Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-28T22:02:10.705Z Has data issue: false hasContentIssue false

Controls on range shifts of coastal Californian bivalves during the peak of the last interglacial and baseline predictions for today

Published online by Cambridge University Press:  21 April 2021

Emily A. Orzechowski
Affiliation:
University of California Museum of Paleontology and Department of Integrative Biology, University of California, Berkeley, Berkeley, California 94720, U.S.A. E-mail: eaorzechowski@berkeley.edu, sethf@berkeley.edu.
Seth Finnegan
Affiliation:
University of California Museum of Paleontology and Department of Integrative Biology, University of California, Berkeley, Berkeley, California 94720, U.S.A. E-mail: eaorzechowski@berkeley.edu, sethf@berkeley.edu.

Abstract

As the most recent time in Earth history when global temperatures were warmer than at present, the peak of the last interglacial (Marine Isotope Substage [MIS] 5e; ~120,000 years ago) can serve as a pre-anthropogenic baseline for a warmer near-future world. Here we use a new compilation of 22 fossil localities in California that have been reliably dated to MIS 5e to establish baseline expectations for contemporary bivalve species movements by identifying and analyzing bivalve species with “extralimital” ranges, that is, species that occupied the California region during MIS 5e but are now restricted to adjacent regions. We find that 15% of species (n = 142) found in MIS 5e localities have extralimital ranges and currently occupy warmer waters to the south of the California region. The majority of extralimital occurrences occur in paleo-embayments, suggesting that these sheltered habitats were more suitable habitats for warm-water species than exposed coasts during the MIS 5e. We further find that extralimital species now tend to occur in cooler, more seasonally productive coastal waters and to occupy more offshore islands when compared with the broader species pool immediately south of California. These findings suggest that high dispersal potential and preexisting tolerances to environmental conditions similar to California's comparatively cool and seasonally productive environments may have enabled extralimital bivalves to colonize the California region during MIS 5e.

Type
Articles
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press on behalf of The Paleontological Society

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.)

Footnotes

Data available from the Dryad Digital Repository: https://doi.org/10.5061/dryad.kprr4xh28

References

Literature Cited

Acker, J. G., and Leptoukh, G.. 2007. Online analysis enhances use of NASA Earth science data. Eos, Transactions of the American Geophysical Union 88:1417.CrossRefGoogle Scholar
Addicott, W. O. 1966. Late Pleistocene marine paleoecology and zoogeography in central California. U.S. Geological Survey Professional Paper 523-C.CrossRefGoogle Scholar
Amante, C., and Eakins, B. W.. 2009. ETOPOI 1 Arc-minute global relief model: procedures, data sources, and analysis. National Oceanographic and Atmospheric Administration Technical Memorandum, National Environment Satellite, Data, and Information Service National Geophysical Data Center-24, Boulder, Colo.Google Scholar
Bakun, A. 1990. Global climate change and intensification of coastal ocean upwelling. Science 247:198201.CrossRefGoogle ScholarPubMed
Bakun, A., Black, B. A., Bograd, S. J., García-Reyes, M., Miller, A. J., Rykaczewski, R. R., and Sydeman, W. J.. 2015. Anticipated effects of climate change on coastal upwelling ecosystems. Current Climate Change Reports 1:8593.CrossRefGoogle Scholar
Belanger, C. L., Jablonski, D., Roy, K., Berke, S. K., Krug, A. Z., and Valentine, J. W.. 2012. Global environmental predictors of benthic marine biogeographic structure. Proceedings of the National Academy of Sciences USA 109:1404614051.CrossRefGoogle ScholarPubMed
Coan, E. V., and Valentich-Scott, P.. 2012. Bivalve seashells of tropical west America. Santa Barbara Museum of Natural History, Santa Barbara, Calif.Google Scholar
Coan, E. V., Scott, P. V., and Bernard, F. R.. 2000. Bivalve seashells of western North America. Santa Barbara Museum of Natural History, Santa Barbara, Calif.Google Scholar
Conrad, T. A. 1855. Report on the fossil shells collected in California by Wm. P. Blake, geologist of the expedition under the command of Lieutenant R. S. Williamson, United States Topographical Engineers. Appendix to the preliminary geological report of William P. Blake, Explorations and surveys for a railroad route from the Mississippi River to the Pacific Ocean, New York.Google Scholar
Dobrowski, S. Z., Thorne, J. H., Greenberg, J. A., Safford, H. D., Mynsberge, A. R., Crimmins, S. M., and Swanson, A. K.. 2011. Modeling plant ranges over 75 years of climate change in California, USA: temporal transferability and species traits. Ecological Monographs 81:241257.CrossRefGoogle Scholar
Elith, J., Leathwick, J. R., and Hastie, T.. 2008 A working guide to boosted regression trees. Journal of Animal Ecology 77:802813.CrossRefGoogle ScholarPubMed
Fenberg, P. B., Menge, B. A., Raimondi, P. T., and Rivandeneira, M. M.. 2014. Biogeographic structure of the northeastern Pacific rocky intertidal: the role of upwelling and dispersal to drive patterns. Ecography 38:8395.CrossRefGoogle Scholar
Grantham, B. A., Eckert, G. L., and Shanks, A. L.. 2003. Dispersal potential of marine invertebrates in diverse habitats. Ecological Applications 13:S108S116.CrossRefGoogle Scholar
Hall, C. A. Jr. 2002. Nearshore marine paleoclimatic regions, increasing zoogeographic provinciality, molluskan extinctions, and paleoshorelines, California: late Oligocene (27 Ma) to late Pliocene (2.5 Ma). Geological Society of America Special Paper 357.Google Scholar
Hansen, T. 1980. Influence of larval dispersal and geographic distribution on species longevity in neogastropods. Paleobiology 6:193207.CrossRefGoogle Scholar
Hellburg, M. E., Deborah, P. B., and Roy, K.. 2001. Climate-driven range expansion and morphological evolution in a marine gastropod. Science 292:17071710.CrossRefGoogle Scholar
Hijmans, R. J., Phillips, S., Leathwick, J., and Elith, J. L.. 2015. dismo: species distribution modeling, R package version 1.0-12. http://CRAN.R-project.org/package=dismo.Google Scholar
Jablonski, D. 1986. Larval ecology and macroevolution in marine invertebrates. Bulletin of Marine Science 39:565597.Google Scholar
Jablonski, D., and Sepkoski, J. J.. 1996. Paleobiology, community ecology, and scales of ecological pattern. Ecology 77:13671378.CrossRefGoogle Scholar
Jackson, S. T., and Blois, J. L.. 2015. Community ecology in a changing environment: perspectives from the Quaternary. Proceedings of the National Academy of Sciences USA 112:49154921.CrossRefGoogle Scholar
Jackson, S. T., and Overpeck, J. T.. 2000. Responses of plant populations and communities to environmental changes of the late Quaternary. Paleobiology 26:194220.CrossRefGoogle Scholar
Kanakoff, G. P., and Emerson, W. K.. 1959. Late Pleistocene invertebrates of the Newport Bay area, California. Los Angeles County Museum, Contributions in Science 31:147.Google Scholar
Kennedy, G. L., Lajoie, K. R., and Wehmiller, J. F.. 1982. Aminostratigraphy and faunal correlations of late Quaternary marine terraces, Pacific Coast, USA. Nature 299:545547.CrossRefGoogle Scholar
Kern, P. 1971. Paleoenvironmental analysis of a Late Pleistocene eustuary in southern California. Journal of Paleontology 45:810823.Google Scholar
Lester, S. E., Gaines, S. D., and Kinlan, B. P.. 2007. Reproduction on the edge: large-scale patterns of individual performance in a marine invertebrate. Ecology 88:22292239.CrossRefGoogle Scholar
Lieberman, B. S., and Saupe, E. E.. 2016. Palaeoniches get stitches: analyses of niches informing macroevolutionary theory. Lethaia 49:145149.CrossRefGoogle Scholar
Lindberg, D. R., and Lipps, J. H.. 1996. Reading the chronicle of Quaternary temperate rocky shore faunas. Pp. 161182 in Jablonski, D., Erwin, D. H., and Lipps, J. H., eds. Evolutionary paleobiology. University of Chicago Press, Chicago.Google Scholar
Lindberg, D. R., Roth, B., Kellogg, M. G., and Hubbs, C. L.. 1980. Invertebrate megafossils of Pleistocene (Sangamon interglacial) age from Isla de Guadalupe, Baja California, Mexico. Pp. 4170 in Powers, D. M., ed. The California islands: proceedings of a multidisciplinary symposium. Santa Barbara Museum of Natural History, Santa Barbara.Google Scholar
Lisiecki, L. E., and Raymo, M. E.. 2005. A Pliocene–Pleistocene stack of 57 globally distributed benthic δ18O records: Pliocene–Pleistocene benthic stack. Paleoceanography 20:317.Google Scholar
Maguire, K. C., Blois, J. L., Nieto-Lugilde, D., Fitzpatrick, M. C., and Williams, J. D.. 2015. Modeling species and community responses to past, present, and future episodes of climatic and ecological change. Annual Review of Ecology, Evolution, and Systematics 46:343368.CrossRefGoogle Scholar
Martinez-Meyer, E., and Peterson, A. T.. 2006. Conservatism of ecological niche characteristics in North American plant species over the Pleistocene-to-recent transition. Journal of Biogeography 33:17791789.CrossRefGoogle Scholar
Martinson, D. G., Pisias, N. G., Hays, J. D., Imbrie, J., Moore, T. C., and Shackleton, N. J.. 1987. Age dating and the orbital theory of the ice ages: development of a high-resolution 0 to 300,000-year chronostratigraphy. Quaternary Research 27:119.CrossRefGoogle Scholar
McLean, J. H., and Coan, E. V.. 1996. Marine mollusks of Rocas Alijos. Monographiae Biologicae 75:305318.CrossRefGoogle Scholar
Muhs, D. R., and Groves, L. T.. 2018. Little islands recording global events: late Quaternary sea level history and paleozoogeography of Santa Barbara and Anacapa Islands, Channel Islands National Park, California. Western North American Naturalist 78:540589.CrossRefGoogle Scholar
Muhs, D. R., and Kyser, T. K.. 1987. Stable isotope compositions of fossil mollusks from southern California: evidence for a cool last interglacial ocean. Geology 15:119122.2.0.CO;2>CrossRefGoogle Scholar
Muhs, D. R., Miller, G. H., Whelan, J. F., and Kennedy, G. L.. 1992. Aminostratigraphy and oxygen isotope stratigraphy of marine-terrace deposits, Palos Verdes Hills and San Pedro areas, Los Angeles County, California. SEPM Special Publications 48:363376.Google Scholar
Muhs, D. R., Simmons, K. R., and Steinke, B.. 2002. Evidence for the timing and duration of the Last Interglacial Period from high-precision uranium-series ages of corals on tectonically stable coastlines. Quaternary Research 58:3640.CrossRefGoogle Scholar
Muhs, D. R., Simmons, K. R., Kennedy, G. L., Ludwig, K. R., and Groves, L. T.. 2006. A cool eastern Pacific Ocean at the close of the Last Interglacial complex. Quaternary Science Reviews 25:235262.CrossRefGoogle Scholar
Muhs, D. R., Simmons, K. R., Schumann, R. R., Groves, L. T., Mitrovica, J. X., and 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.CrossRefGoogle Scholar
Muhs, D. R., Groves, L. T., and Schumann, R. R.. 2014. Interpreting the paleozoogeography and sea level history of thermally anomalous marine terrace faunas: a case study from the Last Interglacial complex of San Clemente Island, California. Monographs of the Western North American Naturalist 7:82108.CrossRefGoogle Scholar
Nogués-Bravo, D. 2009. Predicting the past distribution of species climatic niches. Global Ecology and Biogeography 18:521531.CrossRefGoogle Scholar
Nogués-Bravo, D., Rodríguez-Sánchez, F., Orsini, L., de Boer, E., Jansson, R., Morlon, H., Fordham, D., and Jackson, S. T.. 2018. Cracking the code of biodiversity responses to past climate change. Trends in Ecology and Evolution 33:765776.CrossRefGoogle ScholarPubMed
Ortlieb, L., Diaz, A., and Guzman, N.. 1996. A warm interglacial episode during oxygen isotope stage 11 in northern Chile. Quaternary Science Reviews 15:857871.CrossRefGoogle Scholar
Pearman, P. B., Guisan, A., Broennimann, O., and Randin, C. F.. 2008. Niche dynamics in space and time. Trends in Ecology and Evolution 23:149158.CrossRefGoogle ScholarPubMed
Powell, C. L. II. 2001. Geologic and molluscan evidence for a previously misunderstood late Pleistocene, cool water, open coast terrace at Newport Bay, southern California. The Veliger 44:340347.Google Scholar
Powell, C. L. II, Lajoie, K., and Ponti, D.. 2000. A preliminary chronostratigrpahy based on molluscan biogeography for the Late Quaternary of southern California. Western Society of Malacologists 32:2336.Google Scholar
Powell, C. L. II, Grant, L., and Conkling, S.. 2004. Paleoecologic analysis and age of a Late Pleistocene fossil assemblage at a locality in Newport Beach, Upper Newport Bay, Orange County, California. The Veliger 47:171180.Google Scholar
Powell, C. L., II, R. J. Stanton, M. Jr. Vendrasco, and P. Liff-Grief, . 2009. Warm extralimital fossil mollusks used to recognize the mid-Pliocene warm event in southern California. Western Society of Malacologists 41:7091.Google Scholar
Qiao, H., Saupe, E. E., Soberón, J., Peterson, A. T., and Myers, C. E.. 2016. Impacts of niche breadth and dispersal ability on macroevolutionary patterns. American Naturalist 188:149162.CrossRefGoogle ScholarPubMed
R Core Team. 2019. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.Google Scholar
Ricketts, E. F., and Calvin, J.. Rev. by Hedgpeth, J. W.. 1985. Between Pacific tides. Stanford University Press, Stanford, Calif.Google Scholar
Roy, K., Jablonski, D., and Valentine, J. W.. 1995. Thermally anomalous assemblages revisited: patterns in the extraprovincial latitudinal range shifts of Pleistocene marine mollusks. Geology 23:10711074.2.3.CO;2>CrossRefGoogle Scholar
Roy, K., Valentine, J. W., Jablonski, D., and Kidwell, S. M.. 1996. Scales of climatic variability and time averaging in Pleistocene biotas: implications for ecology and evolution. Trends in Ecology and Evolution 11:458463.CrossRefGoogle ScholarPubMed
Roy, K., Jablonski, D., and Valentine, J. W.. 2001. Climate change, species range limits and body size in marine bivalves. Ecology Letters 4:366370.Google Scholar
Saupe, E. E., Hendricks, J. R., Portell, R. W., Dowsett, H. J., Haywood, A., Hunter, S. J., and Lieberman, B. S.. 2014. Macroevolutionary consequences of profound climate change on niche evolution in marine mollusks over the past three million years. Proceedings of the Royal Society of London B 281:20160654.Google ScholarPubMed
Saupe, E. E., Qiao, H., Hendricks, J. R., Portell, R. W., Hunter, S. J., Soberón, J., and Lieberman, B. S.. 2015. Niche breadth and geographic range size as determinants of species survival on geological time scales: determinants of species survival. Global Ecology and Biogeography 24:11591169.CrossRefGoogle Scholar
Shackleton, N. J., and Opdike, N. D.. 1973. Oxygen isotope and paleomagnetic stratigraphy of equatorial Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 105 year and 106 year scale. Quaternary Research 3:3955.CrossRefGoogle Scholar
Sokal, R., and Rohlf, E. J.. 2012. Biometry: the principles and practice of statistics in biological research. Freeman, New York.Google Scholar
Stigall, A. 2014. When and how do species achieve niche stability over long time scales? Ecography 37:11231132.Google Scholar
Strathmann, M. F. 1987. Reproduction and development of marine invertebrates of the northern Pacific coast: data and methods for the study of eggs, embryos, and larvae. University of Washington Press, Seattle.Google Scholar
Sunday, J. M., Bates, A. E., and Dulvy, N. K.. 2012. Thermal tolerance and the redistribution of animals. Nature Climate Change 2:686690.CrossRefGoogle Scholar
Sydeman, W. J., Garcia-Reyes, M., Schoeman, D. S., Rykaczewski, R. R., Thompson, S. A., Black, B. A., and Bograd, S. J.. 2014. Climate change and wind intensification in coastal upwelling ecosystems. Science 345:7780.CrossRefGoogle ScholarPubMed
Valentine, J. W. 1959. Marine climatic record of northwest American epicontinental Pleistocene. Pp. 295296 in Sears, M., ed. International oceanographic congress. American Association for the Advancement of Science, Washington, D.C.Google Scholar
Valentine, J. W. 1961. Paleoecologic molluskan geography of the Californian Pleistocene. University of California Publications in Geological Sciences 34:309442Google Scholar
Valentine, J. W. 1966. Numerical analysis of marine molluskan ranges on the extratropical northeastern Pacific shelf. Limnology and Oceanography 11:198211.CrossRefGoogle Scholar
Valentine, J. W. 1973. Evolutionary paleoecology of the marine biosphere. Prentice Hall, Englewood Cliffs, N.J.Google Scholar
Valentine, J. W. 1989. How good was the fossil record—clues from the Californian Pleistocene. Paleobiology 15:8394.CrossRefGoogle Scholar
Valentine, J. W., and Emerson, W. K.. 1961. Environmental interpretation of Pleistocene marine species: a discussion. Journal of Geology 69:616618.CrossRefGoogle Scholar
Valentine, J. W., and Jablonski, D.. 1993. Fossil communities: compositional variation at many time scales. Pp. 341349 in Ricklefs, R. E. and Schluter, D., eds. Species diversity in ecological communities: historical and geographic perspectives. University of Chicago Press, Chicago.Google Scholar
Veloz, S. D., Williams, J. W., Blois, J. L., He, F., Otto-Bliesner, B., and Liu, Z.. 2012. No-analogue climates and shifting realized niches during the late Quaternary: implications for 21st-century predictions by species distribution models. Global Change Biology 18:16981713.CrossRefGoogle Scholar
Wang, D., Gouhier, T. C., Menge, B. A., and Ganguly, A. R.. 2015. Intensification and spatial homogenization of coastal upwelling under climate change. Nature 518:390394.CrossRefGoogle ScholarPubMed
Waterson, A. M., Edgar, K. M., Schmidt, D. N., and Valdes, P. J.. 2017. Quantifying the stability of planktic foraminiferal physical niches between the Holocene and Last Glacial Maximum: niche stability of Planktic Foraminifera. Paleoceanography 32:7489.CrossRefGoogle Scholar
Wehmiller, J. F., Lajoie, K. R., Kvenvolden, K. A., Peterson, E., Belknap, D. F., Kennedy, G. L., et al. 1977. Correlation and chronology of Pacific coast marine deposits of continental United States by fossil amino acid stereochemistry-technique evaluation, relative ages, kinetic model ages, and geologic implications. Geological Survey Open-file Report 77–680.CrossRefGoogle Scholar
Williams, J. W., and Jackson, S. T.. 2007. Novel climates, no-analogue communities, and ecological surprises. Frontiers in Ecology and the Environment 5:475482.CrossRefGoogle Scholar
Woodring, W. P. 1951. Basic assumption underlying paleoecology. Science 113: 482483.Google Scholar
Woodring, W. P., Bramlette, M., and Kew, W. S.. 1946. Geology and paleontology of Palos Verdes Hills, California. U.S. Geological Survey Professional Paper 20753–55.Google Scholar
WoRMS Editorial Board. 2018. World Register of Marine Species. http://www.marinespecies.org, accessed 23 April 2017. doi: 10.14284/170.CrossRefGoogle Scholar