearing surface area (m2) using the calculate area function in ARCGIS 10. On a temporal scale, I reviewed oral historical information specific to archaeological study sites for information specific to the sequence of occupation, growth and disruption of settlement to evaluate alongside chronological sequences of archaeological settlement. I focused this comparison on seven study sites within a local group territory on the outer Broken Group Islands (Figure 6.1) where I conducted fieldwork including radiocarbon dating, surface mapping and percussion coring. Historical Background Nuu-chah-nulth peoples were among the first Indigenous communities on the British Columbia coast to make contact with European explorers during the late 18th century. The trade and exchange that took place between Nuu-chah-nulth and visiting European explorer/traders initiated a geopolitically significant period of colonial expansion in the Pacific and the impact of this trade reverberated in communities throughout the coast (Harris 1997; McMillan 1999). This intensive period of trade sustained relationships between particular Nuu-chah-nulth chiefs and European and Eastern North American ship captains and merchants over a 30-year period but rapidly collapsed after sea otter pelts became scarce 22 Basemap generously provided by Ed Gregr (UBC Resource Management & Environmental Studies). \u00E2\u0080\u0093 209 \u00E2\u0080\u0093 and trading ships such as the Boston and the Tonquin were attacked and ransacked by Nuu-chah-nulth groups in the early 19th century (Clayton 2000:143; McMillan 1999:188). Exchanged European goods were widely and rapidly circulated throughout coastal communities and overland through an elaborate set of intermediaries that built on existing social relations and exchange patterns (Galois 2004; Harris 1997). For Nuu-chah-nulth, this period marked an initial process of colonial expansion which was greatly diminished throughout the early 1800s but was then sharply amplified 50 years later during the mid 1850s and 1860s when colonial presence and settlement rapidly expanded amidst disease and violent military attacks from colonial presence combined to reduce and displace Indigenous peoples from much of their former territories (Gough 1984; Harris 1997; Sellers 2013). Individual Nuu-chah-nulth Nations along the west coast differed in their participation in this period of trade and exchange (McMillan 1999). Some groups, particularly the Mowachaht at the village of Yuquot (i.e., Friendly Cove in \u00E2\u0080\u0098Nootka\u00E2\u0080\u0099 Sound) sustained exclusive control over favored anchorages and extracted tribute from or excluded others from direct access. Other Nuu-chah-nulth communities, whether through geographic circumstance or long-standing political conflict and/or alliances, participated indirectly in this exchange, often at a material disadvantage, particularly when guns were involved (Harris 1997). Such is inferred to have been the case in Barkley Sound and the Broken Group Islands where the Tla-o-quiaht chief Wickaninish attempted to exert and maintain control over the distribution of sea otter pelts and direct access to trading ships but this control was tenuous and limited at best (Clayton 2000:143-148; McMillan 1999:188). Nuu-chah-nulth First Nations on western Vancouver Island have persisted through significant political and cultural changes since the early maritime fur trade (ca. AD 1778- \u00E2\u0080\u0093 210 \u00E2\u0080\u0093 1803). In the intervening decades between the maritime fur trade and the 1860s numerous Nuu-chah-nulth groups underwent a series of amalgamations through confederation or as a result of competitive conflict (Arima 1983; Atleo 2011; Clayton 2000; Drucker 1951; Golla 1987; Huu-ay-aht First Nations 2000; Inglis and Haggarty 2000; Kenyon 1980; Marshall 1993; McMillan 1999; St. Claire 1991). After the 1850s, groups were further subjected to devastating demographic changes due to introduced diseases. Thus, while 14 individual Nations are recognized today, this number was far greater during the early contact era and prior. However, this rich mosaic of individual autonomous groups has not been adequately recognized in the broader archaeological literature (McMillan 1999; McMillan, et al. 2008; McMillan 2009). For instance, two widely cited archaeological syntheses note that a limited number of individual groups controlled large territories and exhibited a pattern of seasonal movement between the \u00E2\u0080\u0098inside\u00E2\u0080\u0099 and \u00E2\u0080\u0098outside\u00E2\u0080\u0099 coasts (Dewhirst 1980; Mitchell 1990b). This synthetic impression is based largely on the interpretation of Nuu-chah-nulth territorial structure specific to central and Northern Vancouver Island influentially put forth by Philip Drucker (1951) who detailed patterns of settlement organization that neglected to take into account many of the demographic and political changes during the early contact era (McMillan 2009). This influential viewpoint reflects a lack of engagement with an abundance of primary data pertaining to the early contact era: those recounted and detailed in oral historical accounts. However, many archaeologists reviewing ethnographic literature to provide insight into the archaeological past often project a less complicated and politically and territorially altered circumstance dating from the late 19th century into the more distant past (cf. Grier 2007). Aside from a problematic circumstance for archaeological interpretation, this \u00E2\u0080\u0093 211 \u00E2\u0080\u0093 impression is not an accurate reflection of the rich historical content provided in oral accounts, which often speak of numerous titleholders and groups whose significance may appear diminished relative to the circumstances of the late 19th century and early 20th century, which are nevertheless, among those most relevant to archaeological interpretation of pre-contact history. For instance, on western Vancouver Island, Indigenous oral histories and place names describe a large number of Nuu-chah-nulth communities that occupied spatially distinct territories during the contact era (Golla 1987; Inglis and Haggarty 1986, 2000; Marshall 2000a, b; McMillan and St. Claire 2005; St. Claire 1991). Over the late 18th and 19th centuries, these communities subsequently amalgamated into larger groups as they endured large-scale population reductions from introduced diseases alongside dramatic increases in the colonial presence in the 1850s and 1860s. Archaeological support for larger populations and more numerous but smaller and circumscribed group territories during the precontact era has been observed across western Vancouver Island such as in Hesquiaht Harbour (Calvert 1980; Haggarty 1982) and Nootka Sound (Inglis and Haggarty 2000; Marshall 1993), and particularly in Barkley Sound and the Broken Group Islands where sizable clusters of large shell middens have been inferred to indicate much larger precontact populations (Haggarty and Inglis 1985; McMillan and St. Claire 1982). The archaeological importance of these Indigenous oral histories is that they detail the processes of post-contact political change and amalgamation and show how they unfolded uniquely for individual Nuu-chah-nulth groups at named places that have extensive archaeological deposits. The Broken Group Study Area The Broken Group Islands on the southwest coast of Vancouver Island (Figure 6.1) is an area of the Northwest Coast where there is both a detailed set of inter-generational \u00E2\u0080\u0093 212 \u00E2\u0080\u0093 community knowledge described in Indigenous oral histories and encoded in Indigenous place names, as well as a thoroughly surveyed archaeological record (Haggarty and Inglis 1985) including two large-scale excavation projects at the major village site of Ts\u00E2\u0080\u0099ishaa (McMillan and St. Claire 2005) and ongoing research at Hiikwis (MacLean 2012; Sellers 2013)23. I have conducted further research at a cluster of sites in the southern quadrant of the archipelago (McKechnie 2009, 2010b) (Figure 6.1), which is the pre-contact territory of the Maktl7ii7ath (a Nuu-chah-nulth polity) who amalgamated with Ts\u00E2\u0080\u0099ishaa7ath, the present-day Tseshaht First Nation, in the mid-to-late 18th century (McMillan and St. Claire 2005). 23 This study area is further bounded by extensive research conducted in the adjacent but culturally distinct territories to the south (Mackie and Williamson 2003, n.d.) and North (Brolly and Pegg 1998; McMillan and St. Claire 1992, 1994, 1996; McMillan 1999; Monks 2006, 2011) \u00E2\u0080\u0093 213 \u00E2\u0080\u0093 Figure 6.1 Territorially bounded local group areas in the Broken Group Islands during the 18th century as described in oral historical accounts synthesized by St. Claire (1991) and McMillan and St. Claire (2005). Map modified with permission of St. Claire. \u00E2\u0080\u0093 214 \u00E2\u0080\u0093 Oral Histories A large corpus of oral histories have been transcribed from Nuu-chah-nulth elders in the early 20th century by Edward Sapir and his Tseshaht collaborators Alex Thomas and Frank Williams (Figure 6.2) and subsequently edited and published by Morris Swadesh, and later by Eugene Arima, Denis St. Claire, Katherine Robinson, and others (Arima, et al. 1991; Arima, et al. 2000; Arima, et al. 2004, 2007; Arima, et al. 2009; Sapir and Swadesh 1939, 1955; Sapir n.d.-b, a). These narrative histories were transcribed in the original Barkley Sound dialect of the Nuu-chah-nulth language through a practical orthography collaboratively developed between Thomas, Williams, and Sapir and subsequently translated into English. This occurred in the first few decades of the 20th century when the majority of the Tseshaht First Nation concentrated year-round settlement in Port Alberni on the lower Somass River. This was also the location of the colonial residential school, which mandated attendance for all First Nation children and had a dramatic intergenerational cultural and social impact on Nuu-chah-nulth communities in Barkley Sound and elsewhere in British Columbia. Alexander Thomas and Frank Williams had both been pupils at the residential school and later worked with Sapir, who was then associated with the Anthropological Division of the Geological Survey of Canada. Similar to Franz Boas and George Hunt, after a period of initial fieldwork, Sapir regularly corresponded with Thomas and Williams who worked to further compile oral histories as dictated by elders for a period of over 30 years (Darnell 1990). These narrative histories primarily concern events in the Broken Group Islands and in outer Barkley Sound before and during the contact-era fur trade (ca. 1778-1805) and the 50-year period following its collapse when there were particularly striking changes in the political landscape of Barkley Sound albeit in the absence of extensive European colonial presence (ca. 1805-1860). Rather than detailing the periodic contacts with \u00E2\u0080\u0093 215 \u00E2\u0080\u0093 explorers and colonists, these narratives are overwhelmingly concerned with conflicts, alliances, and ceremonies between Nuu-chah-nulth individuals and groups. This voluminous literature provides a wealth of historical details relevant to archaeological interpretation and have been considered in depth by Alan McMillan and Denis St. Claire in the course of decades of ethnohistorical and archaeological research. Figure 6.2. Alex Thomas (left), Frank Williams and his daughter (centre) in Port Alberni in 1914. Edward Sapir (right) pictured in New Haven in 1934 (Images used with permission from the Canadian Museum of Civilization). CMC Historical Photo Archive numbers 26545, 26574, and 2006-00012 respectively. Indigenous Place Names Within the Broken Group archipelago and Sechart Channel (Figure 6.1), there are 134 documented Tseshaht place names (Figure 6.3). The majority of these (73%) were transcribed and translated during the early 20th century. An additional 36 names were added through the research of Denis St. Claire (1991). Thirty-five of the 134 place names (26%) were not translated, leaving open the possibility for future linguistic analyses. The density of Indigenous place names in the Broken Group and Sechart Channel study area is estimated to be 1.3 names per square kilometer (ca. 100 km2). This greatly \u00E2\u0080\u0093 216 \u00E2\u0080\u0093 exceeds (by orders of magnitude) the density of Indigenous place names recorded in other studies elsewhere in North America (Hunn 1994) and complements other placename syntheses on the Northwest Coast (e.g., Boas 1934; Hilbert, et al. 2001; Richardson and Galloway 2011; Thornton 2012). This high density may be a reflection of the relatively small and tightly bounded study area but it is also an indication of the extensive and sustained research effort to systematize place names, thus making this dataset a significant research contribution as well as a valuable cultural archive. It is also a reflection of the detailed familiarity that generations of Tseshaht had with this landscape, its features, and its history as well as an indication of a high precontact settlement density indicated by archaeological data described below. \u00E2\u0080\u0093 217 \u00E2\u0080\u0093 Figure 6.3. Tseshaht place names as detailed in oral historical accounts shared by community knowledge holders transcribed and translated by Sapir, Thomas, and Williams and later synthesized by St. Claire (1991) who affirmed each translation through interviews with Tseshaht elders in the 1970s and 1980s. Archaeological Site Locations The archaeological landscape of the Broken Group archipelago contains a minimum of 73 shell midden sites with a very broad geographic distribution throughout the archipelago (Figure 6.4). There are also a variety of intertidal sites including fish traps, garden features, and canoe runs. The ubiquitous distribution of shell middens and particularly their range of sizes and proximity to one another demonstrate a rich diversity of human occupation and use has occurred throughout the archipelago. The site sizes range from over 14,000 m2 to less than 5 m2. Collectively there are more than 114,000 m2 of recorded shell midden deposits spread throughout the archipelago. It is significant that all of the 15 largest shell midden sites \u00E2\u0080\u0093 218 \u00E2\u0080\u0093 have surficial evidence of houses, terraces, and/or constructed ridge landforms and all but one are directly associated with Indigenous place names described in oral histories and were noted to be villages occupied prior to and during the contact era (Table 6.1). Figure 6.4 Recorded archaeological sites in the Broken Group Islands and Sechart Channel. The size of each circle represents the horizontal surface area of shell middens as determined by geo-rectifying original site maps in ARCGIS. Also shown are intertidal rock alignments interpreted to be fishtraps by Haggarty and Inglis (1985) (lower left) and sites with canoe runs representing canoe width channels cleared of rocks (right). \u00E2\u0080\u0093 219 \u00E2\u0080\u0093 Table 6.1. The 15 largest shell midden sites in the Broken Group Islands and their associated place names. Bolded sites have been subject to fieldwork as part of this dissertation or previous research. No. Place name Location Site Parks # & Borden # Midden Areaa (m2) House Platform House Depression Back-ridge Radiocarbon data Zooarch data 1 Ts\u00E2\u0080\u0099ishaa, Benson Island 204T&205T DfSi16&17 14,450 X X X X X 2 Maktl7ii Wouwer Island 206T&215T DfSi19& 30 10,040 5 X X X 3 Huumuuwaa Village Island 304T DfSh-4 9,110 15 X X X X X N. Turret Island No place name 198T DfSi-10 6,660 X X X 4 Tl'ihuuw'a Nettle Island 305T DfSh-5 6,630 X X X X X 5 Aalhachmakis Dodd Island 72T DfSh-19 6,150 X X X X X 6 Muk'waa7a Turret Island 195T DfSi-7 4,690 X X X 7 Huts'atswilh Dicebox Is. 83T &129T DfSh31&79 4,520 32 X X X X 8 Kakmakimilh Keith Island 306T DfSh-17 4,840 X X X X X 9 Aasimilh Hand Island 187T DfSh-101 4,030 X ? ? X 10 T\u00E2\u0080\u0099um\u00E2\u0080\u0099aktli Gibraltar Island 80T DfSh-47 3,560 X X X 11 Tlakiyamilh Chalk Island 73T DfSh-20 3,490 X X X 12 Hats'aat'iml Turtle Island 156T DfSh-111 3,300 X ? X 13 Shiwitis Gilbert Island 82T DfSh-29 2,470 X ? ? X X 14 Tl'atl'inkuuwis Effingham Bay DfSh-28 2,060 NA NA NA a \u00E2\u0080\u0093 Shell Midden surface area calculated using the polygon outline of a geo-rectified original site map in ARCGIS 10. Archaeological Site Size Distribution Within the archipelago, the four bounded autonomous local group territories described in oral histories shown in Figure 6.1 occur in physically distinct clusters of islands. A spatial comparison of the number of archaeological sites and the distribution of site sizes within these geographically distinct quadrants shows similar distributional patterning (Figure 6.5), with each quadrant of the archipelago having a similar number of sites and similar numbers of \u00E2\u0080\u0098large\u00E2\u0080\u0099 sites (> 2000 m2) in each cluster. This demonstrates the occurrence of repeating subunits of archaeological site distributions and site size distributions. Given the \u00E2\u0080\u0093 220 \u00E2\u0080\u0093 assumptions that 1) larger sites reflect larger numbers of site occupants and 2) that a range of site sizes reflects a diversity of use and that 3) the sea level history chronologically constrains many of these shoreline proximal sites to within the past 2,000 years, this indicates the contemporaneous use of the archipelago by multiple communities and thus is consistent with the oral historical accounts describing the presence of multiple autonomous groups in this small region (Figure 6.1). Figure 6.5. The size and number of shell midden sites in different quadrants of the Broken Group Islands. Data compiled from scaling and georectifying original site maps in ARCGIS 10 and tracing shell midden site boundaries. \u00E2\u0080\u0093 221 \u00E2\u0080\u0093 The Largest Sites It is additionally significant that of the 73 shell midden sites in the archipelago, the two largest sites are also the oral historically named \u00E2\u0080\u0098origin\u00E2\u0080\u0099 locations for the Tseshaht and Maktl7ii7ath local groups who shared mutual territorial boundaries in the outer Broken Group archipelago. The Indigenous names for these political groups literally translate as: \u00E2\u0080\u0098the people of Ts\u00E2\u0080\u0099ishaa\u00E2\u0080\u0099 and \u00E2\u0080\u0098the people of Maktl7ii.\u00E2\u0080\u0099 The archaeological observation that the two largest sites represent politically foundational groups and are associated with numerous smaller sites in close proximity further supports the oral historical descriptions that these sites were villages occupied year-round and not simply used as seasonally migratory destinations (see also: McMillan and St. Claire 2005; McMillan, et al. 2008). This also indicates a potential relationship between site size and site antiquity of occupation, which is supported by the 5000 year-long record of occupation at the village of Ts\u00E2\u0080\u0099ishaa and the 2,500 year long record at Maktl7ii described in the sections below as well as at the similarly named origin village of Huu7ii nearby in southeastern Barkley Sound (see also Chapter 5 and McMillan and St. Claire 2012). Settlement History of Ts\u00E2\u0080\u0099ishaa, The Ts\u00E2\u0080\u0099ishaa7ath Origin Village As mentioned, the archaeological site of Ts\u00E2\u0080\u0099ishaa is the oral historically named \u00E2\u0080\u0098origin place\u00E2\u0080\u0099 for the Ts\u00E2\u0080\u0099ishaa7ath, the present day Tseshaht First Nation, and has an archaeological settlement history that spans the past 5,000 years (McMillan and St. Claire 2005). The documented occupation from 5,000-3,000 BP is associated with higher mid-Holocene sea levels (McMillan 2003), and is small in comparison to the much more extensive late-Holocene midden deposits. The lower elevation late-Holocene (ca. 2,000-300 BP) deposits contain house platforms and midden terraces. A host of radiocarbon dates \u00E2\u0080\u0093 222 \u00E2\u0080\u0093 provides strong evidence for a village-sized settlement since 1,800 BP (McKechnie 2007a:212; McMillan, et al. 2008:218-219). Oral historical accounts provided by Sayach\u00E2\u0080\u0099apis to Sapir describe the first Ts\u00E2\u0080\u0099ishaa7ath woman and man being \u00E2\u0080\u009Ccreated\u00E2\u0080\u009D at Ts\u00E2\u0080\u0099ishaa by the \u00E2\u0080\u0098Day Chief,\u00E2\u0080\u0099 a transformer/creator who also created a river and a channel for Tseshaht but then \u00E2\u0080\u009Cscattered the river and the channel everywhere. That is why the islands are scattered about now\u00E2\u0080\u009D (Sapir and Swadesh 1955:52). Thereafter, the oral narratives describe that the population of Ts\u00E2\u0080\u0099ishaa expanded to the point where \u00E2\u0080\u009Cthe tribe became numerous, reaching to the other end of the village\u00E2\u0080\u009D (Sapir and Swadesh 1955:52). Additional recounted details about the village include the locations of four lineage households (ushtakimilh), their political rankings relative to one another, as well as the architectural details and descriptions of the figures painted on the housefront of the highest ranked ushtakimilh. The lineage-based households and the individuals who resided in them held various rights to select resources (tuupati) within the Tseshaht local group territory (hahuulhi) owned and managed by the head chief (Taayii Hawilth). The relative order in which the ushtakimilh were ranked reflects the sequence with which they were established; a ranking that was codified in the potlatch seating orders mentioned and detailed in several oral accounts (McMillan and St. Claire 2005:9-10). For instance, the lowest ranking ushtakimilh was located on the southern margins of the village (Figure 6.6) and various accounts mention this was the last to be established (McMillan and St. Claire 2005:9). This area was initially occupied by low-ranking members of the other ushtakimilh but acquired the status of an ushtakimilh after a high-ranking outsider mysteriously arrived after a flood event and was granted permission to establish a \u00E2\u0080\u0093 223 \u00E2\u0080\u0093 lineage household that incorporated the low ranking individuals (Inglis and Haggarty 1986:126; McMillan and St. Claire 2005:9; St. Claire 1991). As part of the Tseshaht Archaeological project, all four of these named ushtakimilh locations were subjected to archaeological excavation and radiocarbon dating down to basal beach sediments (McMillan, et al. 2008:218-219; McMillan 2009:628). Remarkably, this research demonstrated that the oral historically named sequence of village growth parallels the sequence of growth documented archaeologically (Figure 6.6). That is, archaeological excavations in these four named areas follow the oral historically recounted sequence of village expansion whereby the first-ranked and \u00E2\u0080\u0098oldest\u00E2\u0080\u0099 ushtakimilh contains the correspondingly oldest radiocarbon dated deposit (ca. 1,800 BP or AD 250), the second-ranked and second-oldest ushtakimilh contains the second oldest radiocarbon date and so on for the four identified areas. This parallel sequence indicates that this particular oral history, specific to multiple areas within one village site, is archaeologically supported and spans a period of at least 1,800 years. This complementary alignment of archaeological and Indigenous oral historical sequences indicates a chronological depth to the oral historical record that adds robusticity and detail to our understanding of human settlement history at this particular village site. These and related issues are discussed in depth below in the neighboring local group territory of the Maktl7ii7ath where I move beyond an individual site to demonstrate how the oral historical record and archaeological investigations can complement each other within the broader traditional territory of a Nuu-chah-nulth local group. \u00E2\u0080\u0093 224 \u00E2\u0080\u0093 Figure 6.6. Site map of the village of Ts'ishaa (204 and 205T) showing oral historically named locations of household lineages (ushtakimilh) and their numbered sequence of occupation. Horizontal extent of midden deposits and the Tseshaht Archaeological project excavation areas are shown following (McMillan and St. Claire 2005). Inset on the lower right depicts the calibrated ages of the four oldest radiocarbon dates obtained from separate site areas indicating the sequential occupation that accords with oral historical accounts. Dates are presented in McKechnie (2007a:215). Oral Historical Accounts of the Maktl7ii7ath Local Group Territory The Maktl7ii7ath local group territory is the focus of the archaeological fieldwork described in the remainder of this chapter. Multiple oral historical accounts and ethnographic syntheses (Golla 1987; Inglis and Haggarty 1986; Sapir 1910-1914; n.d.-a:34; St. Claire 1991; William 2009) identify the Maktl7ii7ath as a formerly politically autonomous community whose territory encompassed the southwestern portion of the Broken Group Island archipelago (Figure 6.1 and Figure 6.7) during and prior to European contact. Like other \u00E2\u0080\u0098local groups\u00E2\u0080\u0099 on western Vancouver Island (e.g., Arima 1983; Curtis 1916; Drucker 1951), the Maktl7ii7ath are named for the village location from which they took their name: \u00E2\u0080\u0093 225 \u00E2\u0080\u0093 Maktl7ii, which translates as \u00E2\u0080\u0098higher than others\u00E2\u0080\u0099 (St. Claire 1991:143) and is located on southwest Wouwer Island (Figure 6.7). This name refers to the high landform on which houses were located as well as the social standing of the founding and highest ranked lineage in the village. Across Imperial Eagle Channel to the southeast, the Maktl7ii7ath maintained a territorial boundary with the Kiix7in7ath and the Huu7ii7ath in the Deer Group islands and adjacent shoreline. Similarly to Ts\u00E2\u0080\u0099ishaa7ath, these groups later amalgamated to become the Huu7ii7ath, the contemporary Huu-ay-aht First Nations (Huu-ay-aht First Nations 2000; McMillan and St. Claire 2012; Nookmiis 2009). Figure 6.7. Overview of the Maktl7ii7ath study area with individual study sites (view looking west). Photo courtesy of the Royal BC Museum (West Coast Archaeological Project Files). Sapir (n.d.-a:34) describes the Maktl7ii7ath as \u00E2\u0080\u009Cone tribe with four bands\u00E2\u0080\u009D which refers to four named ushtakimilh locations within the village of Maktl7ii (Table 6.2). These households, their leaders and by extension, members, would have held rights (tupaati) to occupy and use resources associated with this territory with the permission of the Taayii Hawilth (head chief). Like at Ts\u00E2\u0080\u0099ishaa, these ushtakimilh have an internal ranking based on Effingham Bay OFFSHORE Coaster Channel _ _ \u00E2\u0080\u0093 226 \u00E2\u0080\u0093 their sequence of establishment (St. Claire 1991:38) with the Maktl7ii7ath holding the highest rank. The lowest ranking group of the Maktl7ii7ath, the Ts\u00E2\u0080\u0099ap\u00E2\u0080\u0099is7ath, was founded late after the daughter of the high ranking chief had a child with a low ranking man. Eventually when the Ts\u00E2\u0080\u0099ap\u00E2\u0080\u0099is7ath held a potlatch at which the flukes of 10 whales were distributed, the group was given a new name, the Nach\u00E2\u0080\u0099imuuwas7ath, and accorded a higher level of status within the Maktl7ii7ath (St. Claire 1991:38). Table 6.2. Lineage household names (ushtakimilh) names at the village of Maktl7ii on Wouwer Island (206T) presented in St. Claire (1991:38-41 and 142-143). Place name Translation Maktl7ii7ath 'higher than others' Also the name for the Entire Village Ts\u00E2\u0080\u0099ap\u00E2\u0080\u0099is7ath later becoming the Nach\u00E2\u0080\u0099imuuwas7ath NA \u00E2\u0080\u009Cpeople who have whale fins all around\u00E2\u0080\u009D (St. Claire 1991:38) Ust\u00E2\u0080\u0099is7ath 'down below (not high up on the beach)' T\u00E2\u0080\u0099imik\u00E2\u0080\u0099ak\u00E2\u0080\u0099is7ath 'people of gooseberry bushes on beach' Oral Historical Events in Relation to Settlement Oral historical accounts of Maktl7ii7ath territory place primacy on the village of Maktl7ii, the namesake of the highest ranking household in the village and take the name of that place. However, oral historical accounts additionally mention two other village settlements in Maktl7ii7ath territory, including the village of Huumuuwaa (Inglis and Haggarty 1986:127; William 2009) the village and fortress of Huts\u00E2\u0080\u0099atswilh on Dicebox Island. Of particular archaeological significance, the large village of Huumuuwaa is identified as \u00E2\u0080\u009Cnot one of the old villages\u00E2\u0080\u009D but rather as one that \u00E2\u0080\u009Cbelongs to Maktl7ii7ath\u00E2\u0080\u009D (William 2009:370). Similarly, Huts\u00E2\u0080\u0099atswilh is identified as a village that was \u00E2\u0080\u009Cinhabited by the Huts\u00E2\u0080\u0099atswilhath, a sept of the Maktl7ii7ath\u00E2\u0080\u009D (ibid.). The reference to \u00E2\u0080\u0098sept\u00E2\u0080\u0099 refers to a lineage group (ushtakimilh) of the Maktl7ii7ath. These accounts provide an indication that \u00E2\u0080\u0093 227 \u00E2\u0080\u0093 villages of Huts\u00E2\u0080\u0099atswilh and Huumuuwaa were established after the village of Maktl7ii and provide a sequential model to evaluate in parallel with the archaeological dataset. The Maktl7ii7ath local group was the first independent local group to politically amalgamate with the Ts\u00E2\u0080\u0099ishaa7ath just prior to or during the early contact-era (ca. late 18th century) (McMillan and St. Claire 2005:16). This amalgamation was one in a sequence of several for the Ts\u00E2\u0080\u0099ishaa7ath, the politically autonomous community that represent the present-day Tseshaht First Nation. Based on several oral historical records, this merger occurred \u00E2\u0080\u009Cprobably in the latter years of the eighteenth century\u00E2\u0080\u009D after a devastating battle with the Hach\u00E2\u0080\u0099aa7ath (McMillan and St. Claire 2005:17). The Hach\u00E2\u0080\u0099aa7ath were also a neighboring local group who occupied the northeastern portion of the archipelago at contact (Figure 6.1) and nearly devastated the Maktl7ii7ath in a particularly formative conflict, which reduced the population to such an extent \u00E2\u0080\u009Cthat only 15 adult men remained\u00E2\u0080\u009D (McMillan and St. Claire 2005:15). The survivors choose to merge politically with the Ts\u00E2\u0080\u0099ishaa7ath but the head chief retained rights to host potlatches and the drift whale rights in their former territory (ibid.). This merger greatly expanded the Ts\u00E2\u0080\u0099ishaa7ath territory and critically included the strategically defensive fortress of Huts\u00E2\u0080\u0099atswilh discussed at length later in this chapter. Prior to and during the early years of contact with European trading ships, the Maktl7ii7ath local group maintained a defined territory encompassing the southwestern portion of the archipelago and shared a mutually recognized boundary with the Ts\u00E2\u0080\u0099ishaa7ath local group (Figure 6.1). The shared boundary was marked by a line of brush cleared across a small island that was visible from the water (iitsmakiis, St Claire 1991:142). The northern boundary of the Maktl7ii7ath was Coaster Channel encompassing \u00E2\u0080\u0098Village Reef,\u00E2\u0080\u0099 the \u00E2\u0080\u0098Faber \u00E2\u0080\u0093 228 \u00E2\u0080\u0093 Islets\u00E2\u0080\u0099 and Wiebe Island (Figure 6.1). This latter portion of Maktl7ii7ath territory including the village of Huumuuwaa on Effingham [Village] Island (Figure 6.7) was violently obtained by Hach\u00E2\u0080\u0099aa7ath local group mentioned above. This conflict was precipitated by Hach\u00E2\u0080\u0099aa7ath \u00E2\u0080\u0098striking\u00E2\u0080\u0099 \u00E2\u0080\u0093 his7ukwt \u00E2\u0080\u0093 to obtain territory in the late 18th century (St. Claire 1991:28-31). Soon thereafter however, Hach\u00E2\u0080\u0099aa7ath were defeated as a fighting force and extinguished as a politically autonomous group by a regional alliance of Nuu-chah-nulth local groups in Barkley Sound and elsewhere (ibid.). Archaeological Case Studies in Maktl7ii7ath Local Group Territory There are 20 recorded archaeological shell midden sites in the area bounded by Maktl7ii7ath territory, including five large shell middens with surface areas greater than 2000 m2 (Figure 6.5). I investigated seven sites in this study area (Figure 6.7) and obtained a total of 56 radiocarbon dates (Table 6.3)24. Combined with detailed surface mapping and percussion coring, these radiocarbon chronologies provide a basis for inferring the long-term settlement history within this territory. Over 40 Indigenous place names are known for this portion of the Broken Group archipelago and each archaeological site is directly or indirectly associated with one or more place name(s) (Figure 6.3). Several oral histories pertaining to events in the territory offer a chance to further evaluate oral historical records and archaeological history prior to the mid-20th century. Below, I describe each of the seven 24 To translate radiocarbon dates into calibrated age estimates, I use the OXCAL calibration program (Bronk Ramsey and Lee 2013) which both calibrates individual dates and provides a way to graphically depict the probability of the calibrated age range for a number of dates. I radiocarbon dated both terrestrial charcoal and marine shellfish, the latter requiring an estimate of the marine reservoir effect. I estimated the temporally variable local marine reservoir using shell-charcoal pairs (Fedje and Southon 2003; McMillan and St. Claire 2005; McKechnie et al. n.d.) and calculated Delta-R values (shown in Table 6.3) following personal communications with Paula Reimer & Maarten Blaauw in July 2013. \u00E2\u0080\u0093 229 \u00E2\u0080\u0093 study sites, focusing first on the archaeological deposits and settlement chronology for the village site of Maktl7ii. Maktl7ii \u00E2\u0080\u0093 Wouwer Island As mentioned, Maktl7ii is a large archaeological shell midden (206T, DfSi-19) that is the oral historically named origin village for the Maktl7ii7ath local group (i.e., \u00E2\u0080\u0098the people of\u00E2\u0080\u0099 Maktl7ii). The site is situated in a sheltered cove on the highly exposed southwestern fringe of Wouwer Island facing south towards the Pacific Ocean (Figure 6.7) and has extensive archaeological deposits consisting of five closely spaced shell middens separated by elevated bedrock promontories. These deposits lie shoreward of several pocket beaches that are sheltered from ocean swell by fringing reefs to the west and south (Figure 6.8). A detailed topographic map of the site created over 10 days with a total station provides spatial context for assessing the shell bearing components, which have a combined 2-D surface area of approximately 10,000 m2 (Figure 6.8). This site includes four very deep \u00E2\u0080\u0098backridges\u00E2\u0080\u0099 (3-5.5 m), that are constructed midden ridges behind house platforms or gradually sloping terraces suitable for house locations. These ridge features are characteristically observed at other village-sized sites throughout Barkley Sound as well as across western Vancouver Island (Mackie and Williamson 2003; Marshall 1993, 2006; McMillan 1999). \u00E2\u0080\u0093 230 \u00E2\u0080\u0093 Figure 6.8. Oblique perspective view of Maktl7ii looking north (with inset satellite image). Black dots represent core-tests locations in cultural deposits and brown- blue transition represents sea level (i.e., barnacle line). Scale in arbitrary metres shown on the bottom of the image. Angular lines on periphery of image represent lack of topographic data. A series of 42 percussion cores, 10 auger samples, and 10 AMS dates, obtained from various locations at Maktl7ii, provide a basis for interpreting the archaeological settlement history. The site can be interpreted to be a village as the three oldest radiocarbon dates are from the base of three separate deep midden ridges situated in the more sheltered northeastern portion of the site (the inner cove). These three dates statistically overlap in calibrated age (ca. 2340-2600 cal BP) and indicate the growth of these separate midden ridges began simultaneously around 2,400 years ago (Figure 6.9). Given the depth and extents of these massive midden features, this contemporaneous deposition represents strong evidence for the establishment of the site as a village location. Moreover, terminal dates from the midpoints and top of two these separate ridge deposits indicates they grew rapidly over the next 1,100 years but stopped accumulating between 1,400 and 1,300 years ago. North Inner Cove Outer Cove \u00E2\u0080\u0093 231 \u00E2\u0080\u0093 Figure 6.9. Calibrated Radiocarbon dates from Maktl7ii (site 206T) at Wouwer Island showing the calibrated age range and probability for individual dates. Dates on marine shell shown in turquoise and dates on charcoal are in grey calibrated on the atmospheric curve. During this same temporal interval, extensive midden deposits began to accumulate in a separate \u00E2\u0080\u0098outer\u00E2\u0080\u0099 cove to the southwest representing a shift in community residence. This area of the site has surficial evidence of historic materials and sub-surface evidence for historic material to a depth of at least 1.8 m in one of four well-defined house platforms on a sharply sloping but terraced shellmidden (Figure 6.10). Here, dates from 3.5m beneath a house platform and at the base of a 5.5m deep backridge indicate rapid midden growth between approximately 1500 and 500 BP and continuing at a slower rate into the historic era (Figure 6.9). This stands in contrast to the older and more extensive site areas discussed above which contain much more heavily vegetated terrain with dense salal that does not contain historic materials on the surface or in the auger samples. \u00E2\u0080\u0093 232 \u00E2\u0080\u0093 Figure 6.10. Terraced midden with house platforms and a 5.5 m deep backridge at site 206T, Outer cove, Wouwer Island. A fourth portion of the site contains a date of approximately 500 BP. This represents a mid-depth estimate of a midden deposit at the head of a cove with ready trail access across to the north side of the island. This deposit lacks ridges or other features and is interpreted to reflect periodic resource use. It provides a constraining age on faunal remains recovered from an auger in this location (see Chapter 7). Huumuuwaa \u00E2\u0080\u0093 Village Island The large shell midden and historic village of Huumuuwaa (306T, DfSh-4) is located along a semi-exposed rocky beach facing Imperial Eagle Channel on eastern Effingham [Village] Island (Figure 6.7). The village area was designated a reserve in 1882 and remains one of only three Tseshaht Reserves in the Broken Group Islands. The shell midden deposits at this site span approximately 275 metres of vegetated shoreline with an estimated surface \u00E2\u0080\u0093 233 \u00E2\u0080\u0093 area of 9,100 m2 (Table 6.1). The cultural topography of the site is remarkably intact with well-defined evidence for 14 rectangular house platforms with mean surface areas of 95.8 m2 (1,060 ft2) and a combined housefloor area of 1379 m2. A number of the house platforms additionally contain the partially preserved remains of 14 house posts and 6 house beams and provide further context inferring the structural dimensions of houses. These house platforms front the shoreward portion of a 90 m long \u00E2\u0080\u0098back-midden ridge\u00E2\u0080\u0099 that contains at least 7.08 m (23 ft) of cultural deposit at its greatest height (Figure 6.11). Shallower shell midden deposits occur at the periphery of this portion of the site and are bisected by two small drainages diverted by the massive shell midden landforms present at the middle of the site. A much more heavily vegetated site area to the south contains less distinct cultural topography overlain by a comparatively thick layer of humic material, and is accessed by canoe via a small south-facing pocket beach. Percussion core tests were conducted in 12 different locations alongside 4 auger tests from multiple site areas (Figure 6.11). Seven radiocarbon dates were obtained from a variety of localities within Huumuuwaa and constrain the age of this large village site to within the past 1,800 years (Figure 6.12). As at Maktl7ii, the three earliest dates come from three separate areas at the base of a very deep 90 m long \u00E2\u0080\u0098backridge\u00E2\u0080\u0099 of shell midden that parallels the shore and represents the inland extent of the massive shell-bearing components of the site. These dates indicate the middle of this massive ridge began to accumulate around 1,800 years ago, expanding 40m to the south over the following century and reaching its northern extent by approximately 1200 BP (Figure 6.12). Another date recovered from the base of a house platform shoreward of the northern extent of the ridge falls a century later (ca. 1070- \u00E2\u0080\u0093 234 \u00E2\u0080\u0093 970 cal BP). As at Maktl7ii, these dates represent good evidence for the initial establishment this site as a village-sized community between approximately 1,800 and 1,200 BP. Figure 6.11. Oblique perspective view of Huumuuwaa showing the numerous house platforms and the large backridge feature. Small circles represent core test locations. Scale in arbitrary metres shown on the margin of the image. Figure 6.12. Calibrated radiocarbon dates from Huumuuwaa on Effingham Island (site 304T) showing the calibrated age range and probability for individual dates. All dates calibrated on the atmospheric curve. \u00E2\u0080\u0093 235 \u00E2\u0080\u0093 Two additional dates from the bottom and top of cultural deposits at the far southern margin of the site (core 20A) indicate a horizontal expansion of the site between approximately 800 years ago (ca. AD 1170-1260) and ca. 400 years ago (ca. AD 1490-1630). If these basal and terminal dates adequately characterize the chronology for this area, it indicates that Huumuuwaa roughly doubled in length and then contracted to its former extent 400 years later (ca. AD 1500-1600). An additional date of approximately 600 years BP, obtained from midway down a small ridge behind a house platform. Shiwitis \u00E2\u0080\u0093 Gilbert Island The archaeological site of Shiwitis (82T, DfSh-29) is a medium-sized (2,400 m2) shell midden on southeastern Gilbert Island facing Effingham Bay, a protected bay overlooking Coaster Channel to the north and east (Figure 6.7). Shell midden deposits are situated on a tombolo landform amidst an elevated promontory to the south and north and parallel the shoreline for nearly 90 m. The deposits extend to a depth of approximately 2 m as determined by multiple percussion core tests. The shell midden deposit includes at least one house-platform associated with a small \u00E2\u0080\u0098backridge\u00E2\u0080\u0099 but cultural topography is generally lacking for much of the rest of the site. Visitors to Pacific Rim National Park Reserve currently use the site as a campground. Numerous bark-stripped culturally modified trees (CMTs) currently grow on the site amidst a series of older hand-logged cedar stumps and likely date to the early or mid-20th century (Sumpter and St. Claire 2007). The Tseshaht placename for the site location translates as \u00E2\u0080\u009Cwhale oil on the beach\u00E2\u0080\u009D (St. Claire 1991:146). This location was used seasonally by Ts\u00E2\u0080\u0099ishaa7ath and Maktl7ii7ath families after the two groups amalgamated (Sumpter and St. Claire 2007). Shiwitis has been classified as one of the 15 village settlements in the Broken Group Islands following the \u00E2\u0080\u0093 236 \u00E2\u0080\u0093 criteria of Haggarty and Inglis (1985:63), who define village sites as \u00E2\u0080\u009Cgenerally exceeding 100 metres in length, with well-defined house platforms, and or house depressions, and a rounded, back midden ridge.\u00E2\u0080\u009D This site represents the smallest site in their \u00E2\u0080\u0098village\u00E2\u0080\u0099 category and provides an interesting contrast to the much larger village sites of Ts\u00E2\u0080\u0099ishaa, Maktl7ii, and Huumuuwaa. The archaeological chronology of Shiwitis begins approximately 1,200 years ago as determined from three charcoal dates on basal cultural sediments from three separate areas along the centre of the midden landform. The two dates from the northern half of the midden landform strongly overlap in time (ca. 1260-1090 cal yr BP or AD 690-870). The third basal date at the south end of the midden began accumulating several decades thereafter around 1,100 years ago (ca. 1170-1060 cal yr BP or AD 780-890). A shell date from near the top of the deposits places the cessation of midden accumulation just after AD 1400 (Figure 6.14). Similar to Huumuuwaa and Maktl7ii, multiple radiocarbon dates from separate areas at the base of a midden ridge landform are closely spaced in time indicating relatively rapid simultaneous occupation of a large area within a 800-year period, from approximately AD 700 up until the mid 1400s (AD). \u00E2\u0080\u0093 237 \u00E2\u0080\u0093 Figure 6.13. Oblique view of the archaeological deposits and landform setting at the site of Shiwitis (82T). Triangles depict auger tests. Contour lines are in metres and the brown- blue transition represents sea level (i.e., barnacle line). Scale in arbitrary metres shown on the margin of the image. Angular lines on periphery of image represent lack of topographic data. Figure 6.14. Calibrated radiocarbon dates for Shiwitis, 82T showing the calibrated age range in years AD and the probability for individual dates. Dates on marine shell shown in turquoise and dates on charcoal are in grey calibrated on the atmospheric curve. Ch'ituukwachisht \u00E2\u0080\u0093 Cree Island Ch'ituukwachisht (Cree or Redonda Island as it is referred to on early marine charts) is a small (7.7-hectare) steeply-sided island in the southernmost periphery of the Broken Group (Figure 6.7 and Figure 6.15). Due to its exposure and lack of suitable landing spots, \u00E2\u0080\u0093 238 \u00E2\u0080\u0093 predominant weather conditions make approaching this island by boat very hazardous except in a very calm wind and minimal ocean swell. Figure 6.15. Southwest facing view of Ch'ituukwachisht from the upper fortress of Huts\u00E2\u0080\u0099atswilh showing the general location of sites 132T and 131T on the far left and far right portion of the island respectively. Photo by the author. Ch'ituukwachisht contains two previously recorded archaeological sites (Figure 6.15), both of which are elevated and have expansive views of the open ocean.. Previous archaeological and ethnohistorical documentation of these sites (131T and 132T) has interpreted the south facing site (131T) as a \u00E2\u0080\u0098lookout\u00E2\u0080\u0099 and the north facing site (132T) as \u00E2\u0080\u0098defensive\u00E2\u0080\u0099 (Haggarty and Inglis 1985:85-86). The interpretation as a lookout for site 132T is supported by its 290-degree view from an elevation of 27 m (90ft), which would enable those with keen eyesight to view at least 18 km out to sea during ideal weather conditions25. Also from this vantage point, one can also see elevated portions of Makah territory 60-90 km to the south (including the Islands off Ozette and Cape Flattery). This fits the Nuu-chah-nulth definition of a lookout, nachowa7a, translated by Sapir as \u00E2\u0080\u009Cwhere one goes to look out to sea\u00E2\u0080\u009D (St. Claire 1991:143 citing [Sapir nd. field notebook I:57]). It is additionally worth 25 http://blogs.discovermagazine.com/badastronomy/2009/01/15/how-far-away-is-the-horizon/ 132T 131T \u00E2\u0080\u0093 239 \u00E2\u0080\u0093 mentioning that there appear to be few other localities in the Broken Group archipelago which have such expansive elevated viewscapes of the offshore environment. Archaeological mapping of the extent of shell-bearing cultural deposits indicate both sites 132T and 131T are relatively small (150 and 1100 m2 respectively) and sub-surface investigation indicates both sites have relatively shallow deposits (less than 2 m) which accumulated over several millennia as determined by six radiocarbon dates (Figure 6.16). Unlike the larger sites discussed previously, the small horizontal extent and the prolonged period of occupation both at sites 131T and 132T are consistent with a small number of people periodically occupying a limited area over long periods time. These observations provide further support for the oral historically documented practice of whalers, sealers, and navigators occupying lookouts to watch for incoming weather and others travelling the coast. Figure 6.16. Calibrated radiocarbon dates on charcoal from deposits at upper Huts\u00E2\u0080\u0099atswilh (site 129T, top), Northern Cree Island (site 132T, middle), and Southern Cree (site 131T, bottom). These plots show the calibrated age range and probability for individual dates. Periodic Use as a Lookout \u00E2\u0080\u0093 240 \u00E2\u0080\u0093 Huts\u00E2\u0080\u0099atswilh \u00E2\u0080\u0093Village and Fortress The final two sites examined in this section are both on Dicebox Island and are named Huts\u00E2\u0080\u0099atswilh. This small 6-hectare island contains extensive archaeological deposits including well-defined house platforms, constructed with shell midden matrix, atop a steep, elevated bluff on the southern side of the island (site 129T, DfSh-79) and a lower elevation shell midden with house depressions covering a tombolo landform (site 83T, DfSh-31) with east and west facing beaches (Figure 6.17). There are three Indigenous place names for the Island. Huts\u00E2\u0080\u0099atswilh is specific to the village and fortress location and is derived from the word for \u00E2\u0080\u0098drift back,\u00E2\u0080\u0099 referring to winter waves which \u00E2\u0080\u009Csplashed through a cave in the middle of the island\u00E2\u0080\u009D (St. Claire 1991:145). A beach on the western side of the island is named 7aats\u00E2\u0080\u0099aatsupshilh, which means, \u00E2\u0080\u009Cwhen you are there it is so beautiful that you don\u00E2\u0080\u0099t want to leave\u00E2\u0080\u009D. A third placename, Ch\u00E2\u0080\u0099iwaakhat, translates as \u00E2\u0080\u009Csomebody holds your hair while you are at it\u00E2\u0080\u009D (ibid.) likely indicating a sheer drop from the elevated site. The site is described in oral histories as a \u00E2\u0080\u009Cvillage\u00E2\u0080\u00A6 on top of the hill inhabited by the Hots\u00E2\u0080\u0099atsswil?ath, a sept of the Ma:ktl?i:?ath\u00E2\u0080\u009D (Sapir 1910-1914; William 2009:370) indicating that the island was owned by the Huts\u00E2\u0080\u0099atswilh7ath, a sub-group of the Maktl7ii7ath (Inglis and Haggarty 1986:127; Sapir 1910-1914). The occurrence of extensive archaeological deposits on the precipitous elevated location of upper Huts\u00E2\u0080\u0099atswilh supports the interpretation that the upper site area was used over a prolonged period that may have occurred during a period of political and/or territorial conflict. \u00E2\u0080\u0093 241 \u00E2\u0080\u0093 Figure 6.17. Perspective view of a topographic model of Huts\u00E2\u0080\u0099atswilh showing the lower village in between two elevated bluffs and the upper defensive site on the sheer-sided elevated area to the right. \u00E2\u0080\u0093 242 \u00E2\u0080\u0093 Figure 6.18. View from upper Huts\u00E2\u0080\u0099atswilh (site 129T) looking west. Photo by Author. Lower Huts\u00E2\u0080\u0099atswilh Lower Huts\u00E2\u0080\u0099atswilh (site 83T, DfSh-31) contains nine surficially recognizable house depressions oriented to the west-facing beach on this tombolo landform (Figure 6.19). Seven radiocarbon dates from throughout this area (Figure 6.19) indicate the age of the cultural deposits falls between AD 1050 and 1620 (2-sigma calibrated range). Five of these dates are from the bottom of cultural sediments just above underlying beach sands and reveal a progressive sequence of horizontal site expansion from south to north over a period of approximately 400 years and spanning a distance of 70 metres. This south-to-north trend is also stratigraphically apparent in that the deepest archaeological deposits, indicating \u00E2\u0080\u0093 243 \u00E2\u0080\u0093 relatively more sustained deposition in the southern portion of the site that has more direct access to the elevated upper defensive site. The timing of the construction of these house depressions is therefore likely sequential as indicated by the greater depth of deposits on the southern portion of the site versus the northern portion. \u00E2\u0080\u0093 244 \u00E2\u0080\u0093 Figure 6.19. Dates from lower Huts\u00E2\u0080\u0099atswilh showing calibrated ages from basal percussion core samples. Note the progressive south to north trend in calibrated age reflecting horizontal growth of the site. Note the date from on lower right (83T11C) is a terminal date from the top of the cultural sediment rather than the other dates which are from basal layers. Dates on marine shell shown in turquoise and calibrated with a Delta-R of 150\u00C2\u00B121. Dates in grey are charcoal and calibrated on the atmospheric curve. Both sets calibrated using the program OxCal. \u00E2\u0080\u0093 245 \u00E2\u0080\u0093 Upper Huts\u00E2\u0080\u0099atswilh Radiocarbon dating was conducted on 19 samples from seven separate house platforms spread throughout upper Huts\u00E2\u0080\u0099atswilh (Figure 6.21). Two of the dates from upper Huts\u00E2\u0080\u0099atswilh are significantly older than the remaining 17, and indicate a highly intermittent sequence of human occupation between 2,500 and 500 calibrated years ago (Figure 6.16). These two dates are from two closely spaced cultural deposits in the western portion of elevated site area (Houses 6 and 3) that have particularly prominent views of the ocean to the west. They also closely overlap in age and are directly inter-visible with two similarly elevated small sites on Cree Island (131T and 132T), 1km to the southwest (Figure 6.15). I interpret the ages of these sites to indicate periodic use as inter-visible lookout locations (nachowa7a, St. Claire 1991:105). As for the remaining occupational sequence on the elevated portion of Huts\u00E2\u0080\u0099atswilh, multiple dates from four separate house platforms (Figure 6.21) indicate that house construction occurred extremely tightly in time between approximately AD 1430 and 1450 (pooled 2-sigma [94.5%] probability). This very narrow temporal range indicates an effectively simultaneous occupation of several areas of the site by multiple household groups. This relatively precise calibrated age estimate representing a span of a few decades (\u00C2\u00B120 years), is well within that of a human generation. This is consistent with a scenario of a relatively large-scale community effort to initiate construction and occupation of a defensive landform at a moment in history. A fifth dated house structure (House 2) situated in the midst of a cluster of these three dated houses appears to have been constructed slightly after this period (AD 1450-1620). Notably, this structure would have much less of a view and be buffered from approach by houses on its periphery. A sixth dated house structure on the \u00E2\u0080\u0093 246 \u00E2\u0080\u0093 southern periphery of the site dates to well within the contact era (ca. 1774-1850) and indicates a later expansion of this portion of the site, as previously interpreted by the 1982 survey and mapping team (Inglis and Haggarty 1986:276). \u00E2\u0080\u0093 247 \u00E2\u0080\u0093 Figure 6.20 Map of the lower village (83T) and upper defensive site (129T) indicating locations of percussion core tests and showing the topography and cluster of house platforms. Numbers following \u00E2\u0080\u0098129T\u00E2\u0080\u0099 indicate the house number while letters indicate the test number within that house. Core and auger tests are in the same locations. Clustered House Platforms Percussion Core Tests \u00E2\u0080\u0093 248 \u00E2\u0080\u0093 Figure 6.21 Dates on cultural deposits on upper Dicebox indicating the rapid construction of houses within the AD 1440-1460 interval and subsequent occupation up until the early 20th century. Discussion The preceding sections have examined archaeological evidence for settlement in seven of the 20 shell midden sites in Maktl7ii7ath local group territory (Figure 6.7). A series of radiocarbon dates document the lateral and vertical growth of these varied archaeological sites, ranging from 150 to 10,400 m2 in extent and from less than 0.5 to more than 7 m in depth (23 ft.). The earliest dates on cultural material from these sites are approximately 2700-2500 years old and come from two elevated sites the southern periphery of the archipelago \u00E2\u0080\u0093 249 \u00E2\u0080\u0093 Ch'ituukwachisht (131T) and the lookout component of Huts\u00E2\u0080\u0099atswilh (129T). Use of both sites at this time and over the following two thousand years is interpreted to be periodic based on the limited surface extent and depths of archaeological deposits. These elevated and directly inter-visible sites likely served as lookouts, nachowa7a, or \u00E2\u0080\u009Cwhere one goes to look out to sea\u00E2\u0080\u009D (St. Claire 1991:143 citing [Sapir nd. field notebook I:57]) and served a logistically vital role of watching for approaching people, animals, and weather. A century or so thereafter, deep midden deposits in three separate areas of the very large site of Maktl7ii began to accumulate simultaneously, which I interpret to be strong evidence for the rapid growth of a \u00E2\u0080\u0098village-sized\u00E2\u0080\u0099 community. Approximately 1,000 years later (ca. 1500-1200 BP), major portions of this site appear to have slowed or stopped accumulating. During this interval the also large site of Huumuuwaa began accumulating rapidly between 1800 and 1200 BP and reached nearly the same size of the Maktl7ii around 800 years ago. By 1200 BP a medium sized midden was established simultaneously at Shiwitis, while 800 years ago the lower village of Huts\u00E2\u0080\u0099atswilh grew rapidly followed by the upper fortress site abruptly 500 years ago. The results of these \u00E2\u0080\u0098site based\u00E2\u0080\u0099 settlement analyses are significant in that the earliest documented evidence for a village-sized settlement in the territory occurs at the named origin place of Maktl7ii. Recall that this placename describes the local group whose territory is recognized in oral historical accounts. The criterion for inferring a \u00E2\u0080\u0098village-sized\u00E2\u0080\u0099 settlement is the rapid horizontal expansion of sites with large midden landforms as determined by multiple radiocarbon dates. At Maktl7ii, two deep midden deposits began to accumulate within the same 300 year period. This site is also the second largest shell midden in the Broken Group archipelago following Ts\u00E2\u0080\u0099ishaa and thus is consistent with a substantial \u00E2\u0080\u0093 250 \u00E2\u0080\u0093 occupation over a long period of time. The use of this location as a village overlaps with the periodic use of three inter-visible sites situated on elevated landforms associated with lookouts (two sites on Cree Island, one on Dicebox). As indicated in the oral history, these were not village locations and the limited extents and depths archaeological deposits confirms their use as periodic. Figure 6.22. Settlement history in Maktl7ii7ath Local Group territory comparing the archaeological settlement sequence and the oral historical sequence (from left to right). Subsequent to this period, the village settlement of Huumuuwaa expands between 1800 and 1200 years ago. This significantly large site is the third largest shell midden in the entire archipelago (Table 6.1) and roughly doubles in size for a brief period of 300 years approximately 800 years ago (Figure 6.12) while the dates from upper Huts\u00E2\u0080\u0099atswilh provide strong evidence for the sudden need for constructing multiple households on a defensive landform suggesting conflict approximately 500 years ago and extended use into the 19th century. The dates at both Huumuuwaa and Huts\u00E2\u0080\u0099atswilh demonstrate these villages came into existence well after Maktl7ii. These observations are sequentially consistent with oral historical accounts identifying these villages as semi-autonomous \u00E2\u0080\u0098septs\u00E2\u0080\u0099 or ushtakimilh within Maktl7ii7ath territory. As indicated in Figure 6.22, when the archaeological chronology is evaluated alongside the oral historical accounts of village establishment, there is complementarity but it is not complete. Such a sequence however provides a chronological \u00E2\u0080\u0093 251 \u00E2\u0080\u0093 architecture for deeper readings of both the archaeological and oral historical sequences of which only an initial sketch has been provided here. Contact-era Political Change and Amalgamation After this millennia-long history of village and fortress establishment within Maktl7ii7ath territory, oral narratives additionally describe how the Maktl7ii7ath became engaged in a devastating violent conflict with the Hach\u00E2\u0080\u0099aa7ath during the \u00E2\u0080\u009Clatter years of the eighteenth century\u00E2\u0080\u009D and were reduced from a robust multi-village community to a small population with \u00E2\u0080\u0098less than 15 men\u00E2\u0080\u0099 (McMillan and St. Claire 2005:17). This fundamentally consequential event for the Maktl7ii7ath occurred during the early European fur trade, a period of widespread political tension and inter-community conflict that persisted into the mid-19th century (Harris 1997; McMillan 1999; Swadesh 1948). After regrouping at their origin village of Maktl7ii, the remaining Maktl7ii7ath members opted to politically amalgamate with their Ts\u00E2\u0080\u0099ishaa7ath neighbors to the north, retaining their potlatch seats and drift whale rights but ceding their territorial ownership to Ts\u00E2\u0080\u0099ishaa7ath (St. Claire 1991:41). Soon thereafter, a political alliance of Nuu-chah-nulth groups within and outside Barkley Sound collectively defeated the Hach\u00E2\u0080\u0099aa7ath (McMillan and St. Claire 2005:17). Regional tensions persisted however, and a series of conflicts unfolded over the next few decades (McMillan 2009; Swadesh 1948) culminating in the \u00E2\u0080\u009CLong War\u00E2\u0080\u009D in the 1840s which also extended beyond the Nuu-chah-nulth world (Sapir and Swadesh 1955:412\u00E2\u0080\u0093439). After this extended and large-scale conflict, the amalgamated Ts\u00E2\u0080\u0099ishaa7ath seasonally resumed settlement at the former Maktl7ii7ath village of Huumuuwaa as recounted in 1921 by Tom Sayach\u00E2\u0080\u0099apis (ca. 1835-1927): \u00E2\u0080\u0093 252 \u00E2\u0080\u0093 \u00E2\u0080\u009Cafter the herring finished spawning. We would go to Huumuuwa (Village Island), the whole Tsishaa Tribe, staying together because the war had ended only recently. We did not want to get separated\u00E2\u0080\u009D (Sapir and Swadesh 1955:39). As this quote indicates, the amalgamated Ts\u00E2\u0080\u0099ishaa7ath remained vigilant and consolidated during the years following the Long War, practicing seasonal residential mobility. As time passed and regional conflict did not re-emerge, the formerly autonomous local groups within the amalgamated Ts\u00E2\u0080\u0099ishaa7ath polity, began to seasonally reoccupy their former village sites within Maktl7ii7ath territory (Sapir and Swadesh 1955:45). Thus, approximately a decade later, when Sayach\u00E2\u0080\u0099apis \u00E2\u0080\u009Cgrew up to be a young man\u00E2\u0080\u009D [ca. 1850], he describes a time when: \u00E2\u0080\u009CWar was not in season. So the Tsishaa moved apart. The Maktlii Tribe went to Maktlii (Storm [Wouwer] Island). The Tsishaa Band was with the Nachimwas at Tsishaa [Benson Island]. The Himayis people went to Himayis. The Wanin people went to Wanin. The Nashas people went to Dutch Harbor [Hiikwis]. The Tlasimyis people went to Tlasimyis. The Hachaa people lived on Village Island [Huumuuwaa], for that was their land. The Hikuuthl people went to Shaahuwis. I used to live at Mokwa'a [Turrett Island]. We would troll when autumn was coming and when the cohoe started going thru the passage in schools.\u00E2\u0080\u009D (Sapir and Swadesh 1955:45). This account describes patterns in seasonal residential settlement that emerged after the Long War, wherein families within the now amalgamated Ts\u00E2\u0080\u0099ishaa7ath returned to their ancestral villages within their former territories. This reversion to pre-amalgamation territorial and residential affiliations, but only on a seasonal basis, indicates an enduring persistence for long-standing cultural identities and place-based lineage affiliations. Yet, it is during this period of a rapidly expanding colonial presence in the 1860s and 1870s (Sellers 2013) amidst the spread of virulent introduced diseases (Harris 1997), on which influential ethnographic observations are based (Grant 1857; Sproat 1868; Swan 1870) and form the foundation for \u00E2\u0080\u0093 253 \u00E2\u0080\u0093 later syntheses (Curtis 1916; Drucker 1951) that can gloss over oral historical sequences and result in the application of less than appropriate archaeological analogues (McMillan 2009). Conclusion This chapter has examined oral historical information and archaeological data specific to the Broken Group Islands to determine whether these distinct ways of understanding the past reveal parallel or comparable sequences of human settlement. I first considered the Broken Group Islands as a whole and examined place names and accounts of local group territories as summarized in the oral historical literature alongside the archaeological survey information for the archipelago. By considering these data for the larger region, I narrowed my focus on the southern cluster of islands within the archipelago as both culturally and archaeologically relevant units of analysis. This island cluster is recognized oral historically as the traditional territory of the Maktl7ii7ath and contains a similar range of archaeological site types, densities, sizes, and size-distributions as is found in other geographically distinct island clusters within the Broken Group associated with other politically autonomous local groups. I then presented a range of archaeological data within the local group territory with a focus on chronological indicators of village growth, expansion, and settlement reorganization. I observed several consistencies between archaeological and oral historical sequences. Specifically, Makt7ii, the named origin site of the Maktl7ii7ath, is the oldest and largest village while Huumuuwaa and Huts\u00E2\u0080\u0099atswilh are respectively more recent. Huumuuwaa exhibits periods of rapid changes in site size, both increasing and decreases over time, which may also correspond with a previous history of conflict between the Maktl7ii7ath and another group. Huts\u00E2\u0080\u0099atswilh expanded rapidly in size during two periods, \u00E2\u0080\u0093 254 \u00E2\u0080\u0093 the latter of which is an indication of the archaeologically simultaneous occupation of multiple houses at a fortress or defensive location (ca. AD 1440-1460). This historical dynamism and complex correspondence between archaeological and oral historical information support the observations of Crowell and Howell (2013), Cruikshank (2001), Green (2008), and Martindale (2006, n.d.), which demonstrate narrative oral accounts of human settlement maintain an enduring historical structure that compliments, and are complemented by, archaeological data. As demonstrated here, archaeological interpretation is not needed to verify the oral historical knowledge but rather, both forms of knowledge offer views of the past, and are made stronger and richer when examined in parallel. \u00E2\u0080\u0093 255 \u00E2\u0080\u0093 Table 6.3 Radiocarbon dates and calibrated age estimates (at 2-sigma) from archaeological sites in the Maktl7ii7ath study area. Site Test Location/Depth Elevation (m abl)* Material 14C Lab number 14C age \u00C2\u00B1 Delta R ** Cal yr BP Upper Cal yr BP Lower 131T 131T1A-64-66cm 11.22 Charcoal UCIAMS-75885 930 20 915 793 131T 131T1A5-87-89cm 10.99 Charcoal UCIAMS-73347 2170 20 2305 2116 131T 131T2A4-98-100cm 11.73 Charcoal UCIAMS-73345 2080 20 2116 1995 132T 132T2A-82-84cm 25.94 Charcoal UCIAMS-73346 2490 25 2719 2467 132T 132T3B-55-60 25.565 Charcoal UCIAMS-75886 1410 20 1344 1292 132T 132T3B-55-60 25.26 Charcoal UCIAMS-73344 1540 20 1518 1374 82T 82T9A14 150-158cm 1.852 Charcoal UCIAMS-105976 1245 15 1161 1093 82T 82T9A2 15-30cm 3.477 Mytilus sp. UCIAMS-105987 1045 15 Delta R=150\u00C2\u00B121 549 463 82T 82T4B 180cm 1.51 Charcoal UCIAMS-73351 1185 20 1173 1059 82T 82T10A-25cm 14.51 Charcoal UCIAMS-73350 95 10 255 32 82T 82T5B-252cm 1.83 Charcoal UCIAMS-73352 1240 20 1263 1083 304T 304T2A-708cm 3.74 Charcoal UCIAMS-72229 1835 15 1819 1720 304T 304T4A-436cm 3.74 Charcoal UCIAMS-72231 1775 15 1771 1616 304T 304T5A-443cm 5.45 Charcoal UCIAMS-72228 1290 15 1281 1179 304T 304T20A 150-153cm 2.985 Charcoal UCIAMS-105974 835 15 779 696 304T 304T20A 25cm 4.25 Charcoal UCIAMS-105975 345 15 396 317 304T 304T3C 135-139cm 5.204 Charcoal UCIAMS-105973 600 15 646 548 304T 304T1A11 183-86cm Charcoal UCIAMS-105979 1125 20 1064 971 206T 206T20B-117-20cm 3.724 Charcoal UCIAMS-105978 470 15 528 504 206T 206T4A-30cm 9.13 Mytilus sp. UCIAMS-72246 1060 15 Delta R=150\u00C2\u00B121 565 467 206T 206T4A-551cm 3.77 Charcoal UCIAMS-72233 1310 15 1290 1183 206T 206T9B-43cm 7.24 Mytilus sp. UCIAMS-72248 1905 15 Delta R=202\u00C2\u00B126 1317 1179 206T 206T19B2-20-40cm 1.976 Mytilus sp. UCIAMS-105977 2015 15 Delta R=202\u00C2\u00B126 1433 1279 206T 206T5A-361cm 1.74 Mytilus sp. UCIAMS-72245 2145 15 Delta R=202\u00C2\u00B126 1581 1394 206T 206T19B-94-96cm 2.702 Charcoal UCIAMS-105977 1785 20 1812 1623 206T 206T12A-335cm 3.65 Charcoal UCIAMS-73348 2380 25 2483 2342 206T 206T9B-365cm 4.02 Mytilus sp. UCIAMS-72247 2960 20 Delta R=220\u00C2\u00B124 2599 2336 206T 206T19A-590cm 2.23 Charcoal UCIAMS- 73349 2370 30 2650 2337 \u00E2\u0080\u0093 256 \u00E2\u0080\u0093 Site Test Location/Depth Elevation (m abl)* Material 14C Lab number 14C age \u00C2\u00B1 Delta R ** Cal yr BP Upper Cal yr BP Lower 129T 129T13F1-3 22.958 Charcoal UCIAMS-58619 440 20 522 483 129T 129T18F1 28.1 Charcoal UCIAMS-72235 150 15 281 0 129T 129T22D-1 13.741 Charcoal UCIAMS-62195 120 20 268 0 129T 129T2B1-5 19.115 Charcoal UCIAMS-61763 340 70 513 155 129T 129T2B1-3 18.255 Charcoal UCIAMS-61762 370 15 498 329 129T 129T3J 17.653 Charcoal UCIAM-75887 2515 20 2732 2494 129T eu129T6J16A 16.8 Charcoal UCIAMS-73410 300 25 455 299 129T eu129T6J16B 16.791 Charcoal UCIAMS-73411 445 25 529 479 129T eu129T6J12C 16.961 Charcoal UCIAMS-73412 190 25 296 -2 129T 129T6G 16.5 Charcoal UCIAMS-72234 1545 15 1517 1384 129T 129T9Z1-3 34cm 20.89 Charcoal UCIAMS-61761 430 20 519 474 129T 129T13F1-5 24.01 Mytilus sp. UCIAMS-58365 955 20 Delta R=150\u00C2\u00B121 503 368 129T 129T13F1-2 22.958 Balanus sp. UCIAMS-58361 880 15 Delta R=150\u00C2\u00B121 443 298 129T 129T25E 29.14 charred seed UCIAMS-72236 50 20 251 0 129T 129T3H 75-80cm 16.528 Charcoal UCIAMS-105980 475 15 530 505 129T 129T2B1-4 19.115 Mytilus sp. UCIAMS-58362 850 20 Delta R=150\u00C2\u00B121 422 279 129T 129T2B1-2 18.255 Mytilus sp. UCIAMS-58363 1000 15 Delta R=150\u00C2\u00B121 520 434 129T 129T9Z1-2 20.89 Mytilus sp. UCIAMS-58364 1035 20 Delta R=150\u00C2\u00B121 545 454 83T 83T11B1-4 7.067 Mytilus sp. UCIAMS-58366 945 25 Delta R=150\u00C2\u00B121 497 333 83T 83T2A2-2 2.189 Mytilus sp. UCIAMS-58368 1140 15 Delta R=150\u00C2\u00B121 633 522 83T 83T6B2-2 2.242 Mytilus sp. UCIAMS-58367 1210 20 Delta R=20\u00C2\u00B121 800 660 83T 83T7D2-2 263.5cm 2.28 Mytilus sp. UCIAMS-58369 1075 15 Delta R=150\u00C2\u00B121 599 484 83T ct-83T6B2-3, 257cm 2.242 Charcoal UCIAMS-61760 825 20 775 689 83T ST83T-ST1 25-32cm 1.51 Charcoal UCIAMS-73413 370 25 501 319 83T ct-83T10B 257cm 2.34 charred seed UCIAMS-72237 865 45 908 691 131T 131T1A97-99cm 10.89 Charcoal UCIAMS-75889 6240 20 7253 7028 131T 131T1A6108-116cm 10.75 Charcoal UCIAMS-75888 8210 35 9282 9031 129T 129T6F-2 58cm 15.961 Charcoal UCIAMS-62196 10765 35 12760 12570 * meters above barnacle line, ** Delta-R values derived from shell-wood pairs following personal communications with Paula Reimer and Maarten Blaauw (July 2013). *** Calibrated upper and lower age estimates shown are at the 2-sigma range and were calibrated using Calib 6.10 (Reimer, et al. 2009). Underlined dates at bottom are from dated landforms and not from definitely cultural deposits. \u00E2\u0080\u0093 257 \u00E2\u0080\u0093 Chapter 7. Zooarchaeological Scales of Variability on the Northwest Coast: A Barkley Sound Case Study Introduction Evaluating difference and similarity between cultural practices and their relationship to ecological factors across space and time is a fundamental challenge in archaeological interpretation. In this chapter, I outline a multi-scalar approach to identifying the most commonly consumed foods at five archaeological sites in the southern Broken Group Islands, falling within the contact-era territory of an autonomous Nuu-chah-nulth local group (Figure 7.1). Rather than seeking to detail the full spectrum of subsistence, I identify the most commonly represented fish and shellfish at the five study sites by examining patterning at multiple scales using several abundance measures applied across numerous depositional contexts spanning a period of 2,500 years (see previous chapter for site context and dates, Table 6.3). The foundation for this analysis are basic units of archaeologically recovered data: a small quantity of cultural sediment (0.5-2 litres) from a series of individual auger sample levels. I argue that attentiveness to such small-scale deposition most closely resembles everyday practices (i.e., regularly disposed or deposited refuse) and provides an enhanced basis for evaluating \u00E2\u0080\u009Chow contemporaneous sites within settlement systems functioned in relation to resource variability\u00E2\u0080\u009D (Moss 2012:6). Using these data, I identify small-scale environmental variation between culturally affiliated settlements through patterning in the consistency of resource use within sites as well as between sites that spans the last several millennia up until the late 19th century. \u00E2\u0080\u0093 258 \u00E2\u0080\u0093 Theoretical and Methodological Context Archaeological research on the Northwest Coast has overwhelmingly been conducted in relation to the compelling and pervasive narrative concerning the \u00E2\u0080\u0098origin\u00E2\u0080\u0099 and development of \u00E2\u0080\u0098intensive\u00E2\u0080\u0099 and/or \u00E2\u0080\u0098complex\u00E2\u0080\u0099 subsistence economies26. The structure of this narrative is progressive, linear, and invariably predicts a series of increasingly complex forms of social and political interaction developing over time (Mackie 2001; Martindale and Letham 2011; Moss 2011a). A host of economic and paleoenvironmental drivers (e.g., storage, trade and exchange, stabilization of sea levels, environmental productivity, population pressure, hereditary inequality, and control over labour) have been proposed to explain the \u00E2\u0080\u0098origins\u00E2\u0080\u0099 of social and political \u00E2\u0080\u0098complexity\u00E2\u0080\u0099 on the Northwest Coast (reviewed in Ames and Maschner 1999; Matson and Coupland 1995). However, archaeological evidence to support this narrative, particularly regionally and temporally coherent socio-economic data on subsistence, remains poorly demonstrated and lacks consensus despite decades of research (Butler and Campbell 2004; Moss 2012). In a recent review, Moss (2012:1) concluded that the \u00E2\u0080\u009Csingle narrative of increasing Northwest Coast cultural complexity is no longer tenable\u00E2\u0080\u009D as an analytical construct and argued for an increased attentiveness and engagement with regional cultural and ecological variability on a smaller scale. This imperative to confront and explain smaller-scale variability builds on Cannon\u00E2\u0080\u0099s insight from a decade previous, who similarly advocated for \u00E2\u0080\u009Cthe pressing need to expand research strategies beyond the individual site to encompass a variety of site locations within specific regions\u00E2\u0080\u009D thereby allowing archaeologists to \u00E2\u0080\u009Cgain a better appreciation of the extent of regional, seasonal, and longer-term temporal variability\u00E2\u0080\u009D (Cannon 2001:185). 26 Even studies that have argued against such a viewpoint have been discursively engaged in a critique of it. This is really the true measure of its pervasiveness and influence. \u00E2\u0080\u0093 259 \u00E2\u0080\u0093 Moss and Cannon\u00E2\u0080\u0099s advocacy for confronting variability is a recognition that the large spatial and temporal scales prevalent in Northwest Coast archaeological interpretation must be refined with stronger datasets that better resolve archaeological patterning and variability and better account for the social processes they represent. They identify the need for studies at an intermediate, sub-regional scale that allow for multiple perspectives on the differences and similarities among spatially and temporally associated sites and at multiple scales of analysis. Such an approach holds the possibility for identifying scales at which archaeological data are most strongly patterned and thus may have the greatest interpretive relevance. It is also a recognition that patterning may be recognizable at certain scales but obscured and/or lost at other scales (cf. Ames 1991; Levin 1992). Fortunately, there have been many sub-regional studies that have investigated spatially and temporally associated sites with the aim to detail such small-scale variability (Acheson 1998; Calvert 1980; Friedman 1976; Maschner 1992; McLay 1999a; Moss 1989; Orchard 2007). More recently, a number of researchers have built on the work of Cannon (2000a, b) to combine percussion coring, auger sampling, and limited excavation at multiple sites to investigate sub-regional patterning in settlement and resource use at coastal shell midden sites (Brewster and Martindale 2011; Cannon, et al. 2008; Cannon, et al. 2011; Letham 2011; Martindale, et al. 2009; Martindale, et al. 2010). Such studies help refine understanding of patterns of indigenous subsistence and settlement by sampling a number of locations within and between coastal shell midden deposits27. 27 Fundamental research is still being conducted on relative sea level change during the Holocene which has immense consequence for where sites dating to certain time periods are located and why sites dating to other time periods are absent from current records (e.g., Dallimore, et al. 2008; Fedje, et al. 2009; Grier, et al. 2009; Mackie, et al. 2011; McLaren, et al. 2011). Stratigraphy, accumulation rates and the temporal resolution of these deposits also remain poorly understood (Taylor, et al. 2011). Even within relatively well known time periods \u00E2\u0080\u0093 260 \u00E2\u0080\u0093 Although this considerable research effort has greatly expanded the basis for interpreting variation, examination of such patterning remains focused on archaeofaunal data from singular sites, thereby collapsing centuries or millennia of subsistence into a singular analytical unit, and if available, a few coarse-scale temporal divisions. An additional limitation is that the spatial scale of comparison rarely examines several sites within a small ecologically similar and ethnographically defined local group territory, which makes it difficult to distinguish between cultural and environmental variability (Moss 2012). Combined, these factors mask understanding of spatial and temporal variation in resource use within individual sites and over time and at the scale that most closely reflects everyday depositional practice. This widespread tendency to \u00E2\u0080\u0098lump rather than split\u00E2\u0080\u0099 is driven by the legitimate concern for adequate sample size, a cautionary awareness of the temporal coarseness of calibrated radiocarbon dates, and a recognition of the complexity of shell midden formation processes. However, these same factors have the corresponding effect of flattening generations of human subsistence and history into singular site-based observations. Here, I use a multitude of small (0.5 - 2 litres) but broadly distributed sediment samples obtained from multiple sites and site areas to examine patterning in depositional practices as a proxy for interpreting the everyday over millennia in a small territory on the Northwest Coast. and regions, there remains a surprising lack of temporal resolution which further limits identification of even the most coarse-scale cultural historical units (Morin 2013). \u00E2\u0080\u0093 261 \u00E2\u0080\u0093 Figure 7.1. Map of the Maktl7ii7ath study area in the Broken Group Islands in Barkley Sound. Figure 7.2. Arial overview of the Maktl7ii7ath study area with individual study sites looking west towards the Pacific Ocean. Photo courtesy of the Royal BC Museum (West Coast Archaeological Project Files). Effingham Bay OFFSHORE Coaster Channel _ _ \u00E2\u0080\u0093 262 \u00E2\u0080\u0093 Analysis of Scale and Variability To establish the taxonomic composition of fish and shellfish in my study area and explore spatial and temporal variability in resource use, I apply four measures of abundance. The measures of abundance include: 1) Relative abundance: the percentage of a particular item (e.g., herring) relative to all other specimens within the same category (e.g., identified fish remains). 2) Density per litre: the number of specimens per litre of excavated volume.28 This measure can be calculated by dividing the frequency of a taxa by the examined volume of cultural sediment (e.g., 40 herring bones per litre). Accordingly, this measure is not subject to the relative abundance of other taxa (e.g., if anchovy abundance goes up, then herring abundance does not necessarily go down). 3) Ubiquity: percentage of contexts in which a certain item is present or absent (frequency of occurrence). For example, herring would have a \u00E2\u0080\u0098ubiquity\u00E2\u0080\u0099 value of 100% if it were present in every level of every excavation unit in a site. 4) Rank order and rank order abundance: Rank order is the numeric rank of taxa based on their relative abundance within a given context (e.g., the most abundant taxa, or that with the highest percentage, receives the rank order of one; the second-most abundant receives a rank of two, etc.). Rank Order Abundance is the calculated average rank order across many examined contexts (e.g., all the individual auger sample levels within a site). Rank order abundance is expressed as a percentage reflecting how frequently a taxon receives a specific ranking (e.g., herring is the highest ranked taxon in 95% of contexts). The first measure, relative abundance, is by far the most common quantitative measure of abundance used in zooarchaeological analysis. Measure 2, density per litre, is much less common but reported with increasing frequency on the Northwest Coast (Brewster and 28 This term is not the same \u00E2\u0080\u0098bone density\u00E2\u0080\u0099 which is an entirely separate form of zooarchaeological analysis that examines the physical properties of bone from a taphonomic perspective (Butler and Chatters 1994). \u00E2\u0080\u0093 263 \u00E2\u0080\u0093 Martindale 2011; Cannon 2000a). Measure 3, ubiquity, has long been a central abundance measure in paleoethnobotany (Lepofsky and Lyons 2003; Popper 1988) but \u00E2\u0080\u009Chas seldom been used\u00E2\u0080\u009D in zooarchaeological datasets (Lyman 2008:84)29. Measure 4, rank order abundance, as a proportion of relative rank order for taxa across the multiple depositional contexts within a site has, to my knowledge, not been applied to the analysis of zooarchaeological datasets. While the use of these multiple measures may initially seem redundant, I employ them here because they have an analytical importance that dramatically expands the archaeological interpretation of variability at different scales. For instance, the first and most widely used measure of relative abundance is commonly the only measure that summarizes the composition of taxa for an entire site and/or stratigraphic/temporal components within a site. This is entirely legitimate and appropriate but has the consequence of limiting the archaeological perspective on variability to a small number of observations. Moreover, such relative percentage data are further influenced by other taxa (e.g., if something goes up, something else has to go down) (Grayson 1984). In contrast, the three other abundance measures are not as limited and collectively offer a greatly-expanded perspective on variability. For instance, density per litre is a measure that can be evaluated alongside relative abundance. Thus, if both measures similarly indicate that one particular taxon has the highest value, this doubly strengthens the observation of abundance as these two measures are quantitatively independent. While such a correspondence is important, it still leaves open the question of depositional variability within a site. Thus, an even more expansive perspective on variability 29 The calculation of ubiquity is not even discussed in two recent syntheses of zooarchaeological methodology (Reitz and Wing 2008; Russell 2012). \u00E2\u0080\u0093 264 \u00E2\u0080\u0093 is provided through the abundance measures of ubiquity and rank order abundance. These measures dramatically expand the archaeological basis on which abundance is evaluated because they give equal weight to the numerous archaeological contexts in which items are found. As a measure of the frequency of occurrence, ubiquity has the virtue (Kowalewski, et al. 1998) of being both spatially averaged across different areas of site and temporally averaged over the time a site has been occupied (e.g., individual auger sample levels from throughout a site). As such, ubiquity is a more direct measure of how frequently a particular taxon is present in an archaeological context and thus, a reflection of how regularly a class of item or animal taxa is used relative to others. It particularly well suited to quantification across small quantities of sediment (0.5-2 litres), and is akin to small scale deposition expected for routinized refuse disposal. As such, ubiquity can be used to evaluate how deposition is structured across a variety of archaeological contexts and thus how such repetitive patterning is relevant to the interpretation of \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 social practices (cf. Garrow 2012). Such a spatially and temporally expanded basis for measuring abundance is appropriate to the Northwest Coast given the many different depositional circumstances with which shell midden deposits accumulate (e.g., as house floors, as dump events, as storage features, etc.). This equal weighting (spatial and temporal averaging) of a large number of distinct depositional contexts (Kowalewski, et al. 1998) provides a significantly more well-rounded basis for assuming the regularity of occurrence, and hence, cultural use within a settlement relative to singularly lumped relative abundance values. It is further applicable to auger sampling which vertically penetrates cultural sediment but often lacks a clear indication of depositional context (housefloor, posthole, hearth feature, etc.). Thus, whereas \u00E2\u0080\u0093 265 \u00E2\u0080\u0093 relative abundance represents a totalized and singular value for a site or temporal component (Lyman 2003a), ubiquity is a proxy for how regularly certain items are used across the many examined contexts within a site and can similarly be scaled up to examine variation among several sites. Here, I use these four measures of abundance to identify the most regularly used taxa in a series of sites within a small ethnographically identified territory. I examine both inter-site and intra-site variation at three scales of analysis. The first scale of analysis applies these multiple abundance measures to all examined contexts in the study area as a single unit representing \u00E2\u0080\u0098resource use in Maktl7ii7ath territory.\u00E2\u0080\u0099 This multi-sited assessment of compositional patterning is legitimate because archaeological sampling was undertaken in multiple areas of each site using similar methods (auger sampling) and the analysis of fauna resulted in relatively similar sample sizes (NISP and shell weights, Table 7.1 Table 7.2). The second and most conventional scale of analysis is site-based, in that I simply compare lumped compositional data between individual sites focusing on the most abundant and most commonly used taxa in the territory as a whole. The third scale examines within-site compositional patterning for individual sites using four abundance measures. Applying the measures outlined above to two different sets of subsistence data (shellfish and fish) allows me to examine the robusticity of archaeological patterning for these regularly deposited foods. Collectively this focus on identifying patterning among those mundane and everyday foods (cf. Garrow 2012) allows for more detailed interpretations of the cultural, environmental, and historical factors that underlie differences in their proportionality. \u00E2\u0080\u0093 266 \u00E2\u0080\u0093 Methods Faunal Identification To explore the archaeological expression of everyday foodways as represented over millennia, I identified animal bones and shellfish from auger samples at seven sites in the southwest Broken Group Islands study area (Table 7.1 and Table 7.2). I examined multiple auger samples collected from separate areas within each site and identified bones systematically collected from 2 mm water screened mesh and shellfish collected from 6.35 mm mesh (\u00C2\u00BC-inch). The greatest auger sampling intensity and examined volume was at two sites on upper and lower Dicebox Island (129T and 83T) while the smallest examined volumes and sample sizes are from two small sites on Cree Island (131T and 132T). Relatively similar numbers of specimens and taxa were identified from five of the seven sites ensuring that comparisons can be made. Vertebrate skeletal identification procedures and quantification protocols followed the methods used in a number of other studies in the region (Frederick and Crockford 2005; Frederick 2012; Wigen 2003b, 2009, 2013) (see also Chapter 5). To further ensure analytical comparability, identifications were conducted using the same comprehensive comparative collection at the University of Victoria. Independent verification of a limited number of morphological identifications was provided by ancient DNA analysis focused on herring, Clupea pallasii (Speller, et al. 2012), and rockfish, Sebastes spp. (Rodrigues, et al. 2013), and indicates that more than 95% of specimens were accurately identified to genus or species. Invertebrate identification was conducted using the comparative shellfish collection housed at the Laboratory of Archaeology facilities at UBC, including a comparative \u00E2\u0080\u0093 267 \u00E2\u0080\u0093 collection donated from Parks Canada Archaeological Services30. Specimens were morphologically identified to the lowest possible taxonomic level and the number of hinges were counted and weighed to generate estimates of the minimum number of individual valves. While there are a variety of approaches to quantifying shellfish (e.g., Giovas 2009; Glassow 2000), the methods employed here are internally consistent and are thus internally comparable for the purposes of evaluating variability within the study sites. Size-at-Harvest Data To provide an additional level of insight into the taxonomic variability of fish within my study area, I examined the size-at-harvest profile for three of the five most abundant fish taxa (herring, rockfish, and greenling) using available skeletal regression formulae (McKechnie and Tollit n.d.; Orchard 2003). This methodology is based on the predictable relationship between the dimensions of particular bones and the known size and weight of modern fish which can then be applied to archaeological specimens to reconstruct size-at-harvest profiles (Casteel 1974). Results Bone and Shell Assemblages The combined vertebrate assemblage from the seven study sites contains 8,307 identified bone specimens representing 25 fish taxa, 6 mammal taxa and 3 bird taxa (Table 7.1). This represents 272 litres of recovered sediment from 58 spatially discrete auger samples containing 366 vertically discrete auger sample levels spanning the past 2,500 years. A total of 22.9 kilograms of shellfish from the 6.35 mm fraction was identified from the 30 Collection formerly managed by Cultural Resource Services staff at the now defunct Parks Canada Northern and Western Centre Victoria office. \u00E2\u0080\u0093 268 \u00E2\u0080\u0093 seven study sites representing 277 litres of recovered sediment from 66 spatially discrete auger samples containing 343 vertically discrete auger sample levels (Table 7.2). Examined volumes and number of examined contexts for vertebrates and invertebrates differ slightly due to identification effort but are based on the same set of samples. For five of the seven study sites, there are relatively similar sample sizes (examined volumes and identified shellfish weights) to provide sufficient basis for inter-site comparisons for both fish and shellfish \u00E2\u0080\u0093 and therefore these five will be analyzed in more detail below. Vertebrate remains are present in the overwhelming majority of all examined auger sample levels from the seven study sites (99.4%). Similarly, the overwhelming majority of the bone that is taxonomically identifiable to species genus or family is fish (Figure 7.3) whereas mammal bones make up only 0.6% of the identified assemblage (NISP=48). Identifiable bird bones are exceedingly rare in comparison, being 12 times less numerous than mammals and roughly 2,000 times less numerous than fish (Table 7.1). This vast difference in the proportionality of vertebrate fauna is commonly observed in other shell midden deposits in Barkley Sound that were screened with fine-screen (\u00E2\u0089\u00A43.2mm) mesh (Wigen 2003b, 2008a, 2009) and the Northwest Coast more generally (Moss and Cannon 2011) (see also Chapters 2 and 5). \u00E2\u0080\u0093 269 \u00E2\u0080\u0093 Figure 7.3 Proportion of fish, birds and mammals identified in the examined assemblage from the five study sites. Table 7.1 Examined vertebrate remains from the seven study sites (2 mm fraction); Maktl7ii, Huumuuwaa (304T), Shitwitis (82T), lower (83T) and upper (129T) Hutsatswilh, and sites on the south (132T) and north (131T) sides of Cree Island (Ch'ituukwachisht). Site Name NISP Fish NISP Mammals NISP Birds N Auger Samples N Auger Levels Volume (Litres)1 Fish per Litre N Fish Taxa Maktl7ii 206T 1939 8 2 6 99 61.8 31 17 Huumuuwaa 304T 1167 8 1 5 62 33.4 35 14 Shiwitis 82T 1374 7 4 37 27.3 50 16 Lwr Hutsatswilh 83T 1999 15 1 12 77 64.1 31 17 Uppr Hutsatswilh 129T 1738 10 27 77 74.0 24 17 N. Cree 131T 27 2 6 4.9 6 6 S. Cree 132T 11 2 8 6.9 2 3 Total 8255 48 4 58 366 272.4 30 25 1 Displaced volume as measured by placing bag of excavated sediment in a bucket with water. \u00E2\u0080\u0093 270 \u00E2\u0080\u0093 Table 7.2 Examined invertebrate remains from the seven study sites (6.35 mm fraction). Site Shell wt. (g) Volume (Litres)12 Mussel (g.) Clam Barnacle All other shell-fish Shell (g.) per litre N Auger Smpls N Auger Levels Maktl7ii 206T 2456.0 40.8 1588.9 319.8 504.2 57.0 60.2 8 62 Huumuuwaa 304T 2048.0 15.9 935.9 676.1 391.3 33.7 129.2 5 33 Shiwitis 82T 1742.7 21.5 775.4 685.5 196.9 61.3 81.1 3 29 Lwr Hutsatswilh 83T 10844.4 77.7 5770.4 2410.7 2331.6 177.7 139.6 13 90 Uppr Hutsatswilh 129T 5082.7 104.5 2485.4 1323.9 1120.1 72.8 48.7 32 109 N. Cree 131T 15.4 3.6 20.68 0 0 0 4.3 2 4 S. Cree 132T 709.0 13.5 636.5 0.27 66.84 3.62 52.5 3 16 Total 22898.3 277.4 12213.2 5416.2 4610.8 406.1 82.6 66 343 1Displaced volume as measured by placing bag of excavated sediment in a bucket with water. 2Examined volumes differ from vertebrate assemblage presented in table above as I examined more contexts for shellfish. Shellfish are present in over 95% of the examined auger sample levels from the seven study sites, a figure that would undoubtedly increase if specimens smaller than the 6.35 mm fraction (\u00C2\u00BC-inch) were included in this ubiquity measure. This demonstrates that shellfish are a pervasive feature of the depositional practices in the study sites and comprise an average of 82 g per litre in the 6.35 mm fraction. The assemblage is overwhelmingly dominated by mussel, clam, and barnacles. The combination of all remaining species amounts to less than 2% of the assemblage by weight (Figure 7.5). Scale 1 \u00E2\u0080\u0093 Multi-sited Assessment of Taxonomic Composition The identified fish assemblage from the seven study sites is numerically dominated by five taxa: herring, anchovy, greenling, rockfish, and salmon (Figure 7.3), followed distantly by a number of other species and genera. As discussed in further detail in the analyses of scales 2 and 3 (below), these same five taxa are also the five most abundant within each of the study sites for which there is sufficient sample size (Figure 7.7) and the five most ubiquitous in the 366 discrete auger sample levels examined from the \u00E2\u0080\u0093 271 \u00E2\u0080\u0093 investigations at each site and spanning all time periods for the study area. These multiple measures, representing a sub-regional, site-specific, and within site variation indicate that an intensive and broadly consistent use of these particular taxa occurred over the 2,500 years in the study area31. Figure 7.4 Proportion of fish taxa by number of identified speciemens (NISP) in the five study sites combined (2 mm fraction). Figure 7.5. Proportion of shellfish by weight from the five study sites combined (6.35 mm fraction). As with the fish remains, the taxonomic proportionality of shellfish reveals that a limited range of taxa dominate the invertebrate assemblage (Figure 7.5). The single-most dominant shellfish species is mussel (likely Mytilus californianus), which is also the most frequently occurring and abundant (by weight) shellfish genus within and throughout the five assemblages (Figure 7.8). Secondarily, various clam species (butter, littleneck, horse, and cockle) are consistently present in the multiple assemblages. Medium-sized barnacles (predominantly Balanus balanus) are another omnipresent invertebrate taxa. Comparatively 31 The remaining taxa are difficult to assess with the same degree of confidence due to the lack of sample size but more intensive sampling in the future will likely provide more on these rarer taxa. \u00E2\u0080\u0093 272 \u00E2\u0080\u0093 rare invertebrate taxa are also present at all five sites including abalone, urchin, and chiton, and turban snails but none are present in sufficient quantities to be evaluated in detail in the scales of analysis described previously. Given that these shellfish remains are from the 6.35 mm fraction, they under-represent smaller and more fragmentary shellfish remains such as mussel and likely over-represent larger and more durable shell taxa such as clam (Muckle 1994; Sumpter 2005:154). The exclusive use of specimens from the 6.35 mm fraction accounts for the lower observed abundance of mussel in the study sites in comparison to the nearby and similarly exposed sites of Ts\u00E2\u0080\u0099ishaa (Sumpter 2005) and Huu7ii (Sumpter 2012). Likewise, the taxonomic representation of shellfish is further skewed against the small, fragmentary, and disproportionately lightweight species such as urchin as well as species such as abalone (e.g., Menzies 2010) which tend to preserve poorly in wet shell midden sediments relative to other taxa. These methodological limitations therefore constrain the taxonomic representation and hence the recognition and resolution of the most regularly harvested shellfish. However, this methodology remains internally comparable for the purposes of this analysis and can be supplemented in future analyses. Size-at-Harvest Data for Fish Using estimates for a total of 214 elements with a relatively even sample size between sites (Table 7.3), I observed similar mean lengths and length distributions for the three fish taxa: rockfish (Sebastes spp.), greenling (Hexagrammos spp.) and herring (Clupea pallasii). Although each taxa exhibits different length variations, the majority of harvested fish are between 20 and 30 cm in length (Figure 7.6). The size-range does not appear to significantly differ between sites (Table 7.3) although a greater sample size would likely reveal a broader \u00E2\u0080\u0093 273 \u00E2\u0080\u0093 size distribution and potentially subtle differences over time or between sites. These observations are consistent with size-at-harvest data for rockfish, greenling, and herring at other sites in Barkley Sound (McKechnie 2007a, b) (see also Chapter 5). This similar size-range observed between sites (Table 7.3) further supports the previously described patterns of taxonomically focused and intensive fish harvesting practices over the past 2,500 years. Figure 7.6. Distribution of fish lengths for three of the five ubiquitous and abundant fish taxa in the study sites showing broadly similar size ranges clustering between 20 and 30 cm, with fewer specimens in the smaller and larger size categories. The faint curved line represents a normal distribution curve applied to each dataset. Table 7.3. Mean lengths and standard deviations in cm for herring (fork length), greenling (total length), rockfish (total length), and all measured fish combined from the five study sites. N indicates sample sizes. Site Herring SD n Greenling SD n Rockfish SD n Combined Mean Combined SD Combined n 129T 24.1 2.7 34 24.4 5.8 14 26.9 5.6 10 24.6 4.2 58 83T 26.1 3.4 23 27.2 7.4 13 25.0 4.6 13 26.1 5.0 49 82T 22.7 3.0 15 24.8 6.4 5 26.1 7.1 5 23.8 4.8 25 304T 22.3 3.2 14 24.8 1.8 2 23.8 7.1 5 22.9 4.2 21 206T 24.2 3.1 27 23.1 5.3 14 26.7 9.3 20 24.8 6.3 61 Total 24.1 3.2 113 24.8 6.1 48 26.0 7.1 53 24.7 5.2 214 \u00E2\u0080\u0093 274 \u00E2\u0080\u0093 Scale 2 \u00E2\u0080\u0093 Site-Based Compositional Differences Site specific spatial variation in use of the most abundant and ubiquitous shellfish and fish taxa across the five study sites indicates a broad similarity among sites in the proportion of species represented (Figure 7.7 and Figure 7.8). This indicates that similar fishing practices were undertaken by the residents of the sites and thus similar items were used as foods over the sites\u00E2\u0080\u0099 histories. Even so, a moderate degree of spatial variation is apparent in the use of the five most abundant and ubiquitous fish and shellfish within this small territory as indicated by differences in the relative and rank order abundance between sites. Among the identified fish (Figure 7.7), anchovy are slightly more abundant than herring in the two sites near Coaster Channel (sites 82T and 304T), while herring are the single-most abundant taxa in the three other sites with a peak at site 129T (the elevated defensive site on Dicebox Island). Other notable differences include a higher proportion of rockfish from site 206T and a high proportion of salmon in site 129T relative to the other study sites (Figure 7.7). The spatial patterning in fish is similarly represented in the shellfish data (Figure 7.8). There is a higher percentage of clams within the two sites in Coaster Channel (82T and 304T) and a higher percentage of mussels and barnacles in the three other sites (206T, 83T, and 129T). A notable difference between the two datasets is evident at site 129T which has a specialized focus on herring but an even use of shellfish taxa. These broadly complementary patterns indicate subtle spatial variation specific to place. \u00E2\u0080\u0093 275 \u00E2\u0080\u0093 Figure 7.7. Identified fish remains from the five study sites representing ca. 2,500 years of history. \u00E2\u0080\u0093 276 \u00E2\u0080\u0093 Figure 7.8. Proportion of shellfish from the five study sites representing ca. 2,500 years of history. Scale 3 \u00E2\u0080\u0093Within-site Patterning To further evaluate the spatial comparison presented above, I examined whether these patterns are present at finer depositional scales and thus, whether there is similarly consistent patterning within a site over the duration of occupation. I focus on anchovy and clams as they demonstrate the greatest variation in abundance at the site-specific scale. Across all four measures of abundance (relative abundance, density per litre, ubiquity, and rank order abundance), both anchovy and clam have their highest values in sites 82T and 304T (Figure 7.9). For clams, which are dominated by butter and littleneck, have their greatest abundance at sites 82T and 304T (Figure 7.9). The significantly greater number of grams per litre of butter clam versus littleneck likely reflects its greater robusticity and weight rather than a \u00E2\u0080\u0093 277 \u00E2\u0080\u0093 differential use between the species as they have such consistent differences across all five study sites. These within-site patterns confirm the spatial differences observed at the site-based spatial scale (Figure 7.7). Anchovy and clams were not present in every sample, but there is a consistently higher ubiquity, density per litre, and relative rank order as well as relative abundance in the samples from 82T and 304T than at the other sites. Figure 7.9. Four measures of abundance (relative abundance, density per litre, ubiquity, and rank order abundance) applied across the five study sites for anchovy (left) and clam (right). Note the consistently higher values for sites 82T and 304T indicate the greater anchovy and clam use throughout the occupation of these sites. Clams are further distinguished by butter and littleneck for two of the four measures. \u00E2\u0080\u0093 278 \u00E2\u0080\u0093 Discussion The previous analyses have shown that particular fish and shellfish resources were consistently and abundantly present in all five study sites among multiple abundance measures and at multiple scales of analysis spanning over two and a half millennia. This provides further support for the interpretation that fishing and shellfishing were vital and regular components of Indigenous subsistence practices in this coastal setting and constitutes strong evidence for regularly practiced aspects of everyday food harvesting and deposition. It further indicates a robusticity and resilience in the social and ecological dimensions of Indigenous foodways in this portion of the archipelago and suggests these particular animals played important economic and nutritional roles in these communities. While the analyses presented here document an enduring continuity in the intensive use of certain archaeologically abundant taxa, they reveal subtle spatial variation between individual sites that is strongly expressed across four measures of abundance. These differences are similarly observed across the multiple scales of analysis, which gives further confidence that these results are long-term reflections of \u00E2\u0080\u0098everyday\u00E2\u0080\u0099 depositional practices. Among the most striking spatial differences in compositional patterning are those evident in the two settlements in the eastern portion of the study area (82T and 304T) relative to the three to the west (206T, 129T and 83T). These two sites contain a consistently higher abundance of anchovy as well as littleneck, and butter clams (Figure 7.9). They are located along relatively protected, lower energy, and sediment-dominated shorelines of Coaster Channel and Effingham Bay \u00E2\u0080\u0093 habitats that are more suitable for clam harvesting and sheltered waters that are more suitable to anchovy schools (Figure 7.2). Conversely, higher percentages of mussel, barnacle, rockfish, and greenling are present on the exposed margin of \u00E2\u0080\u0093 279 \u00E2\u0080\u0093 the archipelago, which provides more suitable habitat for these taxa (Holmes and Tomascik 2003; Lee and Bourne 1977; Markel 2011; Schmidt 1999; Seed and Suchanek 1992). Such strong and persistent spatial differences in a small area and their relation to ecological characteristics near each settlement, indicates that the most abundant and regularly utilized fish and shellfish were likely harvested in close proximity to individual sites (\u00C2\u00B11 km). Moreover, the comparatively greater abundance of rockfish and greenling present in site 206T (Figure 7.7) is consistent with the more exposed offshore environment noted to be a preferred habitat for these genera (Holmes and Tomascik 2003; Markel 2011:27). A sample of rockfish bones identified to species using ancient DNA techniques from the five study sites (Rodrigues, et al. 2013) further substantiates this spatial variation (Figure 7.10). In particular, specimens from site 206T are predominantly composed of widow rockfish (Sebastes entomelas), a deeper-water species (Love, et al. 2002:172) that was not identified at any of the other tested sites. \u00E2\u0080\u0093 280 \u00E2\u0080\u0093 Figure 7.10. Ancient DNA identifications of rockfish in the study area (from Rodrigues, McKechnie, and Yang 2013). Size of pie chart is scaled to sample size. Collectively, these differences indicate that residents of the five study sites regularly used relatively small localized resource gathering areas. Of course, such an observation does not preclude the likelihood that resources from these sites were regularly exchanged and/or shared between site residents or well beyond the Broken Group archipelago, but it suggests that evidence for such practices appears to have been overwhelmed by what was caught and collected locally. It also indicates these small resource harvesting areas were highly productive as they could withstand the intensive use of large village-sized settlements over long periods of time (cf. Campbell and Butler 2010; Cannon, et al. 2008; Reitz 2014; Suttles 1968). \u00E2\u0080\u0093 281 \u00E2\u0080\u0093 This \u00E2\u0080\u0098hyper-localized\u00E2\u0080\u0099 pattern is consistent with the pervasive optimality notion that hunter-gatherers harvest resources in close proximity to their residences, or conversely, situate their residences in close proximity to resources in order to maximize the caloric return for their efforts (Maschner and Stein 1995; Monks 2011). However, this structured approach to resource use also suggests a process whereby intensive, sustained, and community-based use of localized harvesting areas provides an intergenerational incentive to maintain and enhance (where possible) productive resource gathering areas and/or to restrain and coordinate harvests in certain areas for future use, particularly those in close proximity to long-term settlements (Cannon and Burchell 2009; Lepofsky and Caldwell 2013). Such localized patterning in resource use also supports the interpretation that community residents at these sites exhibited territorial control over productive resource gathering areas near their settlements. Considered in relation to the large number of other large settlements elsewhere in the Broken Group archipelago indicating the presence of other autonomous local groups (see Chapter 6), it is reasonable to infer that residents in Maktl7ii7ath territory would be excluded from or have limited access rights to resources in other local group territories, further encouraging cultivation and maintenance of their established harvesting areas. This scenario is supported archaeologically by the spatially patterned use of shellfish and fish in the study area generally. It is particularly illustrated in the intensive use of clams from site 82T, which has the consistently highest clam abundance values across the four abundance measures (see tabular data presented within Figure 7.9) and where the beach fronting this settlement has an 80-m long clam garden feature (Figure 7.11). The feature consists of a boulder alignment paralleling the shore exposed only during the low-low tide (below 0.5m). This boulder alignment is bisected by four canoe runs and fronts an \u00E2\u0080\u0093 282 \u00E2\u0080\u0093 anthropogenically cleared beach containing moderately productive clam habitat. The attributes of this feature are consistent with clam garden features observed elsewhere on the coast (Caldwell, et al. 2012; Cardinal, et al. 2011; Harper, et al. 1995; Harper, et al. 2005) although the boulder wall and finer sediments on the shoreward side of the wall appear to have been dispersed by wave action in the 400 years since the site was intensively occupied (see 82T site chronology in Chapter 6). Combined with the proximity to this feature and the specialized focus on clams at this site over an extended period of time, this pattern indicates that generations of people who inhabited this place created this feature to expand and sustain use of this important resource. Figure 7.11. Anthropogenically altered beach at Shiwitis (Site 82T). View looking south. Note canoe run in foreground. Picture taken at a 0.5m low tide (Photo by author). \u00E2\u0080\u0093 283 \u00E2\u0080\u0093 Such focused, sustained, and patterned use, in combination with physical modifications to habitat, reflects a form of purposeful cultivation and is conducive to long-term resource stewardship and traditional environmental resource management (Cannon and Burchell 2009; Lepofsky and Caldwell 2013; Lertzman 2009; Moss 2011a). For intertidal areas, ecologically consequential management actions could be readily practiced through the periodic removal of shellfish predators such as sea stars (e.g., Paine 1974), the hunting and/or spatial exclusion of sea otters (e.g., Szpak, et al. 2012) as well as the physical alteration and expansion of habitat in order to maximize ecological productivity for specific preferred food items (e.g., Deur 2005; Groesbeck 2013). Thus, rather than a pattern of exploitation, such practices could serve to facilitate and sustain certain ecological interactions (cf. Bruno, et al. 2003; Haggan 2012; Menzies 2012; Smith 2011; Trosper 2009; Turner and Berkes 2006; Whitaker 2008). Numerous ethnographic and oral historical references to Nuu-chah-nulth territorial tenure systems identify that lineage-based households held the rights to resources (tuupati) within local group territories (hahuulhi). Descriptions of Nuu-chah-nulth resource management emphasizes permission to access to resources (excluding others), distributing harvested resources according to long-standing community protocols, and maintaining respectful relations with animals through ritual observances (Arima 1983; Atleo 2004; Curtis 1916; Drucker 1951; Ellis and Swan 1981; George 2003; Inglis and Haggarty 1986; Sapir 1922). The archaeological observations made here suggest a remarkable stability and continuity in Nuu-chah-nulth resource use for a range of regularly utilized fish and shellfish within a small local group territory. Archaeological evidence of resource use at this scale and \u00E2\u0080\u0093 284 \u00E2\u0080\u0093 taxonomic focus does not exhibit clear evidence for a directionality that can be encompassed within \u00E2\u0080\u009Ca single narrative of increasing Northwest Coast cultural complexity\u00E2\u0080\u009D (Moss 2012:1). In contrast, this record of resource harvesting and use provides an indication of the archaeological difficulty of distinguishing subtle cultural and ecological variability within and between sites. However, now that multi-scalar patterning in elements of resource use has been established for this study area, future research can build on these insights to better account for the long-term social, historical, and evolutionary processes they may represent. Conclusion This chapter has explored multiple scales of analysis and abundance measures in zooarchaeological patterning in order to provide more confidence in the archaeological observation of difference and similarity in resource use within a Nuu-chah-nulth local group territory. I have evaluated variation in resource use among several spatially, temporally, and culturally associated sites following Moss\u00E2\u0080\u0099s (2012) advocacy to engage with small-scale cultural and ecological variability in order to better interpret patterning within zooarchaeological datasets. Following Cannon (2001), this research moves beyond generalizing from an ecologically isolated individual site to examine variation in multiple sites that share ecological characteristics. I focused on the most abundant and ubiquitous foods within the politically defined territory of the Maktl7ii7ath local group, spanning 20 km2 with seven study sites in close proximity to each other (\u00C2\u00B1 2 km), all maintaining broadly contemporaneous occupation sequences dating to within the last 2,500 years. From each of these sites I examined faunal remains from multiple auger sample levels and from multiple site areas and observe that herring, anchovy, greenling, rockfish, salmon, mussel, clam and barnacle are the most \u00E2\u0080\u0093 285 \u00E2\u0080\u0093 abundant and ubiquitous species in all sites comprising over 350 depositional contexts. The strength of these observation derives in part from the spatially and temporally averaged sampling strategy, with a focus on the most regularly consumed foods using multiple abundance measures and scales of analysis. This analysis would have been much less representative if samples were taken from a single site area representing a narrow time frame. Future research will seek to broaden the basis for understanding variation and coherence with the goal of including more taxa, such as the important but relatively less regularly practiced activities such as whaling (Cavanagh 1983) and examining such patterning over more refined temporal scales. While the cumulative outcomes of human food harvesting practices in the study sites are clearly influenced by ecological characteristics, they are more accurately a direct outcome human agency in an ecological domain in that they reflect the structured and regularized use of resources in a small territory. The pervasive notion that Indigenous peoples on the Northwest Coast were simply subject to local and environmental availability (Cannon 2011b), masks the culturally situated and historically grounded knowledge about how the territories people maintained and the habitats people managed may be significantly enhanced relative to what was there \u00E2\u0080\u0098naturally\u00E2\u0080\u0099 (Ames 1991:942). As this case study has shown, an increased archaeological attentiveness to structured depositional practices (Garrow 2012) at multiple temporal scales ranging from the everyday to the longue dur\u00C3\u00A9e (Braudel 1970; 1980) indicates that Indigenous peoples did not simply catch the species \u00E2\u0080\u0098that were there\u00E2\u0080\u0099 but did so within a \u00E2\u0080\u009Cspecifically cultural and historical consciousness\u00E2\u0080\u009D (Ortner 2005:34) that reflects a deep knowledge and history of use in this particular marine environment. This is illustrated by the persistent abundance of clams at site 82T across \u00E2\u0080\u0093 286 \u00E2\u0080\u0093 multiple abundance measures and its proximity to a clam garden feature. With further exploration in relation to ecological dimensions of the study area (mussel beds, herring spawning habitats) and durable structures such fish traps and other intertidal rock alignment features recorded in this portion of the Broken Group archipelago, such associations have the possibility of being extended to a variety of other harvesting practices, cultural particularities and social conventions, as well as use of technologies and protocols of collecting food in this environment. Thus, while these patterns are ecologically influenced, they also inescapably cultural and this is an archaeological indication of \u00E2\u0080\u009Chow contemporaneous sites within settlement systems functioned in relation to resource variability\u00E2\u0080\u009D (Moss 2012:6). \u00E2\u0080\u0093 287 \u00E2\u0080\u0093 Chapter 8. Conclusion: Significance, Limitations, and Future Directions This dissertation makes an original contribution by identifying several scales at which two fundamentally important domains of everyday life, food and settlement, are archaeologically patterned and persist and change over millennia on the Northwest Coast. Following the advocacy of Cannon (2001:185), I sought \u00E2\u0080\u009Cto expand research strategies beyond individual sites to encompass a variety of site locations within specific regions.\u00E2\u0080\u009D I examined settlement and subsistence at a variety of scales, including the much of the Northwest Coast (Chapter 2), multiple linguistically and culturally distinct regions in southern British Columbia (Chapter 3), multiple sites in Barkley Sound and Prince Rupert Harbour (Chapter 4, 6, and 7), and between and within individual sites in Nuu-chah-nulth local group territories in Barkley Sound (Chapters 5-7). Collectively, this multi-scalar perspective offers a range of new insights into the temporal and spatial scales of archaeological variation on the Northwest Coast. In contrast to a singular narrative of increasing cultural complexity involving the unidirectional growth, elaboration, and specialization of subsistence economies (Moss 2012), data presented in this dissertation indicate that several aspects of Northwest Coast subsistence and settlement demonstrate remarkable continuity over millennial and centennial time scales and across a large variety of coastal landscapes. This identification of spatially and temporally coherent archaeological patterning is anthropologically significant as it provides a basis for interpreting key dimensions of human lifeways on the Northwest Coast over the \u00E2\u0080\u0098long-term,\u00E2\u0080\u0099 especially since the actions that generated these observed patterns can be reliably assumed to have unfolded within the domain of the \u00E2\u0080\u0098everyday.\u00E2\u0080\u0099 This patterning is \u00E2\u0080\u0093 288 \u00E2\u0080\u0093 an indication of how the culturally embedded practices (the habitus or system of dispositions [Bourdieu 1977]) of dwelling and eating durably persist over the longue dur\u00C3\u00A9e, or \u00E2\u0080\u0098beyond the history of events\u00E2\u0080\u0099 (Braudel 1970). The archaeological identification of such repetitively enacted practices within settlements demonstrates \u00E2\u0080\u009Ccontinuities of cultural categories\u00E2\u0080\u009D (Sahlins 1981:8) and further affirms the interpretive utility of focusing on \u00E2\u0080\u009Cthe little things one hardly notices in time and space\u00E2\u0080\u009D (Braudel 1981:29) such as the secular rituals of food and settlement. This focus on \u00E2\u0080\u009Cthe material signatures of everyday practice\u00E2\u0080\u009D (cf. Garrow 2012) allows for the exploration of patterning at a fine scale but also the integration of these data across larger scales of history. It also provides a productive point of departure for contrasting the equally important, but more infrequently occurring archaeological examples of symbolically laden ritual and ceremonial events that purposefully represent a dramatic departure from everyday practice (cf. Klokler 2008; Luby and Gruber 1999; McNiven 2013). Considering the richness and multi-dimensionality of human history in relation to certain foods, certain land/seascapes, and certain ecological histories (Anderson 2002; Bolster 2012; Mintz and Bois 2002; Wilk and Barbosa 2012), archaeological research refining understanding of long-term foodways is likely to remain a productive avenue for considering how food harvesting has influenced the environments we inhabit and the everyday practices we often take for granted. This dissertation has also demonstrated a simultaneous historical dynamism amidst longer-term patterning in settlement and subsistence (particularly in Chapters 4, 5, and 6). These are historical processes that unfold across an intermediate temporal scale, one that spans the lifetime of individuals and persists in social memory and is passed on through oral \u00E2\u0080\u0093 289 \u00E2\u0080\u0093 history. This intermediate temporal scale provides a fundamentally vital link between Bourdieu\u00E2\u0080\u0099s (1977) concept of practice, as an iterative unfolding of culturally constructed subjectivities and dispositions (Ortner 2005), and the regularized everyday practices of the longue dur\u00C3\u00A9e. Food Four chapters in this dissertation demonstrate a regional and temporal coherence in the archaeological patterning specific to certain animal food remains in coastal settlements (Chapters 2, 3, 5, and 7). These patterns are strongly represented both at centennial and millennial time scales as well as at fine depositional scales that approximate \u00E2\u0080\u0098structured use\u00E2\u0080\u0099 within individual sites (cf. Garrow 2012). This patterning is evident across a large variety of coastal settings and at different archaeological site \u00E2\u0080\u0098types,\u00E2\u0080\u0099 and collectively represents a multitude of linguistically, culturally, and historically distinct Indigenous communities. In chapter 2, I show that coastal settlements spanning the Northwest Coast, from Puget Sound in Washington to southeast Alaska, share a widespread pattern in the intensive and sustained use of Pacific herring, a small and ecologically important forage fish that is a foundation for marine foodwebs on which many other species depend. This contribution provides new insight into a coast-wide focus on this oil-rich fish, which also prolifically spawns on kelp and other substrates. This analysis significantly revises archaeological understanding of Northwest Coast fisheries and appropriately broadens the anthropological and archaeological focus beyond salmon. It further demonstrates regional variation, with particularly intensive use of herring in the northern and western portions of the Salish Sea (northern strait of Georgia) and in Barkley Sound. Temporal analyses of herring abundance data from 50 sites throughout the coast also illustrate a relative lack of chronological \u00E2\u0080\u0093 290 \u00E2\u0080\u0093 variability in herring use and secondarily provide insight into settlement variability with respect to fisheries. Future research on this considerable dataset, integrating a host of recent and neglected primary and grey literature, has the potential to further refine archaeological patterning for other species and also has relevance for fisheries management, historical ecology, and the zooarchaeology of human-animal relationships. Additional fisheries data from Nuu-chah-nulth territories in Barkley Sound presented in Chapters 5 and 7 provide further indication of the importance of herring and can be linked to the earliest documented European observation of a Nuu-chah-nulth village. Upon visiting the village of Yuquot in \u00E2\u0080\u0098Friendly Cove\u00E2\u0080\u0099 in March of 1778, Captain James Cook described herring eggs (herring roe) as the \u00E2\u0080\u009Cwinter bread of these people\u00E2\u0080\u009D and recounted the Mowachaht method for preserving herring in the rafters of houses: \u00E2\u0080\u009CThey hang them on small rods at first about a foot from the fire, afterwards they remove them higher and higher to make room for others till they get to the roof of the house; when dryed they are made up into bales and covered with Mats; thus they are kept till wanting and eat very well, but there is but little meat upon them. In the same manner they cure Cod and other large fish, and some are cured in the air without fire.\u00E2\u0080\u009D (Beaglehole 1967:303-304) This description of herring processing and storage is fulsomely illustrated by the expedition\u00E2\u0080\u0099s artist, John Webber, who sketched the interior of chief Maquinna\u00E2\u0080\u0099s house, including approximately 2,600 small herring-sized fish hanging from the rafters or being smoked or stored along with 26 fish of other species (Figure 8.1). This image provides further insight into daily practices within a Nuu-chah-nulth house and village at contact and can be related to an archaeological history spanning millennia. While it remains to be confirmed whether the small fish depicted are indeed herring, it is notable that the month of the visit, March, is a peak time for herring spawning. \u00E2\u0080\u0093 291 \u00E2\u0080\u0093 Figure 8.1 Illustration by John Webber depicting the inside a portion of the house of Mowachaht Muchalaht Chief Maquinna. Annotations estimate the number and type of fish species (98.8% herring or anchovy). Image courtesy of the Peabody Museum of Archaeology and Ethnology, Harvard University (Peabody ID: 41-72\u00C2\u00B710/499). \u00E2\u0080\u0093 292 \u00E2\u0080\u0093 In chapter 3, I identified long-term stability in marine and terrestrial hunting practice that overlaps with linguistically distinct culture areas (Nuu-chah-nulth and Coast Salish). While this patterning does correlate with environmental factors, it is more directly a reflection of purposeful action undertaken by generations of people who maintained a technologically sophisticated and culturally embedded hunting practice, without which human relations with these particular animals could not be sustained (e.g., pelagic fur seal hunting). These regionally coherent patterns persisted not simply because animals were \u00E2\u0080\u0098out there\u00E2\u0080\u0099 but because of the intergenerational transference of particular skills, ritual magic, and place-based knowledge as well as the long-term resilience of these animals in the face of culturally specialized human harvesting. This analysis provides further insight into the complex regional expression of social and economic relations within and between Nuu-chah-nulth and Coast Salish cultural histories. It also illustrates consistency between archaeological and Indigenous perspectives on food types as exemplified in this quote from Edward Sapir describing Tom Sayach\u00E2\u0080\u0099apis\u00E2\u0080\u0099 unfamiliarity with certain foods: \u00E2\u0080\u009CTom ate very little meat of land mammals in his early days. Indeed, like most of the [west] Coast people, he had a prejudice against deer meat and it was not until as a middle-aged man, he had come into contact with some of the deer hunting tribes of the interior of the island, that he learned to prize it, though to this day, venison has not for him the toothsome appeal of a chunk of whale meat\u00E2\u0080\u009D (Sapir 1922:304). Together, these observations represent continuity in patterning in subsistence over millennia but also exhibit differences in the relative proportionality of commonly used animals at regional, sub-regional, and micro-environmental scales. The identification of such large-scale patterning, as well as the compilation of a diffuse and often obscurely published primary literature that underlies it, provides an opportunity for future researchers to investigate patterning on a spatially and temporally expansive scale as well as on smaller spatial and \u00E2\u0080\u0093 293 \u00E2\u0080\u0093 temporal scales. For example, examining sub-regional patterning in marine mammal hunting can be further evaluated with other datasets (e.g., stable isotopes), provides a basis to infer the differential emphasis on particular mammal species in Indigenous ritual life (e.g., Sapir 1922; Suttles 1952), and provides better estimates of the long-term ecological baselines in relation to recent conditions and management regimes (e.g., Jackson, et al. 2001). This dissertation offers some specific examples of how certain forms of regularly practiced subsistence activities are expressed in the archaeological record on the Northwest Coast. At a coarse scale, I have identified strong evidence for widely shared aspects of everyday cultural practices, stretching across large regions of the coast and emerging from a deep engagement with specific animals and places, over thousands of years. At a finer spatial scale, the observed differences in food remains among sites located close to one another affirms the persistence of \u00E2\u0080\u0098routines of habit\u00E2\u0080\u0099 specific to and influenced by particular places (Mackie 2001). It also indicates that some elements in Indigenous conceptions of \u00E2\u0080\u0098food\u00E2\u0080\u0099 are very broadly understood while others exhibit subtle differences between regions, sites. This continuity suggests this strong repetition underlies the patterning of the longue dur\u00C3\u00A9e as it emerges from culturally situated everyday subsistence practices. This similarity further affirms the practice-based perspective that food is not simply a basis for survival but rather a culturally organized process (Sahlins 1976) that expresses a \u00E2\u0080\u009Cspecifically cultural and historical consciousness\u00E2\u0080\u009D (Bourdieu 1984; Ortner 2005:34). Settlement I have examined the broad-scale patterns and historical dynamics of settlement and place using radiocarbon dates (Chapters 4 and 6) and subsistence in coastal shell midden sites (i.e., Chapters 2, 3, 5, and 7). In Chapter 4, I focused on an area that has been extensively \u00E2\u0080\u0093 294 \u00E2\u0080\u0093 researched with a rich archaeological chronology (Prince Rupert Harbour). In this analysis Eldridge and I applied available marine reservoir correction data to recalibrate two previously disparate archaeological datasets (burial dates and village abandonment dates), allowing us to place them on a single calibrated timescale. This revised analysis makes more explicit the archaeological basis for inferring a large-scale change in the settlement history for the Northern Northwest Coast and established a foundation for further research involving dates on marine or marine influenced organisms. As Ames (2005), Archer (2001), and Martindale and Marsden (2003) have argued, this change resulted from a disruption in settlement as a result of large-scale conflict between the Tsimshian people and culturally distinct groups from the north \u00E2\u0080\u0093 an event recorded in Tsimshian oral history. Despite the relatively sudden abandonment of this dense aggregation of settlements, we observe the chronology of human interments to persist and even peak centuries after the period of \u00E2\u0080\u0098abandonment.\u00E2\u0080\u0099 This contrast in patterning indicates a centuries-long continuation of ancestral, lineage-based affiliations with particular former village locations and suggests a cultural emphasis on maintaining connections with ancestors. The chapter also is methodologically relevant for calibration of marine radiocarbon dates and the integration of regional radiocarbon datasets more broadly. In Chapter 6, I examined the spatial and temporal sequence of archaeological settlement alongside Indigenous oral historical sequences recorded in the early 20th century and observed a complementarity stretching back more than 2,000 years. I show there is enduring consistency in Nuu-chah-nulth occupation and use of landscapes at a site scale and household scale (also evident in Chapter 5) as well as within several broadly contemporaneous sites in a small territory (Chapters 6 and 7). These aspects of settlement \u00E2\u0080\u0093 295 \u00E2\u0080\u0093 and subsistence are consistent with the Nuu-chah-nulth concepts of territory (hahuulhi) in which local resources were used and managed intensively while others were excluded. This is further apparent in the asserted protection of their territory through the sudden construction and large-scale occupation of an elevated, difficult-to-access defensive fortress. This site in particular is a physical manifestation of the assertion of political and territorial autonomy in the face of conflict and contested control of a resource-rich archipelago. Chapter 6 also demonstrates that sequences of oral narratives can describe archaeologically observable construction sequences within a single village (Ts\u00E2\u0080\u0099ishaa) as well as the relative sequences of the establishment of separate village sites nearby (Matkl7ii, Huumuuwaa, and Huts\u00E2\u0080\u0099atswilh). This complementarity is observed in the context of a small named oral historically identified territory (Maktl7ii7ath) in a portion of the Broken Group Islands. This identification of parallel lines of archaeological and oral historical evidence strengthens and enriches archaeological inference (Wylie 1989, 2009). It simultaneously reveals a historical dynamism as well as continuity in the place-based familiarity that Nuu-chah-nulth communities had with their villages and the cultural landscapes surrounding them. This chapter explored how Indigenous oral narratives provide culturally situated insights into the events and causal factors of history and how Indigenous place names provide a spatial architecture for understanding the locations and landscapes in which that history unfolded. This subjective frame on the events of history helps to inform not only \u00E2\u0080\u0098the past\u00E2\u0080\u0099 but the how the past was \u00E2\u0080\u0098understood.\u00E2\u0080\u0099 Although this latter point is beyond my ability to fully explore, this effort to connect this information with archaeological settlement data provides a further indication of the interpretive potential of combining archaeological and Indigenous histories, particularly in the context of place names. \u00E2\u0080\u0093 296 \u00E2\u0080\u0093 Attentiveness to these oral historical accounts additionally details the Indigenous perspective on disruptions to settlement that occurred before and after contact with Europeans in the late 18th century. During the initial contact era, differential access to trading vessels and simmering tensions between politically autonomous local groups resulted in a number of amalgamations and conflicts as detailed in Indigenous oral histories. This latter insight importantly complicates the historical perspective on the events during maritime fur trade on western Vancouver Island (ca. 1778-1805) as a period of undifferentiated \u00E2\u0080\u0098enrichment\u00E2\u0080\u0099 for Nuu-chah-nulth communities participating in the trade (Fisher 1977). Rather, this research has contributed to further contextualizing the varied consequences of the colonial encounter in the late 18th century (Clayton 2000; Martindale 2009; McMillan 2009; Salmond and Salmond 2010; Sellers 2013). The contact-era shifts in territories and political amalgamations further demonstrate how \u00E2\u0080\u0098enrichment\u00E2\u0080\u0099 is an overly simplified perspective that \u00E2\u0080\u009Cignores too much of what was going on along the Northwest Coast during the contact years\u00E2\u0080\u009D (Harris 1997:81). Collectively, these settlement-related case studies illustrate the interpretive utility of considering Indigenous oral history alongside the evaluation of archaeological data. It also highlights the importance of applying a temporally refined \u00E2\u0080\u0098medium-term\u00E2\u0080\u0099 scale of historical analysis to archaeological data to track the dynamism represented over the lived experiences of individuals. Such \u00E2\u0080\u0098medium-term changes\u00E2\u0080\u0099 that span beyond the event-centered history such as Braudel\u00E2\u0080\u0099s (1970) conjunctures, comprise an interpretively rich scale of historical analysis, which, although clearly represented in Indigenous oral traditions, is commonly lacking or rarely specified in many archaeological sequences. Here, I have added chronological resolution to the previously identified phenomena of a peak in the \u00E2\u0080\u0093 297 \u00E2\u0080\u0093 abandonment of 22 village sites in a highly populated locality of Prince Rupert Harbour (Archer 2001; Martindale and Marsden 2003) amidst a more extended history of burial practices. In Barkley Sound, I explored the implications of a sudden reorganization of settlement during the contact era, the outcome of which was historically formative for the persistence of contemporary Tseshaht and Huu-ay-aht First Nations. The archaeological research presented in this dissertation further demonstrates the value of engaging with Indigenous oral narrative on the Northwest Coast. It provides an example of how a detailed consideration of such information can be \u00E2\u0080\u009Ctranslated across cultural epistemologies\u00E2\u0080\u009D (Martindale 2006:160) to strengthen what is often a necessarily more tenuous set of inferences had they rested solely on archaeological observations. It also shifts the evaluation of oral history from a role in ethnographic analogizing towards a more holistic effort to rigorously interrogate human history on the Northwest Coast, one in which archaeology is a part (Atalay 2008; Martindale 2006; Wylie 2009). Data Sources This research has drawn on a wealth of previously existing data compiled by dozens of researchers over several decades as well as my own primary field and laboratory research, which is also set within a long-standing research effort in Barkley Sound. I focused on compiling two types of data (morphologically identified animal bones and shells and radiocarbon dates) across a large range of archaeological scales. These two archaeological datasets comprise a robust basis for archaeological interpretation relative to other common forms of archaeological observation (e.g., artifacts types, architectural features, cultural historical phases). These data are quantitatively comparable, which has made it possible to simultaneously expand the scale of analysis and incorporate a range of data beyond my \u00E2\u0080\u0093 298 \u00E2\u0080\u0093 personal capacity to collect. Such forms of directly comparable data, increasingly synthesized by individual researchers (e.g., appendices in Angelbeck 2009 and Morin 2012) and compiled in interactive data repositories (e.g., Blake, et al. 2012; Gajewski, et al. 2011) allow for a richer collaborative synthesis in Northwest Coast archaeological knowledge. Limitations A fundamental drawback to the interpretation of archaeological data are the challenges of sampling and the problem of time. This dissertation has examined ubiquitous aspects of the archaeological record of food and settlement but has done so without giving equal attention to the issues and data presented in each individual chapter. This logistically necessary constraint can also be considered a virtue in part as it provides a clarity of focus on a limited number of variables at large scales or alternatively, and at the same time, a larger set of variables examined at a smaller spatial and temporal scale. Such a \u00E2\u0080\u0098tradeoff\u00E2\u0080\u0099 is present throughout the dissertation and leaves unexplored gaps in evaluation. For instance, the zooarchaeological analysis of Indigenous fisheries undertaken in Chapter 2 took a broad scale approach to much of the Northwest Coast but was limited to the analysis of a single (albeit numerically dominant) species \u00E2\u0080\u0093 herring. Similarly, the sequences explored in Chapter 5 do not match the level of detail provided in Chapter 7, either in terms of size of the identified assemblage, the temporal resolution, or the quantitative metrics (MNI, MNI per m3, etc.). In spite of this potential shortcoming, Chapter 7 presents data from multiple areas within five separate study sites and includes shellfish in addition to fish remains. Other chapters examined a more wide ranging suite of data, including contextualizing patterning in relation to present day variation and change and identifying a number of important contrasts \u00E2\u0080\u0093 299 \u00E2\u0080\u0093 (Chapters 2 and 3). However, further exploration is warranted, such as a more thorough integration of ecological and archaeologically observed patterning (Moss 2012). It is significant that those food items focused on in this dissertation are only those that were relatively most abundant in zooarchaeological deposits. However, this only represents a small fraction of the total representation of harvested items. In order to assess the abundance of comparatively infrequently occurring (rare) items (e.g., birds, land mammals), it will be necessary to conduct a much more intensive and targeted approach to sampling (Gray 2008). Another limitation is the difficulty of assessing the cryptic influence of cultural and non-cultural taphonomy in structuring the way in which archaeofaunal patterning is expressed. The approach taken in this dissertation has been to identify those elements of the archaeological record that are most abundant and ubiquitous, which reveals key aspects of pervasively used Indigenous foods and does not require as large of a sample to assess patterning. While this dissertation has provided detail on the enduring continuity of a range of cultural practices, it also reveals a historical dynamism at finer temporal and spatial scales. These insights provide a productive reference point for further scrutinizing increasingly refined scales of variability in future research. For instance, at the most refined temporal and spatial scale examined (Chapter 5), I observed spatially and temporally coherent changes in the proportionality of fish taxa within multiple areas of a large household at a single site. This relatively highly resolved record reveals fluctuations in the abundance of regularly consumed taxa that are not as apparent at the coarser scales explored in chapters 2, 3, and 7. Thus, patterning at a coarser scales may mask the patterns we might observe at a finer scale and by extension, obscure the social histories that produced them . \u00E2\u0080\u0093 300 \u00E2\u0080\u0093 Variation and Scale into the Future The challenge of these results is both simple and profound. I have argued that the archaeological data in this thesis is evidence of continuity in Indigenous history over millennia, but that it emerges from cultural conventions in which the volatility of individual agency is contained within what we might think of as tradition. This allows Indigenous history to be both dynamic and patterned, adaptive and responsive, and ultimately to be the history of both purposeful negotiation and under-acknowledged convention. It is this dynamic that links the longue dur\u00C3\u00A9e to the everyday, that allows us to see the constancy of intergenerational transference of convention in routinized momentary acts. The challenge is twofold, empirical and theoretical. The first is to test this consistency by evaluating archaeological patterning in greater detail and at finer scales to query whether the broad patterns of history that I have outlined here are upheld. The second is to consider how the longue dur\u00C3\u00A9e, and thus the long-term patterns of archaeological data, can be reconciled with the everyday moments that make up human history. One key domain explored in this dissertation is the complex ecological influence that human foodways have both exerted on and been shaped by marine environments over multiple human generations. I have sought to explore the small-scale spatial variability in the use of regularly used foods within contemporaneous sites in a small area or region and shown it to be influenced by ecological habitats in proximity to human settlements. This has been observed in patterns of shellfish and rockfish (Chapter 7) and herring and herring spawning areas (Chapter 2) and more broadly with marine and terrestrial mammals (Chapter 3). However, these patterns are also influenced by modification and expansion of habitat such as in the construction and use of clam gardens or the long established practice of burning oak scrubland to facilitate growth of camas and forage for deer. Thus, while the exploration of \u00E2\u0080\u0093 301 \u00E2\u0080\u0093 such patterning presented here, has provided a way to distinguish how \u00E2\u0080\u009Ccontemporaneous sites within settlement systems functioned in relation to resource variability\u00E2\u0080\u009D (Moss 2012:6), we still have a long way to go in our explorations of people\u00E2\u0080\u0099s relationships with everyday food. A final challenge facing future research is grasping the vast scale of the archaeological history on the Northwest Coast that has yet-to-be investigated. A hint of the magnitude of this under-explored historical complexity is suggested Figure 8.2 which depicts over 5,300 recorded coastal shell midden sites on the coast of British Columbia as filed in the provincial archaeological database collectively representing more than 50 years of survey observations by archaeologists and First Nations heritage managers. This map vastly eclipses the number of Indigenous communities on the coast today and yet only represents those shell midden sites located on the shoreline of tidally influenced waterways. However, it is an indication of the richness archaeologists have yet to address as less than 1% of these sites have been subject to sustained excavation let alone zooarchaeological analysis and radiocarbon dating. It also does not reflect the numerous regions which have yet to be surveyed such as those now-underwater or inland sites that were once situated on a former shoreline (Mackie, et al. 2011). Future research into such a large and complex archaeological history will, in the words of Joan Gero (2007), undoubtedly require that \u00E2\u0080\u009Chonoring ambiguity\u00E2\u0080\u009D and \u00E2\u0080\u009Cproblematizing certitude\u00E2\u0080\u009D remains a vitally important part of the interpretive process in the present and into the future. \u00E2\u0080\u0093 302 \u00E2\u0080\u0093 Figure 8.2 Archaeologically recorded shell middens on the British Columbia Coast as represented in the provincial archaeological database (N=5,383, data current as of May 2010). Least well-surveyed area indicated by the lack of green dots along eastern Hecate Strait, south of Prince Rupert, and east of Haida Gwaii. \u00E2\u0080\u0093 303 \u00E2\u0080\u0093 References Acheson, Steven R. 1998 In the Wake of the ya\u00E2\u0080\u0099aats\u00E2\u0080\u0099 xaatgaay [\u00E2\u0080\u0098Iron People\u00E2\u0080\u0099]: A Study of the Changing Settlement Strategies among the Kunghit Haida. International Series 711. British Archaeological Reports, Archaeopress, Oxford. 2003 The Thin Edge: Evidence for Precontact Use and Working of Metal on the Northwest Coast. In Emerging from the Mist: Studies in Northwest Coast Culture History, edited by R. G. Matson, G. Coupland and Q. Mackie, pp. 213\u00E2\u0080\u0093229. University of British Columbia Press, Vancouver. Agostini, Vera N., Robert C. Francis, Anne B. Hollowed, S. D. Pierce, C. Wilson and A.N. Hendrix 2006 The Relationship Between Hake (Merluccius productus) Distribution and Poleward Sub-surface Flow in the California Current System. Canadian Journal of Fisheries and Aquatic Sciences 63(12):2649-265. Allen, S.E., C. Vindeirinho, R.E. Thomson, M.G.G. Foreman and D.L. Mackas 2001 Physical and biological processes over a submarine canyon during an upwelling event. Canadian Journal of Fisheries & Aquatic Sciences 58:671\u00E2\u0080\u0093684. Ames, Kenneth M. 1991 The Archaeology of the Longue Dur\u00C3\u00A9e: Temporal and Spatial Scale in the Evolution of Social Complexity on the Southern Northwest Coast. Antiquity 65(249):935\u00E2\u0080\u0093945. 1998 Economic Prehistory of the Northern British Columbia Coast. Arctic Anthropology 35(1):68\u00E2\u0080\u009387. 2001 Slaves, Chiefs, and Labour on the Northern Northwest Coast. World Archaeology 33(1):1\u00E2\u0080\u009317. 2002 Going by Boat: the Forager-collector Continuum at Sea. In Beyond Foraging and Collecting: Evolutionary Change in Hunter-gatherer Settlement Systems, edited by B. Fitzhugh and J. Habu, pp. 17\u00E2\u0080\u009350. Kluwer/Plenum Press, New York. 2005 The North Coast Prehistory Project Excavations in Prince Rupert Harbour, British Columbia: The Artifacts. International Series 1342. British Archaeological Reports, Oxford. 2012 Radiocarbon Dates and Community Mobility Patterns on the Columbia Plateau. Journal of Northwest Anthropology 7:167\u00E2\u0080\u0093194. Ames, Kenneth M. and Herbert D.G. Maschner 1999 Peoples of the Northwest Coast: Their Archaeology and Prehistory. Thames and Hudson, London. Anderson, Christian N. K., Chih-hao Hsieh, Stuart A. Sandin, Roger Hewitt, Anne B. Hollowed, John Beddington, Robert M. May and George Sugihara 2008 Why Fishing Magnifies Fluctuations in Fish Abundance. Nature 452(7189):835\u00E2\u0080\u0093839. \u00E2\u0080\u0093 304 \u00E2\u0080\u0093 Anderson, Lesleigh, Mark B. Abbott, Bruce P. Finney and Stephen J. Burns 2005a Regional Atmospheric Circulation Change in the North Pacific During the Holocene Inferred from Lacustrine Carbonate Oxygen Isotopes, Yukon Territory, Canada. Quaternary Research 64(1):21\u00E2\u0080\u009335. Anderson, Lesleigh, Mark B. Abbott, Bruce P. Finney and M.E. Edwards 2005b Paleohydrology of the southwest Yukon Territory, Canada, based on multi-proxy analyses of lake sediment cores from a depth transect. The Holocene 15:1172\u00E2\u0080\u00931183. Anderson, Virginia DeJohn 2002 Creatures of Empire: How Domestic Animals Transformed Early America. Oxford University Press, Oxford. Angelbeck, William O. 2009 \"They Recognize No Superior Chief\" : Power, Practice, Anarchism and Warfare in the Coast Salish Past. PhD Dissertation, University of British Columbia. Angelbeck, William O. and Eric McLay 2011 The Battle at Maple Bay: The Dynamics of Coast Salish Political Organization through Oral Histories. Ethnohistory 58(3):359\u00E2\u0080\u0093392. Archer, David J. W. 1992 Results of the Prince Rupert Harbour Radiocarbon Dating Project. Report on submitted to the British Columbia Heritage Trust. On file with the BC Archaeology Branch, Victoria, BC. 2001 Village Patterns and the Emergence of Ranked Society in the Prince Rupert Area. In Perspectives on Northern Northwest Coast Prehistory, edited by J. S. Cybulski, pp. 203\u00E2\u0080\u0093222. Archaeological Survey of Canada Mercury Series. vol. 160. Canadian Museum of Civilization, Hull. Arima, Eugene Y. 1983 The West Coast People: The Nootka of Vancouver Island and Cape Flattery. Special Publication 6. British Columbia Provincial Museum, Victoria, B.C. 1988 Notes on Nootkan Sea Mammal Hunting. Arctic Anthropology 25(1):16\u00E2\u0080\u009327. Arima, Eugene Y., Henry Kammler, Terry Klokeid and Katherine Robinson (editors) 2009 The Whaling Indians : West Coast Legends and Stories: Family Origin Histories, Part 11 of the Sapir-Thomas Nootka Texts. Told by Tyee Bob, Sa:ya:ch'apis, William Qwishanishim, Lot:tisim, Tayi:?a, and Chief Louie Nookmiis and Prepared by Edward Sapir, Morris Swadesh, Hamilton George, Alexander Thomas, Frank Williams, Katies Fraser, and John Thomas. Ethnology 145. Canadian Museum of Civilization, Ottawa. Arima, Eugene Y., Terry Klokeid and Katherine Robinson (editors) 2004 The Whaling Indians. Legendary Hunters. West Coast Legends and Stories. Part 9 of the Sapir-Thomas Nootka Texts. Prepared by Edward Sapir, Morris Swadesh, Alexander Thomas, John Thomas, and Frank Williams. Told by Sa:ya:ch'apis, William, Frank Williams, Big Fred, Captain Bill, and Qwishanishim. Canadian Museum of Civilization, Hull. \u00E2\u0080\u0093 305 \u00E2\u0080\u0093 2007 The Whaling Indians : West Coast Legends and Stories: The Origin of the Wolf Ritual, Part 12 of the Sapir-Thomas Nootka Texts. Prepared by Edward Sapir, Morris Swadesh, Alexander Thomas, John Thomas, and Frank Williams. Told by Sa:ya:ch'apis and Tko:tisim. Ethnology 144. Canadian Museum of Civilization, Ottawa. Arima, Eugene Y., Terry Klokeid, Katherine Robinson, Edward Sapir, Morris Swadesh, Alexander Thomas, John Thomas and Frank Williams (editors) 2000 The Whaling Indians: West Coast Legends and Stories, Tales of Extraordinary Experience: Part 10 Sapir-Thomas Nootka Texts. Told by Tom Sa:yach\u00E2\u0080\u0099apis, William Dick La:maho:s, Captain Bill and Tyee Bob. Ethnology 134. Canadian Museum of Civilization, Hull. Arima, Eugene Y., Denis E. St. Claire, Louis Clamhouse, E. Y. Edgar and Charles Jones 1991 Between Ports Alberni and Renfrew: Notes on West Coast Peoples. Mercury Series 121. Canadian Museum of Civilization, Canadian Ethnology Service, Ottawa. Atalay, Sonya 2008 Multivocality and Indigenous Archaeologies. In Evaluating Multiple Narratives: Beyond Nationalist, Colonialist, Imperialist Archaeologies, edited by J. Habu, C. Fawcett and J. M. Matsunaga, pp. 29\u00E2\u0080\u009344. Humanities, Social Sciences and the Law. Springer, New York. Atalay, Sonya and Christine A. Hastorf 2006 Food, Meals, and Daily Activities: Food Habitus at Neolithic \u00C3\u0087atalh\u00C3\u00B6y\u00C3\u00BCk. American Antiquity 71(2):283\u00E2\u0080\u0093319. Atleo, E. Richard, Umeek 2004 Tsawalk: A Nuu-chah-nulth Worldview. University of British Columbia Press, Vancouver. 2011 Principals of Tsawalk: An Indigenous Approach to Global Crisis. University of British Columbia Press, Vancouver. Bailey, Geoff N. 2007 Time Perspectives, Palimpsests and the Archaeology of Time. Journal of Anthropological Archaeology 26(2):198\u00E2\u0080\u0093223. Baird, Robin W. 2001 Status of Harbour Seals in Canada. Canadian Field Naturalist 115(4):663\u00E2\u0080\u0093675. Bakhtin, Mikhail M. 1981 The Dialogic Imagination: Four Essays, edited by M. Holquist. Translated by C. Emerson and M. Holquist. University of Texas Press, Austin. Bal\u00C3\u00A9e, William 2006 The Research Program of Historical Ecology Annual Review of Anthropology 35:5.1\u00E2\u0080\u00935.24. Bamforth, Douglas B. and Brigid Grund 2012 Radiocarbon calibration curves, summed probability distributions, and early Paleoindian population trends in North America. Journal of Archaeological Science 39(6):1768\u00E2\u0080\u00931774. \u00E2\u0080\u0093 306 \u00E2\u0080\u0093 Banfield, A. W. F. 1974 The Mammals of Canada. National Museum of Natural Science, University of Toronto Press, Toronto. Barrett, James H., Alison M. Locker and Callum M. Roberts 2004 The Origins of Intensive Marine Fishing in Medieval Europe: The English Evidence. Proceedings of the Royal Society of London B 271:2417\u00E2\u0080\u00932421. Basso, Keith 1996 Wisdom Sits in Places: Landscape and Language Among the Western Apache. University of New Mexico Press, Albuquerque. Beaglehole, John Cawte (editor) 1967 The Voyage of the Resolution and Discovery 1776-1780, Part 1 of 2. Vol. 3 of The Journals of Captain James Cook on his Voyages of Discovery. 26. Cambridge University Press, Cambridge. Behrensmeyer, Anna K. and Susan M. Kidwell 1985 Taphonomy\u00E2\u0080\u0099s Contributions to Paleobiology. Paleobiology 11(1):105\u00E2\u0080\u0093119. Benson, Ashleen J., Gordon A. McFarlane, Susan E. Allen and John F. Dower 2002 Changes in Pacific Hake (Merluccius productus) Migration Patterns and Juvenile Growth Related to the 1989 Regime Shift. Canadian Journal of Fisheries and Aquatic Sciences 59:1969\u00E2\u0080\u00931979. Berkeley, Steven A., Colin Chapman and Susan M. Sogard 2004 Maternal Age as a Determinant of Larval Growth and Survival in a Marine Fish, Sebastes melanops. Ecology 85(5):1258\u00E2\u0080\u00931264. Berkes, Fikret and Nancy Turner 2006 Knowledge, Learning and the Evolution of Conservation Practice for Social-Ecological System Resilience. Human Ecology 34(4):479\u00E2\u0080\u0093494. Bigg, Michael A. 1985 Status of Steller sea lion (Eumetopias jubatus) and California sea lion (Zalophus californianus) in British Columbia. Canadian Special Publication of Fisheries and Aquatic Sciences 77. Pacific Biological Station, Deptartment of Fisheries and Oceans, Nanaimo. 1988 Status of the Steller sea lion, Eumetopias jubatus, in Canada. Canadian Field Naturalist 102(2):315\u00E2\u0080\u0093336. Bigg, Michael A. and Ian B. MacAskie 1978 Sea Otters Re-established in British Columbia. Journal of Mammalogy 59(4):874\u00E2\u0080\u0093876. Birkeland, Charles and Paul K. Dayton 2005 The importance in fishery management of leaving the big ones. Trends in Ecology & Evolution 20(7):356-358. Blackman, Margaret B. 1990 Haida: Traditional Culture. In Northwest Coast, edited by W. P. Suttles, pp. 240\u00E2\u0080\u0093260. Handbook of North American Indians. vol. 7. Smithsonian Institution, Washington D.C. \u00E2\u0080\u0093 307 \u00E2\u0080\u0093 Blake, Michael 2004 Fraser Valley Trade and Prestige as Seen from Scowlitz. In Complex Hunter-Gatherers; Evolution and Organization of Prehistoric Communities of the Plateau of Northwestern North America, edited by W. C. Prentiss and I. Kuijt, pp. 103\u00E2\u0080\u0093112. University of Utah Press, Salt Lake. Blake, Michael, Bruce F. Benz, Nicholas Jakobsen, Ryan Wallace, Sue Formosa, Kisha Supernant, Diana Moreiras and Alex Wong 2012 Ancient Maize Map, Version 1.1: An Online Database and Mapping Program for Studying the Archaeology of Maize in the Americas http://en.ancientmaize.com/. Laboratory of Archaeology, University of British Columbia, Vancouver. Blukis Onat, Astrida R. 1985 The Multifunctional Use of Shellfish Remains: From Garbage to Community Engineering. Northwest Anthropological Research Notes 19(1):201\u00E2\u0080\u0093207. 1987 Duwamish No. 1 Site: 1986 Data Recovery. URS Corporation and B.O.A.S., Inc., report submitted to the Department of Archaeology and Historic Preservation, Olympia. Boas, Franz 1887 The Coast Tribes of British Columbia. Science 9(216):288\u00E2\u0080\u0093289. 1932 Bella Bella Tales. 1973 ed. G. E. Stechert, New York. 1934 Geographical Names of the Kwakiutl Indians. 1969 ed. Contributions to Anthropology Vol. 20. Columbia University Press, New York. Bolster, W. Jeffrey 2012 The Mortal Sea: Fishing the Atlantic in the Age of Sail. Belknap Press of Harvard University Press, Cambridge. Bornhold, Brian D., John R. Harper, Duncan McLaren and Richard E. Thomson 2007 Destruction of the First Nations Village of Kwalate by a Rock Avalanche-generated Tsunami. Atmosphere-Ocean 45(2):123\u00E2\u0080\u0093128. Bouchard, Randy and Dorothy Kennedy 1989 The Use of Herring Roe-on-Kelp by the Central Coast People of British Columbia. Report on file, Heiltsuk Tribal Council, Bella Bella, BC. Bourdieu, Pierre 1977 Outline of a Theory of Practice. Cambridge University Press, Cambridge. 1984 Distinction: A Social Critique of the Judgement of Taste. Harvard University Press, Cambridge. Bovy, Kristine M. and Robert Kopperl 2007 Watmough Bay Site Stabilization Project 45-SJ-280 Bone Analysis Summary and Reports. Report Submitted to Bureau of Land Management, Spokane District, WA. \u00E2\u0080\u0093 308 \u00E2\u0080\u0093 Bowie, Kristina M., Jonathan B. Hall, Hayley E. Kanipe, Ryan Spady, Rebecca J. Wigen and Rena Varsakis 2010 Archaeological Monitoring and Data Recovery DdRu-81 Residential Redevelopment Agate Lane, Saanich B.C., Permit 2008-266. I. R. Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Bowie, Kristina M. and Rebecca J. Wigen 2010 Archaeological Impact Assessment DhSb-11, San Malo Crescent Parksville, B.C., Permit 2010-252. IR Wilson Consultants report submitted to the BC Archaeology Branch, Victoria. Bowman, Sheridan 1990 Radiocarbon Dating. University of California Press, Berkeley. Boxberger, Daniel L. 1999 To Fish in Common: The Ethnohistory of Lummi Indian Salmon Fishing. 2nd ed. University of Washington Press, Seattle Boyd, Robert T. 1999 The Coming of the Spirit of Pestilence: Introduced Infectious Diseases and Population Decline Among Northwest Coast Indians, 1774\u00E2\u0080\u00931874. University of Washington Press, Seattle. Braudel, Ferdinand 1970 History and the Social Sciences: The Long Term. Social Science Information 9(1):145\u00E2\u0080\u0093174. Braudel, Fernand 1980 On History. Translated by S. Matthews. University of Chicago Press, Chicago. 1981 The Structures of Everyday Life. Civilization and Capitalism, 15th-18th Century 1. 3 vols. Harper & Row, New York. Brewster, Natalie and Andrew Martindale 2011 An Archaeological History of Holocene Fish Use in the Dundas Island Group, British Columbia. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 247\u00E2\u0080\u0093264. University of Alaska Press, Fairbanks. British Columbia Archaeology Branch n.d. Remote Access to Archaeological Data Application. Accessed April 2011, Victoria BC. Brolly, Richard, Guy Cross, Jackie Holecka, Karla Kusmer, Dana Lepofsky, Heather Pratt, June Ryder and Shannon Wood 1996 Archaeological Investigations at the Beach Grove Site (DgRs-11). Tsawwassen. B.C., Permit 1995-16. Arcas Consulting Archaeologists report submitted to the BC Archaeology Branch, Victoria. Brolly, Richard and Robert J. Muir 1993 Archaeological Investigations at the Locarno Beach Site (DhRj-6), Vancouver, B.C. Arcas Consulting Archaeologists report submitted to the BC Archaeology Branch, Victoria. \u00E2\u0080\u0093 309 \u00E2\u0080\u0093 Brolly, Richard, Robert J. Muir and Rick Schulting 1993 Archaeological Investigations at the Harbour House Site (DjRu-3), Saltspring Island, B.C., Permit 1993-83. Arcas Consulting Archaeologists report submitted to the BC Archaeology Branch, Victoria. Brolly, Richard P. and Brian Pegg 1998 Archaeological Investigations In Ucluelet Traditional Territory, 1994-1997, Permit 1995-048. Arcas Consulting Archaeologists report prepared for the Ucluelet First Nation and the BC Archaeology Branch, Victoria. Bronk Ramsey, Christopher 2008 Radiocarbon Dating: Revolutions in Understanding. Archaeometry 50(2):249\u00E2\u0080\u0093275. Bronk Ramsey, Christopher and Sharen Lee 2013 Recent and Planned Developments of the Program OxCal. Radiocarbon 55(3-4). Brown, Douglas R. 2003 Shell Middens and Midden Burials in Southern Strait of Georgia Prehistory. In Archaeology of Coastal British Columbia Essays in Honour of Professor Philip M. Hobler, edited by R. L. Carlson, pp. 153\u00E2\u0080\u0093164. Archaeology Press Simon Fraser University, Burnaby. Brown, Frank and Y. Kathy Brown 2009 Staying the Course, Staying Alive \u00E2\u0080\u0093 Coastal First Nations Fundamental Truths: Biodiversity, Stewardship and Sustainability. Biodiversity BC., Victoria. Brown, Kendrick J. and Richard J. Hebda 2002 Ancient fires on southern Vancouver Island, British Columbia, Canada: A change in causal mechanisms at about 2000 ybp. Environmental Archaeology 7:1\u00E2\u0080\u009313. Bruno, John F., John J. Stachowicz and Mark D. Bertness 2003 Inclusion of Facilitation Into Ecological Theory. Trends in Ecology and Evolution 18(3):119\u00E2\u0080\u0093125. Brunsden, Jo and Diana Cooper 2008 Genoa Bay Road, Duncan, BC: An Archaeological Impact Assessment, Permit 2008-45. Millennia Research report submitted to the Hul\u00E2\u0080\u0099qumi\u00E2\u0080\u0099num Treaty Group and the BC Archaeology Branch, Victoria. Buchanan, Briggs, Mark Collard and Kevan Edinborough 2008 Paleoindian Demography and the Extraterrestrial Impact Hypothesis. Proceedings of the National Academy of Sciences, USA 105(33):11651\u00E2\u0080\u009311654. Buchanan, Briggs, Marcus Hamilton, Kevan Edinborough, Michael J. O\u00E2\u0080\u0099Brien and Mark Collard 2011 A Comment on Steele\u00E2\u0080\u0099s (2010) \u00E2\u0080\u009CRadiocarbon Dates as Data: Quantitative Strategies for Estimating Colonization Front Speeds and Event Densities\u00E2\u0080\u009D. Journal of Archaeological Science 38(9):2116\u00E2\u0080\u00932122. Budhwa, Rick 2002 Correlations Between Catastrophic Paleoenvironmental Events and Native Oral Traditions of the Pacific Northwest. MA Thesis, Simon Fraser University. \u00E2\u0080\u0093 310 \u00E2\u0080\u0093 Burchell, Meghan, Nadine Hallmann, Andrew Martindale, Aubrey Cannon and Bernd R. Sch\u00C3\u00B6ne 2013 Seasonality and Intensity of Shellfish Harvesting in the North Coast of British Columbia. Journal of Island and Coastal Archaeology 8(2):152\u00E2\u0080\u0093169. Burley, David V. 1989 Senew\u00C3\u00A9lets: Culture History of the Nanaimo Coast Salish and the False Narrows Midden. Royal British Columbia Museum Memoir 2. Royal British Columbia Museum Province of British Columbia, Ministry of Municipal Affairs Recreation and Culture, Victoria, BC. Burton, Robert K., Josh J. Snodgrass, Diane Gifford-Gonzalez, Tom Guilderson, Tom Brown and Paul L. Koch 2001 Holocene Changes in the Ecology of Northern Fur Seals: Insights from Stable Isotopes and Archaeofauna. Oecologia 128:107\u00E2\u0080\u0093115. Busch, Briton C. and Barry M. Gough 1997 Fur Traders From New England: the Boston Men in the North Pacific, 1787\u00E2\u0080\u00931800, The Narratives of William Dane Phelps, William Sturgis and James Gilchrist Swan. Arthur H. Clarke Company, Spokane. Butler, Virginia L. 1987 Fish Remains. In Duwamish No. 1 Site: 1986 Data Recovery, edited by A. R. Blukis Onat, pp. 10.1\u00E2\u0080\u009310.37. URS Corporation and B.O.A.S., Inc. report submitted to the Department of Archaeology and Historic Preservation, Olympia. Butler, Virginia L. and Greg Baker 2002 Appendix 2. Fish Remains From Bay Street Shell Midden (45-KP-115) In Data Recovery Excavations the Bay Street Shell Midden (45-KP-115), Kitsap County, WA, edited by D. E. Lewarch, pp. 201\u00E2\u0080\u0093222. Larson Anthropological Archaeological Services Limited report submitted to the Department of Archaeology and Historic Preservation, Olympia. Butler, Virginia L. and Sarah K. Campbell 2004 Resource Intensification and Resource Depression in the Pacific Northwest of North America: A Zooarchaeological Review. Journal of World Prehistory 18(4):327\u00E2\u0080\u0093405. Butler, Virginia L. and James C. Chatters 1994 The Role of Bone Density in Structuring Prehistoric Salmon Bone Assemblages. Journal of Archaeological Science 21(3):413\u00E2\u0080\u0093424. Caldwell, Megan 2008 A View From the Shore: Interpreting Fish Trap Use in Comox Harbour Through Zooarchaeological Analysis of Fish Remains from the Q\u00E2\u0080\u0099umu?xs Village Site (DkSf-19), Comox Harbour, British Columbia. MA Thesis, University of Manitoba. 2013 PhD Dissertation in Progress, University of Alberta. Caldwell, Megan, Dana Lepofsky, Georgia Combes, Michelle Washington, John R. Welch and John R. Harper 2012 A Bird\u00E2\u0080\u0099s Eye View of Northern Coast Salish Intertidal Resource Management Features. Journal of Island and Coastal Archaeology 7(2):219\u00E2\u0080\u0093233. \u00E2\u0080\u0093 311 \u00E2\u0080\u0093 Calvert, S. Gay 1980 A Cultural Analysis of Faunal Remains from Three Archaeological Sites in Hesquiat Harbor, B.C. Ph.D. dissertation, University of British Columbia. Calvert, S. Gay and Susan J. Crockford 1982 Appendix IV: Analysis of Faunal Remains. In Alberni Prehistory: Archaeological and Ethnographic Investigations on Western Vancouver Island, edited by A. D. McMillan and D. E. St. Claire, pp. 174\u00E2\u0080\u0093219. Theytus Books and Alberni Valley Museum, Penticton and Port Alberni. Campbell, Sarah K. and Virginia L. Butler 2010 Fishes and Loaves: Explaining Sustainable, Long-term Animal Harvesting on the Northwest Coast Using the 'Plant Paradigm'. In The Archaeology of Anthropogenic Environments, edited by R. M. Dean, pp. 175\u00E2\u0080\u0093203. Occasional Paper. vol. 37. Center for Archaeological Investigations, Southern Illinois University, Carbondale. Cannon, Aubrey 1991 The Economic Prehistory of Namu: Patterns in Vertebrate Fauna Publication 19. Department of Archaeology, Simon Fraser University, Burnaby. 1996 Scales of Variability in Northwest Salmon Fishing. In Prehistoric Hunter\u00E2\u0080\u0093Gatherer Fishing Strategies, edited by M. G. Plew, pp. 25\u00E2\u0080\u009340. Monograph Series. Boise State University, Boise. 1997 Permit Report - 1996-115 Investigations at EISx-3, ElSx-6, EISx-8, EISx-10, EISx-16, ElTa-3, EITa-25, and EkSx-12. Report Submitted to the Archaeology Branch, Victoria BC. 1998 Permit Report - 1997-163, Investigations at EISx-4, EISx-5, EISx-6, EISx-8, EISx-10, EISx-18, ElTa-18, EITa-21, EITb-1, and EITb-2. Report Submitted to the Archaeology Branch, Victoria BC. 2000a Assessing Variability in Northwest Coast Salmon and Herring Fisheries: Bucket-Auger Sampling of Shell Midden Sites on the Central Coast of British Columbia. Journal of Archaeological Science 27(8):725\u00E2\u0080\u0093737. 2000b Settlement and Sea Levels on the Central Coast of British Columbia: Evidence from Shell Midden Cores. American Antiquity 65(1):67\u00E2\u0080\u009377. 2001 Was Salmon Important in Northwest Coast Prehistory? In People and Wildlife in Northern North America: Essays in Honor of R. Dale Guthrie, edited by S. C. Gerlach and M. S. Murray, pp. 178\u00E2\u0080\u0093187. International Series. vol. S944. British Archaeological Reports, Oxford, UK. 2002 Sacred Power and Seasonal Settlement on the Central Northwest Coast. In Beyond Foraging and Collecting: Evolutionary Change in Hunter-Gatherer Settlement Systems, edited by B. Fitzhugh and J. Habu, pp. 311\u00E2\u0080\u0093338. Kluwer-Plenum Press, New York. 2003 Long-term Continuity in Central Northwest Coast Settlement Patterns. In Archaeology of Coastal British Columbia: Essays in Honour of Professor Philip M. Hobler, edited by R. L. Carlson, pp. 1\u00E2\u0080\u009312. Archaeology Press, Simon Fraser Univeristy, Burnaby, B.C. \u00E2\u0080\u0093 312 \u00E2\u0080\u0093 2011a Cosmology and Everyday Perception in Northwest Coast Production, Reproduction, and Settlement. In Structured Worlds: The Archaeology of Hunter-Gatherer Thought and Action, edited by A. Cannon, pp. 54\u00E2\u0080\u009368. Equinox Publishing, London. 2011b Introduction. In Structured Worlds: The Archaeology of Hunter-Gatherer Thought and Action, edited by A. Cannon, pp. 1\u00E2\u0080\u009310. Equinox Publishing, London. Cannon, Aubrey and Meghan Burchell 2009 Clam Growth-stage Profiles as a Measure of Harvest Intensity and Resource Management on the Central Coast of British Columbia. Journal of Archaeological Science 36(4):1050\u00E2\u0080\u00931060. Cannon, Aubrey, Meghan Burchell and Rhonda Bathurst 2008 Trends and Strategies in Shellfish Gathering on the Pacific Northwest Coast of North America. In Early Human Impact on Megamolluscs, edited by A. Antczak and R. Cipriani, pp. 7\u00E2\u0080\u009322. International Series. vol. 1865. British Archaeological Reports, Archaeopress, Oxford. Cannon, Aubrey and Dongya Y. Yang 2006 Early Storage and Sedentism on the Pacific Northwest Coast: Ancient DNA Analysis of Salmon Remains from Namu, British Columbia. American Antiquity 71(1):123\u00E2\u0080\u0093140. Cannon, Aubrey, Dongya Y. Yang and Camilla F. Speller 2011 Site-Specific Salmon Fisheries on the Central Coast of British Columbia. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 117\u00E2\u0080\u0093148. University of Alaska Press, Fairbanks. Cardinal, Nathan, Daryl W. Fedje and Nicole F. Smith 2011 Evidence for Clam Gardens in the National Park Reserves in Coastal British Columbia, Paper presented at the International Marine Conservation Congress, Victoria, BC. Symposium Organized by Thomas Thornton, Oxford University. Carlson, Catherine C. 2003 The Bear Cove Fauna and the Subsistence History of Northwest Coast Maritime Culture. In Archaeology of Coastal British Columbia: Essays in Honour of Professor Philip M. Hobler, edited by R. L. Carlson, pp. 65\u00E2\u0080\u009386. Archaeology Press, Simon Fraser University, Burnaby, B.C. Carlson, Roy L. 1991 Appendix B - Namu Periodization and C-14 Chronology. In The Economic Prehistory of Namu, edited by A. Cannon, pp. 85\u00E2\u0080\u009395. Publication 19. Archaeology Press, Simon Fraser University, Burnaby. Carrothers, W. A. 1941 The British Columbia Fisheries. University of Toronto Press. Casey, Edward S. 1993 Getting Back Into Place: Toward a Renewed Understanding of the Place-world. Indiana University Press, Bloomington. \u00E2\u0080\u0093 313 \u00E2\u0080\u0093 Casteel, Richard W. 1972 Some Biases in the Recovery of archaeological Faunal Remains. Proceedings of the Prehistoric Society 38:382\u00E2\u0080\u0093388. 1974 A Method for Estimation of Live Weight of Fish from the Size of Skeletal Elements. American Antiquity 39(1):94\u00E2\u0080\u009398. 1976a Comparison of Column and Whole Unit Samples for Recovering Fish Remains. World Archaeology 8(2):192\u00E2\u0080\u0093196. 1976b Fish Remains From Glenrose. In The Glenrose Cannery Site, edited by R. G. Matson, pp. 82\u00E2\u0080\u009387. Mercury Series 52. Archaeological Survey of Canada, National Museum of Man, Ottawa. Cavanagh, Deborah Mary 1983 Northwest Coast Whaling: A New Perspective. M.A. Thesis, University of British Columbia. Chisholm, Brian S. 1986 Reconstruction of Prehistoric Diet in British Columbia Using Stable-Carbon Isotopic Analysis. Ph.D. Dissertation, Simon Fraser University. Chisholm, Brian S., D. Earle Nelson and Henry P. Schwarcz 1982 Stable Carbon Isotope Ratios as a Measure of Marine Versus Terrestrial Protein in Ancient Diets. Science 216(4550):1131\u00E2\u0080\u00931132. 1983 Marine and Terrestrial Protein in Prehistoric Diets on the British Columbia Coast. Current Anthropology 24(3):396\u00E2\u0080\u0093398. Christensen, Tina and Jim Stafford 2005 Raised Beach Archaeology in Northern Haida Gwaii: Preliminary Results from the Cohoe Creek Site. In Haida Gwaii: Human History and Environment from the Time of Loon to the Time of the Iron People, edited by D. W. Fedje and R. W. Mathewes, pp. 245\u00E2\u0080\u0093273. UBC Press, Vancouver. Christensen, Tina, Jim Stafford and Iain McKechnie 2010 Masset Health Centre Site GaUa-18, Haida Gwaii: Excavation and Monitoring Report, Permit 2005-408. Coast Interior Archaeology report submitted to the Council of the Haida Nation and the BC Archaeology Branch, Victoria. Clayton, Daniel W. 2000 Islands of Truth: The Imperial Fashioning of Vancouver Island. UBC Press, Vancouver. Cleary, Jaclyn S., Sean P. Cox and Jacob F. Schweigert 2010 Performance Evaluation of Harvest Control Rules for Pacific Herring Management in British Columbia, Canada. ICES: Journal of Marine Science 67:2005\u00E2\u0080\u00932011. Collard, Mark, Kevan Edinborough, Stephen Shennan and Mark G. Thomas 2010 Radiocarbon Evidence Indicates that Migrants Introduced Farming to Britain. Journal of Archaeological Science 37(4):866\u00E2\u0080\u0093870. \u00E2\u0080\u0093 314 \u00E2\u0080\u0093 Cooper, Diana 2008 Archaeological Investigations at DdRu-4, Lochside Drive, North Saanich, Permit 2006-343. Millennia Research Limited Report submitted to Tsawout First Nation, Tseycum First Nation, SEN\u00C4\u0086O\u00C5\u00A6EN Alliance, and the BC Archaeology Branch, Victoria. Cooper, H. Kory 2007 The Anthropology of Native Copper Technology and Social Complexity in Alaska and the Yukon Territory: An Analysis Using Archaeology, Archaeometry, and Ethnohistory. PhD Dissertation, University of Alberta. 2012 Innovation and Prestige Among Northern Hunter-Gatherers: Late Prehistoric Native Copper Use in Alaska and Yukon. American Antiquity 77(3):565\u00E2\u0080\u0093590. Cooper, H. Kory, M. John M. Duke, Antonio Simonetti and Guang Cheng Chen 2008 Trace element and Pb isotope provenance analyses of native copper in northwestern North America: results of a recent pilot study using INAA, ICP-MS, and LA-MC-ICP-MS Journal of Archaeological Science 35:1732\u00E2\u0080\u00931747. COSEWIC 2003 COSEWIC assessment and update status report on the Steller sea lion Eumetopias jubatus in Canada, pp. 47. Committee on the Status of Endangered Wildlife in Canada, Ottawa. Coupland, Gary 1989 Archaeological Investigations at the Point Grey Site, DhRt-5 Permit 1988-41. Report prepared for the BC Archaeology Branch, Victoria. 1991 The Point Grey Site: A Marpole Spring Village Component. Canadian Journal of Archaeology 15(1):73\u00E2\u0080\u009396. Coupland, Gary, Craig Bissell and Sarah King 1993 Prehistoric Subsistence and Seasonality at Prince Rupert: Evidence from the McNichol Creek Site. Canadian Journal of Archaeology 17(1):59\u00E2\u0080\u009377. Coupland, Gary, Kathlyn M. Stewart and Katherine Patton 2010 Do You Never Get Tired of Salmon? Evidence for Extreme Salmon Specialization at Prince Rupert Harbor, British Columbia. Journal of Anthropological Archaeology 29(2):189\u00E2\u0080\u0093207. Coupland, Gary, Kathlyn M. Stewart, Katherine Patton and Joan Banahan 2006 Archaeological Investigations in the Prince Rupert Area 2002-2004: Permit 2002-208. Final Report on file at the BC Archaeology Branch. Crockford, Carin Elizabeth 1996 Nuu-chah-nulth Labour Relations in the Pelagic Sealing Industry, 1868\u00E2\u0080\u00931911. MA Thesis, University of Victoria. Crockford, Susan J. 1997 Osteometry of Makah and Coast Salish Dogs. Archaeology Press, Simon Fraser University, Burnaby. \u00E2\u0080\u0093 315 \u00E2\u0080\u0093 Crockford, Susan J., S. Gay Frederick and Rebecca J. Wigen 2002 The Cape Flattery Fur Seal: An Extinct Species of Callorhinus in the Eastern North Pacific? Canadian Journal of Archaeology 26(2):152\u00E2\u0080\u0093174. Croes, Dale R. 2003 Northwest Coast Wet-site Artifacts: A Key to Understanding Resource Procurement, Storage, Management, and Exchange. In Emerging from the Mist: Studies in Northwest Coast Culture History, edited by R. G. Matson, G. Coupland and Q. Mackie, pp. 51\u00E2\u0080\u009375. UBC Press, Vancouver, BC. Croes, Dale R. and Steven Hackenberger 1988 Hoko River Archaeological Complex: Modeling Prehistoric Northwest Coast Economic Evolution. In Prehistoric Economies of the Pacific Northwest Coast, edited by B. L. Issac, pp. 19\u00E2\u0080\u009385. Research in Economic Anthropology. vol. 3. JAI Press, Greenwich. Crowell, Aron L. and Wayne K. Howell 2013 Time, Oral Tradition, and Archaeology at Xakwnoow\u00C3\u00BA, a Little Ice Age Fort in Southeastern Alaska. American Antiquity 78(1):3\u00E2\u0080\u009323. Cruikshank, Julie 1997 Negotiating with Narrative Establishing Cultural Identity at the Yukon International Storytelling Festival. American Anthropologist 99(1):56\u00E2\u0080\u009369. 2001 Glaciers and Climate Change: Perspectives from Oral Tradition. Arctic 54(4):377\u00E2\u0080\u0093393. 2005 Do Glaciers Listen? Local Knowledge, Colonial Encounters, and Social Imagination. University of British Columbia Press, Vancouver. Cui, Yinqiu, John Lindo, Cris E. Hughes, Jesse W. Johnson, Alvaro G. Hernandez, Brian M. Kemp, Jian Ma, Ryan Cunningham, Barbara Petzelt, Joycellyn Mitchell, David Archer, Jerome S. Cybulski and Ripan S. Malhi 2013 Ancient DNA Analysis of Mid-Holocene Individuals from the Northwest Coast of North America Reveals Different Evolutionary Paths for Mitogenomes. PLoS ONE 8(7):e66948. Curtin, A. Joanne 1991 Gabriola Island Burial Recovery Project: Report on the 1989 Field Season, Permit 1989-92. Prepared for the Archaeology Branch, Ministry of Municipal Affairs Recreation and Culture and the British Columbia Heritage Trust, Victoria. Curtis, Edward S. 1916 The North American Indian: Nootka and Haida. 1970 ed 11. Johnson Reprint Corporation, New York. Cybulksi, Jerome S. 2001 Entry to the Canadian Archaeological Radiocarbon Database: http://card.canadianarchaeology.ca/localc14/detail.php?id=5293 and http://card.canadianarchaeology.ca/localc14/detail.php?id=5292. Canadian Museum of Civilization with support of the Canadian Archaeological Association, Gatineau. \u00E2\u0080\u0093 316 \u00E2\u0080\u0093 Cybulski, Jerome S. 1975 Skeletal Variability in British Columbia Coastal Populations: a Descriptive and Comparative Assessment of Cranial Morphology. Mercury Series Paper 30. Archaeological Survey of Canada and the Canadian Museum of Civilization, Hull. 1990 Human Biology. In Northwest Coast, edited by W. P. Suttles, pp. 52\u00E2\u0080\u009359. Handbook of North American Indians. vol. 7. Smithsonian Institution, Washington, D.C. 1992 A Greenville Burial Ground: Human Remains and Mortuary Elements in British Columbia Coast Prehistory. Mercury Series Paper 146. Archaeological Survey of Canada, Canadian Museum of Civilization, Ottawa. 1994 Culture Change, Demographic History, and Health and Disease on the Northwest Coast. In In the Wake of Contact: Biological Responses to Conquest, edited by C. S. Larsen and G. R. Milner, pp. 75\u00E2\u0080\u009385. Wiley-Liss, New York. 1996 Context of Human Remains from the Lachane Site, GbTo-33, Including Evidence for Decapitation. Manuscript MS97-3973 on file, Library Documents Secion (Archaeology) Canadian Museum of Civilization, Hull. 1999 Trauma and Warfare at Prince Rupert Harbour. The Midden 31(2):5\u00E2\u0080\u00937. 2006 Skeletal Biology: Northwest Coast and Plateau. In Environment, Origins, and Population, edited by D. H. Ubelaker, pp. 532\u00E2\u0080\u0093547. Handbook of North American Indians. vol. 3. Smithsonian Institution, Washington D.C. 2010 Labrets And Teeth On The Northwest Coast. In The Crescent Beach Site and the Place of the Locarno Beach Phase, edited by R. G. Matson. vol. http://hdl.handle.net/2429/39218. University of British Columbia, Vancouver. Dallimore, Audrey, Randolph J. Enkin, Reinhard Peinitz, John R. Southon, Judith Baker, Cynthia A. Wright, Tom F. Pedersen, Steve E. Calvert and Richard E. Thomson 2008 Post-glacial evolution of a Pacific Coastal Fjord in British Columbia, Canada: Interactions of Sea-level Change, Crustal Response and Environmental Fluctuations; Results from MONA Core MD02-2494. Canadian Journal of Earth Sciences 45(11):1345\u00E2\u0080\u00931362. Darnell, Regna 1990 Edward Sapir: Linguist, Anthropologist, Humanist. University of California Press, Berkeley. 2000 The Pivotal Role of the Northwest Coast in the History of Americanist Anthropology. BC Studies (125/126):33\u00E2\u0080\u009352. de Laguna, Frederica 1960 The Story of a Tlingit Community: A Problem in the Relationship between Archaeological, Ethnological, and Historical Methods. Bulletin 172. Bureau of American Ethnology, United States Government Printing Office, Washington DC. 1972 Under Mount Saint Elias: The History and Culture of the Yakutat Tlingit: Part One. Smithsonian Contributions to Anthropology 7.1. Smithsonian Institution, Washington DC. \u00E2\u0080\u0093 317 \u00E2\u0080\u0093 de Laguna, Frederica, Francis A. Riddell, Donald F. McGeein, Kenneth S.. Lane and J. Arthur Freed 1964 Archeology of the Yakutat Bay area, Alaska. Bureau of American Ethnology Bulletin 192. Smithsonian Institution, Washington DC. Deo, Jennie N., John O. Stone and Julie K. Stein 2004 Building Confidence in Shell: Variations in the Marine Radiocarbon Reservoir Correction for the Northwest Coast over the past 3,000 Years. American Antiquity 69(4):771-786. Department of Fisheries and Oceans Canada 1999 Petrale Sole British Columbia (Area 3C-5D). DFO Science Stock Status Repor A6-06 (1999). 2002 Northern Anchovy. Can. Sci. Advis. Stock Status Rep. B6-08. Department of Fisheries and Oceans, Naniamo. 2007 Population Assessment for Northern Fur Seals (Callorhinus ursinus). In Science Advisory Report 2007/52. Canadian Science Advisory Secretariat, Ottawa. 2010 Population Assessment Pacific Harbour Seal (Phoca vitulina richardsi). In Science Advisory Report 2009/011. Canadian Science Advisory Secretariat, Ottawa. 2012a Herring Spawning and Catch Records http://www.pac.dfo-mpo.gc.ca/science/species-especes/pelagic-pelagique/herring-hareng/herspawn/pages/default0-eng.html. vol. Accessed December 29th, 2012, Ottawa. 2012b Stock Assessment on Pacific Herring in British Columbia. Stock Status Report 2012/062. 2013 Pacific Region Integrated Fisheries Management Plan for Pacific Herring, Clupea Pallasii, November 7, 2012 To November 6, 2013 Amended February 19, 2013 edited by Fisheries Management Branch. Department of Fisheries & Oceans Canada, Ottawa. Deur, Douglas 2005 Tending the Garden, Making the Soil: Northwest Coast Estuarine Gardens as Engineered Enrivonments. In Keeping It Living: Traditions of Plant Use and Cultivation on the Northwest Coast of North America, edited by D. Deur and N. J. Turner, pp. 296\u00E2\u0080\u0093327. University of Washington Press, Seattle. Dewhirst, John 1980 The Indigenous Archaeology of Yuquot, A Nootkan Outside Village: The Yuquot Project, Volume 1, 39. National Historic Parks and Sites Branch, History and Archaeology, Ottawa. Dick, Lyle 2006 The Maritime Fur Trade in Southern Haida Gwaii (Queen Charlotte Islands), ca. 1787\u00E2\u0080\u00931920. Western and Northern Service Centre, Parks Canada, Vancouver, British Columbia, Canada. \u00E2\u0080\u0093 318 \u00E2\u0080\u0093 Dietler, Michael and Brian Hayden 2001 Digesting the Feast: Good to Eat, Good to Drink, Good to Think: An Introduction. In Feasts: Archaeological and Ethnographic Perspectives on Food, Politics, and Power, edited by M. Dietler and B. Hayden, pp. .1-20. Smithsonian Institution Press, Washington D.C. Dixon, E. James, Timothy H. Heaton, Terence E. Fifield, Thomas D. Hamilton, David E. Putnam and Frederick Grady 1997 Late Quaternary Regional Geoarchaeology of Southeast Alaska Karst: A Progress Report. Geoarchaeology 12(6):689\u00E2\u0080\u0093712. Driver, Jonathan C. 1993 Zooarchaeology in British Columbia. BC Studies 99:77\u00E2\u0080\u0093104. Drucker, Philip 1951 The Northern and Central Nootkan Tribes. Bureau of American Ethnology Bulletin 144. Smithsonian Institution, Washington D.C. Eighmy, Jeffrey L. and Jason M. LaBelle 1996 Radiocarbon Dating of Twenty-seven Plains Complexes and Phases. Plains Anthropologist 41(1):53\u00E2\u0080\u009369. Eldridge, Morley 1987 Mitigative Excavation at DiSe-10, Denman Island, B.C. Rockshelters and Blufftop Hunting Magic, Permit 1986-13. Eldridge Heritage Consulting report prepared for the BC Archaeology Branch, Victoria. 2000 Eagles Nest Archaeological Data Recovery DcRt-71, Permit 1999-289. Millennia Research Limited report prepared for International Integration Inc. and submitted to the Songhees First Nation and the BC Archaeology Branch, Victoria. Eldridge, Morley, Jo Brunsden, Alyssa Parker and Iain McKechnie 2008 Permit 2007-230: Fairview Container Terminal Phase II Archaeological Impact Assessment. Millennia Research Limited report prepared for the Prince Rupert Port Authority and the BC Archaeology Branch, Victoria. Eldridge, Morley, Diana Cooper and D'Ann Owens 2007 Nanaimo Foundry Data Recovery and Monitoring, Permits 2005-417 and 2006-205. Millennia Research Limited report submitted to the BC Archaeology Branch, Victoria. Elliott, Henry Wood 1886 Our Arctic Province: Alaska and the Seal Islands. 1897 ed. Library of Contemporary Exploration and Adventure. Charles Scribner's Sons, New York. Ellis, David W. and Luke Swan 1981 Teachings of the Tides: Uses of Marine Invertebrates by the Manhousaht People. Theytus Books, Nanaimo, BC. Elmendorf, William W. 1960 Twana Narratives: Native Historical Accounts of a Coast Salish Culture. University of Washington Press, Seattle. \u00E2\u0080\u0093 319 \u00E2\u0080\u0093 Environment Canada 2009 Canadian Climate Normals or Averages 1971-2000. In National Climate Data and Information Archive, http://www.climate.weatheroffice.ec.gc.ca/climate_normals/index_e.html. Government of Canada, Ottawa. Erlandson, Jon M. 2001 The Archaeology of Aquatic Adaptations: Paradigms for a New Millennium. Journal of Archaeological Research 9(4):287-350. Erlandson, Jon M. and Madonna L. Moss 2001 Shellfish Feeders, Carrion Eaters, and the Archaeology of Aquatic Adaptations. American Antiquity 66(3):413\u00E2\u0080\u0093432. Ersek, Vasile, Peter U. Clark, Alan C. Mix, Hai Cheng and R. Lawrence Edwards 2012 Holocene Winter Climate Variability in Mid-latitude Western North America. Nature Communications 3(1219). Estes, James A. 1999 Sea otter, Enhydra lutris. In Smithsonian Book of North American Mammals, edited by D. E. Wilson and S. Ruff, pp. 180\u00E2\u0080\u0093183. UBC Press and Smithsonian Institution Press, Vancouver and Washington, D.C. Etnier, Michael A. 2002a The Effects of Human Hunting on Northern Fur Seal (Callorhinus ursinus) Migration and Breeding Distributions in the Late Holocene. Ph.D. Dissertation, University of Washington. 2002b Occurrences of Guadalupe Fur Seals (Arctocephalus townsendi) on the Washington Coast Over the Past 500 Years. Marine Mammal Science 18(2):551\u00E2\u0080\u0093557. 2007 Defining and Identifying Sustainable Harvests of Resources: Archaeological Examples of Pinniped Harvests in the Eastern North Pacific. Journal for Nature Conservation 15(3):196\u00E2\u0080\u0093207. Ewonus, Paul A. 2011 Social Zooarchaeology of a Northwest Coast House. Journal of Island and Coastal Archaeology 6(1):72\u00E2\u0080\u009397. Fedje, Daryl W., Jenny Cohen, Robert G. Commisso and Nicole F. Smith 2011a 2010/11 Gwaii Haanas Archaeology Programme. Report prepared by Cultural Resource Services Western & Northern Canada Service Centre and submitted to the Parks Canada Agency and the Archipelago Management Board, Skidegate. Fedje, Daryl W., Alexander P. Mackie, Rebecca J. Wigen, Quentin Mackie and Cynthia Lake 2005a Kilgii Gwaay: An Early Maritime Site in the South of Haida Gwaii. In Haida Gwaii: Human History and Environment from the Time of Loon to the Time of the Iron People, edited by D. W. Fedje and R. W. Mathewes, pp. 187\u00E2\u0080\u0093203. UBC Press, Vancouver. \u00E2\u0080\u0093 320 \u00E2\u0080\u0093 Fedje, Daryl W., Quentin Mackie, E. James Dixon and Timothy H. Heaton 2004 Late Wisconsin Environments and Archaeological Visibility on the Northern Northwest Coast. In Entering America: Northeast Asia and Beringia before the Last Glacial Maximum, edited by D. B. Madsen, pp. 97\u00E2\u0080\u0093138. University of Utah Press, Salt Lake City. Fedje, Daryl W., Quentin Mackie, Nicole F. Smith and Duncan McLaren 2011b Function, Visibility, and Interpretation of Archaeological Assemblages at the Pleistocene-Holocene Transition in Haida Gwaii. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp. 323\u00E2\u0080\u0093342. Texas A&M University Press, College Station. Fedje, Daryl W., Martin P. Magne and Tina Christensen 2005b Test Excavations at Raised Beach Sites in Southern Haida Gwaii and Their Significance to Northwest Coast Archaeology. In Haida Gwaii: Human History and Environment from the Time of Loon to the Time of the Iron People, edited by D. W. Fedje and R. W. Mathewes, pp. 204\u00E2\u0080\u0093243. UBC Press, Vancouver. Fedje, Daryl W. and Rolf W. Mathewes (editors) 2005 Haida Gwaii: Human History and Environment from the Time of Loon to the Time of the Iron People. UBC Press, Vancouver. Fedje, Daryl W. and Nicole F. Smith 2009 2008/09 Gwaii Haannas Archaeology and Paleoecology. Report on file, Gwaii Haanas National Park Reserve and Parks Canada Agency Northern and Western Service Centre, Skidegate and Calgary. Fedje, Daryl W., Ian D. Sumpter and John R. Southon 2009 Sea-levels and Archaeology in the Gulf Islands National Park Reserve. Canadian Journal of Archaeology 33(2):234\u00E2\u0080\u0093253. Fedje, Daryl W., Rebecca J. Wigen, Quentin Mackie, Cynthia Lake and Ian D. Sumpter 2001 Preliminary Results from Investigations at Kilgii Gwaay: An Early Archaeological Site on Ellen Island, Haida Gwaii, British Columbia. Canadian Journal of Archaeology 25(1-2):98\u00E2\u0080\u0093120. Finney, Bruce P., Irene Gregory-Eaves, Marianne S. V. Douglas and John P. Smol 2002 Fisheries Productivity in the Northeastern Pacific Ocean over the Past 2,200 Years. Nature 416:729\u00E2\u0080\u0093732. Fisher, Robin 1977 Contact and Conflict: Indian-European Relations in British Columbia, 1774\u00E2\u0080\u00931890. UBC Press, Vancouver. Fladmark, Knut R. 1975 A Paleoecological Model for Northwest Coast Prehistory. Mercury Series 43. National Museums of Canada, Ottawa. \u00E2\u0080\u0093 321 \u00E2\u0080\u0093 Fladmark, Knut R., Kenneth M. Ames and Patricia D. Sutherland 1990 Prehistory of the Northern Coast of British Columbia. In Northwest Coast, edited by W. Suttles, pp. 229\u00E2\u0080\u0093239. Handbook of North American Indians. vol. 7. Smithsonian Institution Press, Washington, DC. Fogelson, Raymond D. 1989 The Ethnohistory of Events and Nonevents. Ethnohistory 36(2):133\u00E2\u0080\u0093147. Ford, Pamela J. 1989 Archaeological and Ethnographic Correlates of Seasonality: Problems and Solutions on the Northwest Coast. Canadian Journal of Archaeology 13:133-150. Frederick, Gay 2012 Vertebrate Fauna from the Huu-ay-aht Archaeology Project: Results from the 2006 Huu7ii Village Excavations and Summary of 2004 and 2006 Data. In Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound, edited by A. D. McMillan and D. E. St. Claire, pp. 115\u00E2\u0080\u0093153. Archaeology Press, Simon Fraser University, Burnaby, BC. Frederick, Gay and Susan J. Crockford 2005 Appendix D: Analysis of the Vertebrate Fauna from Ts'ishaa village, DfSi 16, Benson Island. In Ts'ishaa: Archaeology and Ethnography of a Nuu-chah-nulth Origin Site in Barkley Sound, pp. 173\u00E2\u0080\u0093205. by A. D. McMillan and D. E. St. Claire Archaeology Press, Simon Fraser University, Burnaby. Frederick, Gay, Iain McKechnie and Rebecca J. Wigen 2006a Faunal Report for the Huu7ii Archaeology Project: Results from the 2004 Huu7ii Village Excavations. Final report submitted to the Huu-ay-aht First Nation (Anacla, BC), Denis St. Claire (Victoria, BC), and Alan McMillan (Port Moody, BC) October 26, 2006. 2006b Interim Faunal Report for the Huu7ii Archaeology Project: Preliminary Results from the 2004 Huu7ii village excavations. Report submitted to the Huu-ay-aht First Nation (Anacla, BC), Denis St. Claire (Victoria, BC), and Alan McMillan (Port Moody, BC), June 30, 2006. Friedman, Edward 1976 An Archaeological Survey of Makah Territory: a Study of Resource Utilization. Ph.D. Dissertation, Washington State University. Friele, Pierre A. and Ian Hutchinson 1993 Holocene Sea-Level Change and the Central West Coast of Vancouver Island, British Columbia. Canadian Journal of Earth Sciences 30(4):832\u00E2\u0080\u0093840. Funk, Fritz 2010 History of Commercial Herring Fisheries in Southeast Alaska. In Herring Synthesis: Documenting and Modeling Herring Spawning Areas with Socio-ecological Systems over Time in the Southeastern Gulf of Alaska, North Pacific Research Board Project #728, pp. 234\u00E2\u0080\u0093294. \u00E2\u0080\u0093 322 \u00E2\u0080\u0093 Gajewski, Konrad, Samuel E. Munoz, Matthew C. Peros, Andre E. Viau, Richard E. Morlan and Matthew W. Betts 2011 The Canadian Archaeological Radiocarbon Database (CARD): Archaeological 14C Dates in North America and Their Paleoenvironmental Context. Radiocarbon 53(2):371\u00E2\u0080\u0093394. Galdikas-Brindamour, B. 1972 Faunal material from eight archaeological sites: a preliminary report. In Salvage '71: Reports on Salvage Archaeology Undertaken in British Columbia in 1971, edited by R. L. Carlson, pp. 199\u00E2\u0080\u0093209. Department of Archaeology, Simon Fraser University, Burnaby. Galois, Robert M. 2004 A Voyage to the North West Side of America: The Journals of James Colnett, 1786\u00E2\u0080\u009389. UBC Press, Vancouver. Garibaldi, Ann and Nancy J. Turner 2004 Cultural Keystone Species: Implications for Ecological Conservation and Restoration. Ecology and Society 9(3):1\u00E2\u0080\u009318. Garrow, Duncan 2012 Odd Deposits and Average Practice. A Critical History of the Concept of Structured Deposition. Cambridge Archaeological Journal 19 (2):85\u00E2\u0080\u0093115. Gay, James Thomas 1987 American Fur Seal Diplomacy: The Alaskan Fur Seal Controversy. Peter Lang, New York. Geib, Phil R. and Winston B. Hurst 2013 Should Dates Trump Context? Evaluation of the Cave 7 Skeletal Assemblage Radiocarbon Dates. Journal of Archaeological Science 40(6):2754\u00E2\u0080\u00932770. George, Chief Earl Maquinna 2003 Living on the Edge: Nuu-chah-nulth History from an Ahousaht Chief's Perspective. Sono Nis Press, Winlaw, British Columbia. Gero, Joan M. 2007 Honoring Ambiguity/Problematizing Certitude. Journal of Archaeological Method and Theory 14(3):311\u00E2\u0080\u0093327. Gerrard, Anna, Murray Rutherford and Anne K. Salomon n.d. Observed Herring Spawning Activity in Heiltsuk Territory. Conservation Letters. Giddens, Anthony 1984 The Constitution of Society. University of California Press, Berkeley. Gifford, Diane P. 1981 Taphonomy and Paleoecology: A Critical Review of Archaeology's Sister Disciplines. In Advances in Archaeological Method and Theory, edited by M. B. Schiffer, pp. 365\u00E2\u0080\u0093438. vol. 4. Academic Press, Orlando. \u00E2\u0080\u0093 323 \u00E2\u0080\u0093 Gifford-Gonzalez, Diane P. 2008 Thoughts on a Method for Zooarchaeological Study of Quotidian Life. In Engendering Social Dynamics: The Archaeology of Maintenance Activities, edited by S. Mont\u00C3\u00B3n-Sub\u00C3\u00ADas and M. S\u00C3\u00A1nchez-Romero, pp. 15\u00E2\u0080\u009323. vol. 1862. British Archaeological Reports, Oxford. 2011 Holocene Monterey Bay Fur Seals: Distribution, Dates, and Ecological Implications. In Human Impacts on Seals, Sea Lions, and Sea Otters: Integrating Archaeology and Ecology in the Northeast Pacific, edited by T. J. Braje and T. C. RIck, pp. 221\u00E2\u0080\u0093242. University of California Press, Berkeley. Gifford-Gonzalez, Diane P., Seth D. Newsome, Paul L. Koch, Tomas P. Guilderson, Josh J. Snodgrass and Robert K. Burton 2005 Archaeofaunal Insights on Pinniped-Human Interactions in the Northeastern Pacific. In The Exploitation and Cultural Importance of Sea Mammals, edited by G. G. Monks, pp. 19\u00E2\u0080\u009338. Oxbow Books, Oxford. Giovas, Christina M. 2009 The Shell Game: Analytic Problems in Archaeological Mollusc Quantification. Journal of Archaeological Science 36(7):1557\u00E2\u0080\u00931564. Glassow, Michael A. 2000 Weighing vs. Counting Shellfish Remains: A Comment on Mason, Peterson, and Tiffany. American Antiquity 65(2):407\u00E2\u0080\u0093414. Godwin, H. 1962 Half-life of Radiocarbon. Nature 195(4845):984. Golla, Susan 1987 He Has a Name: History and Social Structure Among the Indians of Western Vancouver Island. PhD Dissertation, Columbia University. Gough, Barry M. 1984 Gunboat Frontier: British Maritime Authority and Northwest Coast Indians. UBC Press, Vancouver. Grant, W. Colquhoun 1857 Description of Vancouver Island. Journal of the Royal Geographical Society of London 27:268\u00E2\u0080\u0093320. Gray, Brendan 2008 Sampling Methods in Northwest Coast Household Archaeology: A Simulation Approach Using Faunal Data from the Ozette Site. MA Thesis, University of Victoria. Grayson, Donald K. 1984 Quantitative Zooarchaeology: Topics in the Analysis of Archaeological Faunas. Academic Press, New York. Green, Lesley J. F. 2008 \u00E2\u0080\u0098Indigenous Knowledge\u00E2\u0080\u0099 and \u00E2\u0080\u0098Science\u00E2\u0080\u0099: Reframing the Debate on Knowledge Diversity. Archaeologies 4(1):144\u00E2\u0080\u0093163. \u00E2\u0080\u0093 324 \u00E2\u0080\u0093 Gregory, David A. 2001 An Evaluation of Early Agricultural Period in the Tucson Basin. In Excavation in the Santa Cruz River Floodplain: The Early Argricultural Period Component at Los Pozos, edited by D. A. Gregory. Anthropological Papers. vol. 21. Center for Desert Archaeology, Tucson. Gregr, Edward J., Linda Nichol, John K.B. Ford, Graeme Ellis and Andrew W. Trites 2000 Migration and Population Structure of Northeastern Pacific Whales off Coastal British Columbia: An Analysis of Commercial Whaling Records from 1908-1967. Marine Mammal Science 16(4):699\u00E2\u0080\u0093727. Gregr, Edward J., Linda M. Nichol, Jane C. Watson, John K. B. Ford and Graeme M. Ellis 2008 Estimating Carrying Capacity for Sea Otters in British Columbia. The Journal of Wildlife Management 72(2):382\u00E2\u0080\u0093388. Grier, Colin 2006 Temporality in Northwest Coast Households. In Household Archaeology on the Northwest Coast, edited by E. A. Sobel, D. A. Trieu Gahr and K. M. Ames, pp. 97\u00E2\u0080\u0093119. International Monographs in Prehistory, Ann Arbor. 2007 Consuming the Recent for Constructing the Ancient: The Role of Ethnography in Coast Salish Archaeological Interpretation. In Be Of Good Mind: Essays on the Coast Salish, edited by B. G. Miller, pp. 284\u00E2\u0080\u0093307. UBC Press, Vancouver, B.C. Grier, Colin, Patrick Dolan, Kelly Derr and Eric McLay 2009 Assessing Sea Level Changes in the Southern Gulf Islands of British Columbia Using Archaeological Data from Coastal Spit Locations. Canadian Journal of Archaeology 33(2):254\u00E2\u0080\u0093280. Grier, Colin, Kelli Flanigan, Misa Winters, Leah G. Jordan, Susan Lukowski and Brian M. Kemp 2013 Using Ancient DNA Identification and Osteometric Measures of Archaeological Pacific Salmon Vertebrae for Reconstructing Salmon Fisheries and Site Seasonality at Dionisio Point, British Columbia. Journal of Archaeological Science 40(1):544\u00E2\u0080\u0093555. Groesbeck, Amy Sue 2013 Ancient Clam Gardens Increased Production: Adaptive Strategies from the Past Can Inform Food Security Today. Master of Resource Management, Simon Fraser University. Guilderson, Tom P., Paula J. Reimer and Tom A. Brown 2005 The Boon and Bane of Radiocarbon Dating. Science 307:362-364. Gustafson, Carl E. 1968 Prehistoric Use of Fur Seals: Evidence from the Olympic Coast of Washington. Science 161(3836):49\u00E2\u0080\u009351. Haggan, Nigel 2012 Becoming Indigenous: Measurable and Immeasurable Values in Ecosystem-based Management. PhD Dissertation, University of British Columbia. \u00E2\u0080\u0093 325 \u00E2\u0080\u0093 Haggarty, James C. 1982 The Archaeology of Hesquiat Harbor: The Archaeological Utility of an Ethnographically Defined Social Unit. PhD Dissertation, Washington State University. Haggarty, James C. and Richard I. Inglis 1985 Historical Resources Site Survey and Assessment, Pacific Rim National Park. Report submitted to National Historic Parks and Sites Branch, Parks Service. Haida Marine Traditional Knowledge Study Participants, Janet Winbourne, Haida Oceans Technical Team and Haida Fisheries Program 2011 Haida Marine Traditional Knowledge Study Volume 3: Focal Species Summary. Report prepared for the Council of the Haida Nation with support from the Gordon and Betty Moore Foundation, Gwaii Trust Society, and Fisheries and Oceans Canada through the Aboriginal and Aquatic Resource and Oceans Management program. Hall, Dave, Chris Engisch, Paul Ewonus and Shelley Russell 2003 Kye Bay Water and Sanitary Sewer System Mitigation, Lazo Road, Comox, Vancouver Island, Permit 2003-92. Baseline Archaeological Services report prepared for the BC Archaeology Branch, Victoria. Hamilton, Marcus J. and Briggs Buchanan 2007 Spatial Gradients in Clovis-age Radiocarbon Dates Across North America Suggest Rapid Colonization from the North. Proceedings of the National Academy of Sciences, USA 104(40):15625\u00E2\u0080\u009315630. Hammon, Dimity 1986 Excavations at the Whalen Farm Site (Canadian), DgRs-14, Boundary Bay, B.C., Permit 1985-2. Report prepared for the Delta Museum and Archives and the BC Archaeology Branch, Victoria. Hancock, David 1970 California Sea Lion as a Regular Winter Visitant off the British Columbia Coast. Journal of Mammalogy 51(3):614. Hanson, Diane K 1995 Subsistence During the Late Prehistoric Occupation of Pender Canal, British Columbia (DeRt-1). Canadian Journal of Archaeology 19:29\u00E2\u0080\u009348. Hanson, Diane K. 1991 Late Prehistoric Subsistence in the Strait of Georgia Region of the Northwest Coast. PhD Dissertation, Simon Fraser University. Hanson, Diane K. and Karla D. Kusmer 2001 Sea Otter Scarcity and the Prehistoric Environment of the Strait of Georgia, British Columbia. In People and Wildlife in Northern North America: Essays in Honor of R. Dale Guthrie, edited by S. C. Gerlach and M. S. Murray, pp. 58\u00E2\u0080\u009366. BAR Press, Oxford. Harper, John R., Chief Adam Dick, Donald Mitchell, Daisy Sewid-Smith, Randy Bouchard, Dorothy Kennedy, Mary Morris and Chief Kim Recalma-Clutesi 2005 Clam Gardens of British Columbia. Paper presented at the 38th Annual Canadian Archaeological Association Conference, Nanaimo. \u00E2\u0080\u0093 326 \u00E2\u0080\u0093 Harper, John R., James C. Haggerty and Mary C. Morris 1995 Final Report: Broughton Archipelago Clam Terrace Survey. Report preparted by Coastal and Ocean Resources, Sidney BC. Harris, Cole 1997 Social Power and Cultural Change in Pre-Colonial British Columbia. BC Studies (115/116):45\u00E2\u0080\u009382. Harris, Douglas C. 2000 Territoriality, Aboriginal Rights, and the Heiltsuk Spawn-on-Kelp Fishery. UBC Law Review 34(1):195\u00E2\u0080\u0093238. Hart, John Lawson, Albert L. Tester, Desmond Beall and John P. Tully 1939 Proximate Analysis of British Columbia Herring in Relation to Season and Condition Factor. Journal of the Fisheries Research Board of Canada 4(5):478\u00E2\u0080\u0093490. Hay, Douglas E., P. Bruce McCarter and Kristen S. Daniel 2008 Potential Impacts of the British Columbia Herring Roe Fishery on the Spatial and Temporal Distribution of Herring Spawn: Examination of the Serial Depletion Hypothesis. Canadian Science Advisory Secretariat Research Document 2007/086. Fisheries and Oceans Canada, Ottawa. Hay, Douglas E., P. Bruce McCarter, Kristen S. Daniel and Jacob F. Schweigert 2009 Spatial diversity of Pacific herring (Clupea pallasi) spawning areas. ICES: Journal of Marine Science 66(8):1662\u00E2\u0080\u00931666. Hebda, Richard J. and S. Gay Frederick 1990 History of Marine Resources of the Northeast Pacific Since the Last Glaciation. Transactions of the Royal Society of Canada Series 1(1):319\u00E2\u0080\u0093342. Hebert, Jamie and Thomas F. Thornton 2010 Local And Traditional Ecological Knowledge: Herring Spawning And Massing Observations. In Herring Synthesis: Documenting and Modeling Herring Spawning Areas within Socio-Ecological Systems over Time in the Southeastern Gulf of Alaska, edited by T. F. Thornton, pp. 296\u00E2\u0080\u0093573. Report Submitted to the North Pacific Research Board, Project #728. Hessing-Lewis, Margot, Brittinay Keeling, Anna Gerrard and Anne K. Salomon 2011 Investigating the Social and Ecological Causes and Consequences of Pacific Herring Declines in BC's Central Coast: 2011 Field Report. A Contribution from the Hakai Network for Coastal People, Ecosystems and Management, Simon Fraser University, Burnaby. Hewer, Tony 1999 Archaeological Data Recovery: Blue Heron Road, DeRu-1, North Saanich, British Columbia, Permit 1998-197. IR Wilson Report Prepared for the BC Archaeology Branch, Victoria. Hilbert, Vi, Jay Miller and Zalmai Zahir (editors) 2001 sda\u00CA\u0094da\u00CA\u0094 gwe\u00C9\u00AB dibe\u00C9\u00AB l\u00C9\u0099\u00D3\u0099sucid \u00CA\u0094acaci\u00C9\u00ABtalbixw: Puget Sound Geography by TT Waterman. Lushootseed Press, Federal Way, WA. \u00E2\u0080\u0093 327 \u00E2\u0080\u0093 Hill, Erica 2012 The Nonempirical Past: Enculturated Landscapes and Other-than-Human Persons in Southwest Alaska. Arctic Anthropology 49(2):41\u00E2\u0080\u009357. Hobbs, Larry and Charles W. Fowler 2008 Putting Humans in Ecology: Consistency in Science and Management. Ambio 37(2):119\u00E2\u0080\u0093124. Hobson, Keith A. and Jonathan C. Driver 1989 Archaeological Evidence for Use of the Strait of Georgia by Marine Birds. In The Ecology and Status of Marine and Shoreline Birds in the Strait of Georgia, British Columbia, edited by K. Vermeer and R. W. Butler, pp. 168\u00E2\u0080\u0093173. Canadian Wildlife Service, Ottawa. Hodgetts, Lisa and Farid Rahemtulla 2001 Land and Sea: Use of Terrestrial Mammal Bones in Coastal Hunter\u00E2\u0080\u0093gatherer Communities. Antiquity 75(287):56\u00E2\u0080\u009362. Holmes, Heather and Tomas Tomascik 2003 Underwater Visual Surveys of Inshore Groundfish with Particular Focus on Rockfish, Broken Group Islands, Pacific Rim National Park Reserve. Paper presented at the Fifth Annual International Conference of Science and the Management of Protected Areas, University of Victoria, Victoria, British Columbia. Hourston, Alan S. 1958 Population Studies on Juvenile Herring in Barkley Sound, British Columbia. Journal of the Fisheries Research Board of Canada 15(5):909\u00E2\u0080\u0093960. Howe, Geordie, Robert J. Muir, Randy Bouchard, Dana Lepofsky, Brian Chisholm, Malcolm A. James and Rick Schulting 1994 Archaeological Investigations at the Departure Bay Site (DhRx-16), Nanaimo, BC, Permit 1993-103. Prepared by Arcas Consulting Archaeologists for the City of Nanaimo and submtted to the BC Archaeology Branch, Victoria. Hsieh, Chih-hao, Christian S. Reiss, John R. Hunter, John R. Beddington, Robert M. May and George Sugihara 2006 Fishing Elevates Variability in the Abundance of Exploited Species. Nature 443(7113):859\u00E2\u0080\u0093862. Huber, Harriet R., Jeffrey C. Jorgensen, Virginia L. Butler, Greg Baker and Rebecca Stevens 2011 Can Salmonids (Oncorhynchus spp.) Be Identified to Species Using Vertebral Morphometrics? Journal of Archaeological Science 38(1):136\u00E2\u0080\u0093146. Huelsbeck, David R. 1981 Utilization of Fish at the Ozette Site. Project reports 11. Laboratory of Archaeology and History, Washington State University, Pullman. 1988 Whaling in the Pre-contact Economy of the Central Northwest Coast. Arctic Anthropology 25(1):1\u00E2\u0080\u009315. \u00E2\u0080\u0093 328 \u00E2\u0080\u0093 1994 Mammals and Fish in the Subsistence Economy of Ozette. In Ozette Archaeological Project Research Reports Volume 2, Fauna, edited by S. R. Samuels, pp. 17\u00E2\u0080\u009391. Reports of Investigations. vol. 66. Department of Anthropology, Washington State University, Pullman. Hughen, Konrad A., Mike G. L. Baillie, Edouard Bard, J. Warren Beck, Chanda J. H. Bertrand, Paul G. Blackwell, Caitlin E. Buck, George S. Burr, Kirsten B. Cutler, Paul E. Damon, Richard L. Edwards, Richard G. Fairbanks, Michael Friedrich, Thomas P. Guilderson, Bernd Kromer, Gerry McCormac, Sturt Manning, Christopher Bronk Ramsey, Paula J. Reimer, Ron W. Reimer, Sabine Remmele, John R. Southon, Minze Stuiver, Sahra Talamo, F. W. Taylor, Johannes van der Plicht and Constanze E. Weyhenmeyer 2004 MARINE04 Marine Radiocarbon Age Calibration 0-26 Cal KYR BP. Radiocarbon 46(3):1059\u00E2\u0080\u00931086. Hull, Kathleen L. 2005 Process, Perception, and Practice: Time Perspectivism in Yosemite Native Demography. Journal of Anthropological Archaeology 24(4):354\u00E2\u0080\u0093377. Hunn, Eugene S. 1994 Place-Names, Population Density, and the Magic Number 500. Current Anthropology 35(1):81\u00E2\u0080\u009385. Hunn, Eugene S., Darryll R. Johnson, Priscilla N. Russell and Thomas F. Thornton 2003 Huna Tlingit Traditional Environmental Knowledge, Conservation, and the Management of a \u00E2\u0080\u009CWilderness\u00E2\u0080\u009D Park. Current Anthropology 44(Suppl.):S79-S103. Hutchinson, Ian, Thomas S. James, Paula J. Reimer, Brian D. Bornhold and John J. Clague 2004 Marine and Limnic Radiocarbon Reservoir Corrections for Studies of Late- and Postglacial Environments in Georgia Basin and Puget Lowland British Columbia, Canada and Washington, USA. Quaternary Research 61(2):193\u00E2\u0080\u0093 203. Huu-ay-aht First Nations 2000 Kiix?in Agenda Paper. In Nuu-chah-Nulth Voices, Histories, Objects and Journeys, edited by A. L. Hoover, pp. 33\u00E2\u0080\u009365. Royal British Columbia Museum, Victoria. Inglis, Richard I. and James C. Haggarty 1986 Pacific Rim National Park: Ethnographic History. Microfiche Report Series 257. Parks Service, Environment Canada, Ottawa. 2000 Cook to Jewitt: Three Decades of Change in Nootka Sound. In Nuu-chah-Nulth Voices, Histories, Objects, and Journeys, edited by A. L. Hoover, pp. 92\u00E2\u0080\u0093105. Royal British Columbia Museum, Victoria. Ingold, Tim 1980 Hunters, Pastoralists and Ranchers: Reindeer Economies and their Transformations. Cambridge University Press, Cambridge. 1993 The Temporality of the Landscape. World Archaeology 25(2):152\u00E2\u0080\u0093174. 2000 The Perception of the Environment: Essays on Livelihood, Dwelling and Skill. Routledge Press, London, UK. \u00E2\u0080\u0093 329 \u00E2\u0080\u0093 Ingram, B. Lynn and John R. Southon 1996 Reservoir Ages in Eastern Pacific Coastal and Estuarine Waters. Radiocarbon 38(3):573\u00E2\u0080\u0093582. Ivanochko, Tara S., Stephen E. Calvert, Richard E. Thomson and Thomas F. Pedersen 2008 Geochemical Reconstruction of Pacific Decadal Variability from the Eastern North Pacific During the Holocene. Canadian Journal of Earth Sciences 45(11):1317\u00E2\u0080\u00931329. Jackson, Jeremy B. C., Karen E. Alexander and Enric Sala (editors) 2011 Shifting Baselines: The Past and the Future of Ocean Fisheries. Island Press, Washington DC. Jackson, Jeremy B. C., Michael X. Kirby, Wolfgang H. Berger, Karen A. Bjorndal, Louis W. Botsford, Bruce J. Bourque, Roger H. Bradbury, Richard G. Cooke, Jon Erlandson, James A. Estes, Terence P. Hughes, Susan Kidwell, Carina B. Lange, Hunter S. Lenihan, John M. Pandolfi, Charles H. Peterson, Robert S. Steneck, Mia J. Tegner and Robert R. Warner 2001 Historical Overfishing and the Recent Collapse of Coastal Ecosystems. Science 293(5530):629\u00E2\u0080\u0093638. Jeffries, Steven, Harriet Huber, John Calambokidis and Jefferey Laake 2003 Trends and Status of Harbor Seals in Washington State: 1978-1999. Journal of Wildlife Management 67(1):208\u00E2\u0080\u0093219. Jewitt, John R. 1807 A Journal Kept at Nootka Sound. Printed for the author, Boston. Digital copy of original obtained from Canadiana.org October 16, 2003. Jordan, Christopher 1997 No Bones About It: The Effects of Cooking and Human Digestion on Salmon Bones. Northwest Anthropological Research Notes 31(1 & 2):71\u00E2\u0080\u009374. Kanipe, Hayley E., Rebecca J. Wigen, Ryan A. Spady, Andrew Hickok, Ian R. Wilson and Jennifer Nord 2007 Archaeological Excavation and Monitoring Pat Bay Sewer Main Installation, DdRu-5 North Saanich, B.C., Permits 2006-038 and 2006-107. IR Wilson Consultants report submitted to the BC Archaeology Branch, Victoria. Kemp, Brian M., Ripan S. Malhi, John McDonough, Deborah A. Bolnick, Jason A. Eshleman, Olga Rickards, Cristina Martinez-Labarga, John R. Johnson, Joseph G. Lorenz, E. James Dixon, Terence E. Fifield, Timothy H. Heaton, Rosita Worl and David Glenn Smith 2007 Genetic Analysis of Early Holocene Skeletal Remains from Alaska and its Implications for the Settlement of the Americas. American Journal of Physical Anthropology 132(4):605\u00E2\u0080\u0093621. Kenady, Stephen M. 2000 Archaeological Bulk Sampling and Faunal Analysis at the Flat Point Cable Crossing Upgrade: Lopez Island, Washington. Prepared for Orcas Power and Light Co. and submitted to the Department of Archaeology and Historic Preservation, Olympia. \u00E2\u0080\u0093 330 \u00E2\u0080\u0093 Kennett, Douglas J, B Lynn Ingram, John R Southon and Karen Wise 2002 Differences in 14C Age Between Stratigraphically Associated Charcoal and Marine Shell from the Archaic Period Site of Kilometer 4, Southern Peru: Old Wood or Old Water? Radiocarbon 44(1):53\u00E2\u0080\u009358. Kennett, Douglas J., B. Lynn Ingram, Jon M. Erlandson and Phillip Walker 1997 Evidence for Temporal Fluctuations in Marine Radiocarbon Reservoir Ages in the Santa Barbara Channel, Southern California. Journal of Archaeological Science 24(11):1051\u00E2\u0080\u00931059. Kennett, Douglas J. and James P. Kennett 2000 Competitive and Cooperative Responses to Climatic Instability in Coastal Southern California. American Antiquity 65(2):379\u00E2\u0080\u0093395. Kenyon, Susan M. 1980 The Kyuquot Way: a Study of a West Coast (Nootkan) community. Mercury Series 61. Canadian Ethnology Service, National Museums of Canada, Ottawa. King, Jacquelynne R., Gordon A. McFarlane and Richard J. Beamish 2000 Decadal-Scale Patterns in the Relative Year Class Success of Sablefish (Anoplopoma fimbria). Fisheries Oceanography 9(1):62\u00E2\u0080\u009370. Kittinger, John N., John M. Pandolfi, Jonathan H. Blodgett, Terry L. Hunt, Hong Jiang, Kepa \u00CC\u0084 Maly, Loren McClenachan, Jennifer K. Schultz and Bruce A. Wilcox 2011 Historical Reconstruction Reveals Recovery in Hawaiian Coral Reefs. PLoS One 6(10):e25460. Klokler, Daniela Magalh\u00C3\u00A3es 2008 Food for Body and Soul: Mortuary Ritual in Shell Mounds (Laguna - Brazil). PhD Dissertation, University of Arizona. Kopperl, Robert 2000 Appendix 2: Fish Bone Analysis. In Archaeological Bulk Sampling and Faunal Analysis at the Sperry Peninsula Utilities Upgrade: the South End of Lopez Island Washington, with contributions from Deborah Olson and Robert Kopperl, edited by S. M. Kenady, pp. 25\u00E2\u0080\u009338. Prepared for Orcas Power and Light Co. and submitted to the Department of Archaeology and Historic Preservation, Olympia. Kopperl, Robert E. 2001 Herring Use in Southern Puget Sound: Analysis of Fish Remains at 45-KI-437. Northwest Anthropological Research Notes 35(1):1\u00E2\u0080\u009320. 2007 Appendix B: Analysis of Fish Remains Watmough Bay Site Stabilization Project. In Watmough Bay Site Stabilization Project 45-SJ-280 Bone Analysis Summary and Reports, edited by K. M. Bovy and R. Kopperl, pp. 1\u00E2\u0080\u009317. Report Submitted to Bureau of Land Management, Spokane District and the Department of Archaeology and Historic Preservation, Olympia. \u00E2\u0080\u0093 331 \u00E2\u0080\u0093 Kopperl, Robert E. and Virginia L. Butler 2002 Faunal Analysis: Fish Remains. In Vashon Archaeology: A View from Burton Acres Shell Midden, edited by J. K. Stein and L. S. Phillips, pp. 105\u00E2\u0080\u0093117. Research Report. vol. 8. Burke Museum of Natural History and Culture, Seattle. Kowalewski, Micha\u00C5\u0082, Glen A. Goodfriend and Karl W. Flessa 1998 High Resolution Estimates of Temporal Mixing Within Shell Beds: The Evils and Virtues of Time-averaging. Paleobiology 24(3):287\u00E2\u0080\u0093304. Kristensen, Kira, Jonathan B. Hall, Rebecca J. Wigen and Erin Willows 2009 Archaeological Monitoring: Residential House Construction, DhSb-15, 866 Mariner Way, Parksville, Permit 2009-143. IR Wilson Consultants report prepared for Palladian Development Inc. and submitted to the BC Archaeology Branch, Victoria. Kusmer, Karla D. 1994a Changes in Subsistence Strategies at the Tsawwassen Site, a Southwestern British Columbia Shell Midden. Northwest Anthropological Research Notes 28(2):189\u00E2\u0080\u0093210. 1994b Faunal Remains. In Archaeological Investigations at Tsawwassen, B.C., Vol. II: Archaeology. Arcas Consulting Archaeologists Ltd, Port Moody, BC. Landres, Peter B., Penelope Morgan and Frederick J. Swanson 1999 Overview of the Use of Natural Variability Concepts in Managing Ecological Systems. Ecological Applications 9(4):1179\u00E2\u0080\u00931188. Larson, Lynn L. and Dennis E. Lewarch 1995 The Archaeology of West Point, Seattle, Washington: 4,000 Years of Hunter-Fisher-Gatherer Land Use in Southern Puget Sound. Larson Anthropological/Archaeological Services report prepared for King County and the Department of Archaeology and Historic Preservation, Olympia. Last, Peter Robert, William Toby White and John James Pogonoski 2007 Descriptions of New Dogfishes of the Genus 'Squalus' (Squaloidea: Squalidae). CSIRO. Marine and Atmospheric Research Paper 14. CSIRO Marine and Atmospheric Research, Hobart, Tasmania. Lee, J.C. and N. Bourne 1977 Marine Resource Inventory of Pacific Rim National Park - 1976. Fisheries and Marine Service Manuscript Report No. 1436. Pacific Biological Station, Nanaimo. Lepofsky, Dana, Michael Blake, Doug Brown, S. Morrison, Nichole Oakes and Natasha Lyons 2000 The Archaeology of the Scowlitz site, Southwestern British Columbia. Journal of Field Archaeology 27(4):391\u00E2\u0080\u0093416. Lepofsky, Dana and Megan Caldwell 2013 Indigenous Marine Resource Management on the Northwest Coast of North America. Ecological Processes 2:12. Lepofsky, Dana and Kenneth P. Lertzman 2005 More on Sampling for Richness and Diversity in Archaeobiological Assemblages. Journal of Ethnobiology 25(2):175\u00E2\u0080\u0093188. \u00E2\u0080\u0093 332 \u00E2\u0080\u0093 Lepofsky, Dana, Kenneth P. Lertzman, Douglas J. Hallett and Rolf W. Mathewes 2005 Climate Change and Culture Change on the Southern Coast of British Columbia 2400\u00E2\u0080\u00931200 B.P.: An Hypothesis. American Antiquity 70(2):267\u00E2\u0080\u0093293. Lepofsky, Dana and Natasha Lyons 2003 Modeling Ancient Plant Use on the Northwest Coast: Towards an Understanding of Mobility and Sedentism. Journal of Archaeological Science 30(11):1357\u00E2\u0080\u00931371. Lepofsky, Dana, Teresa Trost and Jesse Morin 2007 Coast Salish Interaction: A View from the Inlets. Canadian Journal of Archaeology 31(2):190\u00E2\u0080\u0093223. Lertzman, Kenneth P. 2009 The Paradigm of Management, Management Systems, and Resource Stewardship. Journal of Ethnobiology 29(2):339\u00E2\u0080\u0093358. Letham, Bryn 2011 The Ancient Cultural Landscape of the Sechelt Inlet System. MA Thesis, University of Toronto. Levin, Simon A. 1992 The Problem of Pattern and Scale in Ecology: The Robert H. MacArthur Award Lecture. Ecology 73(6):1943\u00E2\u0080\u00931967. Lewarch, Dennis E., Leonard A. Forsman, Stephenie K. Kramer, Laura R. Murphy, Lynn L. Larson, David R. Iversen and Amy E. Dugas 2002 Data Recovery Excavations the Bay Street Shell Midden (45-KP-115), Kitsap County, WA. Larson Anthropological Archaeological Services Limited report submitted to the City of Port Orchard, Washington. Libby, William F., E. C. Anderson and J. R. Arnold 1949 Age Determination by Radiocarbon Content: World-wide Assay of Natural Radiocarbon. Science 109(2827):227\u00E2\u0080\u0093228. Lightfoot, Kent G. 1993 Long-term Developments in Complex Hunter-gatherer Societies: Recent Perspectives from the Pacific Coast of North America. Journal of Archaeological Research 1(3):167\u00E2\u0080\u0093201. Lindegren, Martin, David M. Checkley Jr., Tristan Rouyer, Alec D. MacCall and Nils Chr. Stenseth 2013 Climate, Fishing, and Fluctuations of Sardine and Anchovy in the California Current. Proceedings of the National Academy of Sciences, USA 110(33):13672\u00E2\u0080\u009313677. Linse, Angela R. 1992 Is Bone Safe in a Shell Midden? In Deciphering a Shell Midden, edited by J. K. Stein, pp. 327\u00E2\u0080\u0093345. Academic Press, San Diego. Liu, Jianguo, Thomas Dietz, Stephen R. Carpenter, Carl Folke, Marina Alberti, Charles L. Redman, Stephen H. Schneider, Elinor Ostrom, Alice N. Pell, Jane Lubchenco, William W. Taylor, Zhiyun Ouyang, Peter Deadman, Timothy Kratz and William Provencher 2007 Coupled Human and Natural Systems. Ambio: A Journal of the Human Environment 36(8):639\u00E2\u0080\u0093649. \u00E2\u0080\u0093 333 \u00E2\u0080\u0093 Losey, Robert J. 2000 Oral Tradition of Earthquakes and Tsunamis on the Central Cascadia Coast: Variation of Accounts and Relations to Historically Observed Patterns Across the Northwest Coast, Changing Landscapes: Proceedings of the 4th Annual Coquille Cultural Preservation Conference, 3-15. 2002 Communities and Catastrophe: Tillamook Response to the AD 1700 Earthquake and Tsunami, Northern Oregon Coast. PhD Dissertation, University of Oregon. 2005 Earthquakes and Tsunami as Elements of Environmental Disturbance on the Northwest Coast of North America. Journal of Anthropological Archaeology 24(2):101\u00E2\u0080\u0093116. 2010 Animism as a Means of Exploring Archaeological Fishing Structures on Willapa Bay, Washington, USA. Cambridge Archaeological Journal 20(1):17\u00E2\u0080\u009332. Lotze, Heike K. and Boris Worm 2009 Historical Baselines for Large Marine Animals. Trends in Ecology and Evolution 24(5):254\u00E2\u0080\u0093262. Louderback, Lisbeth A., Donald K. Grayson and Marcos Llobera 2011 Middle-Holocene Climates and Human Population Densities in the Great Basin, Western USA. The Holocene 21(2):366\u00E2\u0080\u0093373. Love, Milton S., Mary Yoklavich and Lyman K. Thorsteinson 2002 The Rockfishes of the Northeast Pacific. University of California Press, Berkeley. Luby, Edward M., Clayton D. Drescher and Kent G. Lightfoot 2006 Shell Mounds and Mounded Landscapes in the San Francisco Bay Area: An Integrated Approach. Journal of Island and Coastal Archaeology 1(2):191 - 214. Luby, Edward M. and Mark F. Gruber 1999 The Dead Must be Fed: Symbolic Meanings of the Shell Mounds of the San Francisco Bay Area. Cambridge Archaeological Journal 9(1):95\u00E2\u0080\u0093108. Ludwin, Ruth S., Robert Dennis, Deborah Carver, Alan D. McMillan, Robert Losey, John Clague, Chris Jonientz-Trisler, Janine Bowechop, Jacilee Wray and Karen James 2005 Dating the 1700 Cascadia Earthquake: Great Coastal Earthquakes in Native Stories. Seismological Research Letters 76(2):140\u00E2\u0080\u0093148. Ludwin, Ruth S., Gregory J. Smits, Deborah Carver, K. James, C. Jonientz-Trisler, Alan D. McMillan, Robert Losey, Robert Dennis, J. Rasmussen, A. De Los Angeles, D. Buerge, Colin P. Thrush, John J. Clague, Janine Bowechop and Jacilee Wray 2007 Folklore and Earthquakes: Native American Oral Traditions from Cascadia Compared with Written Traditions from Japan. In Myth and Geology, edited by L. Piccardi and W. B. Masse, pp. 67\u00E2\u0080\u009394. Geological Society Special Publication. vol. 273. Geological Society, London. Lutz, John Sutton 2008 Mak\u00C3\u00BAk: A New History of Aboriginal-white Relations. UBC Press, Vancouver. \u00E2\u0080\u0093 334 \u00E2\u0080\u0093 Lyman, R. Lee 1994 Vertebrate Taphonomy. Cambridge University Press, Cambridge. 2003a The Influence of Time Averaging and Space Averaging on the Application of Foraging Theory in Zooarchaeology. Journal of Archaeological Science 30(5):595\u00E2\u0080\u0093610. 2003b Pinniped Behavior, Foraging Theory, and the Depression of Metapopulations and Nondepression of a Local Population on the Southern Northwest Coast of North America. Journal of Anthropological Archaeology 22(4):376\u00E2\u0080\u009388. 2008 Quantitative Paleozoology. Cambridge Manuals in Archaeology. Cambridge University Press, Cambridge. Lyman, R. Lee and Kenneth M. Ames 2004 Sampling to Redundancy in Zooarchaeology: Lessons from the Portland Basin, Northwestern Oregon and Southwestern Washington. Journal of Ethnobiology 24(2):329\u00E2\u0080\u0093346. 2007 On the Use of Species-area Curves to Detect the Effects of Sample Size. Journal of Archaeological Science 34(12):1985\u00E2\u0080\u00931990. MacCall, Alec D. 1990 Dynamic Geography of Marine Fish Populations. University of Washington Press, Seattle. MacDonald, Brandi Lee 2008 Ochre Procurement and Distribution on the Central Coast of British Columbia. MA Thesis, MacMaster University. MacDonald, Brandi Lee, R.G.V. Hancock, Aubrey Cannon and Alice Pidruczny 2011 Geochemical Characterization of Ochre from Central Coastal British Columbia, Canada. Journal of Archaeological Science 38(12):3620\u00E2\u0080\u00933630. MacDonald, George F. 1983 Prehistoric Art from the Northern Northwest Coast. In Indian Art Traditions of the Northwest Coast, edited by R. L. Carlson, pp. 99\u00E2\u0080\u0093121. Archaeology Press, Simon Fraser University, Burnaby. MacDonald, George F. and Jerome S. Cybulski 2001 Introduction: the Prince Rupert Harbour Project. In Perspectives on Northern Northwest Coast Prehistory, edited by J. S. Cybulski, pp. 1\u00E2\u0080\u009324. Mercury Series, Archaeological Survey of Canada. vol. 160. Canadian Museum of Civilization, Ottawa. MacDonald, George F. and Richard I. Inglis 1981 An Overview of the North Coast Prehistory Project (1966-1980). BC Studies 48:37-63. \u00E2\u0080\u0093 335 \u00E2\u0080\u0093 Mackie, Alexander P. and Laurie Williamson 2003 Nuu-chah-nulth Houses: Structural Remains and Cultural Depressions on Southwest Vancouver Island. In Emerging from the Mist: Studies in Northwest Coast Culture History, edited by R. G. Matson, G. Coupland and Q. Mackie, pp. 105\u00E2\u0080\u0093151. Pacific Rim Archaeology Series. UBC Press, Vancouver. n.d. Ohiaht Ethnoarchaeological Project, FINAL Report, Permit 1984-26. Draft Permit Report in preparation for submission to the Huu-ay-aht First Nation and the BC Archaeology Branch. Mackie, Quentin 2001 Settlement Archaeology in a Fjordland Archipelago: Network Analysis, Social Practice, and the Built Environment of Western Vancouver Island, British Columbia, Canada since 2000 BP. International Series Vol. 926. British Archaeological Reports, Oxford. Mackie, Quentin, Daryl W. Fedje, Duncan McLaren, Nicole Smith and Iain McKechnie 2011 Early Environments and Archaeology of Coastal British Columbia. In Trekking the Shore: Changing Coastlines and the Antiquity of Coastal Settlement, edited by N. F. Bicho, J. A. Haws and L. G. Davis, pp. 51\u00E2\u0080\u0093103. Springer, New York. Mackinson, Susan 1999 Variation in Structure and Distribution of Pre-spawning Pacific Herring Shoals in Two Regions of British Columbia. Journal of Fish Biology 55(5):972\u00E2\u0080\u0093989. MacLean, Kelsey 2012 An Analysis of the Flaked Stone Assemblage from the Hiikwis Site Complex, Barkley Sound, British Columbia. MA Thesis, University of Victoria. Markel, Russell W. 2011 Rockfish Recruitment and Trophic Dynamics on the West Coast of Vancouver Island: Fishing, Ocean Climate, and Sea Otters. PhD Dissertation, University of British Columbia. Marsden, Susan 2001 Defending the Mouth of the Skeena: Perspectives on Tsimshian Tlingit Relations. In Perspectives in Northern Northwest Coast Prehistory, edited by J. S. Cybulski, pp. 61-106. Mercury Series, Archaeological Survey of Canada. vol. 160. Canadian Museum of Civilization, Hull, Quebec. Marshall, Yvonne 1993 A Political History of the Nuu-chah-nulth People: A Case Study of the Mowachaht and Muchalaht Tribes. PhD Dissertation, Simon Fraser University. 2000a The Changing Art and Architecture of Potlach Houses at Yuquot. In Nuu-chah-Nulth Voices Histories, Objects and Journeys, edited by A. L. Hoover, pp. 107\u00E2\u0080\u0093130. Royal British Columbia Museum, Victoria. 2000b Transformations of Nuu-chah-nulth Houses. In Beyond Kinship: Social and Material Reproduction in House Societies, edited by R. A. Joyce and S. D. Gillespie, pp. 73-102. University of Pennsylvania Press, Philadelphia. \u00E2\u0080\u0093 336 \u00E2\u0080\u0093 2006 Houses and Domestication on the Northwest Coast. In Household Archaeology on the Northwest Coast, edited by E. A. Sobel, D. A. Trieu Gahr and K. M. Ames, pp. 37\u00E2\u0080\u009356. International Monographs in Prehistory, Ann Arbor. Martindale, Andrew 2006 Methodological Issues in the Use of Tsimshian Oral Traditions (Adawx) in Archaeology. Canadian Journal of Archaeology 30(2):159\u00E2\u0080\u0093193. 2009 Entanglement and Tinkering: Structural History in the Archaeology of the Northern Tsimshian. Journal of Social Archaeology 9(1):59\u00E2\u0080\u009391. n.d. The Value of Uncertainty: Interdisciplinarity as Federated Knowledge. Martindale, Andrew and Irena Jurakic 2004 Northern Tsimshian Elderberry Use in the Late Pre-contact to Post-contact Era Canadian Journal of Archaeology 28:254\u00E2\u0080\u0093280. Martindale, Andrew and Bryn Letham 2011 Causalities and Models within the Archaeological Construction of Political Order on the Northwest Coast of North America. In The Archaeology of Politics: the Materiality of Political Practice and Action in the Past, edited by P. G. Johansen and A. M. Bauer, pp. 323\u00E2\u0080\u0093353. Cambridge Scholars Press, Newcastle. Martindale, Andrew, Bryn Letham, Natalie Brewster and Angela Ruggles 2010 The Dundas Island Project, Permit 2005-159. Report submitted to Tsimshian Allied Tribes and the BC Archaeology Branch, Victoria. Martindale, Andrew, Bryn Letham, Duncan McLaren, David Archer, Meghan Burchell and Bernd R. Schone 2009 Mapping of Subsurface Shell Midden Components Through Percussion Coring: Examples From the Dundas Islands. Journal of Archaeological Science 36(9):1565\u00E2\u0080\u00931575. Martindale, Andrew and Susan Marsden 2003 Defining the Middle Period (3500 BP to 1500 BP) in Tsimshian History through a Comparison of Archaeological and Oral Records. BC Studies (138):13\u00E2\u0080\u009350. Marty, Gary D., Peter-John F. Hulson, Sara E. Miller, Terrance J. Quinn II, Steve D. Moffitt and Richard A. Merizon 2010 Failure of Population Recovery in Relation to Disease in Pacific Herring. Diseases of Aquatic Organisms 90(1):1\u00E2\u0080\u009314. Maschner, Herbert D. G. 1991 The Emergence of Cultural Complexity on the Northern Northwest Coast. Antiquity 65(249):924\u00E2\u0080\u0093934. 1992 The Origins of Hunter-gatherer Sedentism and Political Complexity: A Case Study from the Northern Northwest Coast. PhD Dissertation, University of California, Santa Barbara. \u00E2\u0080\u0093 337 \u00E2\u0080\u0093 Maschner, Herbert D.G. and Jeffery W. Stein 1995 Multivariate approaches to site location on the Northwest Coast of North America. Antiquity 69:61-73. Mason, Andrew R. and Tanja Hoffmann 1998 Report on Archaeological Mitigation of DjSf-13, Buckley Bay Main Road and Interchange, Vancouver Island Highway Project, Alteration Permit 1997-345. Golder Associates report prepared for the Ministry of Transportation and submitted to the BC Archaeology Branch, Victoria. Mason, Ronald J. 2000 Archaeology and Native North American Oral Traditions. American Antiquity 65(2):239\u00E2\u0080\u0093266. 2006 Inconstant Companions: Archaeology and North American Indian Oral Traditions. University of Alabama Press, Tuscaloosa. Matson, R. G. (editor) 1976 The Glenrose Cannery site. National Museums of Canada, Ottawa. 1992 The Evolution of Northwest Coast Subsistence. In Research in Economic Anthropology, edited by D. R. Croes, R. Hawkins and B. L. Issac, pp. 367\u00E2\u0080\u0093428. vol. 6. JAI Press Inc., Greenwich, CT. 2003 The Coast Salish House: Lessons from Shingle Point,Valdes Island, British Columbia. In Emerging from the Mist: Studies in Northwest Coast Culture History, edited by R. G. Matson, G. Coupland and Q. Mackie, pp. 76-104. Pacific Rim Archaeology Series. UBC Press, Vancouver. 2010 Chapter 4: Stratigraphy and Chronology. In The Crescent Beach site and the place of the Locarno Beach phase, edited by R. G. Matson. UBC Laboratory of Archaeology, Vancouver. Matson, R. G. and Gary Coupland 1995 The Prehistory of the Northwest Coast. Academic Press, New York. Matson, R. G., Joanne Green and Eric McLay 1999 Houses and Households in the Gulf of Georgia: Archaeological Investigations of Shingle Point (DgRv-2), Valdes Island, British Columbia. Report submitted to the Lyacksun First Nation, the Social Science and Humanities Research Council of Canada, Ottawa, and the BC Archaeology Branch, Victoria. McCarter, P. Bruce, Kristen S. Daniel and Douglas E. Hay 2005 Mapping Herring Spawn. Canadian Manuscript Report of Fisheries and Aquatic Sciences 2714. Fisheries and Oceans Canada Science Branch, Pacific Region Pacific Biological Station, Nanaimo. McClenachan, Loren, Francesco Ferretti and Julia K. Baum 2012 From Archives to Conservation: Why Historical Data are Needed to Set Baselines for Marine Animals and Ecosystems. Conservation Letters 5(5):349\u00E2\u0080\u0093359. \u00E2\u0080\u0093 338 \u00E2\u0080\u0093 McFarlane, Gordon A., Jacquelynne R. King and Richard J. Beamish 2000 Have There Been Recent Changes in Climate? Ask the Fish. Progress in Oceanography 47:147\u00E2\u0080\u0093169. McGhee, Robert 2008 Aboriginalism and the Problems of Indigenous Archaeology. American Antiquity 73(4):579\u00E2\u0080\u0093597. McKechnie, Iain 2005a Column Sampling and the Archaeology of Small Fish at Ts'ishaa. In Ts'ishaa: Archaeology and Ethnography of a Nuu-chah-nulth Origin Site in Barkley Sound, edited by A. D. McMillan and D. E. St. Claire, pp. 206\u00E2\u0080\u0093223, R. Carlson, general editor. Archaeology Press, Simon Fraser University, Burnaby. 2005b Five Thousand Years of Fishing at a Shell Midden in the Broken Group Islands, Barkley Sound, British Columbia. MA Thesis, Simon Fraser University. 2007a Investigating the Complexities of Sustainable Fishing at a Prehistoric Village on Western Vancouver Island, British Columbia, Canada. Journal for Nature Conservation 15(3):208\u00E2\u0080\u0093222. 2007b Vertebrate Faunal Analysis at Himayis (205T, DfSi-17) and the Clarke Island Defensive Site (212T, DfSi-26), Broken Group Islands, Pacific Rim National Park Reserve. Report submitted to Parks Canada Cultural Resource Services, Western and Northern Service Centre, Calgary. 2007c Vertebrate Faunal Analysis from Tl\u00E2\u0080\u0099aadiiwa village (39T, DeSf-6), West Coast Trail Unit, Pacific Rim National Park Reserve. Millennia Research Limited report prepared for Cultural Resource Services, Northern and Western Canada Service Centre, Parks Canada Agency, Calgary. 2009 2008 Archaeological Research at Hutsatswilh, A Tseshaht Village and Defensive Fortress in the Broken Group Islands (83T & 129T), PRN Permit 2008-1579. Interim Report Submitted to Tseshaht First Nation, Port Alberni, BC; Pacific Rim National Park Reserve of Canada, Ucluelet, BC; and Parks Canada Cultural Resource Services, Western and Northern Service Centre, Victoria, BC. 2010a An Archaeological Perspective on the Hake Fishery in Barkley Sound. Paper presented at the Presentation at the 1st Barkley Sound Knowledge Symposium, February 9-11, 2010, Bamfield Marine Sciences Centre, Bamfield. 2010b Archaeological Research in the Broken Group Islands, Pacific Rim National Park Reserve, June 2009. Interim Report Submitted to: Tseshaht First Nation, Port Alberni, BC; Pacific Rim National Park Reserve of Canada, Ucluelet, BC; and Parks Canada Cultural Resource Services, Western and Northern Service Centre, Victoria, BC. 2012 Zooarchaeological Analysis of the Indigenous Fishery at the Huu7ii Big House and Back Terrace, Huu-ay-aht Territory, Southwestern Vancouver Island. In Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound, edited by A. D. McMillan and D. E. St. Claire, pp. 154\u00E2\u0080\u0093186. Archaeology Press, Simon Fraser University, Burnaby, BC. \u00E2\u0080\u0093 339 \u00E2\u0080\u0093 McKechnie, Iain, Audrey Dallimore, Gay Frederick and Nicole Smith 2008 Marine Fisheries and Marine Climate Before and During the Little Ice Age on Western Vancouver Island. Paper presented at the 73rd Annual Meeting of the Society for American Archaeology, Vancouver, British Columbia. McKechnie, Iain and Daryl W. Fedje 2010 Archaeological Evidence of a Possible Paleo-tsunami at the Tl\u00E2\u0080\u0099aadiiwa Village. Paper presented at the Poster Presented at the 1st Barkley Sound Knowledge Symposium, February 9-11, 2010, Bamfield Marine Sciences Centre, Bamfield. McKechnie, Iain and Dominic J. Tollit n.d. A Regression-based Method for Estimating the Fork Lengths of Pacific Herring in Archaeological sites in Northwestern North America. McLaren, Duncan 2003 Uncovering Historical Sequences in Central Coast Salish Oral Narratives. In Archaeology of Coastal British Columbia: Essays in Honour of Professor Philip Hobler, edited by R. L. Carlson, pp. 189\u00E2\u0080\u0093202. Archaeology Press, Simon Fraser University, Burnaby. 2008 Sea Level Change and Archaeological Site Locations on the Dundas Island Archipelago of North Coastal British Columbia. PhD Dissertation, University of Victoria. McLaren, Duncan, Andrew Martindale, Daryl W. Fedje and Quentin Mackie 2011 Relict Shorelines and Shell Middens of the Dundas Island Archipelago. Canadian Journal of Archaeology 35(2):86\u00E2\u0080\u0093116. McLay, Eric 1999a The Diversity of Northwest Coast Shell Middens: Late Pre-contact Settlement-subsistence Patterns on Valdes Island, British Columbia. M.A. Thesis, University of British Columbia. 1999b Valdez Archaeological Survey, Permit 1996-125. Report prepared for Lyackson First Nation, British Columbia Heritage Trust, and the BC Archaeology Branch, Victoria. McMillan, Alan D. 1999 Since the Time of the Transformers: The Ancient Heritage of the Nuu-chah-nulth, Ditidaht, and Makah. Pacific Rim Archaeology Series. UBC Press, Vancouver. 2003 The Early Component at Ts'ishaa, An Outer Coast Site on Western Vancouver Island. In Archaeology of Coastal British Columbia: Essays in Honour of Professor Philip M. Hobler, edited by R. L. Carlson, pp. 39\u00E2\u0080\u009344. Archaeology Press, Simon Fraser University, Burnaby. 2009 A Tale of Two Ethnographies: The Contributions of Edward Sapir and Philip Drucker to Nuu-chah-nulth Archaeology In Painting the Past with a Broad Brush: Papers in Honour of James Valliere Wright, edited by D. L. Keenlyside and J.-L. Pilon, pp. 617\u00E2\u0080\u0093646. Mercury Series\u00E2\u0080\u0094Archaeology Paper. vol. 170. Canadian Museum of Civilization, Gatineau. McMillan, Alan D. and Ian Hutchinson 2002 When the Mountain Dwarfs Danced: Aboriginal Traditions of Paleoseismic Events along the Cascadia Subduction Zone of Western North America. Ethnohistory 49(1):41\u00E2\u0080\u009368. \u00E2\u0080\u0093 340 \u00E2\u0080\u0093 McMillan, Alan D., Iain McKechnie, Denis E. St. Claire and S. Gay Frederick 2008 Exploring Variability in Maritime Resource Use on the Northwest Coast: A Case Study from Barkley Sound, Western Vancouver Island. Canadian Journal of Archaeology 32(2):214\u00E2\u0080\u0093238. McMillan, Alan D. and Denis E. St. Claire 1982 Alberni Prehistory: Archaeological and Ethnographic Investigations on Western Vancouver Island. Theytus Books, Penticton. 1992 The Toquart Archaeology Project: Report on the 1992 Excavations, Permit 1992-46. Report submitted to the Archaeology and Outdoor Recreation Branch, Victoria, BC. 1994 The Toquaht Archaeological Project: Report on the 1994 Field Season, Permit 1994-91 Permit Report Submitted to the Toquaht Nation, Ucluelet, and the BC Heritage Trust and BC Archaeology Branch, Victoria. 1996 The Toquaht Archaeological Project: Report on the 1996 Field Season, Permit 1996-120 Permit Report Submitted to the Toquaht Nation, Ucluelet, and the BC Heritage Trust and BC Archaeology Branch, Victoria. 2005 Ts'ishaa: Archaeology and Ethnohistory of a Nuu-chah-nulth Origin Site in Barkley Sound. Archaeology Press, Simon Fraser University, Burnaby. 2012 Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound. Archaeology Press, Simon Fraser University, Burnaby, BC. McNeely, Richard, Arthur S. Dyke and John R. Southon 2006 Canadian Marine Reservoir Ages, Preliminary Data Assessment. Geological Survey of Canada, Open File 5049. McNiven, Ian J. 2013 Ritualized Middening Practices. Journal of Archaeological Method and Theory 20(4):552\u00E2\u0080\u0093587. Menge, Bruce A., Alison C. Iles and Tess L. Freidenburg 2013 Keystone Species. In Encyclopedia of Biodiversity, edited by S. A. Levin, pp. 442\u00E2\u0080\u0093457. 2nd ed. vol. 4. Academic Press, Waltham. Menounos, Brian, Gerald Osborn, John J. Clague and Brian H. Luckman 2009 Latest Pleistocene and Holocene glacier fluctuations in western Canada. Quaternary Science Reviews 28(21-22):2049-2074. Menzies, Charles R. 2010 Dm sibilhaa\u00E2\u0080\u0099nm da Laxyuubm Gitxaa\u00C5\u0082a: Picking Abalone in Gitxaa\u00C5\u0082a Territory. Human Organization 69(3):213\u00E2\u0080\u0093220. 2012 The Disturbed Environment: The Indigenous Cultivation of Salmon. In Keystone Nations: indigenous Peoples and Salmon across the North Pacific, edited by B. J. Colmnbi and J. E. Brooks, pp. 162\u00E2\u0080\u0093182. School for Advanced Research, Santa Fe. \u00E2\u0080\u0093 341 \u00E2\u0080\u0093 Merilees, Bill 1985 Humpbacks in Our Strait. Waters: the Journal of the Vancouver Aquarium 8:7-24. Miller, Bruce G. 1992 Common Sense and Plain Language. BC Studies 95:55\u00E2\u0080\u009365. 2011 Oral History on Trial: Recognizing Aboriginal Narratives in the Courts. UBC Press, Vancouver. Minor, Rick, Debra C. Barner and Ruth L. Greenspan 1986 The Leask Site: a Late Prehistoric Campsite on Annette Island, Southeast Alaska. Heritage Research Associates Report submitted to Bureau of Indian Affairs, Portland and the Department of Archaeology and Historic Preservation, Olympia. Mintz, Sidney W. and Christine M. Du Bois 2002 The Anthropology of Food and Eating. Annual Review of Anthropology 31:99\u00E2\u0080\u0093119. Misarti, Nicole, Bruce Finney, Herbert D.G. Maschner and Matthew J. Wooller 2009 Changes in Northeast Pacific Marine Ecosystems Over the Last 4500 years: Evidence from Stable Isotope Analysis of Bone Collagen from Archeological Middens. The Holocene 19(8):1139\u00E2\u0080\u00931151. Mitchell, Donald H. 1974 Salvage Excavations at the Buckley Bay Site (DjSf-13) Baynes Sound, Vancouver Island, British Columbia, Permit 1974-18. Report Submitted to Mr. Bjorn Simonsen, Provincial Archaeologist Parliament Buildings, Victoria, BC. 1981 DcRu-78: A Prehistoric Occupation of Fort Rodd Hill National Park. Syesis 14:129\u00E2\u0080\u0093150. 1988 The J. Puddleduck Site: A Northern Strait of Georgia Locarno Beach Component and Its Predecessor. Royal British Columbia Museum, Victoria. 1990a Coast Salish Subsistence Studies and a Methodological Barrier. Northwest Anthropological Research Notes 24(2):239\u00E2\u0080\u0093247. 1990b Prehistory of the Coasts of Southern British Columbia and Northern Washington. In Handbook of North American Indians Volume 7: Northwest Coast, edited by W. Suttles, pp. 340-358. Smithsonian Press, Washington, D.C. Mitchell, Mark D. and Laura L. Scheiber 2010 Crossing Divides: Archaeology as Long-Term History. In Across a Great Divide: Continuity and Change in Native North American Societies, 1400-1900, edited by L. L. Scheiber, M. D. Mitchell and K. G. Tregonning, pp. 1\u00E2\u0080\u009321. University of Arizona Press, Tucson. Momaday, N. Scott 1974 Native American Attitudes to the Environment. In Seeing with a Native Eye: Essays on Native American Religion, edited by A. k. Hultkrantz and W. H. Capps, pp. 79\u00E2\u0080\u009385. Harper and Row, New York. \u00E2\u0080\u0093 342 \u00E2\u0080\u0093 Monks, Gregory G. 1987 Prey as Bait: The Deep Bay Example. Canadian Journal of Archaeology 11(1):119\u00E2\u0080\u0093142. 2000 How Much is Enough? An Approach to Sampling Ichthyofaunas. Ontario Archaeology 69:65\u00E2\u0080\u009375. 2001 Quit Blubbering: An Examination of Nuu\u00E2\u0080\u0099chah\u00E2\u0080\u0099nulth (Nootkan) Whale Butchery. International Journal of Osteoarchaeology 11:136-149. 2003 The Cultural Taphonomy of Nuu-chah-nulth Whale Bone Assemblages. In Emerging from the Mist: Studies in Northwest Coast Culture History, edited by R. G. Matson, G. Coupland and Q. Mackie, pp. 188\u00E2\u0080\u0093212. Pacific Rim Archaeology Series. UBC Press, Vancouver. 2006 The Fauna from Ma\u00E2\u0080\u0099acoah (DfSi\u00E2\u0080\u00935), Vancouver Island, British Columbia: An Interpretive Summary. Canadian Journal of Archaeology 30(2):215\u00E2\u0080\u0093244. 2011 Locational Optimization and Faunal Remains in Northern Barkley Sound, Western Vancouver Island, British Columbia. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 129\u00E2\u0080\u0093148. University of Alaska Press, Fairbanks. Morin, Jesse 2012 The Political Economy of Stone Celt Exchange in Pre-Contact British Columbia: The Salish Nephrite/Jade Industry. PhD Dissertation, University of British Columbia. 2013 Book Review: Rewriting Marpole: The Path to Cultural Complexity in the Gulf of Georgia Region by Terence N. Clark BC Studies. Morin, Jesse, Ryan Dickie, Takashi Sakaguchi and Jamie Hoskins 2008/9 Late Prehistoric Settlement Patterns and Population Dynamics along the Mid-Fraser. BC Studies (160):9\u00E2\u0080\u009334. Morita, Shou 1985 History of the Herring Fishery and Review of Artificial Propagation Techiques for Herring in Japan. Canadian Journal of Fisheries and Aquatic Sciences 42(S1):s222\u00E2\u0080\u0093s229. Moss, Madonna L. 1985 Phosphate Analysis of Archaeological Sites, Admiralty Island, Southeast Alaska. Syesis 17:95\u00E2\u0080\u0093100. 1989 Archaeology and Cultural Ecology of the Prehistoric Angoon Tlingit. PhD Dissertation, University of California, Santa Barbara. 2004 Archaeological Investigation of Cape Addington Rockshelter: Human Occupation of the Rugged Seacoast on the Outer Prince of Wales Archipelago, Alaska. University of Oregon Anthropological Papers 63. University of Oregon Museum of Natural and Cultural History, Eugene. 2007a Haida and Tlingit Use of Seabirds from the Forrester Islands, Southeast Alaska. Journal of Ethnobiology 27(1):28\u00E2\u0080\u009345. \u00E2\u0080\u0093 343 \u00E2\u0080\u0093 2007b The Killisnoo Picnicground Midden (49-SIT-124) Revisited: Assessing Archaeological Recovery of Vertebrate Faunal Remains from Northwest Coast Shell Middens. Journal of Northwest Anthropology 41(1):1\u00E2\u0080\u009317. 2008 Outer Coast Maritime Adaptations in Southern Southeast Alaska: Tlingit or Haida? Arctic Anthropology 45(1):41\u00E2\u0080\u009360. 2011a Northwest Coast: Archaeology as Deep History. SAA Contemporary Perspectives. Society for American Archaeology Press, Washington D.C. 2011b Pacific Cod in Southeast Alaska: The \u00E2\u0080\u009CCousin\u00E2\u0080\u009D of the Fish That Changed the World. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 149\u00E2\u0080\u0093169. University of Alaska Press, Fairbanks. 2012 Understanding Variability in Northwest Coast Faunal Assemblages: Beyond Economic Intensification and Cultural Complexity. Journal of Island and Coastal Archaeology 7(1):1\u00E2\u0080\u009322. Moss, Madonna L., Virginia L. Butler and J. Tait Elder 2011 Herring Bones in Southeast Alaska Archaeological Sites. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 281\u00E2\u0080\u0093291. University of Alaska Press, Fairbanks. Moss, Madonna L. and Aubrey Cannon (editors) 2011 The Archaeology of North Pacific Fisheries. University of Alaska Press, Fairbanks. Moss, Madonna L. and Jon M. Erlandson 2001 The Archaeology of Obsidian Cove, Semez Island, Southeast Alaska. Arctic Anthropology 38(1):27\u00E2\u0080\u009347. Moss, Madonna L., Dongya Y. Yang, Seth D. Newsome, Camilla F. Speller, Iain McKechnie, Alan D. McMillan, Robert J. Losey and Paul Koch 2006 Historical Ecology and Biogeography of North Pacific Pinnipeds: Isotopes and Ancient DNA from Three Archaeological Assemblages. Journal of Island and Coastal Archaeology 1(2):165\u00E2\u0080\u0093190. Muckle, Robert J. 1994 Differential Recovery of Mollusk Shell From Archaeological Sites. Journal of Field Archaeology 21(1):129\u00E2\u0080\u0093131. Munro, J. Alex and Wilbert Amie Clemens 1931 Water Fowl in Relation to the Spawning of Herring in British Columbia. Bulletin of the Biological Board of Canada 17. Biological Board of Canada, Toronto. Nagaoka, Lisa 1994 Differential Recovery of Pacific Island Fish Remains: Evidence from the Moturakau Rockshelter, Aitutaki, Cook Islands. Asian Perspectives 33(1):1\u00E2\u0080\u009317. Nelson, Margaret A., Randall F. Schalk, Teresa Trost, Michael Wolverton and Jana Boersema \u00E2\u0080\u0093 344 \u00E2\u0080\u0093 2010a Archaeological Testing At Weaverling Spit, Site 45-SK-43, Skagit County, Washington Final Technical Report. Cascadia Archaeology Report prepared for Samish Indian Nation and the Department of Archaeology and Historic Preservation, Olympia. Nelson, Margaret A., Teresa Trost and Michael Wolverton 2010b Data Recovery Excavations at 45-IS-119, Penn Cove, Whidbey Island, Island County, Washington. Cascadia Archaeology Report prepared for Penn Cove Shelfish and the Department of Archaeology and Historic Preservation, Olympia. Newcombe, Charles F. and W. A. Newcombe 1914 Sea-lions on the Coast of British Columbia. Report of the Commission of Fisheries for 1913, pp. 131\u00E2\u0080\u0093139, Province of British Columbia. Newell, Dianne 1999 Overlapping Territories and Entwined Cultures: Voyage into the Northern BC Spawn-on-Kelp Fishery. In Fishing Places, Fishing People: Traditions and Issues in Canadian Small-Scale Fisheries, edited by D. Newell and R. E. Ommer, pp. 121\u00E2\u0080\u009344. University of Toronto Press, Toronto. Newsome, Seth D., Michael A. Etnier, Diane Gifford-Gonzalez, Donald L. Phillips, Marcel van Tuinen, Elizabeth A. Hadly, Daniel P. Costa, Douglas J. Kennett, Tom P. Guilderson and Paul L. Koch 2007 The Shifting Baseline of Northern Fur Seal Ecology in the Northeast Pacific Ocean. Proceedings of the National Academy of Sciences, USA 104(23):9709\u00E2\u0080\u00939714. Nichol, Linda M., Melissa Boogaards and Robin Abernethy 2009 Recent trends in the abundance and distribution of sea otters (Enhydra lutris) in British Columbia. DFO Canadian Scientific Advisory Secretariat Stock Status Report. vol. 2009/016. NOAA Fisheries Service 2009 California Sea Lion (Zalophus californianus), http://www.nmfs.noaa.gov/pr/species/mammals/pinnipeds/californiasealion.htm. Office of Protected Resources, National Oceanic and Atmospheric Administration, Silver Spring, MD. Nookmiis, Chief Louie 2009 How the Huu-ay-aht Own Their Land\u00E2\u0080\u0093 Translation. In The Whaling Indians : West Coast Legends and Stories: Family Origin Histories, Part 11 of the Sapir-Thomas Nootka Texts. Told by Tyee Bob, Sa:ya:ch'apis, William Qwishanishim, Lot:tisim, Tayi:?a, and Chief Louie Nookmiis and Prepared by Edward Sapir, Morris Swadesh, Hamilton George, Alexander Thomas, Frank Williams, Katies Fraser, and John Thomas, edited by E. Y. Arima, H. Kammler, T. Klokeid, K. Robinson, R. Dennis and K. Neary, pp. 324\u00E2\u0080\u0093328. Mercury Series. vol. Ethnology 145. Canadian Museum of Civilization, Ottawa. O'Neill, Casey 2007 Archaeological Impact Assessment Grouse Nest Resort \u00E2\u0080\u0093 East Sooke, British Columbia, Permit 2007-12. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. O'Neill, Casey and Ian R. Wilson \u00E2\u0080\u0093 345 \u00E2\u0080\u0093 2005 Archaeological Mitigation DdRt-6, Sidney Island, B.C., Alteration Permit 2005-172. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. O\u00E2\u0080\u0099Connell, Jacqueline M. and Verena Tunnicliffe 2001 The Use of Sedimentary Fish Remains for Interpretation of Long-term Fish Population Fluctuations. Marine Geology 174(1-4):177\u00E2\u0080\u0093195. Olesiuk, Peter F. 1999 An Assessment of Status of Harbour Seals (Phoca vitulina) in British Columbia. Canadian Stock Assessment Secretariat Research Document 99/33, Ottawa. 2009a Abundance of Steller sea lions (Eumetopias jubatus) in British Columbia Research Document 2008/063. Canadian Science Advisory Secretariat, Ottawa. 2009b Preliminary Assessment of the Recovery Potential of Northern Fur Seals (Callorhinus ursinus) in British Columbia, Updated February 2009 Research Document 2007/076. Canadian Science Advisory Secretariat, Ottawa. Olesiuk, Peter F., Michael A. Bigg, Graham M. Ellis, Susan J. Crockford and Rebecca J. Wigen 1990 An Assessment of the Feeding Habits of Harbour Seals (Phoca vitulina) in the Strait of Georgia, British Columbia, Based on Scat Analysis. Can. Tech. Rep. Fish. Aquat. Sci: 1730. Oliver, Jeff 2007 Beyond the Water\u00E2\u0080\u0099s Edge: Towards a Social Archaeology of Landscape on the Northwest Coast. Canadian Journal of Archaeology 31(1):1\u00E2\u0080\u009327. Oliver, Lindsay and Julie Park 1997 Archaeological Excavation and Monitoring: Archaeological Site DgRw-4, El Verano Drive, Gabriola Island, B.C., Permit 1997-283. Aegis Associates report submitted to the BC Archaeology Branch, Victoria. Orchard, Trevor J. 2003 An Application of the Linear Regression Technique for Determining Length and Weight of Six Fish Taxa: the Role of Selected Fish Species in Aleut Paleodiet. International Series 1172. British Archaeological Reports, Oxford. 2007 Otters and Urchins: Continuity and Change in Haida Economy during the Late Holocene and Maritime Fur Trade Periods. PhD. Dissertation, University of Toronto. 2008 Mid-to-Late Holocene Haida Subsistence and Settlement in Gwaii Haanas: Final Report on 2007 Field Activities. Report on file, Gwaii Haanas National Park Reserve and Parks Canada Agency Northern and Western Service Centre, Skidegate and Calgary. 2009 Otters and Urchins: Continuity and Change in Haida Economy during the Late Holocene and Maritime Fur Trade Periods. International Series 2027. British Archaeological Reports, Archaeopress, Oxford. 2011a Appendix 1: Mid- to Late-Holocene Subsistence and Settlement in Gwaii Haanas: Analysis of Auger, Column, and Faunal Samples from 2009 and 2010 Fieldwork. In 2010/11 Gwaii Haanas Archaeology Programme, edited by D. W. Fedje, J. Cohen, R. G. Commisso \u00E2\u0080\u0093 346 \u00E2\u0080\u0093 and N. F. Smith, pp. Appendix 1. Prepared by Cultural Resource Services Northern and Western Service Centre, Parks Canada Agency, Calgary. 2011b Late Holocene Fisheries in Gwaii Haanas: Species Composition, Trends in Abundance, and Environmental or Cultural Explanations In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 111\u00E2\u0080\u0093128. University of Alaska Press, Fairbanks. Orchard, Trevor J. and Terence Clark 2005 Multidimensional Scaling of Northwest Coast Faunal Assemblages: A Case Study from Southern Haida Gwaii, British Columbia. Canadian Journal of Archaeology 29(2):88\u00E2\u0080\u0093112. Orchard, Trevor J. and Paul Szpak 2011 Identification of Salmon Species from Archaeological Remains on the Northwest Coast. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 17\u00E2\u0080\u009329. University of Alaska Press, Fairbanks. Orchard, Trevor J. and Rebecca J. Wigen 2008 Halibut Use on the Northwest Coast: Reconciling Ethnographic, Ethnohistoric, and Archaeological Data. Paper presented at the 2008 Northwest Anthropological Conference, Victoria, BC. Ortner, Sherry B. 1984 Theory in Anthropology Since the Sixties. Comparative Studies in Society and History 26(1):126\u00E2\u0080\u0093166. 2005 Subjectivity and Cultural Critique. Anthropological Theory 5(1):31\u00E2\u0080\u009352. Outram, D. N. and R. D. Humphreys 1974 The Pacific Herring in British Columbia Waters. Fisheries and Marine Service, Pacific Biological Station Nanaimo B.C. Circular 100. Paine, Robert T. 1974 Intertidal Community Structure: Experimental Studies on the Relationship between a Dominant Competitor and Its Principal Predator. Oecologia 15(2):93\u00E2\u0080\u0093120. Partlow, Megan A. 2006 Sampling Fish Bones: A Consideration of the Importance of Screen Size and Disposal Context in the North Pacific. Arctic Anthropology 43(1):67-79. Pauketat, Timothy R. 2001 A New Tradition in Archaeology. In The Archaeology of Traditions: Agency and History Before and After Columbus, edited by T. R. Pauketat, pp. 1\u00E2\u0080\u009316. University Press of Florida, Gainesville. Pauly, Daniel 1995 Anecdotes and the Shifting Baseline Syndrome of Fisheries. Trends in Ecology and Evolution 10(10):430. \u00E2\u0080\u0093 347 \u00E2\u0080\u0093 Pauly, Daniel, Ma. Lourdes Palomares, Rainer Froese, Pascualita Sa-a, Michael Vakily, David Preikshot and Scott Wallace 2001 Fishing Down Canadian Aquatic Food Webs. Canadian Journal of Fisheries and Aquatic Sciences 58(1):51\u00E2\u0080\u009362. Pawlowski, Drew 2008 Mill Street, Nanaimo: An Archaeological Impact Assessment, Permit 2008-61. Millennia Research Limited report prepared for the Snuneymuxw First Nation and the BC Archaeology Branch, Victoria. Pearse, Peter H. 1982 Turning the Tide: a New Policy for Canada\u00E2\u0080\u0099s Pacific Fisheries: Final Report, Commission on Pacific Fisheries Policy. Department of Fisheries and Oceans Canada, Vancouver. Pearson, Gordon W. and Minze Stuiver 1986 High-Precision Calibration of the Radiocarbon Time Scale, 500-2500 BC. Radiocarbon 28(2B):839\u00E2\u0080\u0093862. Pegg, Brian 1999 The Taphonomic History of the Vertebrate Faunal Assemblage from British Camp, San Juan Islands, Washington. MA Thesis, Simon Fraser University. Peros, Matthew C., Samuel E. Munoz, Konrad Gajewski and Andre E. Viau 2010 Prehistoric Demography of North America Inferred from Radiocarbon Data. Journal of Archaeological Science 37(3):656\u00E2\u0080\u0093664. Pierson, Nova 2011 Bridging Troubled Waters: Zooarchaeology and Marine Conservation on Burrard Inlet, Southwest British Columbia. MA Thesis, Simon Fraser University. Pikitch, Ellen K., P. Dee Boersma, Ian L. Boyd, David O. Conover, Philippe M. Cury, Timothy E. Essington, Selina S. Heppell, Edward D. Houde, Marc Mangel, Daniel Pauly, \u00C3\u0089va Plag\u00C3\u00A1nyi, Keith Sainsbury and Robert S. Steneck 2012 Little Fish, Big Impact: Managing a Crucial Link in Ocean Food Webs Lenfest Ocean Program. Washington, DC. 108 pp. Pinsky, Malin L., Olaf P. Jensen, Daniel Ricard and Stephen R. Palumbi 2011 Unexpected Patterns of Fisheries Collapse in the World\u00E2\u0080\u0099s Oceans. Proceedings of the National Academy of Sciences, USA 108(20):8317\u00E2\u0080\u00938322. Pitcher, Tony J. 2005 Back-to-the-Future: a Fresh Policy Initiative for Fisheries and a Restoration Ecology for Ocean Ecosystems. Philosophical Transactions of the Royal Society B 360(1453):107\u00E2\u0080\u0093121. Platten, Simon and Thomas Henfrey 2009 The Cultural Keystone Concept: Insights from Ecological Anthropology. Human Ecology 37(4):491\u00E2\u0080\u0093499. \u00E2\u0080\u0093 348 \u00E2\u0080\u0093 Popper, Virginia S. 1988 Selecting Quantitative Measures in Paleoethnobotany. In Current Paleoethnobotany: Analytical Methods and Cultural Interpretations of Archaeological Plant Remains, edited by C. A. Hastorf and V. S. Popper, pp. 53\u00E2\u0080\u009373. University of Chicago Press, Chicago. Poulsen, Bo 2008 Dutch Herring: An Environmental History, C. 1600-1860. Amsterdam University Press, Amsterdam. Powell, Miles 2012 Divided Waters: Heiltsuk Spatial Management of Herring Fisheries and the Politics of Native Sovereignty. Western Historical Quarterly 43(4):489\u00E2\u0080\u0093510. Racerocks.com 2010 http://racerocks.ca/racerock/mirounga2/feb72010newpup/070210pup.htm, Accessed Feb 15th, 2010. Rapp, George Jr., Eiler Henrickson and James Allert 1990 Native Copper Sources of Artifact Copper in Pre-Columbian North America. In Archaeological Geology of North America, edited by N. P. Lasca and J. Donahue, pp. 479\u00E2\u0080\u0093498. Geological Society of America, Boulder. Reimer, Paula J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, P. G. Blackwell, C. Bronk Ramsey, C. E. Buck, G. S. Burr, R. L. Edwards, M. Friedrich, P. M. Grootes, T. P. Guilderson, I. Hajdas, T. J. Heaton, A. G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, F. G. McCormac, S. W. Manning, R. W. Reimer, D. A. Richards, J. R. Southon, S. Talamo, C. S. M. Turney, J. van der Plicht and C. E. Weyhenmeyer 2009 Intcal09 and Marine09 Radiocarbon Age Calibration Curves, 0\u00E2\u0080\u009350,000 Years Cal BP. Radiocarbon 51(4):1111\u00E2\u0080\u00931150. Reimer, Paula J., Mike G. L. Baillie, Edouard Bard, Alex Bayliss, J. Warren Beck, Chanda J. H. Bertrand, Paul G. Blackwell, Caitlin E. Buck, George S. Burr, Kirsten B. Cutler, Paul E. Damon, R. Lawrence Edwards, Richard G. Fairbanks, Michael Friedrich, Thomas P. Guilderson, Alan G. Hogg, Konrad A. Hughen, Bernd Kromer, Gerry McCormac, Sturt Manning, Christopher Bronk Ramsey, Ron W. Reimer, Sabine Remmele, John R. Southon, Minze Stuiver, Sahra Talamo, F. W. Taylor, Johannes van der Plicht and Constanze E. Weyhenmeyer 2004 INTCAL04 Terrestrial Radiocarbon Age Calibration 0-26 Cal KYR BP. Radiocarbon 46(3):1029\u00E2\u0080\u00931058. Reimer, Paula J., Edouard Bard, Alex Bayliss, J. Warren Beck, Paul G. Blackwell, Christopher Bronk Ramsey, Caitlin Buck, R. Lawrence Edwards, Michael Friedrich, Pieter M. Grootes, Thomas P. Guilderson, Haflidi Haflidason, Irka Hajdas, Christine Hatt\u00C3\u00A9, Timonthy J. Heaton, Dirk L. Hoffman, Alan G. Hogg, Konrad A. Hughen, K. Felix Kaiser, Bernd Kromer, Sturt W. Manning, Mu Niu, Ron W. Reimer, David A Richards, E. Marian Scott, John R. Southon, Christian S. M. Turney and Johnannes van der Plicht 2013 IntCal13 and Marine13 radiocarbon age calibration curves 0-50,000 yr cal BP. Radiocarbon 55(4):1869-1887. \u00E2\u0080\u0093 349 \u00E2\u0080\u0093 Reimer/Yumks, Rudy 2011 The Mountains and Rocks are Forever: Lithics and Landscapes of Skwxw\u00C3\u00BA7mesh Uxwumixw Territory. PhD Dissertation, McMaster University. Reitz, Elizabeth J., C. Fred T. Andrus and Daniel H. Sandweiss 2008 Ancient Fisheries and Marine Ecology of Coastal Peru. In Human Impacts on Ancient Marine Ecosystems: A Global Perspective, edited by T. C. Rick and J. M. Erlandson, pp. 125\u00E2\u0080\u0093145. University of California Press, Berkeley. Reitz, Elizabeth J. and Elizabeth S. Wing 2008 Zooarchaeology. 2nd ed. Cambridge University Press, Cambridge. Reitz, Elizabeth. J. 2014 Continuity and resilience in the central Georgia Bight (USA) fishery between 2760 BC and AD 1580. Journal of Archaeological Science in press. Richards, Michael P., Sheila Greer, Lorna T. Corr, Owen Beattie, Alexander P. Mackie, Richard P. Evershed, Al von Finster and John R. Southon 2007 Radiocarbon Dating and Dietary Stable Isotope Analysis of Kwaday Dan Ts'inchi. American Antiquity 72(4):719\u00E2\u0080\u0093734. Richardson, Allan and Brent Galloway 2011 Nooksack Place Names: Geography, Culture, and Language. UBC Press, Vancouver. Rick, Torben C., Jon M. Erlandson, Michael A. Glassow and Madonna L. Moss 2002 Evaluating the Economic Significance of Sharks, Skates, and Rays (Elasmobranchs) in Prehistoric Economies. Journal of Archaeological Science 29:111\u00E2\u0080\u0093122. Rick, Torben C., Gregory A. Henkes, Darrin L. Lowery, Steven M. Colman and Brendan J. Culleton 2012 Marine Radiocarbon Reservoir Corrections (\u00E2\u0088\u0086R) for Chesapeake Bay and the Middle Atlantic Coast of North America. Quaternary Research 77(1):205\u00E2\u0080\u0093210. Rick, Torben C. and Rowan Lockwood 2013 Integrating Paleobiology, Archeology, and History to Inform Biological Conservation. Conservation Biology 27(1):45\u00E2\u0080\u009354. Robinson, Clifford L. K. and Daniel M. Ware 1999 Simulated and Observed Response of the Southwest Vancouver Island Pelagic Ecosystem to Oceanic Conditions in the 1990s. Canadian Journal of Fisheries and Aquatic Sciences 56(12):2433\u00E2\u0080\u00932443. Rodrigues, Antonia, Iain McKechnie and Dongya Y. Yang 2013 Ancient DNA Analysis of Archaeological Rockfish (Sebastes spp.) on the Northwest Coast of North America: Species Preferences and Implications for Marine Protected Area Management. Paper presented at the Society for American Archaeology 78th Annual Meeting, April 4, Honolulu, Hawaii. Rosenberg, Andrew A., W. Jeffrey Bolster, Karen E. Alexander, William B. Leavenworth, Andrew B. Cooper and Matthew G. McKenzie 2005 The History of Ocean Resources: Modeling Cod Biomass Using Historical Records. Frontiers in Ecology and the Environment 3(2):84\u00E2\u0080\u009390. \u00E2\u0080\u0093 350 \u00E2\u0080\u0093 Russell, Nerissa 2012 Social Zooarchaeology: Humans and Animals in Prehistory. Cambridge University Press, New York. Sahlins, Marshall D. 1976 Culture and Practical Reason. University of Chicago Press, Chicago. 1981 Historical Metaphors and Mythical Realities: Structure in the Early History of the Sandwich Islands Kingdom. Association for the Study of Anthropology in Oceania, Special Publication 1. University of Michigan Press, Ann Arbor. 2010 Infrastructuralism. Critical Inquiry 36(3):371\u00E2\u0080\u0093385. 2013 Structural Work: How Microhistories Become Macrohistories and Vice Versa. In Radical Egalitarianism: Local Realities, Global Relations, edited by F. Aulino, M. Goheen and S. J. Tambiah, pp. 161\u00E2\u0080\u0093190. Fordham University Press, Bronx. Salmond, Amiria and Anne Salmond 2010 Artefacts of Encounter. Interdisciplinary Science Reviews 35(3-4):302\u00E2\u0080\u0093317. Samuels, Stephan R. and Richard D. Daugherty 1991 Introduction to the Ozette Archaeological Project. In Ozette Archaeological Project Research Reports, Vol. 1, House Structure and Floor Midden, edited by S. R. Samuels, pp. 1\u00E2\u0080\u009327. Reports of Investigations. vol. 63. Washington State University, Pulman. Sapir, Edward 1910-1914 Nootka Field Notes Vols. I-XXIV. Originals in Boas Collection W2a-18 American Philosophical Society Library, Philadelphia. (Microfilm copy at Provincial Archives of British Columbia, Victoria). 1922 Sayach'apis, a Nootka Trader. In American Indian Life, By Several of Its Students, edited by E. C. Parsons, pp. 297\u00E2\u0080\u0093323. B.W. Huebsch, New York. n.d.-a Unpublished field notes, Notebook IV. In Manuscript held in Provincial Archives of British Columbia, Victoria. n.d.-b Unpublished field notes, Notebook XXIII. In Manuscript held in Provincial Archives of British Columbia, Victoria. Sapir, Edward and Morris Swadesh 1939 Nootka Texts: Tales and Ethnological Narratives with Grammatical Notes and Lexical Materials. 1978 AMS Press, ed. Special Publications of the Linguistic Society of America. Lingusitic Society of America, University of Pennsylvania, Philadelphia. 1955 Native Accounts of Nootka Ethnography. 1978 ed. AMS Press, New York. Schalk, Randall F. and Margaret A. Nelson (editors) 2010 The Archaeology of the Cama Beach Shell Midden, Camano Island Washington. Report Prepared for Washington State Parks and Recreation Commisson, Cascadia Archaeology, Seattle. \u00E2\u0080\u0093 351 \u00E2\u0080\u0093 Schiffer, Michael B. 1986 Radiocarbon dating and the \u00E2\u0080\u009COld Wood\u00E2\u0080\u009D problem: the case of the Hohokam chronology. Journal of Archaeological Science 13(1):13\u00E2\u0080\u009330. 1987 Formation Processes of the Archaeological Record. University of New Mexico Press, Albuquerque, NM. Schmidt, D. 1999 A Review of California Mussel Fisheries Biology and Fisheries Programes. Canadian Stock Assesment Secretariat Research Document 99/187. Schweigert, Jacob F., Jennifer L. Boldt, Linnea Flostrand and Jaclyn S. Cleary 2010 A Review of Factors Limiting Recovery of Pacific Herring Stocks in Canada. ICES: Journal of Marine Sciences 67(9):1903\u00E2\u0080\u00931913. Schweigert, Jacob F. and M. Linekin 1990 The Georgia and Johnstone Straits Herring Bait Fishery in 1986: Results of a Questionnaire Survey. Canadian Technical Report of Fisheries and Aquatic Sciences 1721. Department of Fisheries and Oceans Biological Sciences Branch, Pacific Biological Station, Nanaimo. Seed, R. and T.H. Suchanek 1992 Population and Community Ecology of Mytilus. In The Mussel Mytilus: Ecology, Physiology, Genetics and Culture, edited by E. G. Gosling. Elsevier, Amsterdam. Sellers, Ian 2013 A Historic Archaeology of Nuu-chah-nulth Barkley Sound: Material and Economic Change Through the Nineteenth Century. MA Thesis, Simon Fraser University. Shantry, Kate, Brandy Rinck and Ross E. Smith 2010 Results of Testing at Xaxtl\u00E2\u0080\u0099abish 1 (45-PI-974), Hylebos Creek, Pierce County, WA. Northwest Archaeological Associates report prepared for Wildlands, Inc. and the Department of Archaeology and Historic Preservation, Olympia. Silliman, Stephen W. 2012 Between the Longue Dur\u00C3\u00A9e and the Short Pur\u00C3\u00A9e: Postcolonial Archaeologies of Indigenous History in Colonial North America. In Decolonizing Indigenous Histories: Exploring Prehistoric/Colonial Transitions in Archaeology, edited by M. Oland, S. M. Hart and L. Frink, pp. 113\u00E2\u0080\u0093131. University of Arizona Press, Tucson. Simonsen, Bjorn O. 1991 Results of Archaeological Excavations at Site DkSf-4: The Finneron Property, Permit 1990-135. Bastion Group Heritage Consultants report submitted to the BC Archaeology Branch, Victoria. Smith, Anthony D. M., Christopher J. Brown, Catherine M. Bulman, Fulton. Elizabeth A., Penny Johnson, Isaac C. Kaplan, Hector Lozano-Montes, Steven Mackinson, Martin Marzloff, Lynne J. Shannon, Yunne-Jai Shin and Jorge Tam 2011 Impacts of Fishing Low\u00E2\u0080\u0093Trophic Level Species on Marine Ecosystems. Science 333:1147\u00E2\u0080\u00931150. \u00E2\u0080\u0093 352 \u00E2\u0080\u0093 Smith, Bruce D. 2011 General Patterns of Niche Construction and the Management of \u00E2\u0080\u0098Wild\u00E2\u0080\u0099 Plant and Animal Resources by Small-scale Pre-industrial Societies. Proceedings of the Royal Society of London B 366(1566):836\u00E2\u0080\u0093848. Smith, Michael E. 1992 Braudel\u00E2\u0080\u0099s Temporal Rhythms and Chronology Theory in Archaeology. In Archaeology, Annales, and Ethnohistory, edited by A. B. Knapp, pp. 23\u00E2\u0080\u009334. Cambridge University Press, Cambridge. Smith, Nicole F., Iain McKechnie and Denis E. St. Claire 2012 Kakmakimilh \u00E2\u0080\u0093 Keith Island Wharf and Gazebo Installation: An Archaeological Impact Assessment. Report Prepared for Tseshaht First Nation and Pacific Rim National Park Reserve of Canada. Smith, Ross E., Virginia L. Butler, Shelia Orwoll and Catherine Wilson-Skogen 2011a Pacific Cod and Salmon Structural Bone Density: Implications for Interpreting Butchering Patters in North Pacific Archaeofaunas. In The Archaeology of North Pacific Fisheries, edited by M. L. Moss and A. Cannon, pp. 45\u00E2\u0080\u009356. University of Alaska Press, Fairbanks. Smith, Ross E., Amanda K. Taylor, Brandy Rinck and Robert Kopperl 2011b Results of Archaeological Excavations at 45IS263, Madrona Way, Island County, Washington. Northwest Archaeological Associates Inc. report prepared for Washington State Department of Archaeology and Historic Preservation, Olympia. Smylie, Mike 2004 Herring: A History of the Silver Darlings. Tempus, Stroud, UK. Southon, John R. 2011 Are the Fractionation Corrections Correct: Are the Isotopic Shifts for 14C/12C Ratios in Physical Processes and Chemical Reactions Really Twice Those for 13C/12C? Radiocarbon 53(4):691\u00E2\u0080\u0093704. Southon, John R. and Daryl W. Fedje 2003 A Post-glacial Record of 14C Reservoir Ages for the British Columbia Coast. Canadian Journal of Archaeology 27(1):95\u00E2\u0080\u0093111. Southon, John R., D. Earle Nelson and John S. Vogel 1990 A Record of Past Ocean-Atmosphere Radiocarbon Differences from the Northeast Pacific. Paleoceanography 5(2):197\u00E2\u0080\u0093206. Spady, Ryan and Rebecca J. Wigen 2008 Archaeological Data Recovery and Monitoring DfSj-57 Storm Outfall Construction, Peninsula Road, Spring Cove, Ucluelet, B.C. Permit 2008-88. IR Wilson report submitted to BC Archaeology Branch, Victoria. Speller, Camilla F., Lorenz Hauser, Dana Lepofsky, Daniel Peterson, Jason Moore, Antonia Rodriguez, Madonna Moss, Iain McKechnie and Dongya Y. Yang 2012 High Potential for Using DNA from Ancient Herring Bones to Inform Modern Fisheries Management and Conservation. PLoS ONE 7(11):e51122. \u00E2\u0080\u0093 353 \u00E2\u0080\u0093 Springer, A. M., J. A. Estes, G. B. van Vliet, T. M. Williams, D. F. Doak, E. M. Danner, K. A. Forney and B. Pfisterg 2003 Sequential Megafaunal Collapse in the North Pacific Ocean: An Ongoing Legacy of Industrial Whaling? Proceedings of the National Academy of Sciences, USA 100(21):12223\u00E2\u0080\u009312228. Sproat, Gilbert Malcom 1868 The Nootka: Scenes and Studies of Savage Life. Smith, Elder, & Co., London. St. Claire, Denis E. 1991 Barkley Sound Tribal Territories. In Between Ports Alberni and Renfrew: Notes on West Coast Peoples, edited by E. Y. Arima, D. E. St. Claire, L. Clamhouse, J. Edgar, C. Jones and J. Thomas, pp. 13\u00E2\u0080\u0093202. Mercury Series. vol. 121. Canadian Ethnology Service, Canadian Museum of Civilization, Ottawa. Steele, James 2010 Radiocarbon Dates as Data: Quantitative Strategies for Estimating Colonization Front Speeds and Event Densities. Journal of Archaeological Science 37(8):2017\u00E2\u0080\u00932030. Steffen, Martina L. 2006 Early Holocene Hearth Features and Burnt Faunal Assemblages at the Richardson Island Archaeological Site, Haida Gwaii, British Columbia. MA Thesis, University of Victoria. Stein, Julie K. 1984 Interpreting Stratigraphy in Northwest Coast Shell Middens. Tebiwa: Journal of Idaho Museum of Natural History 21:26\u00E2\u0080\u009334. 1996 Shell Middens. In The Oxford Companion to Archaeology, edited by B. M. Fagan, pp. 638\u00E2\u0080\u0093639. Oxford University Press, Oxford. Stein, Julie K., Jennie N. Deo and Laura S. Phillips 2003 Big Sites\u00E2\u0080\u0094Short Time: Accumulation Rates in Archaeological Sites. Journal of Archaeological Science 30(3):297\u00E2\u0080\u0093316. Stein, Julie K. and Laura S. Phillips 2002 Vashon Island Archaeology: A View from Burton Acres Shell Midden. Burke Museum of Natural History and Culture Research Report 8. University of Washington Press, Seattle. Stewart, Brent S. 1999 Northern Elephant Seal. In Smithsonian Book of North American Mammals, edited by D. E. Wilson and S. Ruff, pp. 217\u00E2\u0080\u0093218. UBC Press and Smithsonian Institution Press, Vancouver and Washington, D.C. Stewart, Kathlyn M., Gary Coupland and D. Naughton 2003 Effects of Screen Size on Fish Element Recovery in Northern Northwest Coast Middens. Canadian Zooarchaeology Suppl. 1.(Transitions in Zooarchaeology):53\u00E2\u0080\u009363. \u00E2\u0080\u0093 354 \u00E2\u0080\u0093 Stewart, Kathlyn M., Frances L. Stewart and Gary Coupland 2009 Boardwalk, Northern Northwest Coast, Canada\u00E2\u0080\u0094 A New Face to an Old Site. Canadian Journal of Archaeology 33(2):205\u00E2\u0080\u0093233. Stewart, Kathlyn M. and Rebecca J. Wigen 2003 Screen Size and the Need for Reinterpretation: A Case Study from the Northwest Coast. Florida Museum of Natural History Bulletin 44(1):27-34. Stick, Kurt C. and Adam Lindquist 2009 2008 Washington State Herring Stock Status Report. Washington Department of Fish and Wildlife, Fish Management Division. Storey, Jennifer 2010 Archaeological Impact Assessment of Proposed Subdivision along Northumberland Channel, Cable Bay (DgRw-46), Southern Nanaimo, BC, Permit 2009\u00E2\u0080\u0093239. Madrone Environmental Services Ltd. report prepared for Island Timberlands and submitted to the BC Archaeology Branch, Victoria. Stryd, Arnoud H., Richard Brolly, Joanne Curtin, John Dewhirst and Diane K. Hanson 1991 Archaeological Investigations at the Little Beach Site, Ucluelet, BC. Arcas Consulting Archaeologists report submitted to the BC Archaeology Branch, Victoria. Stryd, Arnoud H., D. Geordie Howe, Richard P. Brolly, Randy Bouchard, Karla D. Kusmer and Dana Lepofsky 1993 Archaeological Investigations at the Cypress Street Site (DhRx-66), Nanaimo, B.C., Permit 1993-25. Arcas Consulting Archeologists Ltd report prepared for Imperial Oil Limited Toronto and submitted to the BC Archaeology Branch, Victoria. Stryd, Arnoud H., Geordie Howe, A. Joanne Curtin, Randy Bouchard, Dorothy Kennedy, Karla Kusmer, Dana Lepofsky, Karen Preckel, June Ryder, Susan Crockford, Rebecca J. Wigen, Brian Chisholm, Richard Brolly and Beth Bedard 1994 Archaeological Investigations At Tsawwassen, B.C., Volume 2. Arcas Consulting Archeologists Ltd report prepared for the Ministry of Transportation and Highways and submitted to the BC Archaeology Branch, Permits 1989-41 and 1990-2. Stuiver, Minze 1982 A High-Precision Calibration of the AD Radiocarbon Time Scale. Radiocarbon 24(1):1\u00E2\u0080\u009326. Stuiver, Minze and Thomas F. Braziunas 1993 Modeling Atmospheric 14C Influences and Radiocarbon Ages of Marine Samples to 10,000 BC. Radiocarbon 35:137-189. Stuiver, Minze and Gordon W. Pearson 1986 High-Precision Calibration of the Radiocarbon Time Scale, A.D. 1950-500 B.C. and 2000-6000 BC. Radiocarbon 28(2B):805\u00E2\u0080\u0093838. Stuiver, Minze and Henry A. Polach 1977 Discussion: Reporting of 14C Data. Radiocarbon 19(3):355\u00E2\u0080\u0093363. \u00E2\u0080\u0093 355 \u00E2\u0080\u0093 Stuiver, Minze, Paula J. Reimer and Thomas F. Braziunas 1998 High-precision Radiocarbon Age Calibration for Terrestrial and Marine Samples. Radiocarbon 40(3):1127\u00E2\u0080\u00931151. Sumpter, Ian D. 1999 1998 Archaeological Investigations, Site 922T, Gunlai Kin (Hotspring Island), Gwaii Hannas National Park Reserve and Haida Heritage Site. Report on file Cultural Resource Services, Parks Canada Western and Northern Service Centre, Calgary. 2003 Kiix?in Village Mapping and Dating Project, Kiix?in National Historic Site, Revised 2006. Manuscript on file at Cultural Resource Services, Northern and Western Service Centre, Parks Canada Agency, Calgary. 2005 An Analysis of Three Shellfish Assemblages from Ts'ishaa, site DfSi-16 (204T), Benson Island, Pacific Rim National Park Reserve. In Ts'ishaa: Archaeology and Ethnography of a Nuu-chah-nulth Origin Site in Barkley Sound, edited by A. D. McMillan and D. E. St. Claire, pp. 136\u00E2\u0080\u0093172, R. Carlson, general editor. Archaeology Press, Simon Fraser University, Burnaby. 2012 Invertebrate Fauna Analysis \u00E2\u0080\u0093 Huu7ii Village, Diana Island, Barkley Sound. In Huu7ii: Household Archaeology at a Nuu-chah-nulth Village Site in Barkley Sound, edited by A. D. McMillan and D. E. St. Claire, pp. 191\u00E2\u0080\u0093197. Archaeology Press, Simon Fraser University, Burnaby, BC. Sumpter, Ian D. and Daryl W. Fedje 2007 2005-06 Archaeological Resource Management Programme: Gulf Islands National Park Reserve of Canada. Report on file, Cultural Resource Services, Northern and Western Service Centre, Parks Canada Agency, Calgary. Sumpter, Ian D. and Denis E. St. Claire 2007 2006/07 Archaeological Resource Management Programme, Pacific Rim National Park Reserve of Canada. Report on file at Cultural Resource Services, Parks Canada Northern and Western Service Centre, Parks Canada Agency, Calgary. 2009 2008/09 Archaeological Resource Management Programme, Pacific Rim National Park Reserve of Canada. Report on file at Cultural Resource Services, Northern and Western Service Centre, Parks Canada Agency, Calgary. Surovell, Todd A. and P. Jeffrey Brantingham 2007 A Note on the Use of Temporal Frequency Distributions in Studies of Prehistoric Demography. Journal of Archaeological Science 34(11):1868\u00E2\u0080\u00931877. Suttles, Wayne P. 1952 Notes on Coast Salish Sea Mammal Hunting. Anthropology in British Columbia 3:10\u00E2\u0080\u009320. 1968 Coping With Abundance: Subsistence on the Northwest Coast. In Man the Hunter, edited by R. B. Lee and I. DeVore, pp. 55\u00E2\u0080\u009368. Aldine, Chicago. 1987 Variation in Habitat and Culture on the Northwest Coast. In Coast Salish Essays, pp. 26\u00E2\u0080\u009344. University of Washington Press, Seattle. \u00E2\u0080\u0093 356 \u00E2\u0080\u0093 Swadesh, Morris 1948 Motivations in Nootka Warfare. Southwestern Journal of Anthropology 4(1):76\u00E2\u0080\u009393. Swan, James G. 1883 Report of Investigations at Neah Bay, Wash., Respecting the Habits of Fur Seals of that Vicinity, and to Arrange for Procuring Specimens of Skeletons of Cetacea. Bulletin of the United States Fish Commission 3:201\u00E2\u0080\u0093207. 1887 The Fur Seal Industry of Cape Flattery, Washington Territory. In Fisheries and Fishery Industries of the United States, edited by G. B. Goode, pp. 393-400. vol. 7, 8. U. S. Government Printing Office, Washington, D.C. Swan, James Gilchrist 1870 The Indians of Cape Flattery. Smithsonian Contributions to Knowledge 16, Washington D.C. Szaniszlo, Wendy Renee 2005 California Sea Lion (Zalophus californianus) and Steller Sea Lion (Eumetopias jubatus) Interactions with Vessels in Pacific Rim National Park Reserve: Implications for Marine Mammal Viewing Management. MSc Thesis, University of Victoria. Szpak, Paul 2011 Fish Bone Chemistry and Ultrastructure: Implications for Taphonomy and Stable Isotope Analysis. Journal of Archaeological Science 38(12):3358\u00E2\u0080\u00933372. Szpak, Paul, Trevor J. Orchard and Darren R. Gr\u00C3\u00B6cke 2009 A Late Holocene Vertebrate Food Web from southern Haida Gwaii (Queen Charlotte Islands, British Columbia). Journal of Archaeological Science 36:2734\u00E2\u0080\u00932741. Szpak, Paul, Trevor J. Orchard, Iain McKechnie and Darren R. Gr\u00C3\u00B6cke 2012 Historical Ecology of Late Holocene Sea Otters (Enhydra lutris) from Northern British Columbia: Isotopic and Zooarchaeological Perspectives. Journal of Archaeological Science 39(5):1553\u00E2\u0080\u00931571. Szpak, Paul, Trevor J. Orchard, Anne K. Salomon and Darren R. Gr\u00C3\u00B6cke 2013 Regional Ecological Variability and Impact of the Maritime Fur Trade on Nearshore Ecosystems in Southern Haida Gwaii (British Columbia, Canada): Evidence from Stable Isotope Analysis of Rockfish (Sebastes spp.) bone collagen. Archaeological and Anthropological Sciences 5(2):159\u00E2\u0080\u0093182. Tanasichuk, Ron W. 1997 Influence of Biomass and Ocean Climate on the Growth of Pacific Herring (Clupea pallasi) from the Southwest Coast of Vancouver Island. Canadian Journal of Fisheries and Aquatic Sciences 54(12):2782\u00E2\u0080\u00932788. Taylor, Amanda K., Julie K. Stein and Stephanie A. E. Jolivette 2011 Big Sites, Small Sites, and Coastal Settlement Patterns in the San Juan Islands, Washington, USA. Journal of Island and Coastal Archaeology 6(2):287\u00E2\u0080\u0093313. \u00E2\u0080\u0093 357 \u00E2\u0080\u0093 Taylor, F. H. C. 1955 The Pacific Herring (Clupea pallasi) Along the Pacific Coast of Canada. International North Pacific Fisheries Commission Bulletin 1:105\u00E2\u0080\u0093128. 1969 The British Columbia Offshore Herring Survey, 1968-69. Fisheries Research Board of Canada technical report 140. Teeri, J. A. and L. G. Stowe 1976 Climatic patterns and the distribution of C4 grasses in North America. Oecologia 23(1):1\u00E2\u0080\u009312. Telford, Richard J., E. Heegaard and H. J. B. Birks 2004 The Intercept is a Poor Estimate of a Calibrated Radiocarbon Age. The Holocene 14(2):296\u00E2\u0080\u0093298. Tester, Albert L. 1933 Populations of Herring in the Coastal Waters of British Columbia. Transactions of the American Fisheries Society 63:286\u00E2\u0080\u0093289. 1935 The Herring Fishery of British Columbia--Past and Present. Bulletin 47. Biological Board of Canada, Ministy of Fisheries, Ottawa. Therriault, Thomas W., Douglas E. Hay and Jacob F. Schweigert 2009 Biologic Overview and Trends in Pelagic Forage Fish Abundance in the Salish Sea (Strait of Georgia, British Columbia). Marine Ornithology 37:3\u00E2\u0080\u00938. Thomas, David Hurst 1978 The Awful Truth about Statistics in Archaeology. American Antiquity 43(2):231\u00E2\u0080\u0093244. Thomas, Julian 2012 Some Deposits are More Structured Than Others. Archaeological Dialogues 19(2):124\u00E2\u0080\u0093127. Thomson, Richard E. 1981 Oceanography of the British Columbia Coast. Canadian Special Publication of Fisheries and Aquatic Sciences Vol. 56. National Research Council, Ottawa. Thomson, Richard E. , Steven F. Mih\u00C3\u00A1ly and Evgueni A. Kulikov 2007 Estuarine Versus Transient Flow Regimes in Juan de Fuca Strait. Journal of Geophysical Research-Oceans 112:C09022. Thornton, Thomas F. 2008 Being and Place Among the Tlingit. Studies in Anthropology and Environment. University of Washington Press, Seattle. 2012 Haa l\u00C3\u00A9elk'w h\u00C3\u00A1s aan\u00C3\u00AD saax'\u00C3\u00BA: Our grandparents' names on the land. Sealaska Heritage Institute and University of Washington Press Juneau and Seattle. \u00E2\u0080\u0093 358 \u00E2\u0080\u0093 Thornton, Thomas F., Virginia L. Butler, Fritz Funk, Madonna L. Moss, Jamie Hebert and J. Tait Elder 2010a Herring Synthesis: Documenting and Modeling Herring Spawning Areas within Socio-Ecological Systems over Time in the Southeastern Gulf of Alaska. Report Submitted to the North Pacific Research Board, Project #728. Thornton, Thomas F., Madonna L. Moss, Virginia L. Butler, Jamie Herbert and Fritz Funk 2010b Local and Traditional Knowledge and the Historical Ecology of Pacific Herring in Alaska. Journal of Ecological Anthropology 14(1):81\u00E2\u0080\u009388. Trigger, Bruce G. 1991 Distinguished Lecture in Archeology: Constraint and Freedom \u00E2\u0080\u0093 A New Synthesis for Archeological Explanation. American Anthropologist 93(3):551\u00E2\u0080\u0093569. Trosper, Ronald L. 2009 Resilience, Reciprocity and Ecological Economics: Northwest Coast Sustainability. Routledge, London and New York. Trost, Teresa 2005 Forgotten Waters: A Zooarchaeological Analysis of the Cove Cliff Site (DhRr 18), Indian Arm, British Columbia. MA Thesis, Simon Fraser University. Trost, Teresa, Deborah L. Olson and Rebecca J. Wigen 2010 Appendix B: Vertebrate Remains Recovered from the Cama Beach Shell Midden. In The Archaeology of the Cama Beach Shell Midden, Camano Island Washington, edited by R. F. Schalk and M. A. Nelson. Cascadia Archaeology report prepared for Washington State Parks and Recreation Commisson and the Department of Archaeology and Historic Preservation, Olympia. Tunnicliffe, Verena, Jacqueline M. O\u00E2\u0080\u0099Connell and M. R. McQuoid 2001 A Holocene Record of Marine Fish Remains from the Northeastern Pacific. Marine Geology 174:197\u00E2\u0080\u0093210. Turner, Nancy J. and Fikret Berkes 2006 Coming to Understanding: Developing Conservation through Incremental Learning in the Pacific Northwest. Human Ecology 34(4):495\u00E2\u0080\u0093513. Turner, Nancy J., Robin Gregory, Cheryl Brooks, Lee Failing and Terre Satterfield 2008 From Invisibility to Transparency: Identifying the Implications. Ecology and Society 13(2):7. Ulm, Sean 2002 Marine and Estuarine Reservoir Effects in Central Queensland, Australia: Determination of \u00E2\u0088\u0086R Values. Geoarchaeology 17(4):319\u00E2\u0080\u0093348. 2006 Australian Marine Reservoir Effects: A Guide to \u00E2\u0088\u0086R Values. Australian Archaeology 63(December):57\u00E2\u0080\u009360. Valentine, Lisa Philips and Regna Darnell 1999 Introductions: Timely Conversations. In Theorizing the Americanist Tradition, edited by L. P. Valentine and R. Darnell, pp. 3\u00E2\u0080\u009318. University of Toronto Press, Toronto. \u00E2\u0080\u0093 359 \u00E2\u0080\u0093 van Gaalen, Hermina A. 1991 Faunal Remains Analyzed From Sites, DgRw-199 and DgRw-204, Gabriola Island, British Columbia. In Gabriola Island Burial Recovery Project: Report on the 1989 Field Season, Permit 1989-92, edited by A. J. Curtin. Prepared for the British Columbia Heritage Trust and submitted to the BC Archaeology Branch, Victoria. Vansina, Jan M. 1985 Oral Tradition as History. University of Wisconsin Press, Madison. Villagran, Ximena S., Daniela M. Klokler, Silvia Peixoto, Paulo DeBlasis and Paulo C. F. Giannini 2011 Building Coastal Landscapes: Zooarchaeology and Geoarchaeology of Brazilian Shell Mounds. Journal of Island and Coastal Archaeology 6(2):211\u00E2\u0080\u0093234. Walker, Sara L. (editor) 2003 Archaeological Investigations at Sites 45SJ169 and 45SJ165, Decatur Island, San Juan County, Washington. 100\u00C2\u00B7118. Eastern Washington University, Spokane. Ware, Daniel M. and Jacob F. Schweigert 2002 Metapopulation Dynamics of British Columbia Herring During Cool and Warm Climate Regimes. Canadian Science Advisory Secretariat Research Document 2002/107. Fisheries and Oceans Canada, Ottawa. Ware, Daniel M. and Richard E. Thomson 2005 Bottom-Up Ecosystem Trophic Dynamics Determine Fish Production in the Northeast Pacific. Science 308:1280\u00E2\u0080\u00931284. Waterman, Thomas Talbot 1920 Whaling Equipment of the Makah Indians. Publications in Political and Social Science Vol. 1. University of Washington, Seattle. Weathers, Beth, Mikael Larsson and Jennifer Nord 2007 Archaeological Excavation and Monitoring, DdRu-18 Eventide Road Outfall, North Saanich, B.C., Permit 2006-444. IR Wilson Consultants Ltd. report by prepared for District of North Saanich and the BC Archaeology Branch, Victoria. Weathers, Beth, Rebecca J. Wigen and Kira Kristensen 2008 Archaeological Investigation and Data Recovery Peninsula Road upgrade Little Beach Site (DfSj-100). Ucluelet, B.C., Permit 2007-332. IR Wilson Consultants report submitted to the BC Archaeology Branch, Victoria. Webb, R.L. 1988 On the Northwest: Commercial Whaling in the Pacific Northwest (1790-1967). University of British Columbia, Vancouver. Weiser, Andrea L. and Dana Lepofsky 2009 Ancient Land Use and Management of Ebey's Prairie, Whidbey Island, Washington. Journal of Ethnobiology 29(2):184\u00E2\u0080\u0093212. \u00E2\u0080\u0093 360 \u00E2\u0080\u0093 Wessen, Gary C. 2004 A Final Report of Archaeological Site Testing at Potlatch State Park (45MS50), Skokomish Indian Reservation, Washington. Wessen and Associates report prepared for the Skokomish Tribal Historic Preservation Office Skokomish Indian Tribe and the Department of Archaeology and Historic Preservation, Olympia. Whitaker, Adrian R. 2008 Incipient Aquaculture in Prehistoric California?: Long-term Productivity and Sustainability vs. Immediate Returns for the Harvest of Marine Invertebrates. Journal of Archaeological Science 35(4):1114\u00E2\u0080\u00931123. Wigen, Rebecca J. 1990 Identification and Analysis of Vertebrate Fauna from Eighteen Archaeological Sites on the Southern Queen Charlotte Islands. Report Submitted to the British Columbia Heritage Trust, on file at the BC Archaeology Branch, Victoria. 1995 Appendix 5: Fish 45-KI-428 and 45-KI-429. In The Archaeology of West Point, Seattle, Washington: 4000 Years of Hunter-Fisher-Gatherer Land Use in Southern Puget Sound, edited by L. L. Larson and D. E. Lewarch, pp. A5.1\u00E2\u0080\u0093A5.78. Larson Anthropological/Archaeological Services report prepared for KIng County Department of Metropolitan Services, Seattle. 1999a Analysis of Fauna from Second Beach, FhTx-19 Haida Gwaii. Pacific Identifications Inc. report on file at the zooarchaeology Laboratory, Department of Anthropology, University of Victoria. 1999b Report on fauna from EdSv-4, Quattishe, pp. 27. Pacific Identifications report submitted to Sources Archaeological & Heritage Consultants, Vancouver, BC and on file at the Zooarchaeology Laboratory, Dept. of Anthropology, University of Victoria, Victoria, B.C. 2000 Appendix 1. Faunal Analysis: Identification and Analysis of Faunal Remains from Eagle's Nest (DcRt-71). In Eagles Nest Archaeological Data Recovery DcRt-71, Permit 1999-289, edited by M. Eldridge, pp. 37\u00E2\u0080\u009354. Millennia Research Limited report prepared for International Integration Inc. and submitted to the Songhees First Nation and the BC Archaeology Branch, Victoria, Victoria, B.C. 2002 Appendix B: Corporation of Delta Faunal Report, DgRs-14 and DgRs-1. In Corporation of Delta 2001 Capital Works: Archaeological Reviews, Permit 2001-268, edited by L. C. Ham. Report prepared for the Corporation of Delta and submitted to the Tsawwassen First Nation, Semiahmoo First Nation, Stolo Nation (permit #2001-21) and the BC Archaeology Branch, Victoria. 2003a Appendix D. Fish Remains from Sites 45-SJ-169 and 45-SJ-165. In Archaeological Investigations at Sites -SJ-169 and 45-SJ-165, Decatur Island, San Juan County, Washington, edited by S. L. Walker, pp. 393. Reports in Archaeology and History. vol. 100-118. Eastern Washington University, Spokane. \u00E2\u0080\u0093 361 \u00E2\u0080\u0093 2003b Report on Vertebrate Fauna from Kiix7in (DeSh-1 and DeSh-2). In 2002 Kiix7in Village Mapping and Dating Project, Kiix7in National Historic Site, edited by I. D. Sumpter. Pacific Identifications Inc. Report prepared for Cultural Resource Services, Northern and Western Service Centre, Parks Canada Agency, Calgary. 2005 Vertebrate Fauna. In The Hoko River Archaeological Site Complex: The Rockshelter (45-CA-21), 1000--100 B.P., Olympic Peninsula, Washington, edited by D. R. Croes, pp. 71\u00E2\u0080\u0093104. Washington State University Press, Pullman, WA. 2007a Appendix B: Vertebrate Faunal Report for Site 1656T (DeRt-109), North Pender. In 2005-06 Archaeological Resource Management Programme: Gulf Islands National Park Reserve of Canada, edited by I. D. Sumpter and D. W. Fedje, pp. 94\u00E2\u0080\u0093114. Report prepared for Cultural Resource Services, Northern and Western Service Centre, Parks Canada Agency, Calgary. 2007b Pat Bay Project #1836 Faunal Report. Pacific Identifications report submitted to IR Wilson Consultants, July 4, 2007. 2008a Faunal Report for Little Beach, Project #1993 Monitoring Samples. Report Submitted to IR Wilson, October 9, 2008. 2008b Little Beach Faunal Report. Pacific Identifications report submitted to IR Wilson Consultants, March 5, 2008. 2009 Faunal Analysis for the Dodd Island Site (72T). In 2008/09 Archaeological Resource Management Programme, Pacific Rim National Park Reserve of Canada, edited by I. D. Sumpter and D. E. St. Claire, pp. 125\u00E2\u0080\u0093146. Cultural Resource Services, Northern and Western and Service Centre, Parks Canada Agency, Calgary. 2013 Faunal Analysis for Kakmakimilh, Keith Island, 306T (DfSh-17). In Kakmakimilh \u00E2\u0080\u0093 Keith Island Wharf and Gazebo Installation: An Archaeological Impact Assessment., edited by N. F. Smith, I. McKechnie and D. S. Claire. Report Prepared by for Tseshaht First Nation and Pacific Rim National Park Reserve of Canada, Ucluelet. Wigen, Rebecca J., Susan J. Crockford and Ruth L. Greenspan 1990 Comparison of Taphonomic Patterns in Fish Bones from Northwest Coast (North American) Sites. Paper presented at the Sixth International Conference on Archaeozoology, Washington D.C. Wigen, Rebecca J. and Barbara R. Stucki 1988 Taphonomy and Stratigraphy in the Interpretation of Economic Patterns at Hoko River Rockshelter. In Research in Economic Anthropology, Supplement 3: Prehistoric Economies of the Pacific Northwest Coast, edited by B. L. Isaac, pp. 87\u00E2\u0080\u0093146. JAI Press, Greenwich. Wilk, Richard and Livia Barbosa 2012 Rice and Beans: A Unique Dish in a Hundred Places. Berg, New York. Willerton, Ila Moana 2009 Subsistence at Si\u00E2\u0080\u00A2\u00C4\u008D\u00C9\u0099\u00D3\u0099\u00E2\u0080\u0099n\u00C9\u0099\u00D3\u0099\u00C5\u0082: The Willows Beach Site and the Culture History of Southeastern Vancouver Island. MA Thesis, University of Victoria. \u00E2\u0080\u0093 362 \u00E2\u0080\u0093 William 2009 The Origin of the Ts'isha:?ath and the Story of Tlatala:Qokw'ap and Ta:Posh?in?is \u00E2\u0080\u0093 Translation. In The Whaling Indians : West Coast Legends and Stories: Family Origin Histories, Part 11 of the Sapir-Thomas Nootka Texts. Told by Tyee Bob, Sa:ya:ch'apis, William, Qwishanishim, Lot:tisim, Tayi:?a, and Chief Louie Nookmiis and prepared by Edward Sapir, Morris Swadesh, Hamilton George, Alexander Thomas, Frank Williams, Katies Fraser, and John Thomas, edited by E. Y. Arima, H. Kammler, T. Klokeid and K. Robinson, pp. 117\u00E2\u0080\u0093123, 369\u00E2\u0080\u0093370. Mercury Series. vol. Ethnology 145. Canadian Museum of Civilization, Ottawa. Williams, Alan N. 2012 The Use of Summed Radiocarbon Probability Distributions in Archaeology: A Review of Methods. Journal of Archaeological Science 39(3):578\u00E2\u0080\u0093589. 2013 A New Population Curve for Prehistoric Australia. Proceedings of the Royal Society B: Biological Sciences 280(1761):1\u00E2\u0080\u00939. Williams, Raymond 2005 Ideas of Nature. In Culture and Materialism: Selected Essays, pp. 67\u00E2\u0080\u009384. 2nd ed. Verso, London. Willows, Erin 2009 Archaeological Monitoring and Data Recovery DgRw-34 Residential Redevelopment Murdoch Road Cedar, B.C., Alteration Permit 2008-358. IR Wilson Consultants Ltd. report prepared submitted to the BC Archaeology Branch, Victoria. Willows, Erin, Rebecca J. Wigen, Ryan Spady and Andrew Hickok 2012 Archaeological Testing, Data Recovery and Monitoring DgRw-34 Murdoch Road Cedar, B.C., Permits 2010-17, 2010-97, 2010-227. I. R. Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Wilson, Ian R. 1988 Archaeological Test Excavations At DeRu-44, Fulford Harbour, Saltspring Island, Permit 1988-61. IR Wilson Consultants report prepared for the Ministry of Transportation and Highways and submitted to the BC Archaeology Branch, Victoria. 1994 Archaeological Investigations at DgSl-67 Chesterman Beach, West Coast Vancouver Island, Permit 1993-148. IR Wilson Consultants report prepared for Sur La Mer Developments and submitted to the BC Archaeology Branch, Victoria. 2003 Emergency Archaeological Salvage DiSc-26 Qualicum Beach, Vancouver Island, Alteration Permit 2003-224. IR Wilson Consultants Ltd. report prepared for Town of Qualicum Beach and submitted to the BC Archaeology Branch, Victoria. 2005 Archaeological Monitoring at Lochside Drive Utility Trench, DdRu-4, Sidney, B.C., Permit 2004-413. IR Wilson Consultants Ltd. Report prepared for the Town of Sidney and submitted to the BC Archaeology Branch, Victoria. \u00E2\u0080\u0093 363 \u00E2\u0080\u0093 Wilson, Ian R., Shane Bond and Cynthia Lake 2003 Archaeological Excavation & Monitoring DcRu-75, Dallas Road, Victoria, B.C. Permits 2003-187 and 2003-219. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Wilson, Ian R., Shane Bond, Cynthia Lake and Margaret Rogers 2004a Archaeological Excavation and Monitoring DcRt-16, King George Terrace, Victoria, B.C., Permit 2003-232. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Wilson, Ian R. and Terence Clark 2003 Archaeological Mitigation and Site Alteration DkSf-40, Courtenay River Watermain Project, Courtenay, B.C., Permit 2003-120 and Alteration Permit 2003-121. IR Wilson Consultants Ltd. report submitted the the BC Archaeology Branch, Victoria. Wilson, Ian R. and Susan J. Crockford 1994 Public Archaeological Excavations at the Departure Bay Midden, DhRx-16; Permit 1992-29. IR Wilson Consultants report prepared for Nanaimo District Museum and the BC Archaeology Branch, Victoria. Wilson, Ian R., Jennifer Storey, Ian Streeter, Andrew Hickok and Rebecca J. Wigen 2006 Archaeological Excavation and Monitoring DhSb-3, Dogwood Street, Parksville, B.C. Permit 2006-10. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Wilson, Ian R., Rebecca J. Wigen and Pierre Friele 1993 Excavations at EeSu-13: Port Hardy, Vancouver Island, Permit 1992-92. IR Wilson Consultants Ltd. report submitted to the BC Archaeology Branch, Victoria. Wilson, Ian R., Rebecca J. Wigen and Margaret Rogers 2004b Archaeological Column Sample Excavation DiSe-7, Deep Bay Parking Lot, Deep Bay, B.C., Permit 2004-55. IR Wilson Consultants Ltd. report prepared for Department of Fisheries and Oceans, Pacific Region and the BC Archaeology Branch, Victoria. Wolverton, Steve and R. Lee Lyman (editors) 2012 Conservation Biology and Applied Zooarchaeology. University of Arizona Press, Tucson. Worm, Boris and Derek P. Tittensor 2011 Range Contraction in Large Pelagic Predators. Proceedings of the National Academy of Sciences, USA 108(29):11942\u00E2\u0080\u009311947. Wright, Cynthia A., Audrey Dallimore, Richard E. Thomson, R. Timothy Patterson and Daniel M. Ware 2005 Late Holocene Paleofish Populations in Effingham Inlet, British Columbia, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 224(4):367\u00E2\u0080\u0093384. \u00E2\u0080\u0093 364 \u00E2\u0080\u0093 Wylie, Alison 1989 Archaeological Cables and Tacking: The Implications of Practice for Bernstein's Options Beyond Objectivism and Relativism. Philosophy of the Social Sciences 19(1):1\u00E2\u0080\u009318. 2009 Legacies Of Collaboration: Transformative Criticism In Archaeology. Paper presented at the American Anthropological Association as the Patty Jo Watson Distinguished Lecture, San Francisco. Yoneda, Minoru, Yasuyuki Shibata, Masatoshi Morita, Masashi Hirota, Ryo Suzuki, Kazuhiro Uzawa, Naoyuki Ohshima and Yukio Dodo 2004 Interspecies Comparison of Marine Reservoir Ages at the Kitakogane Shell Midden, Hokkaido, Japan. Nuclear Instruments and Methods in Physics Research B 223\u00E2\u0080\u0093224:376\u00E2\u0080\u0093381. \u00E2\u0080\u0093 365 \u00E2\u0080\u0093 Appendices For Chapter 2 Table 2.1 Archaeological data for Pacific herring and associated chronological information for all sites (N = 171). Relative abundance and rank order data based on number of identified skeletal specimens (NISP). Table 2.2 Individual assemblage data for the 50 sites with within site-variability presented in Figure 2.6 Numbers on left correspond with sites listed in in Table 2.1. Table 2.3 Distance between archaeological sites dating to within the last 2,500 years and documented spawning localities mapped by federal fisheries monitoring efforts (ca. 1930-2012) and available data Traditional and Local Ecological Knowledge (TEK/LEK) datasets. Numbers on left column correspond with sites listed in Table 2.1. 366 Table 2.1 Archaeological data for Pacific herring and associated chronological information for all sites (N = 171). Relative abundance and rank order data based on number of identified skeletal specimens (NISP). No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 1 SS SVI DcRu-75 (Wilson, et al. 2003:85, Table 32 ) 682 13 42% 2 300-1200 (Wilson, et al. 2003:93) 2 SS SVI DcRu-78 (Wigen 2009:3, Tables 2 & 3) 61 4 90% 1 Historic/300-1200* (Mitchell 1981) 3 SS SVI DcRu-4 (Stewart and Wigen 2003:29, Table 1) 903 13 64% 1 300-1200 Gajewski et al. 2011 (Gajewski, et al. 2011) 4 SS SVI DcRt-16 (Wilson, et al. 2004a:114, Table 45) 746 13 79% 1 300-1200 Wilson et al. 2004:121 (Wilson, et al. 2004a) 5 SS SVI DcRt-10 (Willerton 2009:150, Table 12.4) 3167 13 77% 1 300-1200/1200-2400 Willerton 2009:36 (Willerton 2009) 6 SS SVI DcRu-71 (Wigen 2000:11, Table 5) 78 6 81% 1 1200-2400/2400-5000* Eldridge 2000:33 (Eldridge 2000) 7 SS SVI DcRw-57 (Bowie and Wigen 2010:42, Table 2 ) 205 10 0.5% 7 300-1200* Bowie & Wigen 2010a:48 (Bowie and Wigen 2010) 8 SS SVI DcRv-21 (O'Neill 2007:40, Table 3) 57 5 77% 1 300-1200* O'Neil 2007:28 (O'Neill 2007) 9 SS SVI DdRu-81 (Bowie, et al. 2010:164, Table 29) 1815 20 75% 1 300-1200 Bowie et al. 2010:184 (Bowie, et al. 2010) 10 SS SVI DdRu-4 (Wilson 2005:154, Table 67 ) 638 14 62% 1 300-1200 Cooper 2008 (Cooper 2008); Wilson 2005 (Wilson 2005) 11 SS SVI DeRu-1 (Hewer 1999:26, Table 4) 1589 17 86% 1 1200-2400 Hewer 1999:52 (Hewer 1999) 12 SS SVI DdRu-5 (Kanipe, et al. 2007:Appendix E, Table 11) 316 7 89% 1 300-1200 Kanipe et al. 2007:283-284 (Kanipe, et al. 2007) 13 SS SVI DdRu-18 Weathers et al. (Weathers, et al. 2007:31, Table 8) 57 10 21% 2 300-1200* Weathers et al. 2007:36 (Weathers, et al. 2007) 14 SS GI-BC DdRt-6 (O'Neill and Wilson 2005:47, Table 6) 597 12 93% 1 300-1200/1200-2400/2400-5000* O'Neil & Wilson 2005:52 (O'Neill and Wilson 2005) 15 SS GI-BC DeRu-44 (Wilson 1988:Appendix 1) 93 4 90% 1 1200-2400 Wilson 1988:44 (Wilson 1988) 16 SS GI-BC DfRu-3 (Brolly, et al. 1993:71, Table 7) 1155 9 88% 1 300-1200/1200-2400 Brolly et al. 1993:91-92 (Brolly, et al. 1993) 17 SS GI-BC DeRt-1 (Hanson 1995:36 & 39) 473 12 75% 1 300-1200 Hanson 1995:32 (Hanson 1995) 18 SS GI-BC DeRt-109 Wigen 2007(Wigen 2007a:103, Table 12) 521 13 85% 1 300-1200 Sumpter & Fedje 2007:62 (Sumpter and Fedje 2007) 19 SS GI-BC DgRu-3 (Ewonus 2011:77-78, Table 1) 12675 29 70% 1 1200-2400 Grier 2006:102 (Grier 2006) 367 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 20 SS GI-BC DgRv-4 (McLay 1999b: Appendix 5) 1681 9 96% 1 300-1200* McLay 1999:67 (McLay 1999b) 21 SS GI-BC DgRv-1 (McLay 1999b: Appendix 5) 246 9 86% 1 300-1200* McLay 1999:67 (McLay 1999b) 22 SS GI-BC DgRv-2 (McLay 1999b: Appendix 5) 404 6 93% 1 300-1200* McLay 1999:67 (McLay 1999b) 23 SS GI-BC DgRv-2 (Matson, et al. 1999:61-71) 2039 14 74% 1 300-1200 Matson 2003:85 (Matson 2003) 24 SS GI-BC DgRw-78 (McLay 1999b: Appendix 5) 59 6 39% 1 300-1200* McLay 1999:67 (McLay 1999b) 25 SS GI-BC DgRw-73 (McLay 1999b: Appendix 5) 646 7 89% 1 300-1200* McLay 1999:67 (McLay 1999b) 26 SS GI-BC DgRw-32 (McLay 1999b: Appendix 5) 80 8 55% 1 300-1200* McLay 1999:67 (McLay 1999b) 27 SS GI-BC DgRw-4 (Oliver and Park 1997:Appendix 3) 552 8 86% 1 300-1200/1200-2400 Burley 1989:33 (Burley 1989) 28 SS GI-BC DgRw-204 (van Gaalen 1991:Table 3) 631 8 72% 1 1200-2400 Curtin 1991:38 (Curtin 1991) 29 SS GI-BC DiSe-10 (Eldridge 1987:5-63-65, Table 5-13-15) 166 8 71% 1 2400-5000* Eldridge 1987:5-14 (Eldridge 1987) 30 SS EVI DjSf-13 (Mason and Hoffmann 1998:AIV-9, Table AIV-8) 221 7 92% 1 300-1200/1200-2400 Mason and Hoffmann 1998:Appendix V & VI (Mason and Hoffmann 1998) and Mitchell 1974 (Mitchell 1974) 31 SS EVI DkSf-26 (Mitchell 1988:17, Table 6) 4577 17 78% 1 1200-2400/2400-5000 Mitchell 1988:16 (Mitchell 1988) 32 SS EVI DkSf-40 (Wilson and Clark 2003:40, Table 10 ) 83 5 83% 1 300-1200 Wilson & Clark 2003:47-48 (Wilson and Clark 2003) 33 SS EVI DkSf-4 (Simonsen 1991:Appendix tables, pg 53 of 59) 2612 10 89% 1 1200-2400 Simonsen 1991:27 (Simonsen 1991) 34 SS EVI DkSf-19 (Caldwell 2008:120, Table 5.39) 2721 8 96% 1 1200-2400 Caldwell 2008:175 (Caldwell 2008) 35 SS EVI DkSf-29 (Hall, et al. 2003:83-93) 909 10 88% 1 300-1200/1200-2400 Hall et al. 2003:33 (Hall, et al. 2003) 36 SS EVI DiSc-26 (Wilson 2003:42 Table 13) 215 6 86% 1 300-1200 Wilson 2003:52-53 (Wilson 2003) 37 SS EVI DiSe-7 (Wilson, et al. 2004b:44, Table 5 ) 5638 14 94% 1 300-1200/1200-2400/2400-5000 Wilson et al. 2004:48-50 (Wilson, et al. 2004b) 38 SS EVI DhSb-3 (Wilson, et al. 2006:60, Table 18) 661 9 97% 1 1200-2400 Wilson et al. 2006:83-84 (Wilson, et al. 2006) 368 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 39 SS EVI DhSb-11 (Bowie and Wigen 2010:41, Table 3) 2464 10 97% 1 300-1200 Kristensen et al. 2009:87, (Kristensen, et al. 2009) 40 SS EVI DhRx-16 (Wilson and Crockford 1994:105, Table 38) 2870 10 97% 1 1200-2400 Wilson & Crockford 1994:v (Wilson and Crockford 1994) 41 SS EVI DhRx-66 (Stryd, et al. 1993:70 ) 2578 9 97% 1 300-1200 Stryd et al. 1993:81 (Stryd, et al. 1993) 171 SS EVI DhRx-101 (Eldridge, et al. 2007:101-102, Table 8) 430 8 90% 1 300-1200/1200-2400 Eldridge et al. 2007:54 (Eldridge, et al. 2007) 42 SS EVI DhRx-104 Pawlowski 2008(Pawlowski 2008:16, Table 4 ) 195 3 89% 1 300-1200* Pawlowski 2008:17 (Pawlowski 2008) 43 SS EVI DgRw-46 (Storey 2010:44, Table 6) 65 3 97% 1 300-1200 Storey 2010:37 (Storey 2010) 44 SS EVI DgRw-34 (Willows 2009:39, Table 3; Willows, et al. 2012:118-119, Tables 14&15) 1668 17 87% 1 300-1200/1200-2400 Willows et al. 2012:138 (Willows, et al. 2012) 45 SS EVI DeRv-68 (Brunsden and Cooper 2008:17, Tables 2 & 3 combined) 456 5 64% 1 300-1200* Brunsden and Cooper 2008:10 (Brunsden and Cooper 2008) 46 SS ESG EaSe-18 (Caldwell 2013) 5715 10 85% 1 300-1200 Caldwell 2013 (Caldwell 2013). 47 SS ESG EaSe-11 (Caldwell 2013) 409 7 74% 1 300-1200 Caldwell 2013 (Caldwell 2013). 48 SS ESG EaSe-76 (Caldwell 2013) 705 9 37% 2 1200-2400 Caldwell 2013 (Caldwell 2013). 49 SS ESG DlSd-3 (Caldwell 2013) 595 10 93% 1 300-1200 Caldwell 2013 (Caldwell 2013). 50 SS ESG DlSd-6 (Caldwell 2013) 1410 8 93% 1 300-1200 Caldwell 2013 (Caldwell 2013). 51 SS ESG DlSe-10 (Caldwell 2013) 64 6 9% 4 300-1200 Caldwell 2013 (Caldwell 2013). 52 SS ESG DlSd-17 (Caldwell 2013) 680 9 93% 1 Historic/300-1200 Caldwell 2013 (Caldwell 2013). 53 SS ESG DjRw-1 (Caldwell 2013) 1791 10 97% 1 300-1200 Coupland et al. in prep. 54 SS FD DhRr-18 (Coupland 1991:88, Table 3) 5429 11 82% 1 1200-2400 Coupland 1989:71 (Coupland 1991) Coupland 1991:89 (Coupland 1989) 55 SS FD DhRt-6 (Brolly and Muir 1993:51, Table 11) 897 11 42% 1 2400-5000 Brolly & Muir 1993:31, (Brolly and Muir 1993) 56 SS FD DhRr-6 (Pierson 2011:77-78, Appendix 1 ) 28 300-1200/1200-2400/2400-5000 Pierson 2011:27 (Pierson 2011) 369 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 57 SS FD DhRr-18 (Trost 2005:111, Appendix C) 4640 19 57% 1 300-1200* Lepofsky et al. 2007:196 (Lepofsky, et al. 2007) 58 SS FD DhRq-1 (Pierson 2011:83-84, Appendix 1) 2098 22 40% 1 300-1200/1200-2400 Pierson 2011:31 (Pierson 2011) 59 SS FD DgRr-6 (Casteel 1976b:85, Table 5-2 ) 304 8 1% 4 1200-2400/2400-5000/5000-10000 Matson 1976 (Matson 1976) 60 SS FD DgRs-14 (Wigen 2002:1) 221 9 32% 1 1200-2400 Hammon 1986:97 (Hammon 1986) 61 SS FD DgRs-2 (Kusmer 1994b:133, Table 5-12; 1994a:197, Table 4) 1244 9 25% 3 300-1200/1200-2400/2400-5000 Kusmer 1994a (Kusmer 1994a) and Stryd et al 1994: volume 2:120 (Stryd, et al. 1994) 62 SS FD DgRs-1 (Brolly, et al. 1996:162 Table 13) 4589 10 22% 2 300-1200/1200-2400/2400-5000 Brolly et al. 1996:156, 180, & 209 (Brolly, et al. 1996) 63 SS FD DgRr-1 (Butler and Campbell 2004:349; Matson 1992: cited in Campbell and Butler 2004) 25659 12 6% 3 1200-2400/2400-5000 Matson 2010:1-2 (Matson 2010) 64 SS PSnd 45-KI-23 (Butler 1987:Table 10-1, pg. 10-2) 3999 16 0.1% 13 300-1200 Blunkis Onat 1987:5-7 (Blukis Onat 1987) 65 SS PSnd 45-KP-115 (Butler and Baker 2002:11, Table 2 cited in Campbell & Butler 2004:349; Butler and Campbell 2004) 806 8 43% 1 300-1200 Lewarch et al. 2002:48 (Lewarch, et al. 2002) 66 SS PSnd 45SK43 (Nelson, et al. 2010a:53, Table 7.5) 14800 11 25% 2 300-1200/1200-2400/2400-5000 Nelson et al. 2010a:33 (Nelson, et al. 2010a) 67 SS PSnd 45-KI-437 (Kopperl and Butler 2002:107, Table 10.1) 5321 16 80% 1 Historic/300-1200 Stein & Phillips 2002:58 (Stein and Phillips 2002) see also Kopperl 2001:3 (Kopperl 2001) 68 SS GI-WA 45SJ24 (Pegg 1999:67, Table 3.1 ) 18654 11 68% 1 300-1200 Pegg 1999:19 (Pegg 1999) and Stein et al. 2003:303 (Stein, et al. 2003) 69 SS GI-WA 45-SJ-169 (Wigen 2003a:283, Table D.7) 3223 24 52% 1 300-1200/1200-2400/2400-5000 Walker 2003:76, Table 6.4 (Walker 2003) 70 SS GI-WA 45-SJ-280 (Kopperl 2007:16, Table 1) 2450 14 6% 3 300-1200/1200-2400/2400-5000 Bovy et al. 2007:20 (Bovy and Kopperl 2007) 71 SS PSnd 45IS2 (Trost, et al. 2010:B8, Table B.10) 16154 31 3.80% 6 300-1200 Shalk and Nelson 2010:102, (Schalk and Nelson 2010) 72 SS PSnd 45IS263 (Smith, RE, et al. 2011b:23, Table 13) 401 15 6% 5 300-1200 Smith et al. 2011:I (Smith, RE, et al. 2011b) 73 SS PSnd 45-IS-119 (Nelson, et al. 2010b:30, Table 6.2) 160 10 1% 7 300-1200 Nelson et al. 2010b:41 370 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context (Nelson, et al. 2010b) 74 SS GI-WA 45SJ200 (Kopperl 2000:27 Table 1) 126 11 25% 1 300-1200* NA 75 SS GI-WA 45JS252 (Kenady 2000:11, Table 9) 373 6 19% 2 300-1200 Kenady 2000:Appdx. A (Kenady 2000) 76 SS PSnd 45KI428 & 29 (Wigen 1995:A5-66, A5-70, Table A5-13) 8057 29 2% 8 300-1200/1200-2400/2400-5000 Larson & Lewarch 1995: Table 6.1, pg 6-4\u00E2\u0080\u00936-5 (Larson and Lewarch 1995) 77 SS PSnd 45PI974 (Shantry, et al. 2010:54, Table 13) 1226 12 43% 1 300-1200 Shantry et al. 2010 (Shantry, et al. 2010) 78 SS PSnd 45MS50 (Wessen 2004:46-47, Tables 14 & 15) 1248 7 2% 7 300-1200/1200-2400 Wessen 2004:Appendix B (Wessen 2004) 79 WCVI WCVI EdSv-4 (Wigen 1999b:Table 3) 183 14 38.80% 1 Historic/300-1200/1200-2400* NA 80 WCVI WCVI DiSo-9 (Calvert 1980:298 & 302, Tables 45 & 49) 19426 22 48% 1 1200-2400 Calvert 1980:123, Table 9 (Calvert 1980) 81 WCVI WCVI DgSl-67 (Wilson 1994:18, Table 2) 68 10 57% 1 300-1200 Wilson 1994:27 (Wilson 1994) 82 WCVI BS DfSj-40 (Brolly and Pegg 1998:167) 757 12 79% 1 300-1200/1200-2400 Brolly & Pegg 1998:156 (Brolly and Pegg 1998) 83 WCVI BS DfSj-100 (Weathers, et al. 2008; Wigen 2008b:16, Table 6) 1021 15 79% 1 2400-5000 Stryd et al. 1991:37 (Stryd, et al. 1991) 84 WCVI BS DfSj-57 (Spady and Wigen 2008:49, Table 8) 1852 14 84% 1 300-1200 Spady & Wigen 2008:60 (Spady and Wigen 2008) 85 WCVI BS DfSi-16 & 17 (McKechnie 2005b:125, Appendix B1) 6979 21 53% 1 300-1200/1200-2400/2400-5000 McMillan & St. Claire 2005:45&77 (McMillan and St. Claire 2005) 86 WCVI BS DfSi-26 (McKechnie 2007b:29, Table 12) 157 7 58% 1 300-1200 (McKechnie 2007b:24) 87 WCVI BS 129T This Study 1738 17 61% 1 Historic/300-1200 Sumpter and St. Claire 2009:112 (Sumpter and St. Claire 2009) 88 WCVI BS 83T This Study 1997 16 47% 1 Historic/300-1200 McKechnie 2010:44 (McKechnie 2010b) 89 WCVI BS 82T This Study 1374 16 39% 2 300-1200 McKechnie 2010:20 (McKechnie 2010b) 90 WCVI BS 306T (Wigen 2013:7, Table 3 and This Study) 482 10 72% 1 Historic/300-1200 Smith et al. 2012 (Smith, et al. 2012) and McKechnie 2010:44 (McKechnie 2010b) 91 WCVI BS 304T This Study 1155 14 41% 2 Historic/300-1200 McKechnie 2010:28 371 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context (McKechnie 2010b) 92 WCVI BS 206T This Study 1939 17 41% 1 Historic/300-1200/1200-2400 McKechnie 2010:20 (McKechnie 2010b) 93 WCVI BS 72T (Wigen 2009:132,Table 6) 3725 18 85% 1 300-1200 McKechnie 2010:44 (McKechnie 2010b) 94 WCVI BS DeSh-1 (Wigen 2003b:Table 6) 700 17 48% 1 Historic/300-1200 Sumpter 2003:18 (Sumpter 2003) 95 WCVI BS DeSh-2 (Wigen 2003b:Table 9) 169 6 18% 2 Historic/300-1200 Sumpter 2003:18 (Sumpter 2003) 96 WCVI BS DfSh-7 (McKechnie 2012: Table 2, or Chapter 5) 32401 33 81% 1 300-1200/1200-2400/2400-5000 McMillan & St. Claire 2012:33&93 (McMillan and St. Claire 2012) 97 WCVI WCVI DeSf-6 (McKechnie 2007c:7) 258 10 11% 3 300-1200 McKechnie & Fedje 2010 (McKechnie and Fedje 2010) 98 CC NVI EeSu-13 (Wilson, et al. 1993:32, Table 6) 1121 15 67% 1 2400-5000 Wilson et al. 1993:69 (Wilson, et al. 1993) 99 CC Namu ElSx-3 (Cannon, et al. 2011:62, Table 5.1) 3356 14 48% 2 300-1200/1200-2400 Cannon 1997:15 (Cannon 1997) 100 CC Namu ElSx-5 (Cannon, et al. 2011:62, Table 5.1) 989 8 61% 1 300-1200/1200-2400/2400-5000 Cannon 2000:74 (Cannon 2000b); Cannon 1998:40 (Cannon 1998) 101 CC Namu ElSx-18 (Cannon, et al. 2011:62, Table 5.1) 567 9 48% 2 300-1200/1200-2400/2400-5000 Cannon 1998:37 (Cannon 1998) 102 CC Namu ElSx-10 (Cannon, et al. 2011:62, Table 5.1) 1433 9 62% 1 300-1200/1200-2400/2400-5000 Cannon 1997:45 (Cannon 1997) and Cannon 2000:72 (Cannon 2000b) 103 CC Namu ElTa-25 (Cannon, et al. 2011:62, Table 5.1) 105 6 35% 2 Historic/300-1200/1200-2400/2400-5000 Cannon 1997:45 (Cannon 1997) and Cannon 2000:72 (Cannon 2000b) 104 CC Namu ElSx-1 (Cannon, et al. 2011:62, Table 5.1) 5870 15 41% 2 300-1200/1200-2400/2400-5000/5000-10000 Carlson 1991 (Carlson 1991) 105 CC Namu EkSx-12 (Cannon, et al. 2011:62, Table 5.1) 1992 18 9.94% 2 300-1200/1200-2400 Cannon 1997:45 (Cannon 1997) 106 CC Namu ElTa-3 (Cannon, et al. 2011:62, Table 5.1) 137 5 12% 3 Historic/300-1200 Cannon 2000:72 (Cannon 2000b); Cannon 1997:45 (Cannon 1997) 107 CC Namu ElTb-1 (Cannon, et al. 2011:62, Table 5.1) 4802 13 86% 1 Historic/300-1200/1200-2400 Cannon 1998:22 (Cannon 1998) Cannon 2000:72 (Cannon 372 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 2000b); 108 CC Namu ElTb-2 (Cannon, et al. 2011:62, Table 5.1) 97 8 38% 2 300-1200/1200-2400 Cannon 1998:26 (Cannon 1998) 109 CC Rinlt EkSt-1 (Cannon, et al. 2011:62, Table 5.1) 2535 5 29% 2 N/A MacDonald et al. 2011:3623 (MacDonald, et al. 2011) 110 CC Rinlt EjSw-1 (Cannon, et al. 2011:62, Table 5.1) 3113 7 47% 1 N/A MacDonald et al. 2011:3623 (MacDonald, et al. 2011) 111 CC Rinlt EjSv-1 (Cannon, et al. 2011:62, Table 5.1) 122 3 35% 1 N/A MacDonald 2008:26 (MacDonald 2008) 112 CC Rinlt EjSv-2 (Cannon, et al. 2011:62, Table 5.1) 374 6 56% 1 N/A MacDonald 2008:26 (MacDonald 2008) 113 CC Rinlt EjSv-3 (Cannon, et al. 2011:62, Table 5.1) 140 3 40% 1 N/A MacDonald et al. 2011:3623 (MacDonald, et al. 2011) 114 CC Rinlt EjSv-4 (Cannon, et al. 2011:62, Table 5.1) 64 5 56% 1 N/A MacDonald 2008:26 (MacDonald 2008) 115 CC Rinlt EjSv-5 (Cannon, et al. 2011:62, Table 5.1) 328 4 66% 1 N/A MacDonald et al. 2011:3623 (MacDonald, et al. 2011) 116 CC Rinlt EjSv-8 (Cannon, et al. 2011:62, Table 5.1) 222 4 48% 1 N/A MacDonald 2008:26 (MacDonald 2008) 117 CC Rinlt EjSv-9 (Cannon, et al. 2011:62, Table 5.1) 551 5 42% 1 N/A MacDonald 2008:26 (MacDonald 2008) 118 CC Rinlt EjSv-10 (Cannon, et al. 2011:62, Table 5.1) 905 7 52% 1 N/A MacDonald 2008:26 (MacDonald 2008) 119 CC Rinlt EjSv-11 (Cannon, et al. 2011:62, Table 5.1) 127 5 42% 2 N/A MacDonald 2008:26 (MacDonald 2008) 120 HG S. HG FaTt-9 (Acheson 1998; Wigen 1990) 23469 15 16% 2 300-1200 Acheson 1998 (Acheson 1998) 121 HG S. HG 1325T (Fedje, et al. 2005a:195, Table 11.2) 751 11 0.10% 7 5000-10000 Fedje et al. 2001 (Fedje, et al. 2001), 2005 (Fedje, et al. 2005a) 122 HG S. HG 2008T (Orchard 2011a:23, Table 12) 254 9 1.60% 5 300-1200* Orchard 2011:22 (Orchard 2011a) 123 HG S. HG 699T (Orchard 2007:236, Table 7.10) 11651 12 28% 2 Historic/300-1200 Orchard 2007:223-227 (Orchard 2007) 124 HG S. HG 740T (Orchard 2007:236, Table 7.10) 112 10 13% 3 300-1200 Orchard 2007:223-227 (Orchard 2007) 125 HG S. HG 717T (Orchard 2007:236, Table 7.10) 6877 10 43% 2 300-1200 Orchard 2007:223-227 (Orchard 2007) 126 HG S. HG 924T (Orchard 2011a:9-10, Table 3&4 and this 35631 13 64.50 1 Historic/300- Fedje et al. 2011:37 (Fedje, 373 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context study) % 1200/1200-2400/2400-5000 et al. 2011a) and Orchard 2007:258, Table 7.16 (Orchard 2007) 127 HG S. HG 1370T (Orchard 2011a:17, Table 8 ) 149 8 24.20% 2 300-1200 Fedje & Smith 2009:5 (Fedje and Smith 2009) 128 HG S. HG 781T (Orchard 2007:236, Table 7.10; 2008:51-67, Tables D1-D10 & C1-C9) 3736 13 80% 1 Historic/300-1200 Orchard 2008:27, Table 3 (Orchard 2007) 129 HG S. HG 922T (Sumpter 1999; Wigen 1999a) 1927 13 2% 6 300-1200/1200-2400 Sumpter 1999:1 (Sumpter 1999) 130 HG S. HG 785T (Orchard 2007:236, Table 7.10 ) 6317 14 3% 4 Historic/300-1200/1200-2400 Orchard 2007:223-227 (Orchard 2007) 131 HG S. HG 1134T (Orchard 2007:237, Table 7.10) 904 5 0.00% absent Historic Orchard 2007:223-227 (Orchard 2007) 132 HG S. HG 1127T (Steffen 2006:65, Table 3.6 ) 1615 11 0.43% 8 5000-10000 Fedje et al. 2005:209 (Fedje, et al. 2005b) 133 HG N. HG FjUb 10 (Christensen and Stafford 2005) 1731 9 2% 6 2400-5000/5000-10000 Christensen & Stafford 2005:251 (Christensen and Stafford 2005) 134 HG N. HG GaUa 18 (Christensen, et al. 2010:69, Table 11) 1020 14 6% 5 300-1200/1200-2400 Christensen et al. 2010:54 (Christensen, et al. 2010) 135 NC Dundas GdTq-1 (Brewster and Martindale 2011:258, Table 15.4) 379 4 4% 3 2400-5000/5000-10000 Brewster and Martindale 2011:256-7 (Brewster and Martindale 2011) 136 NC Dundas GcTr-5 (Brewster and Martindale 2011:258, Table 15.4) 671 4 7% 3 1200-2400/2400-5000 Brewster and Martindale 2011:256-7 (Brewster and Martindale 2011) 137 NC Dundas GcTq-5 (Brewster and Martindale 2011:258, Table 15.4) 896 4 11% 2 1200-2400/2400-5000/5000-10000 Brewster and Martindale 2011:256-7 (Brewster and Martindale 2011) 138 NC Dundas GcTq-7 (Brewster and Martindale 2011:258, Table 15.4) 204 4 12% 3 300-1200/1200-2400/2400-5000* Martindale et al. 2010:73 (Martindale, et al. 2010) 139 NC Dundas GcTr-10 (Brewster and Martindale 2011:258, Table 15.4) 500 4 5% 3 300-1200/1200-2400/2400-5000* Martindale et al. 2010:73 (Martindale, et al. 2010) 140 NC Dundas GcTr-8 (Brewster and Martindale 2011:258, Table 15.4) 347 4 26% 2 1200-2400/2400-5000/5000-10000 Brewster and Martindale 2011:256-7, Table 15.3 141 NC Dundas GcTr-9 (Brewster and Martindale 2011:258, Table 15.4) 75 4 19% 2 300-1200/1200-2400/2400-5000* Martindale and Letham 2008:Figure 14 142 NC Dundas GcTq-1 (Brewster and Martindale 2011:258, Table 15.4) 382 4 13% 3 2400-5000/5000-10000 Brewster and Martindale 2011:256-7 (Brewster and 374 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context Martindale 2011) 143 NC Dundas GcTq-6 (Brewster and Martindale 2011:258, Table 15.4) 82 4 16% 2 2400-5000/5000-10000 Brewster and Martindale 2011:256-7 (Brewster and Martindale 2011) 144 NC Dundas GcTq-4 (Brewster and Martindale 2011:258, Table 15.4) 64 4 36% 2 2400-5000/5000-10000 Brewster and Martindale 2011:256-7 (Brewster and Martindale 2011) 145 NC PRupert GbTo-24 This Study 142 7 4% 4 no dates just yet Martindale in prep. 146 NC PRupert GcTo-6 This Study 155 5 10.30% 2 1200-2400 Coupland et al. 1993 (Coupland, et al. 1993) 147 NC PRupert GbTo-77 (Coupland, et al. 2010:199, Table 12) 83 24.10% 2 1200-2400/2400-5000 Coupland et al. 2010:193 (Coupland, et al. 2010) 148 NC PRupert GbTo-28 (Coupland, et al. 2010:198, Table 8) 742 21.20% 2 1200-2400 Coupland et al. 2010:193 (Coupland, et al. 2010) 149 NC PRupert GbTo-46 (Coupland, et al. 2010:198, Table 8) 419 11.00% 2 1200-2400 Coupland et al. 2010:193 (Coupland, et al. 2010) 150 NC PRupert GbTo-31 (Coupland, et al. 2010:197, Table 6) 374 3.20% 2 300-1200/1200-2400 Stewart et al. 2009 (Stewart, et al. 2009) 151 NC PRupert GbTo-13 (Eldridge, et al. 2008:117, Table 111) 284 7 16% 2 1200-2400 Eldridge et al. 2008:241 (Eldridge, et al. 2008) 152 NC PRupert GbTo-37 (Eldridge, et al. 2008:117, Table 111) 58 4 12% 2 2400-5000 Eldridge et al. 2008:241 153 SEAK Angoon 49-SIT-124 (Moss 1989:242-260; 2007b:8-10, Unit 1) 1655 14 12% 2 300-1200/1200-2400 Moss 1989:249 (Moss 1989) 154 SEAK Angoon 49-SIT-130 (Moss 1989:299-309) 1042 2 0% absent 300-1200 Moss 1989:299 (Moss 1989) 155 SEAK Angoon 49-SIT-132 (Moss 1989:203:Table 5.12 ) 356 6 41% 2 300-1200 Moss 1989:197 (Moss 1989) 156 SEAK Angoon 49-SIT-171 (Moss 1989: 214-224) 384 3 4% 2 300-1200 Moss 1989:217 (Moss 1989) 157 SEAK Angoon 49-SIT-244 (Moss 1989:175-177 ) 1730 9 55% 1 300-1200 Moss 1989:169 (Moss 1989) 158 SEAK Angoon 49-SIT-259 (Moss 1989:267-269 ) 63 5 3% 4 300-1200/1200-2400 Moss 1989:227 (Moss 1989) 159 SEAK Angoon 49-SIT-299 (Moss 1989: 326-334) 1109 9 1% 2 300-1200/1200-2400 Moss 1989:324 (Moss 1989) 160 SEAK Angoon 49-SIT-304 (Moss 1989: 273-282) 920 2 99.70% 1 300-1200 Moss 1989:277 (Moss 1989) 375 No. Reg-ion a Sub-region b Arch. Site No. Zooarchaeological Reference NISPc N Taxa % NISP Rankd Time Interval Years BPe Reference for 14C Date & stratigraphic context 161 SEAK Noyes 49-CRG-188 (Moss 2011b:160) 112 8 36.60% 1 300-1200/1200-2400 Moss 2004 (Moss 2004) 162 SEAK Sumez 49CRG-88 (Moss and Erlandson 2001:36-37, 40-41) 144 9 1% 7 300-1200 Moss & Erlandson 2001:35 (Moss and Erlandson 2001) 163 SEAK Dall 49-DIX-46 (Moss 2008:50, Table 5) 194 12 5% 4 1200-2400/2400-5000 Moss 2008:49 (Moss 2008) 164 SEAK Annette 49-KET-229 (Minor, et al. 1986) 618 8 17% 3 1200-2400 Minor et al. 1986 (Minor, et al. 1986) 165 SEAK PWales 49-PET-67 (Moss 2011b:160) 391 6 23% 2 2400-5000 Moss 2011b:160 (Moss 2011b) 166 SEAK PWales 49-PET-556 (Moss 2011b:160) 224 9 67% 1 1200-2400/2400-5000 Moss 2011b:160 (Moss 2011b) 167 SEAK Kuiu 49-XPA-29 (Maschner 1992:318-320) 10912 9 32% 2 300-1200/1200-2400 Maschner 1992:178 (Maschner 1992) 168 SEAK Kuiu 49-XPA-39 (Maschner 1992:318-320) 11726 7 66% 1 300-1200/1200-2400/2400-5000 Maschner 1992:178 (Maschner 1992) 169 SEAK Kuiu 49-XPA-106 (Maschner 1992:318-320) 515 4 3% 3 300-1200 Maschner 1992:178 (Maschner 1992) 170 SEAK Kuiu 49-XPA-112 (Maschner 1992:318-320) 548 4 1% 2 300-1200 Maschner 1992:178 (Maschner 1992) a -Regions: SS = Salish Sea, Including Puget Sound, Strait of Georgia, & the Strait of Juan de Fuca; WCVI = Western Vancouver Island; CC = Central BC Coast; HG = Haida Gwaii; NC = Northern BC Coast; SEAK = Southeast Alaska. b - Sub-Regions: PSnd = Puget Sound; GI-WA = Washington State Gulf Islands; GI-BC = BC Gulf Islands; SVI = Southern Vancouver Island; EVI = Eastern Vancouver Island; ESG = Eastern Strait of Georgia; FD = Fraser Delta; BS = Barkley Sound; WCVI = Western Vancouver Island; NVI = Northern Vancouver Island; Namu = Namu Area; Rinlt = Rivers Inlet Area; S. HG = Southern Haida Gwaii; N. HG = Northern Haida Gwaii; Dundas = Dundas Archipelago; PRupert = Prince Rupert Harbour; Angoon= Angoon Area; Noyes = Noyes Island; Sumez = Sumez Island; Dall = Dall Island; Pwales = Prince of Wales Island; Kuiu = Kuiu Island c \u00E2\u0080\u0093 NISP = Number of identified specimens. d \u00E2\u0080\u0093 Rank = Rank order of herring based on numbers of identified specimens. e \u00E2\u0080\u0093 Dates fall within these broad temporal intervals. Asterisk indicates no available 14C dates estimated from presence/absence of historic artifacts and /or relative sea level. 376 Table 2.2 Individual assemblage data for the 50 sites with within site-variability presented in Figure 2.6 Numbers on left correspond with sites listed in in Table 2.1. No. Reg-ion Sub-region Site N levels N Taxa Mean % SD % Variance Mean rank SD Reference 168 NC Kuiu Step Island, AK 4 7 58% 24% 5.9% 1.5 0.58 Maschner 1992:318-320 (Maschner 1992) 156 NC Angoon Yaay Shaanoow, AK 7 6 40% 11% 1.1% 1.7 0.49 Moss 1989:203:Table 5.12 (Moss 1989) 150 NC PR Boardwalk, PRH 3 27 3% 4% 0.1% 2.0 0.00 Coupland et al. 2010:197, Table 6 (Coupland, et al. 2010) 131 HG SHG Lyell Bay 5 14 4% 4% 0.2% 4.3 2.22 Orchard 2007:257, Table 7.16 (Orchard 2007) 128 HG SHG Kaidsu 3 13 74% 11% 1.1% 1.0 0.00 Orchard 2007:257, Table 7.16 (Orchard 2007) 126 HG SHG Burnaby Nrrws 16 11 51% 36% 12.7% 1.4 0.67 Orchard 2011:9, 39, Table 3 and B1-7 & this study (Orchard 2011a) 123 HG SHG Benjamin Pt. 5 12 20% 18% 3.4% 2.6 0.89 Orchard 2007:236, Table 7.10 (Orchard 2007) 104 CC Namu Namu 5 15 40% 4% 0.2% 2.0 0.00 Canon et al. 2011:62, Table 5.1 (Cannon, et al. 2011) 107 CC Namu Nulu 43 na 85% 11% 1.2% 1.0 0.00 Cannon n.d., Cannon 2002:320 (Cannon 2002) (Cannon, et al. 2011) 98 CC NVI Port Hardy 7 15 67% 11% 1.3% 1.0 0.00 Wilson 1993:32, Table 6 (Wilson, et al. 1993) 84 WCVI BS Spring Cove 5 14 70% 33% 11.1% 1.2 0.45 Spady and Wigen 2008:49, Table 8 (Spady and Wigen 2008) 83 WCVI BS Little Beach 8 15 69% 17% 2.9% 1.0 0.00 Weathers et al. 2008:46, Tables 7-8 (Weathers, et al. 2008) 82 WCVI BS Itasoo North 5 12 81% 7% 0.5% 1.0 0.00 Brolly and Pegg 1998:167, Table 40 (Brolly and Pegg 1998) 85 WCVI BS Ts'ishaa 38 21 54% 27% 7.2% 1.4 0.64 McKechnie 2005:125, Appendix B1 (McKechnie 2005b) 92 WCVI BS Wouwer 20 17 39% 21% 4.5% 1.6 0.88 this study 88 WCVI BS Lwr Dicebox 8 16 45% 21% 4.3% 1.6 0.74 this study 87 WCVI BS Upper Dicebox 6 17 62% 19% 3.7% 1.2 0.41 this study 91 WCVI BS Huumuuwaa 15 14 43% 20% 3.9% 1.7 0.50 this study 89 WCVI BS Shiwitis 12 16 45% 28% 8.0% 1.5 0.52 this study 93 WCVI BS Dodd 3 19 85% 11% 1.2% 1.0 0.00 Wigen 2009:Table 6 (Wigen 2009) 90 WCVI BS Keith 4 8 72% 12% 1.4% 1.0 0.00 Wigen 2013:7, Table 3 and this Study (Wigen 2013) 94 WCVI BS Kiix7in 6 17 58% 7% 0.5% 1.0 0.00 Wigen 2003:Table 6 (Wigen 2003b) 377 No. Reg-ion Sub-region Site N levels N Taxa Mean % SD % Variance Mean rank SD Reference 96 WCVI BS Huu7ii 115 33 67% 29% 8.2% 1.3 0.74 this study 35 SS EVI Kye Bay, Comox 8 10 87% 12% 1.4% 1.0 0.00 Hall et al. 2003:84-93, Tables 14-23 (Hall, et al. 2003) 34 SS EVI Q'umu?xs, Comox 10 8 96% 4% 0.1% 1.0 0.00 Caldwell 2008:120, Table 5.39 (Caldwell 2008) 33 SS EVI DkSf-4, Comox 7 10 89% 4% 0.1% 1.0 0.00 Simonsen 1991:Appendix tables (pg 53 of 59) (Simonsen 1991) 37 SS EVI Deep Bay 10 14 86% 20% 4.1% 1.1 0.32 Wilson et al. 2004:44, Table 5 (Wilson, et al. 2004b) 40 SS EVI Departure Bay 7 10 94% 4% 0.2% 1.0 0.00 Howe et al. 1994:90, Table 16 (Howe, et al. 1994) & Wilson & Crockford 1994:105, Table 38 (Wilson and Crockford 1994) 171 SS EVI Cypress St. Nanaimo 14 9 97% 3% 0.1% 1.0 0.00 Stryd et al. 1993:70 (Stryd, et al. 1993) 44 SS EVI Murdoch Rd., Cedar 17 17 89% 8% 0.7% 1.0 0.00 Willows et al. 2012:118-119, Tables 14&15 combined (Willows, et al. 2012) 11 SS SVI Blue Heron Rd., Sidney 13 17 85% 6% 0.3% 1.0 0.00 Hewer 1999:26, Table 4 (Hewer 1999) 12 SS SVI Patricia Bay 3 7 87% 9% 0.8% 1.0 0.00 Kanipe et al. 2007:Appendix E, Table 11 (Kanipe, et al. 2007) 5 SS SVI Willows Beach 14 13 82% 16% 2.5% 1.1 0.27 Willerton 2009:129, Table 11.4 (Willerton 2009) 52 SS ESG EaSe-18 12 10 78% 21% 4.2% 1.1 0.29 this study 47 SS ESG DlSd-17 4 9 90% 7% 0.5% 1.0 0.00 this study 46 SS ESG DlSd-3 8 10 94% 2% 0.0% 1.0 0.00 this study 27 SS GI-BC False Narrows 7 8 85% 8% 0.6% 1.0 0.00 Oliver and Park 1997:Appendix 3 (Oliver and Park 1997) 14 SS GI-BC Sidney Island 3 12 87% 12% 1.3% 1.0 0.00 O'Neil and Wilson 2005:47, Table 6 (O'Neill and Wilson 2005) 68 SS GI-WA British Camp 13 11 60% 17% 2.9% 1.1 0.28 Pegg 1999:67, Table 3.1 (Pegg 1999) 69 SS GI-WA Decatur Island 16 24 50% 23% 5.1% 1.3 0.58 Wigen 2003:287, Table D.11 (Wigen 2003a) 54 SS FD Point Grey 9 11 91% 10% 1.0% 1.0 0.00 Coupland 1989:72, Table 4 (Coupland 1989) 56 SS FD Tum-tumay-whuteon 9 23 26% 13% 1.8% 2.0 0.87 Pierson 2011:74-76, Appendix 1 (Pierson 2011) 58 SS FD Noons Creek 12 22 43% 15% 2.3% 1.3 0.45 Pierson 2011:81-84, Appendix 1 (Pierson 2011) 61 SS FD Tsawwassen 4 9 28% 26% 6.6% 2.5 1.29 Kusmer 1994a:197, (Kusmer 1994a), Kusmer 1994b:133, Tbl 5-12 (Kusmer 1994b) 378 No. Reg-ion Sub-region Site N levels N Taxa Mean % SD % Variance Mean rank SD Reference 62 SS FD Beach Grove 4 10 18% 8% 0.7% 2.8 0.50 Brolly et al. 1996:162 Table 13 (Brolly, et al. 1996) 63 SS FD Crescent Beach 3 12 5% 2% 0.0% 3.3 0.58 Matson 1992 (Matson 1992) cited in Campbell & Butler 2004:349 (Butler and Campbell 2004) 66 SS PSnd Weaverling Spit 4 11 32% 26% 6.7% 2.0 0.82 Nelson et al. 2010:53, Table 7.5 (Nelson, et al. 2010a) 71 SS PSnd Cama Beach 6 31 3% 2% 0.0% 6.0 4.15 Trost et al. 2010:B8, Table B.10 (Trost, et al. 2010) 76 SS PSnd West Point 28&29 5 28 1% 1% 0.0% 5.4 4.42 Wigen 1995:A5-62-70, A5-62, Tables A5-2 and A5-13 (Wigen 1995) 67 SS PSnd Burton Acres 7 16 73% 18% 3.4% 1.0 0.00 Kopperl 2001:7, Table 2 (Kopperl 2001) a - Regions: SS = Salish Sea, Including Puget Sound, Strait of Georgia, & the Strait of Juan de Fuca; WCVI = Western Vancouver Island; CC = Central BC Coast; HG = Haida Gwaii; NC = Northern BC Coast; SEAK = Southeast Alaska. b - Sub-Regions: PSnd = Puget Sound; GI-WA = Washington State Gulf Islands; GI-BC = BC Gulf Islands; SVI = Southern Vancouver Island; EVI = Eastern Vancouver Island; ESG = Eastern Strait of Georgia; FD = Fraser Delta; BS = Barkley Sound; WCVI = Western Vancouver Island; NVI = Northern Vancouver Island; Namu = Namu Area; Rinlt = Rivers Inlet Area; S. HG = Southern Haida Gwaii; N. HG = Northern Haida Gwaii; Dundas = Dundas Archipelago; PRupert = Prince Rupert Harbour; Angoon= Angoon Area; Noyes = Noyes Island; Sumez = Sumez Island; Dall = Dall Island; Pwales = Prince of Wales Island; Kuiu = Kuiu Island 379 Table 2.3 Distance between archaeological sites dating to within the last 2,500 years and documented spawning localities mapped by federal fisheries monitoring efforts (ca. 1930-2012) and available data Traditional and Local Ecological Knowledge (TEK/LEK) datasets. Numbers on left column correspond with sites listed in Table 2.1. No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference 1 SS SVI DcRu-75 0.5 Sec. 193 V. Harbour 2 SS SVI DcRu-78 0.5 Sec. 193 V. Harbour 3 SS SVI DcRu-4 1 Sec. 193 V. Harbour 4 SS SVI DcRt-16 unmonitored 2 Sec. 193 V. Harbour 5 SS SVI DcRt-10 unmonitored 5.8 Sec. 193 V. Harbour 6 SS SVI DcRu-71 unmonitored 7 Sec. 193 V. Harbour 7 SS SVI DcRw-57 unmonitored 3.9 Sec. 202 Sooke Inlet 8 SS SVI DcRv-21 1 Sec. 202 Sooke Inlet 9 SS SVI DdRu-81 unmonitored 6.4 Sec. 191-192 Saanich Pen. 10 SS SVI DdRu-4 unmonitored 1.5 Sec. 191-192 Saanich Pen. 11 SS SVI DeRu-1 0.5 Sec. 191-192 Saanich Pen. 12 SS SVI DdRu-5 0.5 Sec. 191-192 Saanich Pen. 13 SS SVI DdRu-18 0.5 Sec. 191-192 Saanich Pen. 14 SS GI-BC DdRt-6 3 Sec. 191-192 Saanich Pen. 15 SS GI-BC DeRu-44 0.5 Sec. 181 Swanson Ch. 16 SS GI-BC DfRu-3 0.5 Sec. 181 Swanson Ch. 17 SS GI-BC DeRt-1 0.5 Sec. 182 Plumper Snd. 18 SS GI-BC DeRt-109 unmonitored 4.3 Sec. 182 Plumper Snd. 19 SS GI-BC DgRu-3 unmonitored 1.6 Sec 173s S Yellow Pt. 20 SS GI-BC DgRv-4 0.5 Sec 173s S Yellow Pt. 21 SS GI-BC DgRv-1 0.5 Sec 173s S Yellow Pt. 380 No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference 22 SS GI-BC DgRv-2 0.5 Sec 173n N Yellow Pt. 23 SS GI-BC DgRv-2 0.5 Sec 173n N Yellow Pt. 24 SS GI-BC DgRw-78 0.5 Sec 173n N Yellow Pt. 25 SS GI-BC DgRw-73 0.5 Sec 173n N Yellow Pt. 26 SS GI-BC DgRw-32 0.5 Sec 173n N Yellow Pt. 27 SS GI-BC DgRw-4 0.5 Sec 173n N Yellow Pt. 28 SS GI-BC DgRw-204 1 Sec 173n N Yellow Pt. 30 SS EVI DjSf-13 0.5 Sec. 142 Baynes Snd. 31 SS EVI DkSf-26 4 Sec. 142 Baynes Snd. 32 SS EVI DkSf-40 3.5 Sec. 142 Baynes Snd. 33 SS EVI DkSf-4 0.5 Sec. 142 Baynes Snd. 34 SS EVI DkSf-19 0.5 Sec. 142 Baynes Snd. 35 SS EVI DkSf-29 0.5 Sec. 142 Baynes Snd. 36 SS EVI DiSc-26 0.5 Sec. 143 Qualicum 38 SS EVI DhSb-3 0.5 Sec. 143 Qualicum 39 SS EVI DhSb-11 0.5 Sec. 143 Qualicum 40 SS EVI DhRx-16 0.5 Sec 142 Nanoose 171 SS EVI DhRx-66 0.5 Sec 142 Nanoose 41 SS EVI DhRx-101 0.5 Sec 142 Nanoose 42 SS EVI DhRx-104 0.5 Sec 142 Nanoose 43 SS EVI DgRw-46 0.5 Sec 173n N Yellow Pt. 44 SS EVI DgRw-34 0.5 Sec 173n N Yellow Pt. 45 SS EVI DeRv-68 1 Sec. 181 Swanson Ch. 381 No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference 52 SS ESG EaSe-18 unmonitored 5.8 Sec. 152a N Powell R. 51 SS ESG EaSe-11 unmonitored 1 Sec. 152a N Powell R. 50 SS ESG EaSe-76 2 Sec. 152a N Powell R. 46 SS ESG DlSd-3 0.5 Sec. 152b S Powell R. 49 SS ESG DlSd-6 unmonitored 2.25 Sec. 152b S Powell R. 48 SS ESG DlSe-10 0.5 Sec. 152b S Powell R. 47 SS ESG DlSd-17 0.5 Sec. 152b S Powell R. 53 SS ESG DjRw-1 0.5 Sec. 165 Sechelt 54 SS FD DhRr-18 unmonitored 5.5 Sec. 280 Howe Snd. 56 SS FD DhRr-6 unmonitored 22 Sec. 280 Howe Snd. 57 SS FD DhRr-18 unmonitored 22 Sec. 280 Howe Snd. 58 SS FD DhRq-1 unmonitored 28 Sec. 280 Howe Snd. 60 SS FD DgRs-14 0.5 Sec. 291& 293 Boundary Bay 61 SS FD DgRs-2 3 Sec. 291& 293 Boundary Bay 62 SS FD DgRs-1 0.5 Sec. 291& 293 Boundary Bay 63 SS FD DgRr-1 0.5 Sec. 291& 293 Boundary Bay 79 WCVI WCVI EdSv-4 unmonitored 6.6 Sec. 271 Quatsino Snd. 80 WCVI WCVI DiSo-9 0.5 Sec. 242 Hesquiat Hrbr. 81 WCVI WCVI DgSl-67 unmonitored 2 Sec 245. Vargas Is. 82 WCVI WCVI DfSj-40 0.5 Sec 232 W Barkley Snd. 84 WCVI BS DfSj-57 0.5 Sec 232 W Barkley Snd. 85 WCVI BS DfSi-16+17 0.5 Sec 232 W Barkley Snd. 382 No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference 86 WCVI BS DfSi-26 0.5 Sec 232 W Barkley Snd. 87 WCVI BS 129T 2 Sec 232 W Barkley Snd. 88 WCVI BS 83T 2 Sec 232 W Barkley Snd. 89 WCVI BS 82T 2 Sec 232 W Barkley Snd. 90 WCVI BS 306T 1.5 Sec 232 W Barkley Snd. 91 WCVI BS 304T 3.5 Sec 232 W Barkley Snd. 92 WCVI BS 206T 2.5 Sec 232 W Barkley Snd. 93 WCVI BS DfSh-19 (72T) 0.5 Sec 232 W Barkley Snd. 94 WCVI BS DeSh-1 1 Sec 231 Trevor Chnl. 95 WCVI BS DeSh-2 1 Sec 231 Trevor Chnl. 96 WCVI BS DfSh-7 unmonitored 3.6 Sec 231 Trevor Chnl. 97 WCVI WCVI DeSf-6 unmonitored 27.3 No records 99 CC Namu ElSx-3 unmonitored 5.9 Sec. 086 Fitzhugh Snd. 0.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 104 CC Namu ElSx-1 0.5 Sec. 086 Fitzhugh Snd. 0.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 105 CC Namu EkSx-12 2 Sec. 085 Kwakshua Ch. 2.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 106 CC Namu ElTa-3 0.5 Sec.076 Kildidt Snd. 0.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 107 CC Namu ElTb-1 0.5 Sec.076 Kildidt Snd. 0.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 108 CC Namu ElTb-2 0.5 Sec.076 Kildidt Snd. 0.5 Gerrard et al. n.d. (Gerrard, et al. n.d.) 109 CC Rinlt EkSt-1 unmonitored 4 Sec. 093 Rivers Inlt Head 110 CC Rinlt EjSw-1 2 Sec. 092 Rivers Inlt. 383 No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference 111 CC Rinlt EjSv-1 4 Sec. 092 Rivers Inlt. 112 CC Rinlt EjSv-2 unmonitored 0.9 Sec. 092 Rivers Inlt. 113 CC Rinlt EjSv-3 unmonitored 1.8 Sec. 092 Rivers Inlt. 114 CC Rinlt EjSv-4 unmonitored 1.3 Sec. 093 Rivers Inlt Head 115 CC Rinlt EjSv-5 6 Sec. 093 Rivers Inlt Head 116 CC Rinlt EjSv-8 unmonitored 4.5 Sec. 092 Rivers Inlt. 117 CC Rinlt EjSv-9 unmonitored 4.6 Sec. 092 Rivers Inlt. 118 CC Rinlt EjSv-10 unmonitored 4.7 Sec. 092 Rivers Inlt. 119 CC Rinlt EjSv-11 unmonitored 5.2 Sec. 092 Rivers Inlt. 120 HG S. HG FaTt-9 2.2 Sec. 006 Louscoone Inlt. 4.4 Haida TEK/LTK Vol 3, pg 67 (Haida Marine Traditional Knowledge Study Participants, et al. 2011) 122 HG S. HG 2008T unmonitored 9 Sec. 006 Louscoone Inlt. 13.2 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 123 HG S. HG 699T unmonitored 15.9 Sec. 006 Louscoone Inlt. 11 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 124 HG S. HG 740T unmonitored 1.7 Sec. 025 Skincuttle Inlt. 0.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 125 HG S. HG 717T 0.5 Sec. 025 Skincuttle Inlt. 0.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 126 HG S. HG 924T 0.5 Sec. 021s S. Juan Perez Snd. 0.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 127 HG S. HG 1370T 0.5 Sec. 021s S. Juan Perez Snd. 0.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine Traditional Knowledge Study Participants, et al. 2011) 128 HG S. HG 781T 2.5 Sec. 021s S. Juan Perez Snd. 0.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et 384 No. Reg-iona Sub-regionb Arch. Site No. Distance to Documented Spawning Locality (\u00C2\u00B1 0.5 km)c Monitoring Status Distance to monitored shoreline DFO Monitoring Section (Department of Fisheries and Oceans Canada 2012a) TEK/LEK Distance to Spawning Locality (\u00C2\u00B10.5 km)c TEK/LEK Reference al. 2011) 129 HG S. HG 922T unmonitored 7.5 Sec. 021n N Juan Perez Snd. 7.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine Traditional Knowledge Study Participants, et al. 2011) 131 HG S. HG 785T unmonitored 10.4 Sec. 024 Laskeek Bay 4.5 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 132 HG S. HG 1134T unmonitored 3.1 Sec. 024 Laskeek Bay 3 Haida TEK/LTK Vol 3, pg 67 (Haida Marine TKS Participants, et al. 2011) 135 HG N. HG GaUa-18 unmonitored 26 Sec. 011 Masset Inlt. 23 Haida TEK/LTK Vol 3, pg 67 (Haida Marine Traditional Knowledge Study Participants, et al. 2011) 153 NC PRupert GbTo-24 0.5 Sec. 042 P. Rupert 146 NC PRupert GcTo-6 0.5 Sec. 042 P. Rupert 147 NC PRupert GbTo-77 0.5 Sec. 042 P. Rupert 148 NC PRupert GbTo-28 2 Sec. 042 P. Rupert 149 NC PRupert GbTo-46 2.5 Sec. 042 P. Rupert 150 NC PRupert GbTo-31 0.5 Sec. 042 P. Rupert 151 NC PRupert GbTo-13 1.5 Sec. 042 P. Rupert a \u00E2\u0080\u0093 Regions: SS = Salish Sea, Including Puget Sound, Strait of Georgia, & the Strait of Juan de Fuca; WCVI = Western Vancouver Island; CC = Central BC Coast; HG = Haida Gwaii; NC = Northern BC Coast; SEAK = Southeast Alaska. b \u00E2\u0080\u0093 Sub-Regions: PSnd = Puget Sound; GI-WA = Washington State Gulf Islands; GI-BC = BC Gulf Islands; SVI = Southern Vancouver Island; EVI = Eastern Vancouver Island; ESG = Eastern Strait of Georgia; FD = Fraser Delta; BS = Barkley Sound; WCVI = Western Vancouver Island; NVI = Northern Vancouver Island; Namu = Namu Area; Rinlt = Rivers Inlet Area; S. HG = Southern Haida Gwaii; N. HG = Northern Haida Gwaii; Dundas = Dundas Archipelago; PRupert = Prince Rupert Harbour; Angoon= Angoon Area; Noyes = Noyes Island; Sumez = Sumez Island; Dall = Dall Island; Pwales = Prince of Wales Island; Kuiu = Kuiu Island c \u00E2\u0080\u0093 measured using georectified map of cumulative spawning location and distance by water in Google Earth. (see materials and methods). "@en . "Thesis/Dissertation"@en . "2014-05"@en . "10.14288/1.0072131"@en . "eng"@en . "Anthropology"@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 . "An archaeology of food and settlement on the Northwest Coast"@en . "Text"@en . "http://hdl.handle.net/2429/45670"@en .