Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-16T03:35:22.606Z Has data issue: false hasContentIssue false

Chapter 11 - The North-East Pacific

Interactions on Subtidal Hard Substrata

Published online by Cambridge University Press:  07 September 2019

Stephen J. Hawkins
Affiliation:
Marine Biological Association of the United Kingdom, Plymouth
Katrin Bohn
Affiliation:
Natural England
Louise B. Firth
Affiliation:
University of Plymouth
Gray A. Williams
Affiliation:
The University of Hong Kong
Get access

Summary

Temperate, nutrient-rich waters along the west coast of North America support highly diverse assemblages of benthic flora and fauna. This chapter summarises recent literature on community composition, species interactions and interactions between species and the environment in rocky subtidal habitats of the north-east Pacific. Subtidal research in the region has long emphasised top-down processes, although these processes vary spatially and may be secondary to abiotic forces in some locations, particularly in the south. Recent research has highlighted the importance of mesoscale oceanographic processes, rapid anthropogenic changes and complex competitive and facilitative interactions in shaping north-east Pacific rocky bottom assemblages. Advances in genetics and chemistry are currently providing insights into paleogeographic history and the interplay of environmental variables and biotic interactions over time; these tools will be essential for anticipating ecosystem response to ongoing, rapid, anthropogenic-induced abiotic changes. Expanded long-term observational studies, as well as field experiments that clarify complex species interactions, are needed to support nearshore managers in the region respond to anthropogenic change.

Type
Chapter
Information
Interactions in the Marine Benthos
Global Patterns and Processes
, pp. 260 - 306
Publisher: Cambridge University Press
Print publication year: 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

Abbott, I. A. and Hollenberg, G. J. (1976). Marine Algae of California. Stanford University Press, Stanford, CA.Google Scholar
Addicott, W. O. (1970). Latitudinal gradients in tertiary molluscan faunas of the Pacific coast. Palaeogeography, Palaeoclimatology, Palaeoecology, 8 (4), 287312, http://dx.doi.org/10.1016/0031-0182(70)90103-3.Google Scholar
Addison, J. A. and Hart, M. W. (2005). Spawning, copulation and inbreeding coefficients in marine invertebrates. Biology Letters, 1 (August), 450–3, http://dx.doi.org/10.1098/rsbl.2005.0353.Google Scholar
Aguilar-Rosas, L. E., Aguilar-Rosas, R., Kawai, H. et al. (2007). New record of Sargassum filicinum Harvey (Fucales, Phaeophyceae) in the Pacific Coast of Mexico. Algae, 22 (1), 1721.Google Scholar
Alberto, F., Raimondi, P. T., Reed, D. C. et al. (2011). Isolation by oceanographic distance explains genetic structure for Macrocystis pyrifera in the Santa Barbara Channel. Molecular Ecology, 20 (12), 2543–54, http://dx.doi.org/10.1111/j.1365-294X.2011.05117.x.CrossRefGoogle ScholarPubMed
Aleem, A. A. (1973). Ecology of a kelp-bed in Southern California. Botanica Marina, 16 (2), 8395, http://dx.doi.org/10.1515/botm.1973.16.2.83.CrossRefGoogle Scholar
Van Alstyne, K. L., Harvey, E. L. and Cataldo, M. (2014). Effects of dopamine, a compound released by the green-tide macroalga Ulvaria obscura (Chlorophyta), on marine algae and invertebrate larvae and juveniles. Phycologia, 53 (2), 195202, http://dx.doi.org/10.2216/13-237.1.Google Scholar
Van Alstyne, K. L., Pelletreau, K. N. and Kirby, A. (2009). Nutritional preferences override chemical defenses in determining food choice by a generalist herbivore, Littorina sitkana. Journal of Experimental Marine Biology and Ecology, 379 (1–2), 8591, http://dx.doi.org/10.1016/j.jembe.2009.08.002.Google Scholar
De Anda-Montañez, J. A., Martínez-Aguilar, S., Balart, E. F. et al. (2016). Spatio-temporal distribution and abundance patterns of red crab Pleuroncodes planipes related to ocean temperature from the Pacific coast of the Baja California Peninsula. Fisheries Science, 82 (1), 115, http://dx.doi.org/10.1007/s12562-015-0938-8.Google Scholar
Andrews, K. S. and Harvey, C. J. (2013). Ecosystem-level consequences of movement: Seasonal variation in the trophic impact of a top predator. Marine Ecology Progress Series, 473, 247–60, http://dx.doi.org/10.3354/meps10095.Google Scholar
Anthony, R. G., Estes, J. A., Ricca, M. A., Miles, A. K. and Forsman, E. D. (2008). Bald eagles and sea otters in the Aleutian Archipelago: Indirect effects of trophic cascades. Ecology, 89 (10), 2725–35, http://dx.doi.org/10.1890/07-1818.1.Google Scholar
Arkema, K. D., Reed, D. D. and Schroeter, S. C. (2009). Direct and indirect effects of giant kelp determine benthic community structure and dynamics. Ecology, 90 (11), 3126–37.Google Scholar
Armstrong, R. (1976). Fugitive species: experiments with fungi and some theoretical considerations. Ecology, 57 (5), 953–63, http://dx.doi.org/10.2307/1941060.Google Scholar
Arreguın-Sánchez, F., Arcos, E. and Chávez, E. A. (2002). Flows of biomass and structure in an exploited benthic ecosystem in the Gulf of California, Mexico. Ecological Modelling, 156 (2–3), 167–83, http://dx.doi.org/10.1016/S0304-3800(02)00159-X.Google Scholar
Azad, A. K., Pearce, C. M. and McKinley, R. S. (2011). Effects of diet and temperature on ingestion, absorption, assimilation, gonad yield, and gonad quality of the purple sea urchin (Strongylocentrotus purpuratus). Aquaculture, 317 (1–4), 187–96, http://dx.doi.org/10.1016/j.aquaculture.2011.03.019.Google Scholar
Bakun, A., Black, B. A., Bograd, S. J. et al. (2015). Anticipated effects of climate change on coastal upwelling ecosystems. Current Climate Change Reports, 1 (2), 8593, http://dx.doi/10.1007/s40641-015-0008-4.Google Scholar
Barnard, P. L., Short, A. D., Harley, M. D. et al. (2015). Coastal vulnerability across the Pacific dominated by El Niño/Southern Oscillation. Nature Geoscience, 8 (10), 801–7, http://dx.doi.org/10.1038/ngeo2539.CrossRefGoogle Scholar
Barth, J. A., Menge, B. A., Lubchenco, J. et al. (2007). Delayed upwelling alters nearshore coastal ocean ecosystems in the northern California current. Proceedings of the National Academy of Sciences of the United States of America, 104 (10), 3719–24, http://dx.doi.org/10.1073/pnas.0700462104.Google Scholar
Basilio, A., Searcy, S. and Thompson, A. R. (2017). Effects of the Blob on settlement of spotted sand bass, Paralabrax maculatofasciatus, to Mission Bay, San Diego, CA. PLoS ONE, 12 (11), e0188449.CrossRefGoogle ScholarPubMed
Behrens Yamada, S. and Gillespie, G. E. (2008). Will the European green crab (Carcinus maenas) persist in the Pacific Northwest? ICES Journal of Marine Science: Journal du Conseil, 65 (5), 725–9, http://dx.doi.org/10.1093/icesjms/fsm191.Google Scholar
Belanger, C. L., Jablonski, D., Roy, K. et al. (2012). Global environmental predictors of benthic marine biogeographic structure. Proceedings of the National Academy of Sciences, 109 (35), 14046–51, http://dx.doi.org/10.1073/pnas.1212381109.Google Scholar
Bell, T. W., Cavanaugh, K. C. and Siegel, D. A. (2015). Remote monitoring of giant kelp biomass and physiological condition: an evaluation of the potential for the Hyperspectral Infrared Imager (HyspIRI) mission. Remote Sensing of Environment, 167, 218–28, http://dx.doi.org/10.1016/j.rse.2015.05.003.CrossRefGoogle Scholar
Benes, K. M. and Carpenter, R. C. (2015). Kelp canopy facilitates understory algal assemblage via competitive release during early stages of secondary succession. Ecology, 96 (1), 241–51, http://dx.doi.org/10.1890/14-0076.1.Google Scholar
Berger, M. S. and Jelinski, D. E. (2008). Spatial patterns of maternal investment in Strongylocentrotus franciscanus along a marine-terrestrial gradient. Marine Ecology Progress Series, 364, 119–27, http://dx.doi.org/10.3354/meps07490.CrossRefGoogle Scholar
Bernardi, G. (2000). Barriers to gene flow in Embiotoca jacksoni, a marine fish lacking a pelagic larval stage. Evolution, 54 (1), 226, http://dx.doi.org/10.1554/0014-3820(2000)054[0226:BTGFIE]2.0.CO;2.Google Scholar
Bernardi, G. (2005). Phylogeography and demography of sympatric sister surfperch species, Embiotoca jacksoni and E. Lateralis along the California coast: historical versus ecological factors. Evolution, 59 (2), 386, http://dx.doi.org/10.1554/04-367.Google Scholar
Bertocci, I., Araújo, R., Oliveira, P. and Sousa‐Pinto, I. (2015). Potential effects of kelp species on local fisheries. Journal of Applied Ecology, 52 (5), 1216–26, http://dx.doi.org/10.1111/1365-2664.12483.CrossRefGoogle Scholar
Bischof, K., Gómez, I., Molis, M. et al. (2007). Ultraviolet radiation shapes seaweed communities. Life in Extreme Environments, 2, 187212, http://dx.doi.org/10.1007/s11157-006-0002-3.Google Scholar
Blanchard, A. L. (2015). Variability of macrobenthic diversity and distributions in Alaskan sub-Arctic and Arctic marine systems with application to worldwide Arctic Systems. Marine Biodiversity, 45 (4), 781–95, http://dx.doi.org/10.1007/s12526-014-0292-6.Google Scholar
Blanchard, A. L., Feder, H. M., Hoberg, M. K. and Knowlton, A. L. (2017). Abiotic/biological interactions in coastal marine communities: insights from an Alaskan fjord. Estuaries and Coasts, 40 (5), 1398–417.CrossRefGoogle Scholar
Blanchette, C. A., Broitman, B. R. and Gaines, S. D. (2006). Intertidal community structure and oceanographic patterns around Santa Cruz Island, CA, USA. Marine Biology, 149 (3), 689701, http://dx.doi.org/10.1007/s00227-005-0239-3.CrossRefGoogle Scholar
Blanchette, C. A., Miner, B. G. and Gaines, S. D. (2002). Geographic variability in form, size and survival of Egregia menziesii around Point Conception, California. Marine Ecology Progress Series, 239, 6982, http://dx.doi.org/10.3354/meps239069.Google Scholar
Blanchette, C. A., Miner, C. M., Raimondi, P. T., Lohse, D., Heady, K. E. K. and Broitman, B. R. (2008). Biogeographical patterns of rocky intertidal communities along the Pacific coast of North America. Journal of Biogeography, 35 (9), 1593–607, http://dx.doi.org/10.1111/j.1365-2699.2008.01913.x.Google Scholar
Bodkin, J. L. (1986). Fish assemblages in Macrocystis and Nereocystis kelp forests off central California. Fishery Bulletin, 84 (4), 799808.Google Scholar
Bodkin, J. L. (1988). Effects of kelp forest removal on associated fish assemblages in central California. Journal of Experimental Marine Biology and Ecology, 117 (3), 227–38, http://dx.doi.org/10.1016/0022-0981(88)90059-7.Google Scholar
Boehm, A. B., Sanders, B. F. and Winant, C. D. (2002). Cross-shelf transport at Huntington Beach. Implications for the fate of sewage discharged through an offshore ocean outfall. Environmental Science and Technology, 36 (9), 1899–906, http://dx.doi.org/10.1021/es0111986.Google Scholar
Bograd, S., Schwing, F., Mendelssohn, R. and Green‐Jessen, P. (2002). On the changing seasonality over the North Pacific. Geophysical Research Letters, 29 (9), 1333, http://dx.doi.org/10.1029/2001GL013790.Google Scholar
Bograd, S. J., Schroeder, I., Sarkar, N. et al. (2009). Phenology of coastal upwelling in the California Current. Geophysical Research Letters, 36 (1), 15, http://dx.doi.org/10.1029/2008GL035933.Google Scholar
Bonaviri, C., Graham, M., Gianguzza, P. and Shears, N. T. (2017). Warmer temperatures reduce the influence of an important keystone predator. Journal of Animal Ecology, 86 (3), 490500.Google Scholar
Bond, N. A., Cronin, M. F., Freeland, H. and Mantua, N. (2015). Causes and impacts of teh 2014 warm anomaly in the north-east Pacific. Geophysical Research Letters, 42, 3414–20, http://dx.doi.org/10.1002/2015GL063306.Google Scholar
Botsford, L. (2001). Physical influences on recruitment to California Current invertebrate populations on multiple scales. ICES Journal of Marine Science, 58 (5), 1081–91, http://dx.doi.org/10.1006/jmsc.2001.1085.CrossRefGoogle Scholar
Botsford, L. W. and Lawrence, C. A. (2002). Patterns of co-variability among California Current chinook salmon, coho salmon, Dungeness crab, and physical oceanographic conditions. Progress in Oceanography, 53 (2–4), 283305, http://dx.doi.org/10.1016/S0079-6611(02)00034-4.Google Scholar
Boudreau, B. P. and Jørgensen, B. B. (2001). The Benthic Boundary Layer: Transport Processes and Biogeochemistry. Oxford University Press, Oxford.CrossRefGoogle Scholar
Breda, V. A. and Foster, M. S. (1985). Composition, abundance, and phenology of foliose red algae associated with two central California kelp forests. Journal of Experimental Marine Biology and Ecology, 94 (1–3), 115–30, http://dx.doi.org/10.1016/0022-0981(85)90053-X.Google Scholar
Briggs, J. C. (1974). Marine Zoogeography. McGraw-Hill, New York.Google Scholar
Briggs, J. C. (2003). Guest editorial: marine centres of origin as evolutionary engines. Journal of Biogeography, 30 (1), 118.Google Scholar
Briggs, J. C. (2007). Marine longitudinal biodiversity: causes and conservation. Diversity and Distributions, 13 (5), 544–55, http://dx.doi.org/10.1111/j.1472-4642.2007.00362.x.Google Scholar
Briggs, J. C. and Bowen, B. W. (2012). A realignment of marine biogeographic provinces with particular reference to fish distributions. Journal of Biogeography, 39 (1), 1230, http://dx.doi.org/10.1111/j.1365-2699.2011.02613.x.Google Scholar
Britton-Simmons, K. H. (2004). Direct and indirect effects of the introduced alga Sargassum muticum on benthic, subtidal communities of Washington State, USA. Marine Ecology Progress Series, 277, 6178, http://dx.doi.org/10.3354/meps277061.Google Scholar
Britton-Simmons, K. H. (2006). Functional group diversity, resource preemption and the genesis of invasion resistance in a community of marine algae. Oikos, 113 (3), 395401.Google Scholar
Britton-Simmons, K. H., Foley, G. and Okamoto, D. (2009). Spatial subsidy in the subtidal zone: Utilization of drift algae by a deep subtidal sea urchin. Aquatic Biology, 5 (3), 233–43, http://dx.doi.org/10.3354/ab00154.Google Scholar
Brodeur, R. D., Hunsicker, M. E., Hann, A. and Miller, T. W. (2018). Effects of warming ocean conditions on feeding ecology of small pelagic fishes in a coastal upwelling ecosystem: a shift to gelatinous food sources. Marine Ecology Progress Series, https://doi.org/10.3354/meps12497.Google Scholar
Brodeur, R., Ralston, S. and Emmett, R. (2006). Anomalous pelagic nekton abundance, distribution, and apparent recruitment in the northern California Current in 2004 and 2005. Geophysical Research Letters, 33 (22), L22S08.Google Scholar
Bromirski, P. D., Cayan, D. R. and Flick, R. E. (2005). Wave spectral energy variability in the northeast Pacific. Journal of Geophysical Research C: Oceans, 110 (3), 115, http://dx.doi.org/10.1029/2004JC002398.Google Scholar
Bromirski, P. D., Miller, A. J., Flick, R. E. and Auad, G. (2011). Dynamical suppression of sea level rise along the Pacific coast of North America: Indications for imminent acceleration. Journal of Geophysical Research: Oceans, 116 (7), 113, http://dx.doi.org/10.1029/2010JC006759.Google Scholar
Brown, D. W., McCain, B. B., Horness, B. H. et al. (1998). Status, correlations and temporal trends of chemical contaminants in fish and sediment from selected sites on the Pacific Coast of the USA. Marine Pollution Bulletin, 37 (1–2), 6785, http://dx.doi.org/10.1016/S0025-326X(98)00133-7.Google Scholar
Brown, N. E. M., Therriault, T. W. and Harley, C. D. G. (2016). Field-based experimental acidification alters fouling community structure and reduces diversity. Journal of Animal Ecology, 85 (5), 1328–39, http://dx.doi.org/10.1111/1365-2656.12557.Google Scholar
Brzezinski, M. A. and Washburn, L. (2011). Phytoplankton primary productivity in the Santa Barbara Channel: Effects of wind-driven upwelling and mesoscale eddies. Journal of Geophysical Research: Oceans, 116 (12), http://dx.doi.org/10.1029/2011JC007397.Google Scholar
Burd, B. J., Barnes, P. A. G., Wright, C. A. and Thomson, R. E. (2008). A review of subtidal benthic habitats and invertebrate biota of the Strait of Georgia, British Columbia. Marine Environmental Research, 66, S338, http://dx.doi.org/10.1016/j.marenvres.2008.09.004.Google Scholar
Burt, J. M., Tinker, M. T., Okamoto, D. K., Demes, K. W., Holmes, K. and Salomon, A. K. (2018). Sudden collapse of a mesopredator reveals its complementary role in mediating rocky reef regime shifts. Proceedings of the Royal Society B: Biological Sciences, 285 (1883), 20180553.Google Scholar
Burton, R. S. (1998). Intraspecific phylogeography across the Point Conception biogeographic boundary. Evolution, 52 (3), 734–45.Google Scholar
Busch, D. S., Harvey, C. J. and McElhany, P. (2013). Potential impacts of ocean acidification on the Puget Sound food web. ICES Journal of Marine Science, 70 (4), 823–33, http://dx.doi.org/10.1093/icesjms/fst176.CrossRefGoogle Scholar
Byers, J. E. (2002). Physical habitat attribute mediates biotic resistance to non-indigenous species invasion. Oecologia, 130, 146–56, http://dx.doi.org/10.1007/s004420100777.Google Scholar
Bylhouwer, B., Ianson, D. and Kohfeld, K. (2013). Changes in the onset and intensity of wind-driven upwelling and downwelling along the North American Pacific coast. Journal of Geophysical Research: Oceans, 118 (5), 2565–80, http://dx.doi.org/10.1002/jgrc.20194.Google Scholar
Byrnes, J. E., Reed, D. C., Cardinale, B. C. et al. (2011). Climate-driven increases in storm frequency simplify kelp forest food webs. Global Change Biology, 17 (8), 2513–24, http://dx.doi.org/10.1111/j.1365-2486.2011.02409.x.Google Scholar
Calvin, N. I. and Ellis, R. J. (1978). Quantitative and qualitative observations on Laminaria dentigera and other subtidal kelps of Southern Kodiak Island, Alaska. Marine Biology, 47, 331–6.Google Scholar
Carey, A. G. (1991). Ecology of North American Arctic continental shelf benthos: a review. Continental Shelf Research, 11 (8–10), 865–83, http://dx.doi.org/10.1016/0278-4343(91)90083-I.Google Scholar
Carlton, J. T. (1987). Patterns of transoceanic marine biological invasions in the Pacific-Ocean. Bulletin of Marine Science, 41 (2), 452–65.Google Scholar
Carlton, J. T. (1996). Pattern, process, and prediction in marine invasion ecology. Biological Conservation, 78 (1–2), 97106, http://dx.doi.org.10.1016/0006-3207(96)00020-1.Google Scholar
Carrington, E., Moeser, G. M., Dimond, J., Mello, J. J. and Boller, M. L. (2009). Seasonal disturbance to mussel beds: Field test of a mechanistic model predicting wave dislodgment. Limnology and Oceanography, 54 (3), 978–86, http://dx.doi.org/10.4319/lo.2009.54.3.0978.Google Scholar
Carter, S. K. and VanBlaricom, G. R. (2002). Effects of experimental harvest on red sea urchins (Strongylocentrotus franciscanus) in northern Washington. Fishery Bulletin, 100 (4), 662–73.Google Scholar
Carter, S. K., VanBlaricom, G. R. and Allen, B. L. (2007). Testing the generality of the trophic cascade paradigm for sea otters: A case study with kelp forests in northern Washington, USA. Hydrobiologia, 579 (1), 233–49, http://dx.doi.org/10.1007/s10750-006-0403-x.Google Scholar
Castilla, J. C. and Camus, P. A. (1992). The Humboldt- El Niño scenario: coastal benthic resources and anthropogenic influences, with particular reference to the 1982/83 ENSO. South African Journal of Marine Science, 12 (1), 703–12, http://dx.doi.org/10.2989/02577619209504735.Google Scholar
Castorani, M. C. N., Reed, D. C., Alberto, F. et al. (2015). Connectivity structures local population dynamics: a long-term empirical test in a large metapopulation system. Ecology, 96 (12), 3141–52, http://dx.doi.org/10.1890/15-0283.1.Google Scholar
Cavanaugh, K. C., Siegel, D. A., Kinlan, B. P. and Reed, D. C. (2010). Scaling giant kelp field measurements to regional scales using satellite observations. Marine Ecology Progress Series, 403, 1327, http://dx.doi.org/10.3354/meps08467.Google Scholar
Cavanaugh, K. C., Siegel, D. A., Reed, D. C. and Dennison, P. E. (2011). Environmental controls of giant-kelp biomass in the Santa Barbara Channel, California. Marine Ecology Progress Series, 429, 117, http://dx.doi.org/10.3354/meps09141.Google Scholar
Cavole, L. M., Demko, A. M., Diner, R. E. et al. (2016). Biological impacts of the 2013–2015 warm-water anomaly in the Northeast Pacific: winners, losers, and the future. Oceanography, 29 (2), 273–85.Google Scholar
Chauvet, P., Metaxas, A., Hay, A. E. and Matabos, M. (2018). Annual and seasonal dynamics of deep-sea megafaunal epibenthic communities in Barkley Canyon (British Columbia, Canada): a response to climatology, surface productivity and benthic boundary layer variation. Progress in Oceanography, 169, 89105.Google Scholar
Chavez, F. P., Strutton, P. G., Friederich, G. E. et al. (1999). Biological and chemical response of the equatorial pacific ocean to the 1997-98 El Nino. Science, 286 (5447), 2126–31, http://dx.doi.org/10.1126/science.286.5447.2126.Google Scholar
Checkley, D. M. and Barth, J. A. (2009). Patterns and processes in the California Current System. Progress in Oceanography, 83 (1–4), 4964, http://dx.doi.org/10.1016/j.pocean.2009.07.028.Google Scholar
Chenelot, H., Iken, K., Konar, B. and Edwards, M. (2006). Spatial and Temporal Distribution of Echinoderms in Rocky Nearshore Areas of Alaska. In The Nagisa World Congress, 11–28.Google Scholar
Chenillat, F., Rivière, P., Capet, X., Di Lorenzo, E. and Blanke, B. (2012). North Pacific gyre oscillation modulates seasonal timing and ecosystem functioning in the California current upwelling system. Geophysical Research Letters, 39 (1), 16, http://dx.doi.org.10.1029/2011GL049966.Google Scholar
Clark, R. P., Edwards, M. S. and Foster, M. S. (2004). Effects of shade from multiple kelp canopies on an understory algal assemblage. Marine Ecology Progress Series, 267, 107–19, http://dx.doi.org/10.3354/meps267107.Google Scholar
Clasen, J. L. and Shurin, J. B. (2015). Previously published works. Ecology, 96 (3), 862–72, http://dx.doi.org.10.1121/1.4929899.Google Scholar
Cohen, A. N. and Carlton, J. T. (1998). Accelerating invasion rate in a highly invaded estuary. Science, 279 (5350), 555–8, http://dx.doi.org/10.1126/science.279.5350.555.Google Scholar
Colombo-Pallotta, M. F., García-Mendoza, E. and Ladah, L. B. (2006). Photosynthetic performance, light absorption, and pigment composition of Macrocystis pyrifera (Laminariales, Phaeophyceae) blades from different depths. Journal of Phycology, 42 (6), 1225–34, http://dx.doi.org/10.1111/j.1529-8817.2006.00287.x.Google Scholar
Connell, J. H. and Keough, M. J. (1985). Disturbance and Patch Dynamics of Subtidal Marine Animals on Hard Substrata. In Pickett, S. T. and White, P. S., eds. The Ecology of Natural Disturbance and Patch Dynamics. Academic Press, Orlando, FL, 125–51.Google Scholar
Connell, J. H. and Slatyer, R. (1977). Mechanisms of succession in natural communities and their role in community stability and organization. The American Naturalist, 111 (982), 1119–44.Google Scholar
Connolly, S. R., Menge, B. A. and Roughgarden, J. (2001). A latitudinal gradient in recruitment of intertidal invertebrates in the northeast Pacific Ocean. Ecology, 82 (7), 17991813.Google Scholar
Connolly, S. R. and Roughgarden, J. (1998). A latitudinal gradient in northeast Pacific intertidal community structure: evidence for an oceanographically based synthesis of marine community theory. The American Naturalist, 151 (4), 311–26, http://dx.doi.org/10.1086/286121.Google Scholar
Cope, J. M. and Haltuch, M. A. (2012). Temporal and spatial summer groundfish assemblages in trawlable habitat off the west coast of the USA, 1977 to 2009. Marine Ecology Progress Series, 451, 187200, http://dx.doi.org/10.3354/meps09595.Google Scholar
Cordell, J. R., Levy, C. and Toft, J. D. (2013). Ecological implications of invasive tunicates associated with artificial structures in Puget Sound, Washington, USA. Biological Invasions, 15 (6), 1303–18, http://dx.doi.org/10.1007/s10530-012-0366-y.Google Scholar
Courtney-Hogue, C. (2016). Heavy metal accumulation in Lacistorhynchus dollfusi (Trypanorhyncha: Lacistorhynchidae) infecting Citharichthys sordidus (Pleuronectiformes: Bothidae) from Santa Monica Bay, Southern California. Parasitology, 143 (6), 794–9, http://dx.doi.org/10.1017/S003118201600038X.Google Scholar
Cowen, R. K. (1985). Large scale pattern of recruitment by the labrid, Semicossyphus pulcher: Causes and implications. Journal of Marine Research, 43 (3), 719–42, http://dx.doi.org/10.1357/002224085788440376.Google Scholar
Cowen, R. K., Agegian, C. R. and Foster, M. S. (1982). The maintenance of community structure in a central California giant kelp forest. Journal of Experimental Marine Biology and Ecology, 64 (2), 189201, http://dx.doi.org/10.1016/0022-0981(82)90152-6.Google Scholar
Crossett, K., Culliton, T. J., Wiley, P. C. and Goodspeed, T. R. (2004). Population trends along the coastal United States: 1980–2008, Coastal Trends Series Report, 1980–2008. US Department of Commerce, Washington, DC.Google Scholar
Cummins, P. F. and Masson, D. (2014). Climatic variability and trends in the surface waters of coastal British Columbia. Progress in Oceanography, 120, 279–90, http://dx.doi.org/10.1016/j.pocean.2013.10.002.Google Scholar
Dafforn, K. A., Glasby, T. M., Airoldi, L., Rivero, N. K., Mayer-Pinto, M. and Johnston, E. L. (2015). Marine urbanization: an ecological framework for designing multifunctional artificial structures. Frontiers in Ecology and the Environment, 13 (2), 8290, http://dx.doi.org/10.1890/140050.Google Scholar
Datsky, A. V. (2015). Ichthyofauna of the Russian exclusive economic zone of the Bering Sea: 1. Taxonomic diversity. Journal of Ichthyology, 55 (6), 792826, http://dx.doi.org/10.1134/S0032945215060065.Google Scholar
Davenport, A. C. and Anderson, T. W. (2007). Positive indirect effects of reef fishes on kelp performance: the importance of mesograzers. Ecology, 88 (6), 1548–61.Google Scholar
Davis, B. J., De Martini, E. E. and McGee, K. (1981). Gene flow among population of a teleost (painted greenling, Oxylebius pictus) from Puget Sound to southern California. Marine Biology, 65 (1), 1723, http://dx.doi.org/10.1007/BF00397063.Google Scholar
Dawson, A. G. (1992). Ice Age Earth: Late Quaternary Geology and Climate. Routledge, New York.Google Scholar
Dawson, M. (2001). Phylogeography of the tidewater goby, Eucyclogobius newberryi (Teleostei, Gobiidae), in Coastal California. Evolution, 55 (6), 1167–79.Google Scholar
Dawson, M. N., Waples, R. S. and Bernardi, G. (2006). Phylogeography Chapter. In Allen, L. G., Pondella, D. J. and Horn, M. H., eds. The Ecology of Marine Fishes: California and Adjacent Waters. University of California Press, Berkeley, pp. 2654.Google Scholar
Dayton, P. K. (1971). Competition, disturbance, and community organization: the provision and subsequent utilization of space in a rocky intertidal community. Ecological Monographs, 41 (4), 351–89.Google Scholar
Dayton, P. K. (1975). Experimental studies of algal canopy interactions in a sea otter dominated kelp community at Amchitka-Island, Alaska. Fishery Bulletin, 73 (2), 230–7.Google Scholar
Dayton, P. K. (1985). Ecology of kelp communities. Annual Review of Ecology and Systematics, 16, 215–45.Google Scholar
Dayton, P. K., Currie, V. and Gerrodette, T. I. M. (1984). Patch dynamics and stability of some california kelp communities. Ecological Monographs, 54 (3), 253–89.Google Scholar
Dayton, P. K. and Tegner, M. J. (1984). Catastrophic storms, El Niño, and patch stability in a southern California kelp community. Science, 224 (4646), 283–5.Google Scholar
Dayton, P. K., Tegner, M. J., Edwards, P. B. and Riser, K. L. (1999). Temporal and spatial scales of kelp demography: the role of oceanographic climate. Ecological Monographs, 69 (2), 219–50, http://dx.doi.org/10.1890/0012-9615(1999)069[0219:TASSOK]2.0.CO;2.Google Scholar
Dayton, P. K., Tegner, M. J., Parnell, P. E. and Edwards, P. B. (1992). Temporal and spatial patterns of disturbance and recovery in a kelp forest community. Ecological Society of America, 62 (3), 421–45.Google Scholar
Dean, T. A., Haldorson, L., Laur, D. R., Jewett, S. C. and Blanchard, A. (2000). The distribution of nearshore fishes in kelp and eelgrass communities in PrinceWilliam Sound, Alaska: associations with vegetation and physical habitat characteristics. Environmental Biology of Fishes, 57, 271–87.Google Scholar
Debenham, P., Brzezinski, M., Foltz, K. and Gaines, S. (2000). Genetic structure of populations of the red sea urchin, Strongylocentrotus franciscanus. Journal of Experimental Marine Biology and Ecology, 253 (1), 4962, http://dx.doi.org/10.1016/S0022-0981(00)00242-2.Google Scholar
Denny, M. W. (1987). Lift as a mechanism of patch initiation in the mussel beds. Journal of Experimental Marine Biology and Ecology, 113, 231–45.CrossRefGoogle Scholar
Dethier, M. N., Brown, A. S., Burgess, S. et al. (2014). Degrading detritus: Changes in food quality of aging kelp tissue varies with species. Journal of Experimental Marine Biology and Ecology, 460, 72–9, http://dx.doi.org/10.1016/j.jembe.2014.06.010.Google Scholar
Dever, E. P. (2004). Objective maps of near-surface flow states near Point Conception, California. Journal of Physical Oceanography, 34, 444–61.Google Scholar
Devinny, J. S. and Volse, L. A. (1978). Effects of sediments on the development of Macrocystis pyrifera gametophytes. Marine Biology, 48 (4), 343–8, http://dx.doi.org/10.1007/BF00391638.Google Scholar
Dominik, C. and Zimmerman, R. (2006). Dynamics of carbon allocation in a deep-water population of the deciduous kelp Pleurophycus gardneri (Laminariales). Marine Ecology Progress Series, 309, 143–57, http://dx.doi.org/10.3354/meps309143.Google Scholar
Doney, S. C., Ruckelshaus, M., Duffy, J. E. et al. (2012). Climate change impacts on marine ecosystems. Annual Review of Marine Science, 4, 1137, http://dx.doi.org/10.1146/annurev-marine-041911-111611.Google Scholar
O’Donnell, M. J., George, M. N. and Carrington, E. (2013). Mussel byssus attachment weakened by ocean acidification. Nature Climate Change, 3 (4), 14, http://dx.doi.org/10.1038/nclimate1846.Google Scholar
O’Donnell, M. J., Hammond, L. M. and Hofmann, G. E. (2009). Predicted impact of ocean acidification on a marine invertebrate: elevated CO2 alters response to thermal stress in sea urchin larvae. Marine Biology, 156 (3), 439–46, http://dx.doi.org/10.1007/s00227-008-1097-6.Google Scholar
Drew, E. A. (1983). Physiology of Laminaria. Marine Ecology, 4 (3), 227–50, http://dx.doi.org/10.1111/j.1439-0485.1983.tb00298.x.Google Scholar
Druehl, L. D. (1970). The pattern of Laminariales distribution in the northeast Pacific. Phycologia, 9 (3–4), 237–47, http://dx.doi.org/10.2216/i0031-8884-9-3-237.1.Google Scholar
Druehl, L. D. (1978). The distribution of Macrocystis integrifolia in British Columbia as related to environmental parameters. Canadian Journal of Botany, 56 (1), 6979.Google Scholar
Duggins, D. O. (1980). Kelp beds and sea otters: an experimental approach. Ecology, 61 (3), 447–53.Google Scholar
Duggins, D. O. (1983). Starfish predation and the creation of mosaic patterns in a kelp-dominated community. Ecology, 64 (6), 1610–19.Google Scholar
Duggins, D. Eckman, J., Siddon, C. and Klinger, T. (2003). Population, morphometric and biomechanical studies of three understory kelps along a hydrodynamic gradient. Marine Ecology Progress Series, 265, 5776, http://dx.doi.org/10.3354/meps265057.Google Scholar
Duggins, D. O. and Eckman, J. E. (1997). Is kelp detritus a good food for suspension feeders? Effects of kelp species, age and secondary metabolites. Marine Biology, 128 (3), 489–95, http://dx.doi.org/10.1007/s002270050115.Google Scholar
Duggins, D. O., Eckman, J. E., Siddon, C. E. and Klinger, T. (2001). Interactive roles of mesograzers and current flow in survival of kelps. Marine Ecology Progress Series, 223, 143–55, http://dx.doi.org/10.3354/meps223143.Google Scholar
Duggins, D. O., Gómez-Buckley, M. C., Buckley, R. M., Lowe, A. T., Galloway, A. W. E. and Dethier, M. N. (2016). Islands in the stream: kelp detritus as faunal magnets. Marine Biology, 163 (1), 110, http://dx.doi.org/10.1007/s00227-015-2781-y.Google Scholar
Duggins, D. O., Simenstad, C. A. and Estes, J. A. (1989). Magnification of secondary production by kelp detritus in coastal marine ecosystems. Science, 245 (4914), 170–3.Google Scholar
Dungan, M. L., Miller, T. E. and Thompson, D. A. (1982). Catastrophic decline of a top carnivore in the Gulf of California rocky intertidal zone. Science, 216 (4549), 989–91.Google Scholar
Dunn, R. P. and Hovel, K. A. (2019). Experiments reveal limited top‐down control of key herbivores in southern California kelp forests. Ecology, e02625.Google Scholar
Ebeling, A. W., Laur, D. R. and Rowley, R. J. (1985). Severe storm disturbances and reversal of community structure in a southern California kelp forest. Marine Biology, 84 (3), 287–94, http://dx.doi.org/10.1007/BF00392498.Google Scholar
Eberl, R., Mateos, M., Grosberg, R. K., Santamaria, C. A. and Hurtado, L. A. (2013). Phylogeography of the supralittoral isopod Ligia occidentalis around the Point Conception marine biogeographical boundary. Journal of Biogeography, 40 (12), 2361–72, http://dx.doi.org/10.1111/jbi.12168.Google Scholar
Ebert, T. A. and Russell, M. P. (1988). Latitudinal variation in size structure of the west coast purple sea urchin: a correlation with headlands. Limnology and Oceanography, 33 (2), 286–94, http://dx.doi.org/10.4319/lo.1988.33.2.0286.Google Scholar
Eckman, J. E. (1990). A model of passive settlement by planktonic larvae onto bottoms of differing roughness. Limnology and Oceanography, 35 (4), 887901, http://dx.doi.org/10.4319/lo.1990.35.4.0887.Google Scholar
Eckman, J. E., Duggins, D. O. and Sewell, A. T. (1989). Ecology of under story kelp environments. I. Effects of kelps on flow and particle transport near the bottom. Journal of Experimental Marine Biology and Ecology, 129 (2), 173–87, http://dx.doi.org/10.1016/0022-0981(89)90055-5.Google Scholar
Eckman, J. E., Duggins, D. O. and Siddon, C. E. (2003). Current and wave dynamics in the shallow subtidal: implications to the ecology of understory and surface-canopy kelps. Marine Ecology Progress Series, 265, 4556, http://dx.doi.org/10.3354/meps265045.Google Scholar
Edwards, M. S. and Estes, J. A. (2006). Catastrophe, recovery and range limitation in north-east Pacific kelp forests: a large-scale perspective. Marine Ecology Progress Series, 320, 7987, http://dx.doi.org/10.3354/meps320079.Google Scholar
Edwards, M. S. and Hernández-Carmona, G. (2005). Delayed recovery of giant kelp near its southern range limit in the North Pacific following El Niño. Marine Biology, 147 (1), 273–9, http://dx.doi.org/10.1007/s00227-004-1548-7.Google Scholar
Eisenlord, M. E., Groner, M. L., Yoshioka, R. M. et al. (2016). Ochre star mortality during the 2014 wasting disease epizootic: role of population size structure and temperature. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 371 (1689), http://dx.doi.org/10.1098/rstb.2015.0212.Google Scholar
Elahi, R., Birkeland, C., Sebens, K. P., Turner, K. R. and Dwyer, T. R. (2013). Limited change in the diversity and structure of subtidal communities over four decades. Marine Biology, 160 (12), 3209–19, http://dx.doi.org/10.1007/s00227-013-2308-3.Google Scholar
Elahi, R., Dwyer, T. R. and Sebens, K. P. (2014). Mesoscale variability in oceanographic retention sets the abiotic stage for subtidal benthic diversity. Marine Ecology Progress Series, 498, 117–32, http://dx.doi.org/10.3354/meps10642.Google Scholar
Elahi, R. and Sebens, K. P. (2012). Consumers mediate natural variation between prey richness and resource use in a benthic marine community. Marine Ecology Progress Series, 452, 131–43, http://dx.doi.org.10.3354/meps09603.Google Scholar
Elahi, R. and Sebens, K. P. (2013). Experimental removal and recovery of subtidal grazers highlights the importance of functional redundancy and temporal context. PLoS ONE, 8 (11), http://dx.doi.org/10.1371/journal.pone.0078969.Google Scholar
Emery, W. J. and Hamilton, K. (1985). Atmospheric forcing of interannual variability in the northeast Pacific Ocean: connections with El Nino. Journal of Geophysical Research, 90, 857–68.Google Scholar
Emmett, R., Llansó, R., Newton, J. et al. (2000). Geographic signatures of North American west coast estuaries. Estuaries, 23 (6), 765, http://dx.doi.org/10.2307/1352998.Google Scholar
Epelbaum, A., Herborg, L. M., Therriault, T. W. and Pearce, C. M. (2009). Temperature and salinity effects on growth, survival, reproduction, and potential distribution of two non-indigenous botryllid ascidians in British Columbia. Journal of Experimental Marine Biology and Ecology, 369 (1), 4352, http://dx.doi.org/10.1016/j.jembe.2008.10.028.CrossRefGoogle Scholar
Erlandson, J. M., Graham, M. H., Bourque, B. J., Corbett, D., Estes, J. A. and 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 (2), 161–74, http://dx.doi.org/10.1080/15564890701628612.Google Scholar
Estes, J. A., Danner, E. M., Doak, D. F. et al. (2004). Complex trophic interactions in kelp forest ecosystems. Bulletin of Marine Science, 74 (3), 621–38, www.ingentaconnect.com/content/umrsmas/bullmar/2004/00000074/00000003/art00010.Google Scholar
Estes, J. A. and Duggins, D. O. (1995). Sea otters and kelp forests in Alaska: generality and variation in a community ecological paradigm. Ecological Monographs, 65 (1), 75100, http://dx.doi.org/10.2307/2937159.Google Scholar
Estes, J. A., Lindberg, D. R. and Wray, C. (2005). Evolution of large body size in abalone (Haliotis): patterns and implications. Paleobiology, 31 (4), 591606, http://dx.doi.org/10.1666/0094-8373(2005)031[0591:EOLBSI]2.0.CO;2.Google Scholar
Estes, J. A. and Palmisano, J. F. (1974). Sea otters: their role in structuring nearshore communities. Science, 185, 1058–60, http://dx.doi.org/10.1126/science.185.4156.1058.Google Scholar
Estes, J. A., Smith, N. S. and Palmisano, J. F. (1978). Sea otter predation and community organization in the Western Aleutian Islands, Alaska. Ecology, 59 (4), 822–33.Google Scholar
Estes, J. A., Terborgh, J., Brashares, J. S. et al. (2011). Trophic downgrading of planet Earth. Science, 333 (6040), 301–7.Google Scholar
Estes, J. A., Tinker, M. T., Williams, T. M. and Doak, D. F. (1998). Killer whale predation on sea otters linking oceanic and nearshore ecosystems. Science, 282, 473–6, http://dx.doi.org/10.1126/science.282.5388.473.Google Scholar
Etherington, L. L., Hooge, P. N., Hooge, E. R. and Hill, D. F. (2007). Oceanography of Glacier Bay, Alaska: implications for biological patterns in a glacial fjord estuary. Estuaries and Coasts, 30 (6), 927–44, http://dx.doi.org/10.1007/BF02841386.Google Scholar
Fautin, D., Dalton, P., Incze, L. S. et al. (2010). An overview of marine biodiversity in United States waters. PLoS ONE, 5 (8), http://dx.doi.org/10.1371/journal.pone.0011914.Google Scholar
Feehan, C. J., Ludwig, Z., Yu, S. and Adams, D. K. (2018). Synergistic negative effects of thermal stress and altered food resources on echinoid larvae. Scientific Reports, 8 (1), 12229.Google Scholar
Feely, R. A., Alin, S. R., Newton, J. et al. (2010). The combined effects of ocean acidification, mixing, and respiration on pH and carbonate saturation in an urbanized estuary. Estuarine, Coastal and Shelf Science, 88 (4), 442–9, http://dx.doi.org/10.1016/j.ecss.2010.05.004.Google Scholar
Feely, R. A., Sabine, C. L., Hernandez-Ayon, J. M., Ianson, D. and Hales, B. (2008). Evidence for upwelling of corrosive “acidified” water onto the continental shelf. Science, 320 (5882), 1490–2, http://dx.doi.org/10.1126/science.1155676.Google Scholar
Fiedler, P. C., Methot, R. D. and Hewitt, R. P. (1986). Effects of California El Niño 1982–1984 on the northern anchovy. Journal of Marine Research, 44 (2), 317–38, http://dx.doi.org/10.1357/002224086788405365.Google Scholar
Forchhammer, M. C. and Post, E. (2004). Using large-scale climate indices in climate change ecology studies. Population Ecology, 46 (1), 112, http://dx.doi.org/10.1007/s10144-004-0176-x.Google Scholar
Foreman, M. G. G., Pal, B. and Merryfield, W. J. (2011). Trends in upwelling and downwelling winds along the British Columbia shelf. Journal of Geophysical Research: Oceans, 116 (10), 111, http://dx.doi.org/10.1029/2011JC006995.Google Scholar
Foreman, R. E. (1977). Benthic community modification and recovery following intensive grazing by Strongylocentrotus droebachiensis. Helgoländer Wissenschaftliche Meeresuntersuchungen, 30 (1–4), 468–84, http://dx.doi.org/10.1007/BF02207855.Google Scholar
Foreman, R. E. (1984). Studies on Nereocystis growth in British Columbia, Canada. Hydrobiologia, 116/117, 325–32, http://link.springer.com/chapter/10.1007/978-94-009-6560-7_65.Google Scholar
Foster, M. S. (1990). Organization of macroalgal assemblages in the Northeast Pacific: the assumption of homogeneity and the illusion of generality. Hydrobiologia, 192, 2133.Google Scholar
Foster, M. S. (1992). How Important Is Grazing to Seaweed Evolution and Assemblage Structure in the North-East Pacific. In John, D. M., Hawkins, S. J. and Price, J. H., eds. Plant–Animal Interactions in the Marine Benthos. Oxford University Press, Oxford, pp. 6185.Google Scholar
Foster, M. S. and Schiel, D. R. (1985). The Ecology of Giant Kelp Forests in California: A Community Profile. US Fish and Wildlife Service, Washington, D.C.Google Scholar
Foster, M. S. and Schiel, D. R. (1988). Kelp Communities and Sea Otters: Keystone Species or Just Another Brick in the Wall?’ In The Community Ecology of Sea Otters. Springer, Berlin, pp. 92115, http://dx.doi.org/10.1007/978-3-642-72845-7_5.Google Scholar
Foster, M. S. and Schiel, D. R. (2010). Loss of predators and the collapse of southern California kelp forests (?): alternatives, explanations and generalizations. Journal of Experimental Marine Biology and Ecology, 393 (1–2), 5970, http://dx.doi.org/10.1016/j.jembe.2010.07.002.Google Scholar
Fram, J. P., Stewart, H. L., Brzezinski, M. A. et al. (2008). Physical pathways and utilization of nitrate supply to the giant kelp, Macrocystis pyrifera. Limnology and Oceanography, 53 (4), 1589–603, http://dx.doi.org/10.4319/lo.2008.53.4.1589.Google Scholar
Francis, R. C., Hare, S. R., Hollowed, A. B. and Wooster, W. S. (1998). Effects of interdecadal climate variability on the oceanic ecosystems of the north-east Pacific. Fisheries Oceanography, 7 (1), 121, http://dx.doi.org/10.1046/j.1365-2419.1998.00052.x.Google Scholar
Freeland, H. J. (2006). What proportion of the north pacific current finds its way into the Gulf of Alaska?Atmosphere-Ocean, 44 (4), 321–30, http://dx.doi.org/10.3137/ao.440401.Google Scholar
Freeland, H. J. (2013). Evidence of change in the winter mixed layer in the northeast Pacific Ocean: a problem revisited. Atmosphere-Ocean, 5900, 3741, http://dx.doi.org/10.1080/07055900.2012.754330.Google Scholar
Freeland, H. J. and Denman, K. L. (1982). A topographically controlled upwelling center off southern Vancouver Island. Journal of Marine Research, 40 (4), 1069–93.Google Scholar
Fritchman, H. K. (1961). A study of the reproductive cycle in the California Acmaeidae (Gastropoda). Veliger, 4, 4147, www.sciencedirect.com/science/article/pii/0022098173900579.Google Scholar
Funes-Rodríguez, R., Ruíz‐Chavarría, J. A., González‐Armas, R., Durazo, R. and Guzmán‐del Proó, S. A. (2015). Influence of hydrographic conditions on the distribution of spiny lobster larvae off the west coast of Baja California. Transactions of the American Fisheries Society, 144 (6), 1192–205, http://dx.doi.org/10.1080/00028487.2015.1083474.Google Scholar
Gaines, S. and Roughgarden, J. (1985). Larval settlement rate: a leading determinant of structure in an ecological community of the marine intertidal zone. Proceedings of the National Academy of Sciences of the United States of America, 82 (11), 3707–11, http://dx.doi.org/10.1073/pnas.82.11.3707.Google Scholar
Gaitán-Espitia, J. D., Hancock, J. R., Padilla-Gamiño, J. L. et al. (2014). Interactive effects of elevated temperature and pCO2 on early-life-history stages of the giant kelp Macrocystis pyrifera. Journal of Experimental Marine Biology and Ecology, 457, 51–8, http://dx.doi.org/10.1016/j.jembe.2014.03.018.Google Scholar
Galloway, A. W. E., Lowe, A. T., Sosik, E. A., Yeung, J. S. and Duggins, D. O. (2013). Fatty acid and stable isotope biomarkers suggest microbe-induced differences in benthic food webs between depths. Limnology and Oceanography, 58 (4), 1451–62, http://dx.doi.org/10.4319/lo.2013.58.4.1451.Google Scholar
Garbary, D. J., Kim, K. Y., Klinger, T. and Duggins, D. (1999). Preliminary observations on the development of kelp gametophytes endophytic in red algae. Hydrobiologia, 398, 247–52.Google Scholar
Gasbarro, R., Wan, D. and Tunnicliffe, V. (2018). Composition and functional diversity of macrofaunal assemblages on vertical walls of a deep northeast Pacific fjord. Marine Ecology Progress Series, 597, 4764.Google Scholar
Gattuso, J.-P., Gentili, B., Duarte, C. M., Kleypas, J. A., Middelburg, J. J. and Antoine, D. (2006). Light availability in the coastal ocean: impact on the distribution of benthic photosynthetic organisms and contribution to primary production. Biogeosciences Discussions, 3 (4), 895959, http://dx.doi.org/10.5194/bgd-3-895-2006.Google Scholar
Gaylord, B. and Gaines, S. D. (2000). Temperature or transport? Range limits in marine species mediated solely by flow. The American Naturalist, 155 (6), 769–89, http://dx.doi.org/10.1086/303357.Google Scholar
Gaylord, B., Reed, D. C., Washburn, L. and Raimondi, P. T. (2004). Physical-biological coupling in spore dispersal of kelp forest macroalgae. Journal of Marine Systems, 49 (1–4), 1939, http://dx.doi.org/10.1016/j.jmarsys.2003.05.003.Google Scholar
Gaylord, B., Rosman, J. H., Reed, D. C. et al. (2007). Spatial patterns of flow and their modification within and around a giant kelp forest. Limnology and Oceanography, 52 (5), 1838–52, http://dx.doi.org/10.4319/lo.2007.52.5.1838.Google Scholar
Gaylord, B. P., Nickols, K. J. and Jurgens, L. (2012). Roles of transport and mixing processes in kelp forest ecology. Journal of Experimental Biology, 215 (6), 9971007, http://dx.doi.org/10.1242/jeb.059824.Google Scholar
Genovese, S. J. and Witman, J. D. (1999). Interactive effects of flow speed and particle concentration on growth rates of an active suspension feeder. Limnology and Oceanography, 44 (4), 1120–31.Google Scholar
Gerard, V. A. (1982). Growth and utilization of internal nitrogen reserves by the giant kelp Macrocystis pyrifera in a low-nitrogen environment. Marine Biology, 66 (1), 2735, http://dx.doi.org/10.1007/BF00397251.Google Scholar
Gladenkov, A. Y., Oleinik, A. E., Marincovich, L. Jr and Barinov, K. B. (2002). A refined age for the earliest opening of Bering Strait. Palaeogeography, Palaeoclimatology, Palaeoecology, 183 (3–4), 321–8, http://dx.doi.org/10.1016/S0031-0182(02)00249-3.Google Scholar
Glynn, P. W. (1988). El Nino-southern oscillation 1982–1983: nearshore population, community, and ecosystem responses. Annual Review of Ecology and Systematics, 19 (1), 309–46, http://dx.doi.org/10.1146/annurev.es.19.110188.001521.Google Scholar
Goddard, J. H. R., Treneman, N., Pence, W. E. et al. (2016). Nudibranch range shifts associated with the 2014 warm anomaly in the north-east Pacific. Bulletin of the Southern California Academy of Sciences, 115 (1), 1540.Google Scholar
Gottscho, A. D. (2016). Zoogeography of the San Andreas Fault system: great Pacific fracture zones correspond with spatially concordant phylogeographic boundaries in western North America. Biological Reviews, 91 (1), 235–54, http://dx.doi.org/10.1111/brv.12167.Google Scholar
Graham, K. R. and Sebens, K. P. (1996). The distribution of marine invertebrate larvae near vertical surfaces in the rocky subtidal zone. Ecology, 77 (3), 933–49.Google Scholar
Graham, M. H. (2004). Effects of local deforestation on the diversity and structure of southern California giant kelp forest food webs. Ecosystems, 7 (4), 341–57, http://dx.doi.org/10.1007/s10021-003-0245-6.Google Scholar
Graham, M. H., Kinlan, B. P. and Grosberg, R. K. (2009). Post-glacial redistribution and shifts in productivity of giant kelp forests. Proceedings of the Royal Society B, 277 (1680), 399406, http://dx.doi.org/10.1098/rspb.2009.1664.Google Scholar
Graham, W. M. and Largier, J. L. (1997). Upwelling shadows as nearshore retention sites: the example of northern Monterey Bay. Continental Shelf Research, 17 (5), 509–32.Google Scholar
Grantham, B. A., Eckert, G. L. and Shanks, A. L. (2003). Dispersal potential of marine invertebrates in diverse habitats. Ecological Applications, 13 (1), 108–16, http://dx.doi.org/10.1890/1051-0761(2003)013[0108:DPOMII]2.0.CO;2.Google Scholar
Grason, E. W. and Buhle, E. R. (2016). Comparing the influence of native and invasive intraguild predators on a rare native oyster. Journal of Experimental Marine Biology and Ecology, 479, 18, http://dx.doi.org/10.1016/j.jembe.2016.02.012.Google Scholar
Grason, E. W. and Miner, B. G. (2012). Behavioral plasticity in an invaded system: Non-native whelks recognize risk from native crabs. Oecologia, 169 (1), 105–15, http://dx.doi.org/10.1007/s00442-011-2188-5.Google Scholar
Gruber, N., Hauri, C., Lachkar, Z., Loher, D., Frölicher, T. L. and Plattner, G.-K. (2012). Rapid progression of ocean acidification in the California Current System. Science, 337 (6091), 220–3, http://dx.doi.org/10.1126/science.1216773.Google Scholar
Gudenkauf, B. M. and Hewson, I. (2015). Metatranscriptomic analysis of Pycnopodia helianthoides (Asteroidea) affected by sea star wasting disease. PLoS ONE, 10 (5), e0128150, http://dx.doi.org/10.1371/journal.pone.0128150.Google Scholar
Gutt, J. (2001). On the direct impact of ice on marine benthic communities, a review. Polar Biology, 24 (8), 553–64, http://dx.doi.org/10.1007/s003000100262.Google Scholar
Haggerty, M. B., Anderson, T. W. and Long, J. D. (2018). Fish predators reduce kelp frond loss via a trait‐mediated trophic cascade. Ecology, 99 (7), 1574–83.Google Scholar
Haigh, R., Ianson, D., Holt, C. A., Neate, H. E. and Edwards, A. M. (2015). Effects of ocean acidification on temperate coastal marine ecosystems and fisheries in the northeast pacific. PLoS ONE, 10 (2), 146, http://dx.doi.org/10.1371/journal.pone.0117533.Google Scholar
Hall, C. A. (1964). Shallow-water marine climates and molluscan. Ecology, 45 (2), 226–34.Google Scholar
Hallett, T. B., Coulson, T., Pilkington, J. G., Clutton-Brock, T. H., Pemberton, J. M. and Grenfell, B. T. (2004). Why large-scale climate indices seem to predict ecological processes better than local weather. Nature, 430 (6995), 71–5, http://dx.doi.org/10.1038/nature02708.Google Scholar
Hamilton, J. and Konar, B. (2007). Implications of substrate complexity and kelp variability for south-central Alaskan nearshore fish communities. Fishery Bulletin, 105 (2), 189–96.Google Scholar
Hamilton, S. L., Caselle, J. E., Malone, D. P. and Carr, M. H. (2010). Incorporating biogeography into evaluations of the Channel Islands marine reserve network. Proceedings of the National Academy of Sciences of the United States of America, 107 (43), 18272–7, http://dx.doi.org/10.1073/pnas.0908091107.Google Scholar
Hammond, L. M. and Hofmann, G. E. (2010). Thermal tolerance of Strongylocentrotus purpuratus early life history stages: Mortality, stress-induced gene expression and biogeographic patterns. Marine Biology, 157 (12), 2677–87, http://dx.doi.org/10.1007/s00227-010-1528-z.Google Scholar
Hare, S. R. and Mantua, N. J. (2000). Empirical evidence for North Pacific regime shifts in 1977 and 1989. Progress in Oceanography, 47 (2–4), 103–45, http://dx.doi.org/10.1016/S0079-6611(00)00033-1.Google Scholar
Harley, C. D. G., Anderson, K. M., Demes, K. W. et al. (2012). Effects of climate change on global seaweed communities. Journal of Phycology, 48 (5), 1064–78, http://dx.doi.org/10.1111/j.1529-8817.2012.01224.x.Google Scholar
Harley, C. D. G. and Connell, S. D. (2009). Shifts in Abiotic Variables and Consequences for Diversity. In Wahl, M., ed. Marine Hard Bottom Communities: Patterns, Dynamics, Diversity, and Change. Springer, New York, pp. 257–68, http://dx.doi.org/10.1007/978-3-540-92704-4.Google Scholar
Harley, C. D. G., Hughes, A. R., Hultgren, K. M. et al. (2006). The impacts of climate change in coastal marine systems: Climate change in coastal marine systems. Ecology Letters, 9 (2), 228–41, http://dx.doi.org/10.1111/j.1461-0248.2005.00871.x.Google Scholar
Harley, C. D. G. and Lopez, J. P. (2005). The natural history, thermal physiology, and ecological impacts of intertidal mesopredators, Oedoparena spp. (Diptera: Dryomyzidae). Invertebrate Biology, 122 (1), 6173, http://dx.doi.org/10.1111/j.1744-7410.2003.tb00073.x.Google Scholar
Harrold, C. and Reed, D. C. (1985). Food availability, sea urchin grazing, and kelp forest community structure. Ecology, 66 (4), 1160–9.Google Scholar
Harvell, C. D., Montecino-Latorre, D., Caldwell, J. M. et al. (2019). Disease epidemic and a marine heat wave are associated with the continental-scale collapse of a pivotal predator (Pycnopodia helianthoides). Science Advances, 5 (1), eaau7042.Google Scholar
Harvey, J. B. J., Hoy, M. S. and Rodriguez, R. J. (2009). Molecular detection of native and invasive marine invertebrate larvae present in ballast and open water environmental samples collected in Puget Sound. Journal of Experimental Marine Biology and Ecology, 369 (2), 93–9, http://dx.doi.org/10.1016/j.jembe.2008.10.030.Google Scholar
Hastings, A. (1988). Food web theory and stability. Ecology, 69 (6), 1665–8, http://dx.doi.org/10.2307/1941143.Google Scholar
Hawkes, M. W. (1981). Porphyra nereocystis and P. thuretii (Rhodophyta): gametophyte morphology, distribution, and occurrence. Syesis, 14, 97108.Google Scholar
Hendler, G. (2013). Recent mass mortality of Strongylocentrotus purpuratus (Echinodermata: Echinoidea) at Malibu and a review of purple sea urchin kills elsewhere in California. Bulletin. Southern California Academy of Sciences, 112 (1), 1937, www.bioone.org/doi/abs/10.3160/0038-3872-112.1.19.Google Scholar
Hernández-Carmona, G., Robledod, D. and Serviere-Zaragozab, E. (2001). Effect of nutrient availability on Macrocystis pyrifera recruitment and survival near its southern limit off Baja California. Botanica Marina, 44, 221–9.Google Scholar
Hewson, I., Button, J. B., Gudenkauf, B. M. et al. (2014). Densovirus associated with sea-star wasting disease and mass mortality. Proceedings of the National Academy of Sciences of the United States of America, 111 (48), 17278–83, http://dx.doi.org/10.1073/pnas.1416625111.Google Scholar
Hickerson, M. J. and Ross, J. R. P. (2001). Post-glacial population history and genetic structure of the northern clingfish (Gobbiesox maeandricus), revealed from mtDNA analysis. Marine Biology, 138 (2), 407–19, http://dx.doi.org/10.1007/s002270000465.Google Scholar
Hickey, B. M. and Banas, N. S. (2003). Oceanography of the U.S. Pacific northwest coastal ocean and estuaries with application to coastal ecology. Estuaries, 26 (4), 1010–31, http://dx.doi.org/10.1007/BF02803360.Google Scholar
Hobday, A. J. (2000). Persistence and transport of fauna on drifting kelp (Macrocystis pyrifera (L.) C. Agardh) rafts in the Southern California Bight. Journal of Experimental Marine Biology and Ecology, 253 (1), 7596, http://dx.doi.org/10.1016/S0022-0981(00)00250-1.Google Scholar
Hoegh-Guldberg, O. and Bruno, J. (2010). The impact of climate change on the world’s marine ecosystems. Science, 328 (5985), 1523–8, http://dx.doi.org/10.1126/science.1189930.Google Scholar
Hogue, C. and Swig, B. (2007). Habitat quality and endoparasitism in the Pacific sanddab Citharichthys sordidus from Santa Monica Bay, southern California. Journal of Fish Biology, 70 (1), 231–42, http://dx.doi.org/10.1111/j.1095-8649.2006.01298.x.Google Scholar
Holbrook, S. J., Schmitt, R. J., Carr, M. H. and Coyer, J. A. (1990). Effect of giant kelp on local abundance of reef fishes: the importance of ontogenetic resource requirements. Bulletin of Marine Science, 47 (1), 104–14.Google Scholar
Holbrook, S. J., Schmitt, R. J. and Stephens, J. S. (1997). Changes in an assemblage of temperate reef fishes associated with a climate shift. Ecological Applications, 7 (4), 1299–310, http://dx.doi.org/10.1890/1051-0761(1997)007.Google Scholar
Holsman, K. K., McDonald, P. S. and Armstrong, D. A. (2006). Intertidal migration and habitat use by subadult Dungeness crab Cancer magister in a north-east Pacific estuary. Marine Ecology Progress Series, 308, 183–95, http://dx.doi.org/10.3354/meps308183.Google Scholar
Hood, E., Fellman, J., Spencer, R. G. M. et al. (2009). Glaciers as a source of ancient and labile organic matter to the marine environment. Nature, 462 (7276), 1044–7, http://dx.doi.org/10.1038/nature08580.Google Scholar
Horn, M. H., Allen, L. G. and Lea, R. N. (2006). Biogeography. In Allen, L. G., Pondella, D. J. and Horn, M. H., eds. The Ecology of Marine Fishes: California and Adjacent Waters. University of California Press, Berkeley, pp. 325, http://dx.doi.org/10.1525/gfc.2008.8.4.24.Google Scholar
Hornberger, M. I., Luoma, S. N., van Geen, A., Fuller, C. and Anima, R. (1999). Historical trends of metals in the sediments of San Francisco Bay, California. Marine Chemistry, 64 (1–2), 3955, http://dx.doi.org/10.1016/S0304-4203(98)80083-2.Google Scholar
Hoshijima, U. and Hofmann, G. (2019). Variability of seawater chemistry in a kelp forest environment is linked to in situ transgenerational effects in the purple sea urchin, Strongylocentrotus purpuratus. Frontiers in Marine Science, 6, 62.Google Scholar
Hubbard, C. B., Garbary, D. J., Kim, K. Y. and Chiasson, D. M. (2004). Host specificity and growth of kelp gametophytes symbiotic with filamentous red algae (Ceramiales, Rhodophyta). Helgoland Marine Research, 58 (1), 1825, http://dx.doi.org/10.1007/s10152-003-0162-2.Google Scholar
Hurd, C. L. (2000). Water motion, marine macroalgal physiology, and production. Journal of Phycology, 36 (3), 453–72, http://dx.doi.org/10.1046/j.1529-8817.2000.99139.x.Google Scholar
Huyer, A. (1983). Coastal upwelling in the California Current System. Progress in Oceanography, 12 (3), 259–84, http://dx.doi.org/10.1016/0079-6611(83)90010-1.Google Scholar
Huyer, A., Smith, R. L. and Fleischbein, J. (2002). The coastal ocean off Oregon and northern California during the 1997–98 El Niño. Progress in Oceanography, 54, 311–41.Google Scholar
Jackson, G. A. (1977). Nutrients and production of giant kelp, Macrocystis pyrifera, off southern California. Limnology and Oceanography, 22, 979–95, http://dx.doi.org/10.4319/lo.1977.22.6.0979.Google Scholar
Jacobs, D. K., Haney, T. A. and Louie, K. D. (2004). Genes, diversity, and geologic process on the Pacific Coast. Annual Review of Earth and Planetary Sciences, 32 (1), 601–52, http://dx.doi.org/10.1146/annurev.earth.32.092203.122436.Google Scholar
Jacox, M. G., Hazen, E. L., Zaba, K. D. et al. (2016). Impacts of the 2015-2016 El Niño on the California current system: early assessment and comparison to past events. Geophysical Research Letters, 58 (1), 1825, http://dx.doi.org/10.1002/2016GL069716.Google Scholar
Jensen, M. M. and Denny, M. W. (2015). Experimental determination of the hydrodynamic forces responsible for wave impact events. Journal of Experimental Marine Biology and Ecology, 469, 123–30, http://dx.doi.org/10.1016/j.jembe.2015.04.013.Google Scholar
Jensen, M. M. and Denny, M. W. (2016). Life in an extreme environment: characterizing wave-imposed forces in the rocky intertidal zone using high temporal resolution hydrodynamic measurements. Limnology and Oceanography, 61 (5), 1750–61, http://dx.doi.org/10.1002/lno.10327.Google Scholar
Jeong, Y., Grant, S. B., Ritter, S. et al. (2005). Identifying pollutant sources in tidally mixed systems: case study of fecal indicator bacteria from marinas in Newport Bay, Southern California. Environmental Science and Technology, 39 (23), 9083–93, http://dx.doi.org/10.1021/es0482684.Google Scholar
Jessee, W. N., Carpenter, A. L. and Carter, J. W. (1985). Distribution patterns and density estimates of fishes on a southern California artificial reef with comparisons to natural kelp-reef habitats. Bulletin of Marine Science, 37 (1), 214–26.Google Scholar
Johansson, M. L., Alberto, F., Reed, D. C. et al. (2015). Seascape drivers of Macrocystis pyrifera population genetic structure in the northeast Pacific. Molecular Ecology, 24 (19), 4866–85, http://dx.doi.org/10.1111/mec.13371.Google Scholar
Johnson, G., Attrill, M. and Sheehan, E. (2007). Recovery of meiofauna communities following mudflat disturbance by trampling associated with crab-tiling. Marine Environmental Research, 64 (4), 409–16, www.sciencedirect.com/science/article/pii/S0141113607000438.Google Scholar
Kagley, A. N., Snider, R. G., Kardong, K. E. and Casillas, E. (2014). Effects of chemical contaminants on growth, age-structure, and reproduction of mytilus edulis complex from Puget Sound, Washington. Bulletin of Environmental Contamination and Toxicology, 93 (1), 712.Google Scholar
Kahn, A. S., Ruhl, H. A. and Smith, K. L. (2012). Temporal changes in deep-sea sponge populations are correlated to changes in surface climate and food supply. Deep-Sea Research Part I: Oceanographic Research Papers, 70, 3641, http://dx.doi.org/10.1016/j.dsr.2012.08.001.Google Scholar
Kavanaugh, M. T., Nielsen, K. J., Chan, F. T. et al. (2009). Experimental assessment of the effects of shade on an intertidal kelp: do phytoplankton blooms inhibit growth of open coast macroalgae? Limnology and Oceanography, 54 (1), 276–88, http://dx.doi.org/10.4319/lo.2009.54.1.0276.Google Scholar
Kawamata, S. (1998). Effect of wave-induced oscillatory flow on grazing by a subtidal sea urchin Strongylocentrotus nudus (A. Agassiz). Journal of Experimental Marine Biology and Ecology, 224 (1), 3148, http://dx.doi.org/10.1016/S0022-0981(97)00165-2.Google Scholar
Keever, C. C., Sunday, J., Puritz, J. B. et al. (2009). Discordant distribution of populations and genetic variation in a sea star with high dispersal potential. Evolution, 63 (12), 3214–27, http://dx.doi.org/10.1111/j.1558-5646.2009.00801.x.Google Scholar
Kelly, R. P. and Palumbi, S. R. (2010). Genetic structure among 50 species of the northeastern pacific rocky intertidal community. PLoS ONE, 5 (1), http://dx.doi.org/10.1371/journal.pone.0008594.Google Scholar
King, J. R., Agostini, V. N., Harvey, C. J. et al. (2011). Climate forcing and the California current ecosystem. ICES Journal of Marine Science, 68 (6), 1199–216, http://dx.doi.org/10.1093/icesjms/fsr009.Google Scholar
Klinger, T., Padilla, D. K. and Britton-Simmons, K. (2006). Two invaders achieve higher densities in reserves. Aquatic Conservation: Marine and Freshwater Ecosystems, 16 (3), 301–11, http://dx.doi.org/10.1002/aqc.717.Google Scholar
Koehl, M. A. R., Powell, T. M. and Dairiki, G. (1993). Measuring the Fate of Patches in the Water: Larval Dispersal. In Levin, S. A., Powell, T. M. and Steele, J. H., eds. Patch Dynamics. Springer-Verlag, Berlin, pp. 5060.Google Scholar
Koenigs, C., Miller, R. and Page, H. (2015). Top predators rely on carbon derived from giant kelp Macrocystis pyrifera. Marine Ecology Progress Series, 537, 18, http://dx.doi.org/10.3354/meps11467.Google Scholar
Kohl, W. T., McClure, T. I. and Miner, B. G. (2016). Decreased temperature facilitates short-term sea star wasting disease survival in the keystone intertidal sea star Pisaster ochraceus. PLoS ONE, 11 (4), e0153670, http://dx.doi.org/10.1371/journal.pone.0153670.Google Scholar
Kokkinakis, S. A. and Wheeler, P. A. (1987). Nitrogen uptake and phytoplankton growth in coastal upwelling regions. Limnology and Oceanography, 32 (5), 1112–23, http://dx.doi.org/10.4319/lo.1987.32.5.1112.Google Scholar
Konar, B., Iken, K. and Edwards, M. (2009). Depth-stratified community zonation patterns on Gulf of Alaska rocky shores. Marine Ecology, 30 (1), 6373, http://dx.doi.org/10.1111/j.1439-0485.2008.00259.x.Google Scholar
Konotchick, T., Dupont, C. L., Valas, R. E., Badger, J. H. and Allen, A. E. (2013). Transcriptomic analysis of metabolic function in the giant kelp, Macrocystis pyrifera, across depth and season. The New phytologist, 198 (2), 398407, http://dx.doi.org/10.1111/nph.12160.Google Scholar
Kordas, R. L., Harley, C. D. G. and O’Connor, M. I. (2011). Community ecology in a warming world: the influence of temperature on interspecific interactions in marine systems. Journal of Experimental Marine Biology and Ecology, 400 (1–2), 218–26, http://dx.doi.org/10.1016/j.jembe.2011.02.029.Google Scholar
Kosro, P. M. (2002). A poleward jet and an equatorward undercurrent observed off Oregon and northern California, during the 1997–98 El Niño. Progress in Oceanography, 54 (1–4), 343–60, http://dx.doi.org/10.1016/S0079-6611(02)00057-5.Google Scholar
Kusumo, H. T., Pfister, C. A. and Wootton, J. T. (2006). Small-scale genetic structure in the sea palm Postelsia palmaeformis Ruprecht (Phaeophyceae). Marine Biology, 149 (4), 731–42, http://dx.doi.org/10.1007/s00227-006-0254-z.Google Scholar
Ladah, L. B. and Zertuche-González, J. A. (1999). Giant kelp (Macrocystis pyrifera, Phaeophyceae) recruitment near its southern limit in Baja California after mass disappearance during ENSO. Journal of Phycology, 35, 1106–12.Google Scholar
Ladah, L. B. and Zertuche-González, J. A. (2004). Giant kelp (Macrocystis pyrifera) survival in deep water (25-40 m) during El Niño of 1997–1998 in Baja California, Mexico. Botanica Marina, 47 (5), 367–72, http://dx.doi.org/10.1515/BOT.2004.054.Google Scholar
Lang, M. A., Marinelli, R. L., Roberts, S. J. and Taylor, P. R., eds. (2013) Research and Discoveries: The Revolution of Science through Scuba, Smithsonian Contributions to the Marine Sciences. Smithsonian Institution Press, Washington, DC, www.vliz.be/imisdocs/publications/256301.pdf#page=141.Google Scholar
Lauerman, L. M. L., Kaufmann, R. S. and Smith, K. L. (1996). Distribution and abundance of epibenthic megafauna at a long time-series station in the abyssal northeast Pacific. Deep-Sea Research Part I: Oceanographic Research Papers, 43 (7), 1075–103, http://dx.doi.org/10.1016/0967-0637(96)00045-3.Google Scholar
Laur, D., Ebeling, A. and Coon, D. (1988). Effects of Sea Otter Foraging on Subtidal Reef Communities off Central California. In VanBlaricom, G. R. and Estes, J. A., eds. The Community Ecology of Sea Otters. Springer-Verlag, Berlin, pp. 151–68, http://link.springer.com/chapter/10.1007/978–3-642–72845-7_7.Google Scholar
LeClair, L., Pleus, A. and Schultz, J. (2009) 2007-2009 Biennial Report: Invasive Species Tunicate Response. Puget Sound Partnership, Olympia, WA.Google Scholar
Lesser, M. P., Witman, J. D. and Sebens, K. P. (1994). Effects of flow and seston availability on scope for growth of benthic suspension-feeding invertebrates from the Gulf of Maine. Biological Bulletin, 187 (3), 319–35, http://dx.doi.org/10.2307/1542289.Google Scholar
Lester, S. E., Tobin, E. D. and Behrens, M. D. (2007). Disease dynamics and the potential role of thermal stress in the sea urchin, Strongylocentrotus purpuratus. Canadian Journal of Fisheries and Aquatic Sciences, 64 (2), 314–23, http://dx.doi.org/10.1139/f07-010.Google Scholar
Levenbach, S. (2008). Community-wide ramifications of an associational refuge on shallow rocky reefs. Ecology, 89 (10), 2819–28, http://dx.doi.org/10.1890/07.0656.1.Google Scholar
Levings, C. D., Cordell, J. R., Ong, S. and Piercey, G. E. (2004). The origin and identity of invertebrate organisms being transported to Canada’s Pacific coast by ballast water. Canadian Journal of Fisheries and Aquatic Sciences, 61, 111, http://dx.doi.org/10.1139/f03-135.Google Scholar
Levings, C. D., Foreman, R. E. and Tunnicliffe, V. J. (1983). Review of the benthos of the Strait of Georgia and contiguous fjords. Canadian Journal of Fisheries and Aquatic Sciences, 40 (7), 1120–41, http://dx.doi.org/10.1139/f83-131.Google Scholar
Liggan, L. (2016) Under Pressure: Biomechanics of Buoyancy in Bull Kelp (Nereocystis leutkeana). University of British Columbia, Vancouver, http://dx.doi.org/10.14288/1.0300227.Google Scholar
Lindstrom, S. C. (2009). The biogeography of seaweeds in Southeast Alaska. Journal of Biogeography, 36 (3), 401–9, http://dx.doi.org/10.1111/j.1365-2699.2007.01855.x.Google Scholar
Lindstrom, S. C. and Foreman, R. E. (1978). Seaweed associations of the Flat Top Islands, British Columbia – comparison of community methods. Syesis, 11, 171–85.Google Scholar
Lindstrom, S. C., Olsen, J. L. and Stam, W. T. (1997). Postglacial recolonization and the biogeography of Palmaria mollis (Rhodophyta) along the northeast Pacific coast. Canadian Journal of Botany, 75 (11), 1887–96, http://dx.doi.org/10.1139/b97-900.Google Scholar
Litzow, M. A., Mueter, F. J. and Hobday, A. J. (2013). Reassessing regime shifts in the North Pacific: incremental climate change and commercial fishing are necessary for explaining decadal-scale biological variability. Global Change Biology, 20 (1), 3850, http://dx.doi.org/10.1111/gcb.12373.Google Scholar
Long, E. R., Dutch, M., Weakland, S., Chandramouli, B. and Benskin, J. P. (2013). Quantification of pharmaceuticals, personal care products, and perfluoroalkyl substances in the marine sediments of Puget Sound, Washington, USA. Environmental Toxicology and Chemistry, 32 (8), 1701–10, http://dx.doi.org/10.1002/etc.2281.Google Scholar
Di Lorenzo, E., Cobb, K. M., Furtado, J. C. et al. (2010). Central Pacific El Niño and decadal climate change in the North Pacific Ocean. Nature Geoscience, 3, 762–5, http://dx.doi.org/10.1038/ngeo984.Google Scholar
Di Lorenzo, E., Schnieder, N., Cobb, K. M. et al. (2008). North Pacific gyre oscillation links ocean climate and ecosystem change. Geophysical Research Letters, 35 (8), 27, http://dx.doi.org/10.1029/2007GL032838.Google Scholar
Love, M. S. and Larson, R. J. (1978). Geographic variation in the occurrence of tympanic spines and possible genetic differentiation in the kelp rockfish (Sebastes atrovirens). Copeia, 1978 (1), 53–9.Google Scholar
Love, M. S., Saiki, M. K., May, T. W. and Yee, J. L. (2013). Whole-body concentrations of elements in three fish species from offshore oil platforms and natural areas in the Southern California Bight, USA. Bulletin of Marine Science, 89 (3), 717–34, http://dx.doi.org/10.5343/bms.2012.1078.Google Scholar
Love, M. S., Schroeder, D. M., Lenarz, W., MacCall, A., Bull, A. S. and Thorsteinson, L. (2006). Potential use of offshore marine structures in rebuilding an overfished rockfish species, bocaccio (Sebastes paucispinis). Fishery Bulletin, 104 (3), 383–90.Google Scholar
Low, N. H. and Micheli, F. (2018). Lethal and functional thresholds of hypoxia in two key benthic grazers. Marine Ecology Progress Series, 594, 165–73.Google Scholar
Lowe, A. T., Galloway, A. W. E., Yeung, J. S., Dethier, M. N. and Duggins, D. O. (2014). Broad sampling and diverse biomarkers allow characterization of nearshore particulate organic matter. Oikos, 123 (11), 1341–54, http://dx.doi.org/10.1111/oik.01392.Google Scholar
Lowe, A. T., Whippo, R., Galloway, A. W. E., Britton‐Simmons, K. H. and Dethier, M. N. (2015). Sedentary urchins influence benthic community composition below the macroalgal zone. Marine Ecology, 36 (2), 129–40, http://dx.doi.org/10.1111/maec.12124.Google Scholar
Lubchenco, J., Navarette, S. A., Tissot, B. N. and Castilla, J. C. (1993). Possible Ecological Responses to Global Climate Change: Nearshore Benthic Biota of Northeastern Pacific Coastal Ecosystems. In Mooney, H. A., Fuentes, E. and Kronberg, B. I., eds. Earth System Responses To Global Change: Contrasts between North and South America. Academic Press, San Diego, CA, pp. 147–66.Google Scholar
Lucas, A. J., Dupont, C. L., Tai, V., Largier, J. L., Palenik, B. and Franks, P. J. S. (2011). The green ribbon: multiscale physical control of phytoplankton productivity and community structure over a narrow continental shelf. Limnology and Oceanography, 56 (2), 611–26, http://dx.doi.org/10.4319/lo.2011.56.2.0611.Google Scholar
Lüning, K. and Freshwater, W. (1988). Temperature tolerance of northeast Pacific marine algae. Journal of Phycology, 24 (3), 310–15, http://dx.doi.org/10.1111/j.1529-8817.1988.tb04471.x.Google Scholar
Lüning, K., Yarish, C. and Kirkman, H. (1990) Seaweeds: Their Environment, Biogeography, and Ecophysiology. Wiley, New York.Google Scholar
Lyle, M., Koizumi, I., Delaney, M. L. and Barron, J. A. (2000). Sedimentary record of the California current system, middle Miocene to Holocene: a synthesis of leg 167 results. Proceedings of the Ocean Drilling Program, Scientific results, 167, 341–76.Google Scholar
Lynn, R. J. and Simpson, J. J. (1987). The California current system: the seasonal variability of its physical characteristics. Journal of Geophysical Research: Oceans, 92 (C12), 12947–66, http://dx.doi.org/10.1029/JC092iC12p12947.Google Scholar
Mace, A. J. and Morgan, S. G. (2006). Biological and physical coupling in the lee of a small headland: contrasting transport mechanisms for crab larvae in an upwelling region. Marine Ecology Progress Series, 324, 185–96, http://dx.doi.org/10.3354/meps324185.Google Scholar
Mackas, D., Peterson, W. and Ohman, M. (2006). Zooplankton anomalies in the California current system before and during the warm ocean conditions of 2005. Geophysical Research Letters, 33 (22), L22S07, http://onlinelibrary.wiley.com/doi/10.1029/2006GL027930/full.Google Scholar
Mann, K. H., Kenneth, H. and Lazier, J. R. N. (2006). Dynamics of Marine Ecosystems Biological–Physical Interactions in the Oceans, third edn. Blackwell Publishing, Hoboken, NJ.Google Scholar
Mantua, N. J. and Hare, S. R. (2002). The Pacific decadal oscillation. Journal of Oceanography, 58, 3544.Google Scholar
Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M. and Francis, R. C. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society, 78 (6), 1069–79, http://dx.doi.org/10.1175/1520-0477(1997)078<1069:APICOW>2.0.CO;2.Google Scholar
Marchinko, K. B. and Palmer, A. R. (2003). Feeding in flow extremes: dependence of cirrus form on wave-exposure in four barnacle species. Zoology, 106 (2), 127–41, http://dx.doi.org/10.1078/0944-2006-00107.Google Scholar
Markel, R. W. and Shurin, J. B. (2015). Indirect effects of sea otters on rockfish (Sebastes spp.) in giant kelp forests. Ecology, 96 (11), 2877–90, http://dx.doi.org/10.1121/1.4929899.Google Scholar
Marko, P. B. (2004). ‘“What”s larvae got to do with it?’ Disparate patterns of post-glacial population structure in two benthic marine gastropods with identical dispersal potential. Molecular Ecology, 13 (3), 597611, http://dx.doi.org/10.1046/j.1365-294X.2004.02096.x.Google Scholar
Matson, P. G. and Edwards, M. S. (2007). Effects of ocean temperature on the southern range limits of two understory kelps, Pterygophora californica and Eisenia arborea, at multiple life-stages. Marine Biology, 151 (5), 1941–9, http://dx.doi.org/10.1007/s00227-007-0630-3.Google Scholar
Mazariegos-Villarreal, A., Casas-Valdez, M., Siqueiros-Beltrones, D. A., Piñon-Gimate, A. and Serviere-Zaragoza, E. (2012). During the 1997 to 1998 El Niño Event in Baja California Sur, Mexico. Journal of Shellfish Research, 31 (3), 795800, http://dx.doi.org/10.2983/035.031.0325.Google Scholar
McCann, L. D., Holzer, K. K., Davidson, I. C. et al. (2013). Promoting invasive species control and eradication in the sea: options for managing the tunicate invader Didemnum vexillum in Sitka, Alaska. Marine Pollution Bulletin, 77 (1–2), 165–71, http://dx.doi.org/10.1016/j.marpolbul.2013.10.011.Google Scholar
McClain, C. R., Lundsten, L., Ream, M., Barry, J. and DeVogelaere, A. (2009). Endemicity, biogeography, composition, and community structure on a northeast Pacific seamount. PLoS ONE, 4 (1), http://dx.doi.org/10.1371/journal.pone.0004141.Google Scholar
McCormick-Ray, J., Warwick, R. M. and Ray, G. C. (2011). Benthic macrofaunal compositional variations in the northern Bering Sea. Marine Biology, 158 (6), 1365–76, http://dx.doi.org/10.1007/s00227-011-1655-1.Google Scholar
McGowan, J. A., Cayan, D. R. and Dorman, L. M. (1998). Climate-ocean variability and ecosystem response in the northeast Pacific. Science, 281 (5374), 210–17, http://dx.doi.org/10.1126/science.281.5374.210.Google Scholar
Menge, B. A., Chan, F., Nielsen, K. J., Di Lorenzo, E. and Lubchenco, J. (2009). Climatic variation alters supply-side ecology: impact of climate patterns on phytoplankton and mussel recruitment. Ecological Monographs, 79 (3), 379–95, http://dx.doi.org/10.1890/08-2086.1.Google Scholar
Menge, B. A., Daley, B. A., Wheeler, P. A., Dahlhoff, E., Sanford, E. and Strub, P. T. (1997). Benthic-pelagic links and rocky intertidal communities: bottom-up effects on top-down control? Proceedings of National Academy of Sciences, 94 (26), 14530–5.Google Scholar
Miller-Rushing, A. J., Høye, T. T., Inouye, D. W. and Post, E. (2010). The effects of phenological mismatches on demography. Philosophical Transactions of the Royal Society of London B, 365, 3177–86, http://dx.doi.org/10.1098/rstb.2010.0148.Google Scholar
Miller, B. A. and Emlet, R. B. (1997). Influence of nearshore hydrodynamics on larval abundance and settlement of sea urchins Strongylocentrotus franciscanus and S. purpuratus in the Oregon upwelling zone. Marine Ecology Progress Series, 148 (1–3), 8394, http://dx.doi.org/10.3354/meps148083.Google Scholar
Miller, E. F. and McGowan, J. A. (2013). Faunal shift in southern California’s coastal fishes: a new assemblage and trophic structure takes hold. Estuarine, Coastal and Shelf Science, 127, 2936, http://dx.doi.org/10.1016/j.ecss.2013.04.014.Google Scholar
Miller, K., Engle, J. M., Uwai, S. et al. (2007). First report of the Asian seaweed Sargassum filicinum Harvey (Fucales) in California, USA. Biological Invasions, 9 (1), 926–9, http://dx.doi.org/10.1007/s10530-006-9060-2.Google Scholar
Miller, K. and Estes, J. (1989). Western range extension for Nereocystis luetkeana in the North Pacific Ocean. Botanica Marina, 32 (6), 535–8, www.degruyter.com/view/j/botm.1989.32.issue-6/botm.1989.32.6.535/botm.1989.32.6.535.xml.Google Scholar
Miller, K. A. and Engle, J. M. (2009). The natural history of Undaria pinnatifida and Sargassum filicinum at the California channel islands: non-native seaweeds with different invasion styles. In Proceedings of the 7th California Islands Symposium, 131–40.Google Scholar
Miller, K. B. (2016). Forecasting at the edge of the niche: Didemnum vexillum in southeast Alaska. Marine Biology, 163 (2), 112, http://dx.doi.org/10.1007/s00227-015-2799-1.Google Scholar
Miller, L. P. and Gaylord, B. (2007). Barriers to flow: the effects of experimental cage structures on water velocities in high-energy subtidal and intertidal environments. Journal of Experimental Marine Biology and Ecology, 344 (2), 215–28, http://dx.doi.org/10.1016/j.jembe.2007.01.005.Google Scholar
Miller, M. A., Byrne, B. A., Jang, S. S. et al. (2010). Enteric bacterial pathogen detection in southern sea otters (Enhydra lutris nereis) is associated with coastal urbanization and freshwater runoff. Veterinary Research, 41 (1), http://dx.doi.org/10.1051/vetres/2009049.Google Scholar
Miller, R. J., Harrer, S. and Reed, D. C. (2012). Addition of species abundance and performance predicts community primary production of macroalgae. Oecologia, 168 (3), 797806, http://dx.doi.org/10.1007/s00442-011-2143-5.Google Scholar
Miller, R. J. and Page, H. M. (2012). Kelp as a trophic resource for marine suspension feeders: A review of isotope-based evidence. Marine Biology, 159 (7), 1391–402, http://dx.doi.org/10.1007/s00227-012-1929-2.Google Scholar
Miller, R. J., Page, H. M. and Brzezinski, M. A. (2013). δ13C and δ15N of particulate organic matter in the Santa Barbara Channel: Drivers and implications for trophic inference. Marine Ecology Progress Series, 474, 5366, http://dx.doi.org/10.3354/meps10098.Google Scholar
Miller, R. J., Reed, D. C. and Brzezinski, M. A. (2011). Partitioning of primary production among giant kelp (Macrocystis pyrifera), understory macroalgae, and phytoplankton on a temperate reef. Limnology and Oceanography, 56 (1), 119–32, http://dx.doi.org/10.4319/lo.2011.56.1.0119.Google Scholar
Miller, S. H. and Morgan, S. G. (2013). Interspecific differences in depth preference: Regulation of larval transport in an upwelling system. Marine Ecology Progress Series, 476, 301–6, http://dx.doi.org/10.3354/meps10150.Google Scholar
Montecino-Latorre, D., Eisenlord, M. E., Turner, M. et al. (2016). Devastating transboundary impacts of sea star wasting disease on subtidal asteroids. PLoS ONE, 11 (10), e0163190.Google Scholar
Morales-Zárate, M. V., Arreguín-Sánchez, F., López-Martínez, J. and Lluch-Cota, S. E. (2004). Ecosystem trophic structure and energy flux in the northern Gulf of California, Mexico. Ecological Modelling, 174 (4), 331–45, http://dx.doi.org/10.1016/j.ecolmodel.2003.09.028.Google Scholar
Morgan, S. G., Fisher, J. L., Mace, A. J., Akins, L., Slaughter, A. M. and Bollens, S. M. (2009). Cross-shelf distributions and recruitment of crab postlarvae in a region of strong upwelling. Marine Ecology Progress Series, 380, 173–85, http://dx.doi.org/10.3354/meps07913.Google Scholar
Morzaria-Luna, H. N., Ainsworth, C. H., Kaplan, I. C., Levin, P. S. and Fulton, E. A. (2013). Indirect effects of conservation policies on the coupled human-natural ecosystem of the upper Gulf of California. PLoS ONE, 8 (5), http://dx.doi.org/10.1371/journal.pone.0064085.Google Scholar
Muñoz-Barbosa, A. and Huerta-Diaz, M. A. (2013). Trace metal enrichments in nearshore sediments and accumulation in mussels (Modiolus capax) along the eastern coast of Baja California, Mexico: environmental status in 1995. Marine Pollution Bulletin, 77 (1–2), 7181, http://dx.doi.org/10.1016/j.marpolbul.2013.10.030.Google Scholar
Murray, S. N., Littler, M. M. and Abbott, I. A. (1980). Biogeography of the California marine algae with emphasis on the Southern California Islands. In The California Islands: Proceedings of a Multidiciplinary Symposium, 35–6.Google Scholar
Muth, A. F., Graham, M. H., Lane, C. E. and Harley, C. D. (2019). Recruitment tolerance to increased temperature present across multiple kelp clades. Ecology, e02594.Google Scholar
Myers, M. S., Johnson, L. L. and Collier, T. K. (2003). Establishing the causal relationship between polycyclic aromatic hydrocarbon (PAH) exposure and hepatic neoplasms and neoplasia-related liver lesions in English Sole (Pleuronectes vetulus). Human and Ecological Risk Assessment, 9 (1), 6794.Google Scholar
Nakamura, H., Lin, G. and Yamagata, T. (1997). Decadal climate variability in the North Pacific during the recent decades. Bulletin of the American Meteorological Society, 78 (10), 2215–25, http://dx.doi.org/10.1175/1520-0477(1997)078<2215:DCVITN>2.0.CO;2.Google Scholar
O’Neel, S., Hood, E., Bidlack, A. L. et al. (2015). Icefield-to-ocean linkages across the northern pacific coastal temperate rainforest ecosystem. BioScience, 65 (5), 499512, http://dx.doi.org/10.1093/biosci/biv027.Google Scholar
Neushul, M. (1967). Studies of subtidal marine vegetation in western Washington. Ecology, 48 (1), 8394.Google Scholar
Neushul, M. and Haxo, F. (1963). Studies on the giant kelp, Macrocystis. I. Growth of young plants. American Journal of Botany, 50 (4), 349–53.Google Scholar
Newman, M., Alexander, M. A., Ault, T. R. et al. (2016). The Pacific decadal oscillation, revisited. Journal of Climate, 29, 4399–427.Google Scholar
Nidzieko, N. J. and Largier, J. L. (2013). Inner shelf intrusions of offshore water in an upwelling system affect coastal connectivity. Geophysical Research Letters, 40 (20), 5423–8, http://dx.doi.org/10.1002/2013GL056756.Google Scholar
Nishizaki, M. T. and Carrington, E. (2014a). Temperature and water flow influence feeding behavior and success in the barnacle Balanus glandula. Marine Ecology Progress Series, 507, 207–18, http://dx.doi.org/10.3354/meps10848.Google Scholar
Nishizaki, M. T. and Carrington, E. (2014b). The effect of water temperature and flow on respiration in barnacles: patterns of mass transfer versus kinetic limitation. Journal of Experimental Biology, 217 (12), 2101–9, http://dx.doi.org/10.1242/jeb.101030.Google Scholar
Nishizaki, M. T. and Carrington, E. (2015). The effect of water temperature and velocity on barnacle growth: quantifying the impact of multiple environmental stressors. Journal of Thermal Biology, 54, 3746, http://dx.doi.org/10.1016/j.jtherbio.2015.02.002.Google Scholar
North, W. J. and Hubbs, C. L. (1968) Utilization of Kelp-Bed Resources in Southern California. Department of Fish and Game, Sacramento, CA.Google Scholar
North, W. J., Jackson, G. A. and Manley, S. L. (1986). Macrocystis and its environment, knowns and unknowns. Aquatic Botany, 26, 926.Google Scholar
North, W. J. and Pearse, J. S. (1970). Sea urchin population explosion in southern California coastal waters. Science, 167 (3915), 209.Google Scholar
Nowell, A. R. M. and Jumars, P. A. (1984). Flow environments of aquatic benthos. Annual Review of Ecology and Systematics, 15 (1), 303–28, http://dx.doi.org/10.1146/annurev.es.15.110184.001511.Google Scholar
Okamoto, D. K., Reed, D. C. and Schroeter, S. C. (2018). Geographically opposing responses of sea urchin recruitment to changes in ocean climate. BioRxiv, 387282.Google Scholar
Okamoto, D. K., Stekoll, M. S. and Eckert, G. L. (2013). Coexistence despite recruitment inhibition of kelps by subtidal algal crusts. Marine Ecology Progress Series, 493, 103–12, http://dx.doi.org/10.3354/meps10505.Google Scholar
Okey, T. A., Alidina, H. M. and Agbayani, S. (2015). Mapping ecological vulnerability to recent climate change in Canada’s Pacific marine ecosystems. Ocean and Coastal Management, 106, 3548, http://dx.doi.org/10.1016/j.ocecoaman.2015.01.009.Google Scholar
Okey, T. A., Alidina, H. M., Lo, V. and Jessen, S. (2014). Effects of climate change on Canada’s Pacific marine ecosystems: a summary of scientific knowledge. Reviews in Fish Biology and Fisheries, 24 (2), 519–59, http://dx.doi.org/10.1007/s11160-014-9342-1.Google Scholar
Pace, M., Cole, J. J., Carpenter, S. R. and Kitchell, J. F. (1999). Trophic cascades revealed in diverse ecosystems. Trends in Ecology & Evolution, 14 (12), 483–8, www.ncbi.nlm.nih.gov/pubmed/10542455.Google Scholar
Page, H. M., Dugan, J. E., Culver, C. S. and Hoesterey, J. C. (2006). Exotic invertebrate species on offshore oil platforms. Marine Ecology Progress Series, 325, 101–7, http://dx.doi.org/10.3354/meps325101.Google Scholar
Page, H. M., Reed, D. C., Brzezinski, M. A., Melack, J. M. and Dugan, J. E. (2008). Assessing the importance of land and marine sources of organic matter to kelp forest food webs. Marine Ecology Progress Series, 360, 4762, http://dx.doi.org/10.3354/meps07382.Google Scholar
Paine, R. T. and Levin, S. A. (1981). Intertidal landscapes: disturbance and the dynamics of pattern. Ecological Monographs, 51 (2), 145–78.Google Scholar
Palardy, J. E. and Witman, J. D. (2011). Water flow drives biodiversity by mediating rarity in marine benthic communities. Ecology Letters, 14 (1), 63–8, http://dx.doi.org/10.1111/j.1461-0248.2010.01555.x.Google Scholar
Palumbi, S. R. (2003). Population genetics, demographic connectivity, and the design of marine reserves. Ecological Applications, 13 (1), 146–58.Google Scholar
Palumbi, S. R. and Wilson, A. C. (1990). Mitochondrial DNA diversity in the sea urchins Strongylocentrotus purpuratus and S. droebachiensis. Evolution, 44 (2), 403–15.Google Scholar
Parnell, P. E., Miller, E. F., Lennert Cody, C. E., Dayton, P. K., Carter, M. L. and Stebbins, T. D. (2010). The response of giant kelp (Macrocystis pyrifera) in southern California to low-frequency climate forcing. Limnology and Oceanography, 55 (6), 2686–702, http://dx.doi.org/10.4319/lo.2010.55.6.2686.Google Scholar
Patterson, M. R. (1980). Hydromechanical Adaptations in Alcyonium Sidereum (Octocorallia). In Schneck, P. J., ed. Biofluid Mechanics 2. Springer, Boston, 183201, http://dx.doi.org/10.1007/978-1-4757-4610-5_10.Google Scholar
Pearse, J. S. and Hines, A. H. (1979). Expansion of a central California kelp forest following the mass mortality of sea urchins. Marine Biology, 51 (1), 8391, http://dx.doi.org/10.1007/BF00389034.Google Scholar
Pelc, R. A., Warner, R. R. and Gaines, S. D. (2009). Geographical patterns of genetic structure in marine species with contrasting life histories. Journal of Biogeography, 36 (10), 1881–90, http://dx.doi.org/10.1111/j.1365-2699.2009.02138.x.Google Scholar
Pespeni, M. H., Sanford, E., Gaylord, B. et al. (2013). Evolutionary change during experimental ocean acidification. Proceedings of the National Academy of Sciences, 110 (17), 6937–42, http://dx.doi.org/10.1073/pnas.1220673110.Google Scholar
Peterson, W., Robert, M. and Bond, N. (2015). The warm blob – conditions in the northeastern Pacific Ocean. PICES Press, 23 (1), 36–8, http://search.proquest.com/docview/1665110669?pq-origsite=gscholar.Google Scholar
Pfister, C. A., Berry, H. D. and Mumford, T. (2018). The dynamics of kelp forests in the northeast Pacific Ocean and the relationship with environmental drivers. Journal of Ecology, 106 (4), 1520–33.Google Scholar
Piacenza, S. E., Barner, A. K., Benkwitt, C. E. et al. (2015). Patterns and variation in benthic biodiversity in a large marine ecosystem. PLoS ONE, 10 (8), 123, http://dx.doi.org/10.1371/journal.pone.0135135.Google Scholar
Pineda, J., Reyns, N. and Lentz, S. J. (2018). Reduced barnacle larval abundance and settlement in response to large‐scale oceanic disturbances: temporal patterns, nearshore thermal stratification, and potential mechanisms. Limnology and Oceanography, 63 (6), 2618–29.Google Scholar
Pirtle, J. L., Ibarra, S. N. and Eckert, G. L. (2012). Nearshore subtidal community structure compared between inner coast and outer coast sites in southeast Alaska. Polar Biology, 35 (12), 1889–910, http://dx.doi.org/10.1007/s00300-012-1231-2.Google Scholar
Pister, B. (2009). Urban marine ecology in southern California: the ability of riprap structures to serve as rocky intertidal habitat. Marine Biology, http://link.springer.com/article/10.1007/s00227–009-1130-4.Google Scholar
Pondella, D. J., Gintert, B. E., Cobb, J. R. and Allen, L. G. (2005). Biogeography of the nearshore rocky-reef fishes at the southern and Baja California Islands. Journal of Biogeography, 32 (2), 187201.Google Scholar
Rassweiler, A., Schmitt, R. J. and Holbrook, S. J. (2010). Triggers and maintenance of multiple shifts in the state of a natural community. Oecologia, 164 (2), 489–98, http://dx.doi.org/10.1007/s00442-010-1666-5.Google Scholar
Reed, D. C., Anderson, T. W., Ebeling, A. W. and Anghera, M. (1997). The role of reproductive synchrony in the colonization potential of kelp. Ecology, 78 (8), 2443–57.Google Scholar
Reed, D. C. and Foster, M. S. (1984). The effects of canopy shadings on algal recruitment and growth in a giant kelp forest. Ecology, 65 (3), 937–48.Google Scholar
Reed, D. C., Laur, D. R. and Ebeling, A. W. (1988). Variation in algal dispersal and recruitment: the importance of episodic events. Ecological Monographs, 58 (4), 321–35.Google Scholar
Reed, D. C., Rassweiler, A., Carr, M. H., Cavanaugh, K. C., Malone, D. P. and Siegel, D. A. (2011). Wave disturbance overwhelms top-down and bottom-up control of primary production in California kelp forests. Ecology, 92 (11), 2108–16.Google Scholar
Reed, D. C., Rassweiler, A. R. and Arkema, K. D. (2008). Biomass rather than growth rate determines variation in net primary production by giant kelp. Ecology, 89 (9), 2493–505, http://dx.doi.org/10.1890/07-1106.1.Google Scholar
Reed, D. C., Rassweiler, A. R., Miller, R. J., Page, H. M. and Holbrook, S. J. (2016a). The value of a broad temporal and spatial perspective in understanding dynamics of kelp forest ecosystems. Marine and Freshwater Research, 67 (1), 1424, http://dx.doi.org/10.1071/MF14158.Google Scholar
Reed, D., Washburn, L., Rassweiler, A., Miller, R., Bell, T. and Harrer, S. (2016b). Extreme warming challenges sentinel status of kelp forests as indicators of climate change. Nature Communications, 7, 13757.Google Scholar
Reeves, R. L., Grant, S. B., Mrse, R. D., Copil Oancea, C. M., Sanders, B. F. and Boehm, A. B. (2004). Scaling and management of fecal indicator bacteria in runoff from a coastal urban watershed in southern California. Environmental Science and Technology, 38 (9), 2637–48, http://dx.doi.org/10.1021/es034797g.Google Scholar
Renner, M., Arimitsu, M. L. and Piatt, J. F. (2012). Structure of marine predator and prey communities along environmental gradients in a glaciated fjord. Canadian Journal of Fisheries and Aquatic Sciences, 69, 2029–45, http://dx.doi.org/10.1139/f2012-117.Google Scholar
Riosmena-Rodríguez, R., Hinojosa-Arango, G., López-Vivas, J. M. and León-Cisneros, K. (2005). Caracterización espacial y biogeográfica de las asociaciones de macroalgas de Bahía del Rincón, Baja California Sur, México. Revista de Biologia Tropical, 53 (1–2), 97109.Google Scholar
Robinson, H. E., Finelli, C. M. and Koehl, M. A. R. (2013). Interactions between benthic predators and zooplanktonic prey are affected by turbulent waves. Integrative and Comparative Biology, 53 (5), 810–20, http://dx.doi.org/10.1093/icb/ict092.Google Scholar
Robles, C., Sweetnam, D. A. and Dittman, D. (1989). Diel variation of intertidal foraging by Cancer productus L. in British Columbia. Journal of Natural History, 23 (5), 1041–9, http://dx.doi.org/10.1080/00222938900770951.Google Scholar
Robles, C. D., Alvarado, M. A. and Desharnais, R. A. (2001). The shifting balance of littoral predator–prey interaction in regimes of hydrodynamic stress. Oecologia, 128 (1), 142–52, http://dx.doi.org/10.1007/s004420100638.Google Scholar
Rocha-Olivares, A. and Vetter, R. D. (1999). Effects of oceanographic circulation on the gene flow, genetic structure, and phylogeography of the rosethorn rockfish (Sebastes helvomaculatus). Canadian Journal of Fisheries and Aquatic Sciences, 56 (5), 803–13, http://dx.doi.org/10.1139/cjfas-56-5-803.Google Scholar
Rodriguez, G. E., Rassweiler, A., Reed, D. C. and Holbrook, S. J. (2013). The importance of progressive senescence in the biomass dynam of giant kelp (Macrocystis pyrifera). Ecology, 94 (8), 1848–58, http://dx.doi.org/10.1890/12-1340.1.Google Scholar
Rogers-Bennett, L. (2007). Is climate change contributing to range reductions and localized extinctions in northern (Haliotis kamtschatkana) and flat (Haliotis walallensis) abalones? Bulletin of Marine Science, 81 (2), 283–96.Google Scholar
Rogers-Bennett, L., Dondanville, R. F., Moore, J. D. and Ignacio Vilchis, L. (2010). Response of red abalone reproduction to warm water, starvation, and disease stressors: implications of ocean warming. Journal of Shellfish Research, 29 (3), 599611, http://dx.doi.org/10.2983/035.029.0308.Google Scholar
Rosenfeld, L. K., Schwing, F. B., Garfield, N. and Tracy, D. E. (1994). Bifurcated flow from an upwelling center: a cold water source for Monterey Bay. Continental Shelf Research, 14 (9), 931–64, http://dx.doi.org/10.1016/0278-4343(94)90058-2.Google Scholar
Rosenthal, R. J., Clarke, W. D. and Dayton, P. K. (1974). Ecology and natural history of a stand of giant kelp, Macrocystis pyrifera, off Del Mar, California. Fishery Bulletin, 72 (3), 670–84.Google Scholar
Roughgarden, J., Gaines, S. and Possingham, H. (1988). Recruitment dynamics in complex life cycles. Science, 1460–6, http://dx.doi.org/10.1126/science.11538249.Google Scholar
Royer, T. C. (1982). Coastal fresh water discharge in the northeast Pacific. Journal of Geophysical Research, 87 (1), 2017, http://dx.doi.org/10.1029/JC087iC03p02017.Google Scholar
Royer, T. C. (1983). Observations of the Alaska Coastal Current. In Coastal Oceanography. Springer, Boston, pp. 930, http://dx.doi.org/10.1007/978-1-4615-6648-9_2.Google Scholar
Ruesink, J. L., Hong, J.-S., Wisehart, L. et al. (2010). Congener comparison of native (Zostera marina) and introduced (Z. japonica) eelgrass at multiple scales within a Pacific Northwest estuary. Biological Invasions, 12 (6), 1773–89, http://dx.doi.org/10.1007/s10530-009-9588-z.Google Scholar
Sackmann, B., Mack, L., Logsdon, M. and Perry, M. J. (2004). Seasonal and inter-annual variability of SeaWiFS-derived chlorophyll a concentrations in waters off the Washington and Vancouver Island coasts, 1998–2002. Deep-Sea Research Part II: Topical Studies in Oceanography, 51 (10–11), 945–65, http://dx.doi.org/10.1016/j.dsr2.2003.09.004.Google Scholar
Sagarin, R. D., Barry, J. P.,Gilman, S. E. and Baxter, C. H. (1999). Climate-related change in an intertidal community over short and long time scales. Ecological Monographs, 69 (4), 465–90.Google Scholar
Sanford, E. and Menge, B. A. (2001). Spatial and temporal variation in barnacle growth in a coastal upwelling system. Marine Ecology Progress Series, 209, 143–57, http://dx.doi.org/10.3354/meps209143.Google Scholar
Saunders, G. W. (2014). Long distance kelp rafting impacts seaweed biogeography in the northeast Pacific: the kelp conveyor hypothesis. Journal of Phycology, 50 (6), 968–74, http://dx.doi.org/10.1111/jpy.12237.Google Scholar
Scagel, R. F. (1963). Distribution of Attached Marine Algae in Relation to Oceanographic Conditions in the Northeast Pacific. In Dunbar, M. J., ed. Marine Distributions. University of Toronto Press, Toronto, pp. 3750.Google Scholar
Schiel, D. R. and Foster, M. S. (2015) The Biology and Ecology of Giant Kelp Forests. University of California Press, Berkeley.Google Scholar
Schiel, D. R., Steinbeck, J. R. and Foster, M. S. (2004). Ten years of induced ocean warming causes comprehensive changes in marine benthic communities. Ecology, 85 (7), 1833–9.Google Scholar
Schmitz, K. and Srivastava, L. M. (1976). The fine structure of sieve elements of Nereocystis lutkeana. American Journal of Botany, 63 (5), 679–93.Google Scholar
Schroeder, D. M. and Love, M. S. (2004). Ecological and political issues surrounding decommissioning of offshore oil facilities in the Southern California Bight. Ocean and Coastal Management, 47 (1–2), 2148, http://dx.doi.org/10.1016/j.ocecoaman.2004.03.002.Google Scholar
Schroeter, S. C., Reed, D. C. and Raimondi, P. T. (2015). Effects of reef physical structure on development of benthic reef community: a large-scale artificial reef experiment. Marine Ecology Progress Series, 540, 4355, http://dx.doi.org/10.3354/meps11483.Google Scholar
Schultz, J. A., Cloutier, R. N. and Côté, I. M. (2016). Evidence for a trophic cascade on rocky reefs following sea star mass mortality in British Columbia. PeerJ, 4, e1980, http://dx.doi.org/10.7717/peerj.1980.Google Scholar
Sebens, K. P. (1984). Water flow and coral colony size: Interhabitat comparisons of the octocoral Alcyonium siderium. Proceedings of the National Academy of Sciences of the United States of America, 81 (17), 5473–7, http://dx.doi.org/10.1073/pnas.81.17.5473.Google Scholar
Sebens, K. P. (1985a). Community Ecology of Vertical Rock Walls in the Gulf of Maine, USA: Small-Scale Processes and Alternative Community States. In Moore, P. G. and Seed, R., eds. The Ecology of Rocky Coasts. Hodder and Stoughton Educational Press, London, pp. 346–71.Google Scholar
Sebens, K. P. (1985b). The ecology of the rocky subtidal zone. American Scientist, 73 (6), 548–57.Google Scholar
Sebens, K. P. (1986). Spatial relationships among encrusting marine organisms in the New England subtidal zone. Ecological Monographs, 56 (1), 7396.Google Scholar
Sebens, K. P. (1991). Habitat Structure and Community Dynamics in Marine Benthic Systems. In Bell, S. S., McCoy, E. D. and Mushinsky, H. R., eds. Habitat Structure: The Physical Arrangement of Objects in Space, first edn. Chapman and Hall, London, pp. 211–34.Google Scholar
Secretaria de Desarrollo Social: Catálogo de Localidades (2014). Mexico City, www.microrregiones.gob.mx/catloc/LocdeMun.aspx?tipo=clave&campo=loc&ent=02&mun=003.Google Scholar
See, K. E. and Feist, B. E. (2010). Reconstructing the range expansion and subsequent invasion of introduced European green crab along the west coast of the United States. Biological Invasions, 12 (5), 1305–18, http://dx.doi.org/10.1007/s10530-009-9548-7.Google Scholar
Selkoe, K. A., Gaines, S. D., Caselle, J. E. and Warner, R. R. (2006). Current shifts and kin aggregation explain genetic patchiness in fish recruits. Ecology, 87 (12), 3082–94.Google Scholar
Seymour, R. J., Tegner, M. J., Dayton, P. K. and Parnell, P. E. (1989). Storm wave induced mortality of giant kelp, Macrocystis pyrifera, in Southerm California. Estuarine, Coastal and Shelf Science, 28, 277–92, http://dx.doi.org/10.1016/0272-7714(89)90018-8.Google Scholar
Shaffer, J. A. (2000). Seasonal variation in understory kelp bed habitats of the Strait of Juan de Fuca. Journal of Coastal Research, 16 (3), 768–75.Google Scholar
Shelton, A. O., Harvey, C. J., Samhouri, J. F. et al. (2018). From the predictable to the unexpected: kelp forest and benthic invertebrate community dynamics following decades of sea otter expansion. Oecologia, 188 (4), 1105–19.Google Scholar
Shivji, M., Parker, D., Hartwick, B., Smith, M. J. and Sloan, N. A. (1983). Feeding and distribution study of the sunflower sea star Pycnopodia helianthoides (Brandt, 1835). Pacific Science, 37 (2), 133–40, http://scholarspace.manoa.hawaii.edu/handle/10125/653.Google Scholar
Siddon, C. E. and Witman, J. D. (2003). Influence of chronic, low-level hydrodynamic forces on subtidal community structure. Marine Ecology Progress Series, 261, 99110, http://dx.doi.org/10.3354/meps261099.Google Scholar
Siddon, E. C., Siddon, C. E. and Stekoll, M. S. (2008). Community level effects of Nereocystis luetkeana in southeastern Alaska. Journal of Experimental Marine Biology and Ecology, 361 (1), 815, http://dx.doi.org/10.1016/j.jembe.2008.03.015.Google Scholar
Simenstad, C. A., Estes, J. A. and Kenyon, K. W. (1978). Aleuts, sea otters, and alternate stable-state communities. Science, 200 (4340), 403–11, http://dx.doi.org/10.1126/science.200.4340.403.Google Scholar
Simkanin, C., Davidson, I. C., Dower, J. F., Jamieson, G. and Therriault, T. W. (2012). Anthropogenic structures and the infiltration of natural benthos by invasive ascidians. Marine Ecology, 33 (4), 499511, http://dx.doi.org/10.1111/j.1439-0485.2012.00516.x.Google Scholar
Simkanin, C., Dower, J. F., Filip, N., Jamieson, G. and Therriault, T. W. (2013). Biotic resistance to the infiltration of natural benthic habitats: examining the role of predation in the distribution of the invasive ascidian Botrylloides violaceus. Journal of Experimental Marine Biology and Ecology, 439, 7683, http://dx.doi.org/10.1016/j.jembe.2012.10.004.Google Scholar
Sirenko, B. I. and Gagaev, S. Y. (2007). Unusual abundance of macrobenthos and biological invasions in the Chukchi Sea. Russian Journal of Marine Biology, 33 (6), 355–64, http://dx.doi.org/10.1134/S1063074007060016.Google Scholar
Sivasundar, A. and Palumbi, S. R. (2010). Life history, ecology and the biogeography of strong genetic breaks among 15 species of Pacific rockfish, Sebastes. Marine Biology, 157 (7), 1433–52, http://dx.doi.org/10.1007/s00227-010-1419-3.Google Scholar
Small, C. and Nicholls, R. (2003). A global analysis of human settlement in coastal zones. Journal of Coastal Research, 19 (3), 584–99, www.jstor.org/stable/4299200?seq=1#page_scan_tab_contents.Google Scholar
Smith, K. L., Baldwin, R. J., Ruhl, H. A., Kahru, M., Mitchell, B. G. and Kaufmann, R. S. (2006). Climate effect on food supply to depths greater than 4,000 meters in the northeast Pacific. Limnology and Oceanography, 51 (1), 166–76, http://dx.doi.org/10.4319/lo.2006.51.1.0166.Google Scholar
Smith, R. L., Huyer, A. and Fleischbein, J. (2001). The coastal ocean off Oregon from 1961 to 2000: Is there evidence of climate change or only of Los Niños?. Progress in Oceanography, 49 (1–4), 6393, http://dx.doi.org/10.1016/S0079-6611(01)00016-7.Google Scholar
Sousa, W. P. (1984). The role of disturbance in natural communities. Annual Review of Ecology and Systematics, 15, 353–91.Google Scholar
Spalding, H., Foster, M. S. and Heine, J. N. (2003). Composition, distribution, and abundance of deep-water (>30m) macroalgae in Central California. Journal of Phycology, 284, 273–84.Google Scholar
Springer, Y. P., Hays, C. G., Carr, M. H. and Mackey, M. R. (2010). Toward ecosystem-based management of marine macroalgae – the Bull Kelp, Nereocystis luetkeana. Oceanography and Marine Biology: An Annual Review, 48, 142.Google Scholar
Spurkland, T. and Iken, K. (2011). Kelp bed dynamics in estuarine environments in subarctic Alaska. Journal of Coastal Research, 275, 133–43, http://dx.doi.org/10.2112/JCOASTRES-D-10-00194.1.Google Scholar
Stabeno, P. J., Bond, N. A., Hermann, A. J., Kachel, N. B., Mordy, C. W. and Overland, J. E. (2004) Meteorology and oceanography of the Northern Gulf of Alaska. Continental Shelf Research, 24 (7–8), 859–97, http://dx.doi.org/10.1016/j.csr.2004.02.007.Google Scholar
Statistics Canada. (2011). 2011 Census. Statistics Canada Catalogue no. 98-316-XWE. (2012). Ottawa.Google Scholar
Steneck, R., Graham, M. H., Bourque, B. J. et al. (2002). Kelp forest ecosystems: biodiversity, stability, resilience and future. Environmental Conservation, 29 (4), 436–59, http://digitalcommons.library.umaine.edu/sms_facpub.Google Scholar
Stenseth, N. C., Mysterud, A., Ottersen, G., Hurrell, J. W., Chan, K. S. and Lima, M. (2002). Ecological effects of climate fluctuations. Science, 297, 1292–6.Google Scholar
Stephens, J. S., Larson, R. J. and Pondella, D. J. (2006). Rocky Reefs and Kelp Beds. In Allen, L. G., Pondella, D. J. and Horn, M. H., eds. Ecology of Marine Fishes: California and Adjacent Waters. University of California Press, Berkeley, pp. 227–52.Google Scholar
Stewart, H. L., Fram, J. P., Reed, D. C. et al. (2009). Differences in growth, morphology and tissue carbon and nitrogen of Macrocystis pyrifera within and at the outer edge of a giant kelp forest in California, USA. Marine Ecology Progress Series, 375, 101–12, http://dx.doi.org/10.3354/meps07752.Google Scholar
Stoker, S. W. (1978) Benthic invertebrate macrofauna of the Eastern Continental Shelf of the Bering and Chukchi Seas. PhD, University of Alaska Fairbanks.Google Scholar
Strom, S. L., Olson, M. B., Macri, E. L. and Mordy, C. W. (2006). Cross-shelf gradients in phytoplankton community structure, nutrient utilization, and growth rate in the coastal Gulf of Alaska. Marine Ecology Progress Series, 328, 7592, http://dx.doi.org/10.3354/meps328075.Google Scholar
Strub, P. T., Combes, V., Shillington, F. A. and Pizarro, O. (2013) Currents and processes along the eastern boundaries. International Geophysics, 103, 339–84, http://dx.doi.org/10.1016/B978-0-12-391851-2.00014-3.Google Scholar
Sunday, J. M., Popovic, I., Palen, W. J., Foreman, M. G. G. and Hart, M. W. (2014). Ocean circulation model predicts high genetic structure observed in a long-lived pelagic developer. Molecular Ecology, 23 (20), 5036–47, http://dx.doi.org/10.1111/mec.12924.Google Scholar
Swanson, A. and Druehl, L. (2002). Induction, exudation and the UV protective role of kelp phlorotannins. Aquatic Botany, 73 (3), 241–53, www.sciencedirect.com/science/article/pii/S0304377002000359.Google Scholar
Swanson, A. K. and Druehl, L. D. (2000). Differential meiospore size and tolerance of ultraviolet light stress within and among kelp species along a depth gradient. Marine Biology, 136 (4), 657–64, http://dx.doi.org/10.1007/s002270050725.Google Scholar
Sydeman, W., Bradley, R. and Warzybok, P. (2006). Planktivorous auklet Ptychoramphus aleuticus responses to ocean climate, 2005: unusual atmospheric blocking? Geophysical Research Letters, 33 (22), L22S09, http://onlinelibrary.wiley.com/doi/10.1029/2006GL026736/full.Google Scholar
Sydeman, W. J., Thompson, S. A., Field, J. C. et al. (2011). Does positioning of the North Pacific current affect downstream ecosystem productivity?. Geophysical Research Letters, 38 (12), 16, http://dx.doi.org/10.1029/2011GL047212.Google Scholar
Szpak, P., Orchard, T. J., Salomon, A. K. and Gröcke, D. R. (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), 159–82, http://dx.doi.org/10.1007/s12520-013-0122-y.Google Scholar
Tegner, M., Dayton, P. K., Edwards, P. B. and Riser, K. L. (1997). Large scale, low frequency oceanographic effects on kelp forest succession: a tale of two cohorts. Marine Ecology Progress Series, 146, 117–34, http://dx.doi.org/10.3354/meps146117.Google Scholar
Tegner, M. J. and Dayton, P. K. (1981). Population structure, recruitment and mortality of two sea urchins (Strongylocentrotus franciscanus and S. purpuratus) in a kelp forest. Marine Ecology Progress Series, 5, 255–68, http://dx.doi.org/10.3354/meps005255.Google Scholar
Tegner, M. J. and Dayton, P. K. (1987). El Nino effects on southern California kelp forest communities. Advances in Ecological Research, 17, 243–79, http://dx.doi.org/10.1016/S0065-2504(08)60247-0.Google Scholar
Tegner, M. J. and Dayton, P. K. (1991). Sea urchins, El Ninos, and the long term stability of Southern California kelp forest communities. Marine Ecology Progress Series, 77, 4963, http://dx.doi.org/10.3354/meps077049.Google Scholar
Tegner, M. J. and Dayton, P. K. (2000). Ecosystem effects of fishing in kelp forest communities. ICES Journal of Marine Science, 57 (3), 579–89, http://dx.doi.org/10.1006/jmsc.2000.0715.Google Scholar
Tegner, M. J. and Levin, L. A. (1983). Spiny lobsters and sea urchins: analysis of a predator–prey interaction. Journal of Experimental Marine Biology and Ecology, 73 (2), 125–50, http://dx.doi.org/10.1016/0022-0981(83)90079-5.Google Scholar
Thiel, M. and Gutow, L. (2005). The ecology of rafting in the marine environment. II. The rafting organisms and community. An Annual Review, 43, 279418.Google Scholar
Thomas, A. and Brickley, P. (2006). Satellite measurements of chlorophyll distribution during spring 2005 in the California Current. Geophysical Research Letters, 33 (22), L22S05, http://onlinelibrary.wiley.com/doi/10.1029/2006GL026588/full.CrossRefGoogle Scholar
Thomson, D. A., Findley, L. T. and Kerstitch, A. N. (2000) Reef Fishes of the Sea of Cortez: The Rocky-Shore Fishes of the Gulf of California. University of Texas Press, Austin.Google Scholar
Thornber, C. (2007). Associational resistance mediates predator-prey interactions in a marine subtidal system. Marine Ecology, 28 (4), 480–6, http://dx.doi.org/10.1111/j.1439-0485.2007.00187.x.Google Scholar
Thornber, C. S., Graham, M. H., Kinlan, B. P. and Stachowicz, J. J. (2004). Population ecology of the invasive kelp Undaria pinnatifida in California: environmental and biological controls on demography. Marine Ecology Progress Series, 268, 6980.Google Scholar
Tilman, D. (1994). Competition and biodiversity in spatially structured habitats. Ecology, 75 (1), 216.CrossRefGoogle Scholar
tom Dieck, I. (1993). Temperature tolerance and survival in darkness of kelp gametophytes (Laminariales, Phaeophyta) – ecological and biogeographical implications. Marine Ecology Progress Series, 100 (3), 253–64, http://dx.doi.org/10.3354/meps100253.Google Scholar
Torres-Moye, G. and Escofet, A. (2014). Land–sea interactions in Punta China (Baja California, México): addressing anthropic and natural disturbances in a retrospective context. Journal of Environmental Protection, 5, 1520–30.Google Scholar
Trebilco, R., Dulvy, N. K., Stewart, H. and Salomon, A. K. (2015). The role of habitat complexity in shaping the size structure of a temperate reef fish community. Marine Ecology Progress Series, 532, 197211, http://dx.doi.org/doi:10.3354/meps11330.CrossRefGoogle Scholar
Tsurumi, M. and Tunnicliffe, V. (2001). Characteristics of a hydrothermal vent assemblage on a volcanically active segment of Juan de Fuca Ridge, northeast Pacific. Canadian Journal of Fisheries and Aquatic Sciences, 58 (3), 530–42, http://dx.doi.org/10.1139/f01-005.Google Scholar
United States Census Bureau. (2016). Annual Estimates of the Resident Population: April 1, 2010 to July 1, 2015. United States Census Bureau, Washington, DC.Google Scholar
Vadas, R. L. (1972). Ecological implications of culture studies on Nereocystis luetkeana. Journal of Phycology, 8 (2), 196203, http://dx.doi.org/10.1111/j.1529-8817.1972.tb04025.x.CrossRefGoogle Scholar
Valentine, J. W. (1966). Numerical analysis of marine molluscan ranges on the extratropical Northeastern Pacific shelf. Limnology and Oceanography, 11 (2), 198211.Google Scholar
Valentine, J. W., Roy, K. and Jablonski, D. (2002). Carnivore/non-carnivore ratios in northeastern Pacific marine gastropods. Marine Ecology Progress Series, 228, 153–63, http://dx.doi.org/10.3354/meps228153.CrossRefGoogle Scholar
Vance, R. (1988). Ecological succession and the climax community on a marine subtidal rock wall. Marine Ecology Progress Series, 48, 125–36, http://dx.doi.org/10.3354/meps048125.CrossRefGoogle Scholar
Vermeij, G. J. (1991). When biotas meet: understanding biotic interchange. Science, 253 (5024), 1099–104, http://dx.doi.org/10.1126/science.253.5024.1099.Google Scholar
Vermeij, G. J. (2012). The evolution of gigantism on temperate seashores. Biological Journal of the Linnean Society, 106 (4), 776–93, http://dx.doi.org/10.1111/j.1095-8312.2012.01897.x.Google Scholar
Vetter, E. W. (1998). Population dynamics of a dense assemblage of marine detritivores. Journal of Experimental Marine Biology and Ecology, 226 (1), 131–61, http://dx.doi.org/10.1016/S0022-0981(97)00246-3.Google Scholar
Walters, S. P., Thebo, A. L. and Boehm, A. B. (2011). Impact of urbanization and agriculture on the occurrence of bacterial pathogens and stx genes in coastal waterbodies of central California. Water Research, 45 (4), 1752–62, http://dx.doi.org/10.1016/j.watres.2010.11.032.Google Scholar
Wares, J. P. and Schiebelhut, L. M. (2016). What doesn’t kill them makes them stronger: an association between elongation factor 1-α overdominance in the sea star Pisaster ochraceus and “sea star wasting disease”. PeerJ, 4, e1876, http://dx.doi.org/10.7717/peerj.1876.Google Scholar
Watanabe, J. M. and Harrold, C. (1991). Destructive grazing by sea urchins Strongylocentrotus sin a central California kelp forest: potential roles of recruitment, depth, and predation. Marine Ecology Progress Series, 71 (2), 125–41, http://dx.doi.org/10.3354/meps071125.Google Scholar
Watson, J. and Estes, J. A. (2011). Stability, resilience, and phase shifts in rocky subtidal communities along the west coast of Vancouver Island, Canada. Ecological Monographs, 81 (2), 215–39, http://dx.doi.org/10.1890/10-0262.1.Google Scholar
Watson, J. R., Mitarai, S., Siegel, D. A., Caselle, J. E., Dong, C. and McWilliams, J. C. (2010). Realized and potential larval connectivity in the Southern California Bight. Marine Ecology Progress Series, 401, 3148, http://dx.doi.org/10.3354/meps08376.Google Scholar
Weise, M., Costa, D. and Kudela, R. (2006). Movement and diving behavior of male California sea lion (Zalophus californianus) during anomalous oceanographic conditions of 2005 compared to those of. Geophysical Research Letters, 33 (22), L22S10, http://onlinelibrary.wiley.com/doi/10.1029/2006GL027113/full.Google Scholar
West, J. E., Helser, T. E. and Neill, S. M. O. (2014). Variation in quillback rockfish (Sebastes maliger) growth patterns from oceanic to inland waters of the Salish Sea. Bulletin of Marine Science, 90 (3), 747–61, http://dx.doi.org/10.5343/bms.2013.1044.Google Scholar
Wheeler, W. N. and Neushul, M. (1981). The Aquatic Environment. In Physiological Plant Ecology I. Springer, Berlin, pp. 229–47, http://dx.doi.org/10.1007/978-3-642-68090-8_9.Google Scholar
Williams, J. P., Williams, C. M., Blanchette, C. A., Claisse, J. T., Pondella, D. J. and Caselle, J. E. (2018). Where the weird things are: a collection of species range extensions in the Southern California Bight. Bulletin, Southern California Academy of Sciences, 117 (3), 189203.Google Scholar
Wilson, A. B. (2006). Genetic signature of recent glaciation on populations of a near-shore marine fish species (Syngnathus leptorhynchus). Molecular Ecology, 15 (7), 1857–71, http://dx.doi.org/10.1111/j.1365-294X.2006.02911.x.Google Scholar
Wing, S. R., Botsford, L. W., Largier, J. L. and Morgan, L. E. (1995). Spatial variability in settlement of benthic invertebrates in a northern California upwelling system. Marine Ecology Progress Series, 128, 199211.Google Scholar
Wing, S. R., Botsford, L. W., Largier, J. L. and Morgan, L. E. (1996). Spatial structure of relaxation events and crab settlement in the northern California upwelling system. Oceanographic Literature Review, 7 (43), 701.Google Scholar
Witman, J. D., Etter, R. J. and Smith, F. (2004). The relationship between regional and local species diversity in marine benthic communities: a global perspective. Proceedings of National Academy of Sciences, 101 (44), 15664–9.Google Scholar
Witman, J. D., Leichter, J. J., Genovese, S. J. and Brooks, D. A. (1993). Pulsed phytoplankton supply to the rocky subtidal zone: influence of internal waves. Proceedings of the National Academy of Sciences of the United States of America, 90 (5), 1686–90, http://dx.doi.org/10.1073/pnas.90.5.1686.Google Scholar
Wonham, M. and Carlton, J. (2005). Trends in marine biological invasions at local and regional scales: the northeast Pacific Ocean as a model system. Biological Invasions, 7 (3), 369–92, http://dx.doi.org/10.1007/s10530-004-2581-7.Google Scholar
Worm, B. and Paine, R. T. (2016). Humans as a hyperkeystone species. Trends in Ecology & Evolution, 31 (8), 600–7, http://dx.doi.org/10.1016/J.TREE.2016.05.008.Google Scholar
Yamada, S. B., Peterson, W. T. and Kosro, P. M. (2015). Biological and physical ocean indicators predict the success of an invasive crab, Carcinus maenas, in the northern California current. Marine Ecology Progress Series, 537, 175–89, http://dx.doi.org/10.3354/meps11431.Google Scholar
Yang, B., Emerson, S. R. and Peña, M. A. (2018). The effect of the 2013–2016 high temperature anomaly in the subarctic Northeast Pacific (the “Blob”) on net community production. Biogeosciences, 15 (21), 6747–59.Google Scholar
Yorke, C. E., Miller, R. J., Page, H. M. and Reed, D. C. (2013). Importance of kelp detritus as a component of suspended particulate organic matter in giant kelp Macrocystis pyrifera forests. Marine Ecology Progress Series, 493, 113–25, http://dx.doi.org/10.3354/meps10502.Google Scholar
Young, C. M. and Braithwaite, L. E. E. F. (1980). Orientation and current-induced flow in the stalked Ascidian Styela montereyensis. Biological Bulletin, 159 (2), 428–40.Google Scholar
Young, M. A., Cavanaugh, K., Bell, T. et al. (2016). Environmental controls on spatial patterns in the long-term persistence of giant kelp in central California. Ecology, 86 (1), 4560, http://dx.doi.org/10.1890/15-0267.1.Google Scholar
Zaba, K. D. and Rudnick, D. L. (2016). The 2014–2015 warming anomaly in the southern California current system observed by underwater gliders. Geophysical Research Letters, 43 (3), 1241–8, http://dx.doi.org/10.1002/2015GL067550.Google Scholar
Zabin, C. J., Ashton, G. V., Brown, C. W. and Ruiz, G. M. (2009). Northern range expansion of the asian kelp Undaria pinnatifida (Harvey) suringar (Laminariales, Phaeophyceae) in Western North America. Aquatic Invasions, 4 (3), 429–34, http://dx.doi.org/10.3391/ai.2009.4.3.1.Google Scholar
Zacharias, M. A. and Kushner, D. J. (2006). Sea temperatureand wave height as predictors of population size structure and density of Megastraea (Lithopoma) undosa: implications for fishery management. Bulletin of Marine Science, 79 (1), 7182, www.ingentaconnect.com/content/umrsmas/bullmar/2006/00000079/00000001/art00004.Google Scholar
Zahn, L. A., Claisse, J. T., Williams, J. P., Williams, C. M. and Pondella, D. J. II (2016). The biogeography and community structure of kelp forest macroinvertebrates. Marine Ecology, 37, 770–85, http://dx.doi.org/10.1111/maec.12346.Google Scholar
Zhan, A., Darling, J. A., Bock, D. G. et al. (2012). Complex genetic patterns in closely related colonizing invasive species. Ecology and Evolution, 2 (7), 1331–46, http://dx.doi.org/10.1002/ece3.258.Google Scholar
Zimmerman, R. C. and Kremer, J. N. (1986). In situ growth and chemical composition of the giant kelp, Macrocystis pyrifera, response to temporal changes in ambient nutrient availability. Marine Ecology Progress Series, 27 (2), 277–85.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×