Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-18T19:35:35.465Z Has data issue: false hasContentIssue false

Submerged paleoshoreline mapping using high-resolution Chirp sub-bottom data, Northern Channel Islands platform, California, USA

Published online by Cambridge University Press:  18 October 2019

Alexander W. Laws*
Affiliation:
Department of Geological Sciences, San Diego State University, San Diego, California 92182, USA
Jillian M. Maloney
Affiliation:
Department of Geological Sciences, San Diego State University, San Diego, California 92182, USA
Shannon Klotsko
Affiliation:
Department of Geological Sciences, San Diego State University, San Diego, California 92182, USA
Amy E. Gusick
Affiliation:
Natural History Museum of Los Angeles County, Los Angeles, California 90007, USA
Todd J. Braje
Affiliation:
Department of Anthropology, San Diego State University, San Diego, California, 92182, USA
David Ball
Affiliation:
United States Department of the Interior, Bureau of Ocean Energy Management, Camarillo, California 93010, USA
*
*Corresponding author e-mail address: alexanderlaws@outlook.com (A.W. Laws)

Abstract

High-resolution Chirp sub-bottom data were obtained offshore from the Northern Channel Islands (NCI), California, to image submerged paleoshorelines and assess local uplift rates. Although modern bathymetry is often used for modeling paleoshorelines, Chirp data image paleoshorelines buried beneath sediment that obscures their seafloor expression. The NCI were a unified landmass during the last glacial maximum (LGM; ~20 ka), when eustatic sea level was ~120 m lower than present. We identified seven paleoshorelines, ranging from ~28 to 104 m in depth, across this now-submerged LGM platform. Paleoshoreline depths were compared to local sea-level curves to estimate ages, which suggest that some were reoccupied over multiple sea-level cycles. Additionally, previous studies determined conflicting uplift rates for the NCI, ranging from 0.16 to 1.5 m/ka. Our results suggest that a rate on the lower end of this range better fits the observed submerged paleoshorelines. Using the uplift rate of ~0.16 m/ka, we estimate that paleoshorelines formed during Marine Oxygen Isotope Stage 3, the LGM, and the Younger Dryas stade are preserved on the NCI platform. These results help clarify uplift rates for the NCI and illustrate the importance of sub-bottom data for mapping submerged paleoshorelines.

Type
Research Article
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2019 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Anderson, R.S., Densmore, A.L. and Ellis, M.A., 1999. The generation and degradation of marine terraces. Basin Research 11, 719.Google Scholar
Atwater, T, 1989. Plate tectonic history of the northeast Pacific and western North America. The Eastern Pacific Ocean and Hawaii N, 5572.Google Scholar
Braje, T.J., Dillehay, T.D., Erlandson, J.M., Klein, R.G., Rick, T.C., 2017. Finding the first Americans. Science 358, 592594.Google Scholar
Braje, T.J., Erlandson, J.M., Rick, T.C., 2013. Points in space and time: the distribution of Paleocoastal points and crescents on California's Northern Channel Islands. In: Jazwa, C., Perry, J. (Eds.), Small Islands, Big Implications: The California Channel Islands and Their Archaeological Contributions. University of Utah Press, Salt Lake City, pp. 72106.Google Scholar
Cattaneo, A., Steel, R.J., 2003. Transgressive deposits: a review of their variability. Earth-Science Reviews 62, 187228.Google Scholar
Chaytor, J.D., Goldfinger, C., Meiner, M.A., Huftile, G.J., Romos, C.G., Legg, M.R., 2008. Measuring vertical tectonic motion at the intersection of the Santa Cruz-Catalina Ridge and Northern Channel Islands platform, California Continental Borderland, using submerged paleoshorelines. Geological Society of America Bulletin 120, 10531071.Google Scholar
Clark, J., Mitrovica, J.X., Alder, J., 2014. Coastal paleogeography of the California-Oregon-Washington and Bering Sea continental shelves during the latest Pleistocene and Holocene: implications for the archaeological record. Journal of Archaeological Science 52, 1223.Google Scholar
Crouch, J.K., 1979. Neogene tectonic evolution of the California Continental Borderland and western Transverse Ranges. Geological Society of America Bulletin 90, 338345.Google Scholar
Darigo, N.J., Osborne, R.H., 1986. Quaternary stratigraphy and sedimentation of the inner continental shelf, San Diego County, California. In: Knight, R.J., McLean, J.R. (Eds.), Shelf Sands and Sandstones: Canadian Society of Petroleum Geologists, Memoir II. Ronalds Printing, Calgary, pp. 7398.Google Scholar
Dickinson, W.R., 2001. Paleoshoreline record of relative Holocene sea levels on Pacific islands. Earth-Science Reviews 55, 191234.Google Scholar
Dillehay, T.D., Ramirez, C., Pino, M., Collins, M.B., Rossen, J., Pino-Navarro, J.D., 2008. Monte Verde: seaweed, food, medicine, and the peopling of South America. Science 320, 784786.Google Scholar
Emery, K.O., 1958. Shallow submerged marine terraces of southern California. Bulletin of the Geological Society of America 69, 3960.Google Scholar
Erlandson, J.M., 2002. Anatomically modern humans, maritime voyaging, and the Pleistocene colonization of the Americas. In: Jablonski, N.G. (Ed.), The First Americans: The Pleistocene Colonization of the New World. California Academy of Sciences, San Francisco, pp. 5992.Google Scholar
Erlandson, J.M., Braje, T.J., 2011. From Asia to the Americas by boat? Paleogeography, paleoecology, and stemmed points of the northwest Pacific. Quaternary International 239, 2837.Google Scholar
Erlandson, J.M., Graham, M.H., Bourque, B.J., Corbett, D., Estes, J.A., Steneck, R.S., 2007. The kelp highway hypothesis: marine ecology, the coastal migration theory, and the peopling of the Americas. The Journal of Island and Coastal Archaeology 2, 161174.Google Scholar
Fagundes, N.J., Kanitz, R., Eckert, R., Valls, A.C., Bogo, M.R., Salzano, F.M., Smith, D.G., et al. , 2008. Mitochondrial population genomics supports a single pre-Clovis origin with a coastal route for the peopling of the Americas. The American Journal of Human Genetics 82, 583592.Google Scholar
Fedje, D.W., Josenhans, H., Clague, J.J., Barrie, J.V., Archer, D.J., Southon, J.R., 2005. Hecate Strait paleoshorelines. In: Fedje, D.W., Mathewes, R.W. (Eds.), Haida Gwaii: Human History and Environment from the Time of Loon to the Time of the Iron People. University of British Columbia Press, Vancouver, pp. 2137.Google Scholar
Graham, M.H., Dayton, P.K., Erlandson, J.M., 2003. Ice-ages and ecological transitions on temperate coasts. Trends in Ecology and Evolution 18, 3340.Google Scholar
Grant, L.B., Mueller, K.J., Gath, E.M., Cheng, H., Edwards, R.L., Munro, R., Kennedy, G.L., 1999. Late Quaternary uplift and earthquake potential of the San Joaquin Hills, southern Los Angeles Basin, California. Geology 27, 10311034.Google Scholar
Gusick, A.E., Erlandson, J.M., 2019. Paleocoastal landscapes, marginality, and initial settlement of California's islands. In: Gill, K., Erlandson, J., Fauvelle, M., An Archaeology of Abundance: Re-evaluating the Marginality of California's Islands. University of Florida Press, Gainesville, pp. 5997.Google Scholar
Gusick, A.E., Faught, M.K., 2011. Prehistoric underwater archaeology: a nascent subdiscipline critical to understanding early coastal occupations and migration routes. In: Bicho, N., Haws, J., Davis, L.G. (Eds.), Trekking the Shore: Changing Coastlines and the Antiquity of Coastal Settlement. Springer, New York, pp. 145157.Google Scholar
Haaker, E.C., Rockwell, T.K., Kennedy, G.L., Ludwig, L.G., Freeman, S.T., Zumbro, J.A., Mueller, K.J., Edwards, R.L., 2016. Style and rate of long-term uplift of the southern California coast between San Diego and Newport Beach with potential implications for assessing blind thrust models. Applied Geology of California 26, 679719.Google Scholar
Hammond, W.C., Burgette, R.J., Johnson, K.M., Blewitt, G., 2018. Uplift of the Western Transverse Ranges and Ventura area of southern California: a four-technique geodetic study combining GPS, InSAR, leveling, and tide gauges. Journal of Geophysical Research: Solid Earth, 836858.Google Scholar
Hogarth, L.J., Babcock, J., Driscoll, N.W., Le Dantec, N., Haas, J.K., Inman, D.L., Masters, P.M., 2007. Long-term tectonic control on Holocene shelf sedimentation offshore La Jolla, California. Geology 35, 275.Google Scholar
Hogarth, L.J., Driscoll, N.W., Babcock, J.M., Orange, D.L., 2012. Transgressive deposits along the actively deforming Eel River Margin, Northern California. Marine Geology 303, 99114.Google Scholar
Johnson, J.R., Stafford, T.W. Jr., Ajie, H.O., Morris, D.P., 2002. Arlington Springs Revisited. In: Browne, D., Mitchell, K., Chaney, H.. (Eds.), Proceedings of the Fifth California Islands Symposium. Santa Barbara, CA: Santa Barbara Museum of Natural History, 541545.Google Scholar
Junger, A., 1979. Maps and seismic profiles showing geologic structure of the Northern Channel Islands platform. California Continental Borderland, Report MF-991. United States Geological Survey, Department of Interior. doi: 10.3133/mf991Google Scholar
Kelsey, H.M., Bockheim, J.G., 1994. Coastal landscape evolution as a function of eustasy and surface uplift rate, Cascadia margin, southern Oregon. Geological Society of America Bulletin 106, 840854.Google Scholar
Kennett, D.J., Kennett, J.P., West, J., Erlandson, J.M., Johnson, J.R., Hendy, I.L., West, A., Jones, T.L., 2008. Wildfire and abrupt ecosystem disruption on California's Northern Channel Islands at the Ållerød-Younger Dryas boundary (13.0–12.9 ka). Quaternary Science Reviews 27, 25302545.Google Scholar
Kern, J.P., Rockwell, T.K., 1992. Chronology and deformation of Quaternary marine shorelines, San Diego County, California. In: Fletcher, C.H. and Wehmiller, J.F. (Eds.) Quaternary Coasts of the United States: Marine and Lacustrine Systems. Society of Economic Paleontologists and Mineralogists Special Publication, No. 48. SEPM, Tulsa, pp. 377382.Google Scholar
Klotsko, S., Driscoll, N., Kent, G., Brothers, D., 2015. Continental shelf morphology and stratigraphy offshore San Onofre, California: the interplay between rates of eustatic change and sediment supply. Marine Geology 369, 116126.Google Scholar
Ku, T.-L., Kern, J.P., 1974. Uranium-series age of the upper Pleistocene Nestor terrace, San Diego, California. Geological Society of America Bulletin 85, 17131716.Google Scholar
Lajoie, K. R., 1986, Coastal tectonics. In: Wallace, R., (Ed.), Active Tectonics: Washington, D.C., National Academy Press, Studies in Geophysics Series, Geophysics Research Forum, p. 95124.Google Scholar
Lambeck, K., Yokoyama, Y., Purcell, T., 2002. Into and out of the Last Glacial Maximum: sea level change during Oxygen Isotope Stages 3 and 2. Quaternary Science Reviews 21, 343360.Google Scholar
Lantzsch, H., Hanebuth, T.J., Bender, V.B., Krastel, S., 2009. Sedimentary architecture of a low-accumulation shelf since the Late Pleistocene (NW Iberia). Marine Geology 259, 4758.Google Scholar
Le Dantec, N., Hogarth, L.J., Driscoll, N.W., Babcock, J.M., Barnhardt, W.A., Schwab, W.C., 2010. Tectonic controls on nearshore sediment accumulation and submarine canyon morphology offshore La Jolla, Southern California. Marine Geology, v. 268, no. 1–4, p. 115128, doi: 10.1016/j.margeo.2009.10.026Google Scholar
Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20. http://dx.doi.org/10.1029/2004PA001071.Google Scholar
Lonsdale, P., 1991. Structural patterns of the Pacific floor offshore of Peninsular California. In: Dauphin, J. P. and Simoneit, B. R. T., (Eds.), The Gulf and Peninsular Province of the Californias, AAPG Memoir 47, p. 87125.Google Scholar
Mackie, D.W.F.Q., Heaton, E.J.D.T.H., 2004. Late Wisconsin environments and archaeological visibility on the northern Northwest Coast. In: Madsen, D.B., Entering America: Northeast Asia and Beringia Before the Last Glacial Maximum. Canadian Archaeological Association, Calgary, p.97.Google Scholar
Maloney, J.M., Driscoll, N.W., Kent, G.M., Duke, S., Freeman, T., Bormann, J., 2016. Segmentation and step-overs along strike-slip fault systems in the inner California borderlands: Implications for fault architecture and basin formation. In: Anderson, R., Ferriz, H. (Eds.), Association of Environmental and Engineering Geologists Special Publication 26: Applied Geology in California. Star Publishing Company, Inc., Redwood City, pp. 655677.Google Scholar
Mandryk, C.A., Josenhans, H., Fedje, D.W., Mathewes, R.W., 2001. Late Quaternary paleoenvironments of northwestern North America: implications for inland versus coastal migration routes. Quaternary Science Reviews 20, 301314.Google Scholar
McLaren, D., Fedje, D., Hay, M., Mackie, Q., Walker, I.J., Shugar, D.H., Eamer, J.B.R., Lian, O.B., Neudorf, C., 2014. A post-glacial sea level hinge on the central Pacific coast of Canada. Quaternary Science Reviews 97, 148169.Google Scholar
Mueller, K., Kier, G., Rockwell, T., Jones, C.H., 2009. Quaternary rift flank uplift of the Peninsular Ranges in Baja and southern California by removal of mantle lithosphere. Tectonics 28, TC5003. http://dx.doi.org/10.1029/2007TC002227.Google Scholar
Muhs, D.R., Simmons, K.R., Schumann, R.R., Groves, L.T., Devogel, S.B., Minor, S.A., Laurel, D., 2014. Coastal tectonics on the eastern margin of the Pacific Rim: late Quaternary sea level history and uplift rates, Channel Islands National Park, California, USA. Quaternary Science Reviews 105, 209238.Google Scholar
Muhs, D.R., Simmons, K.R., Schumann, R.R., Groves, L.T., Mitrovica, J.X., Laurel, D., 2012. Sea level history during the last interglacial complex on San Nicolas Island, California: implications for glacial isostatic adjustment processes, paleozoogeography and tectonics. Quaternary Science Reviews 37, 125.Google Scholar
Nicholson, C., Sorlien, C.C., Atwater, T., Crowell, J.C., Luyendyk, B.P., 1994. Microplate capture, rotation of the western Transverse Ranges, and initiation of the San Andreas transform as a low-angle fault system. Geology 22, 491495.Google Scholar
Orr, P. C. 1962. The Arlington Spring Site, Santa Rosa Island, California. American Antiquity 27:417419Google Scholar
Orr, P.C., 1968. Prehistory of Santa Rosa Island. Santa Barbara Museum of Natural History, Santa Barbara.Google Scholar
Patterson, R.H., 1979. Tectonic geomorphology and neotectonics of the Santa Cruz Island fault Santa Barbara County, California. Master's thesis, University of California, Santa Barbara.Google Scholar
Pinter, N., Johns, B., Little, B., Vestal, W.D., 2001. Fault-related folding in California's Northern Channel Islands documented by rapid-static GPS positioning. GSA Today 11, 49.Google Scholar
Pinter, N., Lueddecke, S.B., Keller, E.A., Simmons, K.R., 1998. Late Quaternary slip on the Santa Cruz Island Fault, California. Geological Society of America Bulletin 110, 711722.Google Scholar
Pinter, N., Sorlien, C., 1991. Evidence for latest Pleistocene to Holocene movement on the Santa Cruz Island Fault, California. Geology 19, 909912.Google Scholar
Pinter, N., Sorlien, C.C., Scott, A.T., 2003. Fault-related growth and isostatic subsidence, California Channel Islands. American Journal of Science 303, 300318.Google Scholar
Posamentier, H.W., 2002. Ancient shelf ridges—a potentially significant component of the transgressive system tract: case study from offshore northwest Java. AAPG Bulletin 86, 75106.Google Scholar
Posamentier, H.W., Allen, G.P., 1993. Variability of the sequence stratigraphic model: effects of local basin factors. Sedimentary Geology 86, 91109.Google Scholar
Posamentier, H.W., Allen, G.P., 1999. Siliciclastic Sequence Stratigraphy: Concepts and Applications. SEPM Concepts in Sedimentology and Paleontology, Vol. 7. Society for Sedimentary Geology, Tulsa, p. 210.Google Scholar
Reeder-Myers, L., Erlandson, J.M., Muhs, D.R., Rick, T.C., 2015. Sea level, paleogeography, and archaeology on California's Northern Channel Islands. Quaternary Research 83, 263272.Google Scholar
Reynolds, L.C., Simms, A.R., 2015. Late Quaternary relative sea level in Southern California and Monterey Bay. Quaternary Science Reviews 126, 5766.Google Scholar
Rockwell, T.K., Clark, K., Gamble, L., Oskin, M.E., Haaker, E.C., Kennedy, G.L., 2016. Large Transverse Range earthquakes cause coastal upheaval near Ventura, southern California. Bulletin of the Seismological Society of America 106, 27062720.Google Scholar
Ryan, H.F., Conrad, J.E., Paull, C.K., McGann, M., 2012. Slip rate on the San Diego Trough fault zone, inner California borderland, and the 1986 Oceanside earthquake swarm revisited. Bulletin of the Seismological Society of America 102, 23002312.Google Scholar
Sahakian, V., Bormann, J., Driscoll, N., Harding, A., Kent, G., Wesnousky, S., 2017. Seismic constraints on the architecture of the Newport-Inglewood/Rose Canyon fault: implications for the length and magnitude of future earthquake ruptures. Journal of Geophysical Research: Solid Earth 122, 20852105.Google Scholar
Schumann, R.R., Minor, S., Muhs, D.R., Pigati, J., 2014. Landscapes of Santa Rosa Island, Channel Islands National Park, California. Monographs of the Western North American Naturalist 7, 4867.Google Scholar
Schwab, W.C., Baldwin, W.E., Denny, J.F., Hapke, C.J., Gayes, P.T., List, J.H., Warner, J.C., 2014. Modification of the Quaternary stratigraphic framework of the innercontinental shelf by Holocene marine transgression: an example offshore of Fire Island, New York. Marine Geology 355, 346360.Google Scholar
Seeber, L., Sorlien, C.C., 2000. Listric thrusts in the western Transverse Ranges, California. Geological Society of America Bulletin 112, 10671079.Google Scholar
Shaw, J.H., Suppe, J., 1994. Active faulting and folding in the eastern Santa Barbara Channel, California. Geological Society of America Bulletin 106, 607626.Google Scholar
Simms, A.R., Rouby, H., Lambeck, K. 2016. Marine terraces and rates of vertical tectonic motion: the importance of glacio-isostatic adjustment along the Pacific coast of central North America. Geological Society of America Bulletin 128, 8193.Google Scholar
Sorlien, C.C., 1994. Faulting and uplift of the northern Channel Islands, California. In: Halvorson, W.L., Maender, G.J. (Eds.), The Fourth California Islands Symposium: Update on the Status of Resources. Santa Barbara Museum of Natural History, Santa Barbara, pp. 281296.Google Scholar
Sorlien, C.C., Kamerling, M.J., Seeber, L., Broderick, K., 2006. Restraining segments and reactivation of the Santa Monica-Dume-Malibu Coast fault system, offshore Los Angeles, California. Journal of Geophysical Research 111. http://dx.doi.org/10.1029/2005JB003632.Google Scholar
Spratt, R.M., Lisiecki, L.E., 2016. A Late Pleistocene sea level stack. Climate of the Past 12, 10791092.Google Scholar
ten Brink, U.S., Zhang, J., Brocher, T.M., Okaya, D.A., Klitgord, K.D., Fuis, G.A., 2000. Geophysical evidence for the evolution of the California inner continental borderland as a metamorphic core complex. Journal of Geophysical Research 105, 58355857.Google Scholar
Teng, L.S., Gorsline, D.S., 1991. Stratigraphic framework of the California Continental Borderland basins, southern California. In: Dauphin, J.P., Simoneit, B.R.T. (Eds.), The Gulf and Peninsular Province of the Californias. American Association of Petroleum Geologists Memoir 47. American Association of Petroleum Geologists, Tulsa, pp. 127143.Google Scholar
Vedder, J.G., 1987. Regional geology and petroleum potential of the southern California Borderland. In: Gantz, A. (Ed.), Geology and Resource Potential of the Continental Margin of Western North America and Adjacent Ocean Basins, Beaufort Sea to Baja California. Circum-Pacific Council for Energy and Mineral Resources, Houston, pp. 403447.Google Scholar
Wright, T.L., 1991. Structural geology and tectonic evolution of the Los Angeles Basin, California. In: Biddle, K.T. (Ed.), Active Margin Basins. American Association of Petroleum Geologists Memoir 52. American Association of Petroleum Geologists, Tulsa, p. 35134.Google Scholar