User:LinusTheScientist/California Current

From Wikipedia, the free encyclopedia

The California Current is an eastern boundary current in the Pacific Ocean that flows southward along the western coast of North America, from the Strait of Juan de Fuca to the southern tip of Baja California. It carries the southward flow of the North Pacific Gyre and is one of six major coastal currents affiliated with strong upwelling zones (including the Humboldt Current, the Benguela Current, the Canary Current, the Portugal Current, and the Somali Current).[1] The current moves relatively fresh and cool surface water from northern latitudes southward and plays a key role in nutrient and sediment transport as well as plankton community composition.[2][1][3][4] The current also displays strong seasonal, interannual, and decadal variability due to changes in atmospheric pressure, winds, and ocean properties.[5]

Upwelling brings nutrients to the surface in the California Current region, fueling high productivity and supporting a diverse ecosystem.[6][7] The current is economically important to coastal communities, linked to livelihoods through fisheries as well as shipping ports and ecotourism.[8][9] Climate change is expected to impact the California Current in many ways, including intensified upwelling,[10][11] increased seasonal hypoxia,[12] acidification,[13] increased harmful algal blooms,[14] and disruptions to fisheries, marine mammals, and seabirds.[15][16]

This diagram shows the southward-flowing California Current, which ranges from the Strait of Juan de Fuca to Baja California. The California Undercurrent (subsurface) and Davidson Current (surface) flow northward, with the surface flow limited to winter months. The Southern California Eddy/Countercurrent occurs in the Southern California Bight.
The California Current is part of a system of currents along the west coast of North America. It flows southward, from the Strait of Juan de Fuca to the southern tip of Baja California.

Physical properties[edit]

The California Current is 50-100 kilometers wide and flows at a maximum surface velocity of 40 to 80 centimeters per second, extending to a depth of about 300 meters.[1][2] The current is made up of narrow areas of fast flow called jets within a broad area of slow flow, with two prominent zones of faster flow, a main branch offshore and a seasonal coastal jet.[5][17][18] On average, the main branch of the current is located 200-300 kilometers offshore.[1] The characteristics of the California Current vary along the coast, and it can be separated into a northern region north of Cape Mendocino, a southern region south of Point Conception, and a central region between the two.[5] The transport of water by the current is a few Sverdrups, or million cubic meters per second.[1]

The current has two components, both of which are ultimately driven by atmospheric pressure and winds. The first component is the southward flow of the North Pacific subtropical gyre, a large clockwise flow in the North Pacific Ocean.[1] The California Current is fed by the North Pacific Current on the north side of the gyre.[5] This part of the flow is driven by large-scale winds in the Pacific, which cause water to build up in the center of the gyre. This pressure distribution causes a surface flow to the south on the eastern side of the gyre through the geostrophic balance, which describes how pressure and the rotating earth combine to drive ocean currents. The buildup of water in the center of the gyre also drives downward motion of water called Ekman pumping. The squeezing of water columns by this downward motion causes southward Sverdrup transport throughout the gyre as the water columns move southward to maintain their angular momentum.[1]

The second component of the California Current is the coastal upwelling system, driven by alongshore winds. These winds occur because the westerlies, the prevailing winds in the midlatitudes, are deflected to the south by the North American continent. In the Northern Hemisphere, water is transported to the right of the wind direction due to the rotation of the Earth in a process called Ekman transport. When winds blow to the south along the coast, this process moves water to the right (offshore). The movement of the water creates a pressure gradient with lower pressure along the coast, driving geostrophic surface flow to the south. The movement of water offshore also pulls water up from below, causing upwelling.[1]

The movement of water from the north down the west coast of the United States, combined with upwelling in summer, results in much cooler ocean temperatures than at comparable latitudes on the East Coast, where ocean currents bring water from the tropics. Ocean surf temperatures are rarely above 66 °F (19 °C) during the summer along the California coast south to San Diego, whereas they are often above 80 °F (27 °C) on the East Coast from North Carolina southward.[19] Cool surface water, along with the offshore movement of warm air on land, contributes to fog production in the coastal zone.[20]

Related currents[edit]

The California Current is part of a system of currents that occur along the west coast of North America, including the California Current, the coastal jet, the California Undercurrent, the Davidson Current (sometimes called the Inshore Countercurrent), and the Southern California Eddy (also called the Southern California Countercurrent).[18][5] The coastal jet is a strongly seasonal current occurring 5 to 25 kilometers from shore that contributes to the southward flow of the California Current.[18][5] The jet flows southward in the spring and summer, with the strongest flow in spring and the maximum speed moving progressively offshore. In winter, flow along the coast tends to be northward. The strength and direction of the coastal jet are closely linked to winds. Especially in winter, the flow can vary greatly, even reversing, on a timescale of days as winds switch from north to southward.[18]

The California Undercurrent flows northward underneath and inshore of the California Current, carrying tropical Pacific water with a warm, salty, low-oxygen signature. The undercurrent flows at speeds of 10 centimeters per second or faster, and it is about 20 kilometers wide. Its core is located at 250 meters depth and it extends to 1000 meters depth. The undercurrent occurs because the water along the coast has layers of density that change the pressure gradient at depth, leading to a reversal in the flow direction through a process called the thermal wind.[1]

Inshore of the California Current is the Davidson Current, which flows to the north in the fall and winter. This current is the surface expression of the California Undercurrent, which moves closer to the surface during the winter and deepens in summer.[2] The Southern California Eddy/Countercurrent originates at an eastward bend in the California Current near 32°N, where a northward flow splits off from the main current and moves into the Southern California Bight. This flow is seasonally variable, but there is northward flow near the shore most of the year in the northern part of the bight.[18] The flow tends to recirculate within the bight in summer, when it is called the Southern California Eddy, whereas it can connect to the Davidson Current in winter, when it is called the Southern California Countercurrent.[21]

Upwelling[edit]

Alongshore winds drive upwelling in the California Current system. Upwelled water comes from 150-200 meters depth and is defined by its high nutrient content and low temperature, producing a region of cool surface water 80-300 kilometers from shore and fueling high productivity in the coastal zone. The offshore motion of water is not uniform, but rather a complicated, swirling pattern of eddies and jets. These jets take the form of squirts, which end offshore, and meanders, which circle back to the coast.[1]

This satellite image shows the temperature of the ocean surface along the coastline of California. Cooler temperatures occur along the coast, with warmer temperatures offshore and south of 34.5°N. The cooler temperatures are associated with upwelling.
Water temperature along the coast of California reveals the presence of cooler, upwelled water. The upwelled water moves offshore in filaments that occur near points in the coastline.

Two wind-driven processes control upwelling in the current. The first process is the Ekman transport of water offshore, driven by southward flow along the coast. Because the coast is a solid boundary, this transport requires that water move up from below to replace the water moved offshore. Similarly, if winds blow northward along the coast, water is transported towards the shore, causing downwelling. The second process leading to upwelling is the curl of the wind stress, which describes the degree of rotation of the winds. When the speed of the southward wind increases moving offshore, the curl of the wind is positive and the Ekman transport increases moving away from the coast, which enhances upwelling.[1]

The extent and strength of upwelling varies strongly with the seasons, especially in the northern part of the current, with shifts in the atmospheric pressure patterns that drive winds. In the winter, north of Cape Mendocino, winds change from southward to northward and there is downwelling instead of upwelling. South of Cape Mendocino, winds support upwelling during the entire year. The upwelling system is more sensitive to local winds in the northern part of the current compared to the southern part.[22] Overall, the strongest upwelling occurs from April to July, with the maximum occurring near 34°N, offshore of Point Conception.[1] The shape of the coastline and the ocean bottom, including features like capes and canyons, also affects upwelling by causing jets of water to move offshore.[5]

Upwelling indices are used to assess the strength of upwelling, and they are calculated based on winds or associated Ekman transport.[1] The Bakun Index is based on the monthly mean Ekman transport, calculated from atmospheric pressure.[23] The Coastal Upwelling Transport Index and the Biologically Effective Upwelling Transport Index provide improved estimates of vertical transport and vertical nitrate flux.[24]

Variability[edit]

The California Current varies on multiple timescales, from seasonal to interannual to decadal. The major driver of seasonal variability is changes in winds and wind stress curl.[18] In winter, the current is located farther offshore, with minimal upwelling and northward flow along the coast associated with the Davidson Current.[1][2] In summer, an upwelling front develops at the coast, where there is a sharp change in water properties at the junction between upwelled water at the coast and surface water offshore. This front moves offshore as the season progresses.[1] The onset of upwelling, called the "spring transition," and the change to downwelling, called the "fall transition," vary from year to year with winds.[5][25] When upwelling starts later than usual, the water is warmer than usual and impacts are felt throughout the California Current ecosystem.[26]

On interannual time scales, the El-Niño Southern Oscillation (ENSO) also impacts the California Current. During ENSO events, coastal Kelvin waves propagate northward along the coast from the equator. They change the vertical locations of sharp changes in properties like water temperature, described by the thermocline, and nutrients.[27][28] This, in turn, changes the properties of upwelled water. When nutrient content is lower than usual, productivity is also lower than usual.[29] El Niño events tend to cause warmer, saltier water and higher sea level than usual, with stronger flow to the north along the coast in the fall and winter and lower productivity.[28][27] La Niña produces the opposite effect.[30] ENSO can also change the California Current by affecting atmospheric pressure in the North Atlantic and by causing more "tropical" water to be carried northward along the coast.[31][32] El Niño events vary in their impact on the current, depending on when they arrive and pre-existing water conditions.[28][29]

Moving north along the coast, more of the variability in the current can be explained by decadal changes, although the dynamics of these changes are not fully understood.[5] The Pacific Decadal Oscillation (PDO) and the North Pacific Gyre Oscillation (NPGO) describe patterns of variability in sea surface temperature and height that account for some of the changes in the California Current system. During the positive phase of the PDO, the California Current tends to be weaker, whereas the current is stronger and upwelling is enhanced during the positive phase of the NPGO.[33][34]

Biogeochemical properties[edit]

In the California Current and other eastern boundary currents, upwelling brings water with a signature of high nutrients, low oxygen, and high dissolved inorganic carbon to the surface. The interactions between physical upwelling and advection of water and biological processes such as photosynthesis and respiration make the California Current a biogeochemically dynamic region with high variability.[35][36]

Nutrient flux[edit]

Upwelling and downwelling in the California Current are important drivers of nutrient transport, and the biological response within the current depends on the interaction between these physical processes and underlying biogeochemistry.[5][37] Water transported by upwelling is rich in nutrients that are essential for the growth of ocean algae, so upwelling supports productive ecosystems.[38] Changes in nutrient fluxes on seasonal timescales cause distinct regimes based on atmospheric conditions and ocean circulation. Flux of nutrients also varies along the length of the current, because the shape of the coastline strongly influences the intensity of upwelling. This creates an inherent patchiness to the distribution of nutrients and oxygen along the coast.[39]

Additional sources of nutrients from river and estuarine inputs along the US West Coast fuel ecosystem productivity.[40] Notably, the Columbia River discharges an average of 3600 cubic meters of water per second.[41] The nutrient signature from the Columbia River, which drains a total area of 660,480 square kilometers, can be tracked throughout the northern extent of the California Current.[42]

This diagram shows the processes leading to hypoxia on the continental shelf. Water that is upwelled along the coast is low in oxygen but high in nutrients. Once the water reaches the surface, phytoplankton grow and die along the coast. When they sink to the bottom and decompose, more oxygen is used up and the water on the shelf becomes hypoxic.
Upwelling of low oxygen water can produce hypoxic and acidic conditions on the shelf due to the decomposition of plankton after a bloom. This process consumes oxygen and releases carbon dioxide in shelf waters that are already low in oxygen and high in carbon. [36]

Limiting nutrients[edit]

Nitrogen, phosphorus, and iron are important limiting nutrients for the California Current. Macronutrients such as nitrate, phosphate, and silicate, along with trace metals like iron, are supplied to the coastal region by upwelling.[43][44] Iron availability governs nitrate drawdown in many coastal upwelling systems and is strongly influenced by physical drivers and bathymetry along the California Current. Locations with narrow continental shelves can become iron-limited due to low levels of suspended sediment and high nitrate concentrations from upwelling.[6] Further offshore, away from the freshly upwelled waters, there are regions that are high in nitrate, but iron-limited. These areas are designated as high nutrient low chlorophyll regions, where given the abundance of nitrate, higher chlorophyll levels may be expected.[39]

Oxygen[edit]

Dissolved oxygen (DO) dynamics play a significant role in the biogeochemical processes of the California Current. The oxygen minimum zone intersects the continental slope at more than 600 meters depth.[12] Oxygen minimum zones exist between depths of 100 and 900 meters, with minimum values between 300 and 500 meters. These zones are formed due to a combination of poor ventilation with surface waters and respiration.[45]

The biogeochemical cycling of many important inorganic compounds is highly oxygen-dependent.[46] The California Current experiences significant variation in oxygen, and there is recent evidence of strong hypoxia (DO < 1.4 milliliters per liter) or even anoxia (DO = 0 ml l-1) over large expanses of the current.[47][12] Oxygen concentrations are not uniform across the current; general ocean circulation, wind forcing, and bathymetry cause patchiness.[37] Oxygen also changes on a seasonal basis, with low DO typically observed during the summer, between the spring and fall transition of prevailing winds. There is evidence of oxygen changing on longer time scales as well, because decadal variations in ocean gyres create conditions that transport more upwelled water onto large portions of the continental shelf.[48]

Carbonate chemistry[edit]

Coastal water in the California Current system is naturally more acidic and susceptible to undersaturation of carbonate minerals like aragonite than other ocean regions because upwelling moves comparatively low pH waters onto the continental shelf.[13][49] Aragonite saturation state, which depends on the relative abundance of carbonate ions in the water, impacts the ability of calcium carbonate shell-forming organisms such as coccolithophores to produce their shells.[50]

Surface waters also exchange carbon dioxide (CO2) with the atmosphere, and the amount of CO2 in ocean water affects its pH and saturation state.[51] Water that upwells from depth tends to be high in CO2 because these waters have been isolated from the surface, and CO2 produced by respiration builds up over time. When this water is brought to the surface, CO2 is released to the atmosphere, but the availability of nutrients also supports high productivity. This productivity takes up CO2, and can cause the coastal region to become a net carbon sink rather than a source.[52] The carbonate chemistry of surface water is quite variable because of these dynamics.[35][13]

Geological properties[edit]

Geologic context[edit]

The complex bathymetry and coastal geomorphology of the Southern California Bight region influence the path of ocean currents.

The geological setting of the California Current helps shape its flow dynamics as well as seafloor structures and habitats. Events like submarine earthquakes can alter the seabed, disturbing biological communities and altering food webs.[53]

The California Current flows over a tectonically active region with multiple plate boundaries, including the Cascadia Subduction Zone, San Andreas Fault, and the Juan de Fuca and Gorda Ridges. The Cascadia Subduction Zone sits between Vancouver Island and Cape Mendocino, where the Juan de Fuca Plate slides under the North American Plate, creating an active continental margin characterized by frequent earthquakes and tsunamis.[54] The San Andreas Fault is a continental transform boundary along the coast of Southern and Central California, where the Pacific Plate grinds against the North America Plate.[55][56] The Juan de Fuca and Gorda Ridges are divergent plate boundaries where the Pacific Plate spreads away from the smaller Juan de Fuca and Gorda Plates.[57]

The active Cascadia Subduction Zone creates a relatively narrow continental shelf along the US West Coast, with depth rapidly increasing away from the coast.[57] This quick drop-off is in contrast to passive continental margins where there is no subduction and depth drops off more gradually, as is the case along the East Coast of the US. The steep seafloor topography along the US West Coast contributes to characteristic upwelling along the California Current by forcing deep, nutrient-rich water upward.[22]

Plate tectonics also created the Southern California Bight, the portion of the California coastline that runs east to west, south of Point Conception. This bend in the coastline creates a large eddy off the California Current called the Southern California Eddy.[58] The Channel Islands off the coast also impact the flow of surface currents in the Southern California Bight region.

Finally, features like capes, submarine canyons, and banks influence the flow of the California Current near the coast, as well as the intensity of upwelling.[5] Cape Blanco causes a jet of water to separate from the coast, carrying coastal water across the shelf.[59] Heceta Bank produces an offshore movement of the coastal jet, recirculation inshore of the bank, and a "retention" region with higher productivity.[60][40]

The Eel River, located in Northern California, discharged sediment into the California Current after a series of heavy storms in the fall of 2012.

Sediment transport[edit]

The complex interactions between plate tectonics, waves, discharge of sediments from rivers, coastal erosion, and sea level change define the geomorphological shape of the coastline that parallels the California Current. Coastal geology determines the boundaries of features called littoral cells,[61] which are complete systems of sediment cycling.[62]

Particles of rock, sand, and silt make their way into these oceanic littoral cells from sources including crumbling sea cliffs and rivers. In the ocean, the fate of these particles depends on factors like the angle of incoming waves relative to the coastline, the bathymetry of the sea floor, and the relative position of sediment discharge to major geological features like headlands and submarine canyons.[63] Particles will often flow along the coastline in the direction of longshore drift, forced by incoming waves, until they are blocked from moving further by headlands.[64] Particles may escape littoral cell systems by sinking into deep submarine canyons, or be carried further offshore by periods of intensified river discharge.[3] Sediments that make it into the California Current can be carried southward for great distances, depending on factors such as sediment grain size and the seasonal intensity of winds.

These sediment inputs to the sea are important to physical and biogeochemical ocean processes. Riverine inputs are an important link between the ocean and the land, transporting nutrients and sediment that settles on continental shelves. Major rivers such as the Russian River and Columbia River carry large volumes of sediment from land into the California Current system. Local waves, bottom currents, and storms influence the volume of sediment delivery, as well as the suspension of particles near the bottom.[65] The movement of sediments offshore is affected by the California and Davidson Currents, upwelling, seasonality, and wave patterns.[66]

Sediment transport processes impact coastal bathymetry, and they are important to multiple fields including geology and civil engineering. Studying past sedimentation events and processes in the geologic record gives scientists insight into what the California Current and broader region looked like thousands of years ago, including past levels of biological productivity, upwelling force, and changes to atmospheric circulation and current flow.[67]

Biology[edit]

This satellite image shows the coast of California and the California Current system from space. Phytoplankton blooms are visible as intricate swirls of green in the blue ocean along the coast.
A satellite image of the California Current system reveals phytoplankton blooms along the coast fueled by upwelling.

The California Current is a highly biologically active region, home to a range of species diversity spanning many taxa. Eastern boundary currents like the California Current support many different species, a quality termed species richness,[7] as well as high productivity relative to other ocean regions.[68] The scale of organisms in the current spans from single-celled bacteria up to the largest animal on earth, the blue whale.[69][70]

Despite their small size, microscopic organisms have a large impact on the region. Primary productivity in the current is influenced by microbial organisms, including marine viruses, bacteria, and grazers. Although they occur on microscopic scales, the interactions between these groups constitute important ecosystem controls for the broader California Current that change on daily, seasonal, and interannual time scales.[69]

Basin-scale climatic forces such as El Niño, La Niña, and Pacific Decadal Oscillation (PDO) events combine with seasonal variations to create dynamic changes in water properties to which the biology of the California Current responds.[71][72] For example, biological communities in the North Pacific experience shifts on a 50-year timescale related to the PDO. A warm "sardine" regime favors these fish during periods when the California Current is weaker, and a cool "anchovy" regime results when the current is stronger and upwelling is enhanced.[73]

Primary producers[edit]

Multiple species of the algal groups diatoms, coccolithophores, and dinoflagellates form diverse and dynamic communities throughout the California Current. The size and composition of these communities is considered an essential ecosystem metric.[74] Phytoplankton are microscopic ocean plants (smaller than 200 micrometers) that shape food web dynamics by providing food for other species, influencing available nutrient concentrations, and fueling primary production through photosynthesis.[75]

Microscopic algae like diatoms are an essential part of the food web in the California Current system and beyond.

Ecosystem models show that large phytoplankton like diatoms comprise 90% of the California Current system's primary productivity.[74] Rapid, significant periods of phytoplankton growth, or blooms, fuel ecosystem productivity. Blooms have also been linked to ocean oxygen-depletion events, which occur seasonally in the California Current system.[74] Some phytoplankton species found in the California Current, such as Pseudo-nitzschia and Karenia brevis, produce toxins harmful to other organisms, like domoic acid and brevetoxins.[14]

Macroscopic brown algae of the family Phaeophyceae are abundant throughout the California Current. Common brown algae include bull kelp and giant kelp. Large aggregations of kelp form dense forests along the coastal region of the current. These kelp forests are important habitat for thousands of marine species.[71]

Consumers[edit]

Consumers, or organisms that are unable to produce their own food, must rely on predation to acquire energy. They exist across a range of trophic levels, which describe an organism's position in the food web. The following sub-sections outline a few ecologically-important consumers found in the California Current.

Zooplankton[edit]

Zooplankton such as copepods shape the waters they live in by grazing on phytoplankton and providing food for larger animals in the California Current.

Zooplankton in the California Current include copepods, jellyfish, euphausiids, and crustacean larvae. By consuming phytoplankton and other zooplankton, they facilitate the transfer of energy and nutrients through the food web and ecosystem.[74] Environmental and climatic factors like upwelling and the PDO favor different types of zooplankton, resulting in seasonal shifts in zooplankton community structures that further impact the food web.[76] The copepod community, for example, includes more cold-water species when the California Current is strong, and more warm-water species when it is weak.[4]

Zooplankton community structure also changes significantly along the length of the California Current because of changes in water characteristics.[74] In addition, the formation and maintenance of large eddies can isolate zooplankton and influence survival of their larvae, as has been observed at Point Conception in southern California and Punta Eugenia in Baja California.[77]

Upwelling dynamics also impact the transport of zooplankton, including the larval stages of marine organisms. The distribution of transported spawn is an important component of recruitment (the process by which new individuals are added to a population) for many species, and it has implications for future fish abundance throughout the California Current.[78] The relative timing of spawning events and the spring and fall transitions in upwelling cause significant variations in the ranges of many species.[79]

Shellfish[edit]

Shellfish live in a variety of habitats within the California Current, such as kelp forests, rocky intertidal zones, sandy bottoms, and even open water habitats. Echinoderms, including sea urchins and sea stars, commonly inhabit California Current kelp forests and rocky intertidal zones, although sea star wasting disease has ravaged populations in recent decades.[80] Filter-feeding mollusks like abalone, snails, oysters, clams, and mussels are also widespread within the California Current region, as are shrimp and crab species.[81] Many of these shellfish species are commercially valuable, most notably Dungeness crab and Oregon pink shrimp (Pandalus jordani).[82]

Fish[edit]

Fish at lower trophic levels feed on zooplankton and transport energy up the food chain. Examples of forage fish with varying ranges throughout the California Current include Whitebait smelt, Pacific sardine, Pacific herring, and Northern anchovy.[83] These species constitute important food sources for higher tropic levels in the ecosystem.[84]

Pacific sardines and other forage fish species are an important trophic link and fisheries resource in the California Current system.

The California Current is also home to many types of predatory fish. Rockfish (Sabastes) is a genus with over 100 species present along the US West Coast. These commercially-important fish (see Fisheries section below) rely on kelp forests for habitat. Other demersal (groundfish) predatory fishes important to the California Current ecosystem include hake, cod, and spiny dogfish.[83] Predators also include a number of shark species including the salmon shark, which migrate throughout the California Current.[85]

Seabirds[edit]

A wide variety of seabirds take advantage of the productive waters within the California Current. Notable species include Cassin's auklets, Rhinoceros auklets, common murres, pigeon guillemots, brown boobies, Brown pelicans, sooty shearwaters, and several species of cormorants and gulls.[72] Sooty shearwaters migrate 40,000 miles annually from their breeding grounds in New Zealand to feed in the California Current.[86]

Marine mammals[edit]

High biological productivity in the California Current provides food for humpback whales, seabirds, and other animals.

High levels of nutrient upwelling and primary productivity within the California Current create food for large marine mammals, including whales, seals, and sea lions. Some of the most visible groups of organisms within the current fall into the order Cetacea, which includes dolphins, porpoises, toothed whales, and baleen whales.

Dolphins are by far the most numerous cetaceans, with an estimated population size of 540,000 individuals within the California Current ecosystem. Species that prefer warmer conditions, including common dolphins and short-finned pilot whales, tend to stay in the southern section of the Current off the coast of California. Cold-loving species like Pacific white-sided dolphins tend to stay farther north, off the coast of Northern California, Oregon, and Washington. Dall’s porpoises also inhabit the colder regions of the current. The “cosmopolitan” dolphins, including bottlenose dolphins and killer whales, travel the entire range of the California Current.[87] Individual species ranges vary from year to year as temperatures fluctuate, especially during El Niño and La Niña events.[88]

Baleen and toothed whales also inhabit the California Current ecosystem, including humpback, gray, minke, blue, and sperm whales.[89] Many species of baleen whales, including gray whales and humpbacks, migrate thousands of miles each year between seasonal feeding grounds and mating/birthing grounds. These migration paths take many whales directly through the California Current ecosystem.[90]

Seals and sea lions are also common in the current region. The six most frequently occurring pinniped species are California sea lions, Stellar sea lions, northern fur seals, Guadalupe fur seals, harbor seals, and northern elephant seals.[91] Whales, seabirds, and pinnipeds often congregate around schooling fish, creating "feeding frenzies" in coastal waters, especially during periods of intense upwelling.[92]

Human dimensions[edit]

The productive nature of the California Current creates numerous opportunities for industry, recreation, and tourism. Human habitation on the west coast of North America dates back as far as 130,000 years.[93] The remains of marine species native to California Current ecosystems have been found in pre-historic archaeological sites, inland and distant from their natural species ranges.[94] Such remains suggest that humans have actively harvested organisms from the California Current and relied on them as a food source and means of trade since at least pre-historic times.

In modern times, the California Current plays an important role in the livelihoods of many people, enabling active fishing, tourism, and shipping industries. These interactions between humans and the California Current have also led to significant human impacts on the current (see Climate Change section below).

Fisheries[edit]

Commercial fishing in the California Current region generates over $558 million in annual revenue while employing over 232,000 people.[8][95] Shellfish like crabs and shrimp make up the dominant source of commercial revenue, whereas pelagic fishes make up the dominant catch by tonnage.[96] Fishes like groundfish, salmon, and tunas are common recreational catches, with over 1.97 million recreational fishers participating in harvest in 2006. Recreational anglers stimulate coastal economies by spending money on charter trips, boat rentals, gear, and other personal costs. Along with commercial and recreational harvests, aquaculture and subsistence fishing also contribute greatly to non-harvest fishery levels.[8] The makeup and success of fishery catches are influenced by regional processes and interannual, annual, and decadal climatic variability as well as long-term climate changes.[96]

The productive waters of the California Current support lucrative fisheries off the coast of Monterey and elsewhere throughout the region.

Harvest of California Current species is managed by various organizations such as the Pacific Fishery Management Council and the National Oceanographic and Atmospheric Administration. Management strategies are often organized by the location and life histories of the species, such as groundfishes, highly migratory species, and coastal pelagic species that live in the water column. In recent years, effort has been made to implement ecosystem-based management approaches.[97]

Some species are more strictly managed due to historic exploitation in the region, or based on life history and desirability.[98] For example, the California market squid has a short life-span and high fecundity, so fishing pressure has little effect on the population.[99] Comparatively, the Pacific sardine has experienced historic overexploitation, and the population is subject to large natural fluctuations.[100][101] As a result, the fishery is heavily monitored and frequently closed.[100] In many instances, management strategies involve numerous stakeholder groups because of the important roles the species play in communities and ecosystems. For example, salmon populations fuel key commercial, recreational, and subsistence fisheries, and salmon also play a critical role in California Current and freshwater ecosystems. Adaptive management strategies have been implemented to reduce human impact on salmon and associated environments.[102]

Shipping, ports, and other marine operations[edit]

The California Current region supports industry and ship traffic through hubs such as the Port of Long Beach.

Along with fishing, marine operations on the west coast of the United States and Mexico are impacted by the properties of the California Current and vice versa. Some of the busiest ports in the world, such as the Ports of Los Angeles and Long Beach, bring heavy ship traffic through the California Current region. While this generates income for the region, it also impacts the area and current itself. For example, major ports often require dredging and widening to enable deep draft vessels to enter. This impacts sediment transport processes, as well as bottom-dwelling producers and consumers.[103] Researchers and environmental groups have highlighted the impacts of marine operations on migratory animals like whales and seabirds that visit the region, and the negative impacts of invasive species and pollutants brought into the California Current by large vessels.[104]

Marine oil and gas operations, like oil drilling rigs, have a long history within the offshore California Current. Some places, like Oregon, have banned drilling, while others still have active industries, such as Baja Mexico and California. While active drilling and oil exploration generate income, they may also have many negative environmental impacts on the region.[105]

Ecotourism[edit]

The high biodiversity of the California Current lends itself to ecotourism ventures including whale watching, fishing charters, SCUBA diving, and sea kayaking.[106][9] In some places, decommissioned oil rigs have been transformed into artificial reef habitats, which now serve as popular diving destinations.

In popular culture[edit]

In the Disney/Pixar animated films Finding Nemo and Finding Dory, the California Current is portrayed as a superhighway that fish and sea turtles use to travel to California. The characters Marlin, Nemo, and Dory join Crush, Squirt, and a group of baby and adult sea turtles in using the California Current to help them travel to Morro Bay, California to find Dory's parents.

Climate change[edit]

Anthropogenic climate change has introduced additional variability in the physical and geochemical properties of the California Current system. These changes have ramifications for the organisms and coastal communities supported by the current.[107][10][16]

Physical changes[edit]

Increasing atmospheric CO2 and temperature will alter the characteristic physical properties of the California Current region in multiple ways.[108][36][109] Climate change has already influenced the seasonality of upwelling, and climate models predict that upwelling will begin, peak, and end later, with a delay in its onset of up to a month.[11] Additionally, shifts in atmospheric circulation and temperature will cause regional shifts in upwelling throughout the California Current system.[16][110] Upwelling in northern regions of the current, from Southern British Columbia to northern California, may be enhanced by stronger southward winds.[16][111][112] These increased winds are caused by the disproportionate warming of land relative to the ocean, especially at higher latitudes where temperature differences are most pronounced.[16][111] Upwelling along the northern and central California Current may also be further enhanced by eddy activity.[10] In contrast, the southern California Current region may experience decreased upwelling due to intensified ocean stratification. Induced by warming of the upper ocean, intensified stratification may counteract the effects of upwelling by reducing mixing in this region.[113][114]

As ocean temperatures and stratification increase, lower dissolved oxygen levels are already occurring due to decreased solubility and reduced ocean mixing.[115] Increased hypoxia and water column anoxia observed throughout portions of the California Current are also closely tied to increased upwelling of deoxygenated water, as well as the respiration of increased organic matter introduced by this upwelling.[12][115][10] Because oxygen plays a key role in the cycling of other elements, such as carbon, nitrogen, and iron, deoxygenation will have widespread biogeochemical effects.[116]

Pyrosomes are a type of gelatinous animal that is more tolerant to warming conditions and has experienced a range expansion in the California Current region. A notable bloom occurred in the Northern California Current in the early to mid 2010's following anomalous warming.

Human-induced increases in atmospheric CO2 since the Industrial Revolution, primarily from the burning of fossil fuels, have greatly increased the uptake of CO2 by the ocean. This process has lowered seawater pH in various regions, a phenomenon known as ocean acidification.[109] This process is exacerbated in upwelled waters, like those that occur in the California Current, which are typically higher in CO2 than overlying waters. Ocean acidification will likely have the greatest effect in northern regions of the current, where upwelling is greater.[117] Increased ocean acidification also results in reduced carbonate saturation state.[49]

Biological ramifications[edit]

The combination of increased temperature, ocean acidification, decreased oxygen, and nutrient availability changes will have widespread impacts on living things throughout the California Current system, from microbial life all the way up the food web to humans.[16]

As a result of upwelling-affiliated nutrient availability, primary production is predicted to increase most along the central and northern California Current, supporting higher zooplankton abundance in these areas.[10][15] Increasingly favorable conditions for phytoplankton blooms may also increase the frequency of harmful algal blooms and hypoxic events, as has already been reported off of Monterey, California and along the Oregon coast.[38][118] In contrast, intensified stratification along the southern California Current may limit nutrient transport, causing potential phytoplankton declines near Baja California.[10][114] Changes in upwelling-favorable winds may shift phytoplankton blooms further offshore in some regions of the current, influencing food web dynamics.[16]

Many organisms in the California Current ecosystem have evolved the timing of their developmental and reproductive stages, known as phenology, to align with optimal seasonal upwelling and prey availability.[16][15][119] As upwelling conditions change, the mismatch between evolutionary timing and actual seasonal upwelling and prey availability will cause ecological shifts with cascading effects on upper trophic levels including invertebrates, fishes, marine mammals, and seabirds.[16][120][121] For example, small forage fishes throughout the current will be negatively impacted by changing upwelling and prey conditions.[122] Declines and shifts in these important prey species will in turn influence the success of larger marine predators like seabirds and marine mammals.[119][121]

Shifts to upwelling and prey species will be compounded by the effects of ocean acidification, temperature changes, and oxygen availability, inducing diverse ecological shifts in the California Current.[16] Ocean acidification will impact the ability of many marine invertebrates and phytoplankton to produce protective calcium carbonate shells because of reduced carbonate saturation state.[109][16][49] This threatens prey availability, water quality, and contribution of calcium carbonate shells to seafloor habitat.[109][16] Reduced oxygen levels will likely support an increase in organisms that are tolerant of lower oxygen conditions.[16]

Changing biogeochemical conditions in the California Current have also already induced range shifts in a number of species.[16] For example, the range of the Humboldt Squid has recently expanded, while seabird and marine mammal ranges have shifted latitudinally with temperature, leading to a decline in biodiversity.[120][121] In comparison, gelatinous organisms such as jellyfish or pyrosomes, which have lower nutrient and oxygen demands and are more tolerant of warming conditions, have experienced population increases in recent years.[123] Given the socioeconomic link between the California Current ecosystem and adjacent coastal communities, these ecological shifts and their cascading trophic effects may render coastal communities vulnerable to the impacts of climate change.[124][125][126]

References[edit]

  1. ^ a b c d e f g h i j k l m n o Talley, Lynne D.; Pickard, George L.; Emery, William J.; Swift, James H. (2011). Descriptive physical oceanography : an introduction (6th ed.). Amsterdam: Academic Press. ISBN 978-0-7506-4552-2. OCLC 720651296.
  2. ^ a b c d Lynn, Ronald J.; Simpson, James J. (1987). "The California Current system: The seasonal variability of its physical characteristics". Journal of Geophysical Research. 92 (C12): 12, 947–12, 966. doi:10.1029/jc092ic12p12947. ISSN 0148-0227.
  3. ^ a b Thornton, Scott E. (1981). "Suspended sediment transport in surface waters of the California Current off southern California: 1977–78 floods". Geo-Marine Letters. 1 (1): 23–28. doi:10.1007/bf02463297. ISSN 0276-0460.
  4. ^ a b Keister, J. E.; Di Lorenzo, E.; Morgan, C. A.; Combes, V.; Peterson, W. T. (2011). "Zooplankton species composition is linked to ocean transport in the Northern California Current". Global Change Biology. 17 (7): 2498–2511. doi:10.1111/j.1365-2486.2010.02383.x.
  5. ^ a b c d e f g h i j k Checkley, David M.; Barth, John A. (2009). "Patterns and processes in the California Current System". Progress in Oceanography. 83: 49–64. doi:10.1016/j.pocean.2009.07.028.
  6. ^ a b Firme, Giselle F.; Rue, Eden L.; Weeks, Debra A.; Bruland, Kenneth W.; Hutchins, David A. (2003). "Spatial and temporal variability in phytoplankton iron limitation along the California coast and consequences for Si, N, and C biogeochemistry". Global Biogeochemical Cycles. 17 (1). doi:10.1029/2001GB001824.
  7. ^ a b Reese, D. C.; Brodeur, R. D. (2006-02-01). "Identifying and characterizing biological hotspots in the northern California Current". Deep Sea Research Part II: Topical Studies in Oceanography. Top predator "hot spots" in the North Pacific. 53 (3): 291–314. doi:10.1016/j.dsr2.2006.01.014. ISSN 0967-0645.
  8. ^ a b c U.S. Department of Commerce, NOAA, NMFS (2017). Fisheries Economics of the United States 2015. NOAA.{{cite book}}: CS1 maint: multiple names: authors list (link)
  9. ^ a b Sydeman, WJ; Mills, KL; Santora, JA; Thompson, SA; Bertram; Morgan, JH; Hipfner, JM; Wells, BK; Wolfe, SG (2009). "Seabirds and climate in the California Current-a synthesis of change". California Cooperative Oceanic Fisheries Investigation Reports. 50: 82–104.
  10. ^ a b c d e f Xiu, Peng; Chai, Fei; Curchitser, Enrique N.; Castruccio, Frederic S. (2018). "Future changes in coastal upwelling ecosystems with global warming: The case of the California Current System". Scientific Reports. 8 (1): 2866. doi:10.1038/s41598-018-21247-7. ISSN 2045-2322. PMC 5809506. PMID 29434297.{{cite journal}}: CS1 maint: PMC format (link)
  11. ^ a b Snyder, Mark A.; Sloan, Lisa C.; Diffenbaugh, Noah S.; Bell, Jason L. (2003). "Future climate change and upwelling in the California Current". Geophysical Research Letters. 30 (15). doi:10.1029/2003GL017647.
  12. ^ a b c d Chan, F.; Barth, J. A.; Lubchenco, J.; Kirincich, A.; Weeks, H.; Peterson, W. T.; Menge, B. A. (2008). "Emergence of Anoxia in the California Current Large Marine Ecosystem". Science. 319 (5865): 920–920. doi:10.1126/science.1149016. ISSN 0036-8075.
  13. ^ a b c Hauri, C., Gruber, N., Platnner, G., Alin, S., Feely, R., Hales, B., & Wheeler, P. (2009). Ocean acidification in the California Current system. Oceanography, 22(4), 60-71. Retrieved November 18, 2020, from http://www.jstor.org/stable/24861024
  14. ^ a b Smith, Jayme; Connell, Paige; Evans, Richard H.; Gellene, Alyssa G.; Howard, Meredith D. A.; Jones, Burton H.; Kaveggia, Susan; Palmer, Lauren; Schnetzer, Astrid; Seegers, Bridget N.; Seubert, Erica L. (2018-11-01). "A decade and a half of Pseudo-nitzschia spp. and domoic acid along the coast of southern California". Harmful Algae. Domoic acid 30 years on. 79: 87–104. doi:10.1016/j.hal.2018.07.007. ISSN 1568-9883.
  15. ^ a b c Asch, Rebecca G. (2015-07-28). "Climate change and decadal shifts in the phenology of larval fishes in the California Current ecosystem". Proceedings of the National Academy of Sciences. 112 (30): E4065–E4074. doi:10.1073/pnas.1421946112. ISSN 0027-8424. PMC 4522805. PMID 26159416.{{cite journal}}: CS1 maint: PMC format (link)
  16. ^ a b c d e f g h i j k l m n Bakun, A.; Black, B. A.; Bograd, S. J.; García-Reyes, M.; Miller, A. J.; Rykaczewski, R. R.; Sydeman, W. J. (2015). "Anticipated Effects of Climate Change on Coastal Upwelling Ecosystems". Current Climate Change Reports. 1 (2): 85–93. doi:10.1007/s40641-015-0008-4. ISSN 2198-6061.
  17. ^ Huyer, A.; Barth, J.A.; Kosro, P.M.; Shearman, R.K.; Smith, R.L. (1998). "Upper-ocean water mass characteristics of the California current, Summer 1993". Deep Sea Research Part II: Topical Studies in Oceanography. 45 (8–9): 1411–1442. doi:10.1016/s0967-0645(98)80002-7. ISSN 0967-0645.
  18. ^ a b c d e f Hickey, Barbara M. (1979). "The California Current System -- hypotheses and facts". Progress in Oceanography. 8: 191–279.
  19. ^ Mann, K. H. (2006). Dynamics of marine ecosystems : biological-physical interactions in the oceans. Lazier, J. R. N. (3rd ed ed.). Malden, MA: Blackwell Pub. ISBN 1-4051-1118-6. OCLC 57675858. {{cite book}}: |edition= has extra text (help)
  20. ^ Koračin, Darko; Dorman, Clive E.; Lewis, John M.; Hudson, James G.; Wilcox, Eric M.; Torregrosa, Alicia (2014). "Marine fog: A review". Atmospheric Research. 143: 142–175. doi:10.1016/j.atmosres.2013.12.012.
  21. ^ Hickey, B.M.; Royer, T.C.; Amos, C.M. (2019), "California and Alaska Currents", Encyclopedia of Ocean Sciences, vol. 3, Academic Press, pp. 318–329, ISBN 978-0-12-813082-7
  22. ^ a b Huyer, Adriana (1983). "Coastal upwelling in the California current system". Progress in Oceanography. 12 (3): 259–284. doi:10.1016/0079-6611(83)90010-1. ISSN 0079-6611.
  23. ^ Bakun, Andrew (1973). "Coastal Upwelling Indices, West Coast of North America, 1946-71". NOAA Technical Report. NMFS SSRF-671: 1–96.
  24. ^ Jacox, Michael G; Edwards, Christopher A; Hazen, Elliott A; Bograd, Steven J (2018). "Coastal Upwelling Revisited: Ekman, Bakun, and Improved Upwelling Indices for the U.S. West Coast". JGR Oceans. 123: 7332–7350.
  25. ^ Strub, P. Ted; James, Corinne (1988). "Atmospheric conditions during the spring and fall transitions in the coastal ocean off western United States". Journal of Geophysical Research: Oceans. 93 (C12): 15561–15584. doi:10.1029/JC093iC12p15561. ISSN 2156-2202.
  26. ^ Barth, J. A.; Menge, B. A.; Lubchenco, J.; Chan, F.; Bane, J. M.; Kirincich, A. R.; McManus, M. A.; Nielsen, K. J.; Pierce, S. D.; Washburn, L. (2007-03-06). "Delayed upwelling alters nearshore coastal ocean ecosystems in the northern California current". Proceedings of the National Academy of Sciences. 104 (10): 3719–3724. doi:10.1073/pnas.0700462104. ISSN 0027-8424. PMC 1805484. PMID 17360419.{{cite journal}}: CS1 maint: PMC format (link)
  27. ^ a b Chavez, Francisco P. (1996). "Forcing and biological impact of onset of the 1992 El Niño in central California". Geophysical Research Letters. 23 (3): 265–268. doi:10.1029/96GL00017.
  28. ^ a b c Huyer, A.; Smith, R.L.; Fleischbein, J. (2002). "The coastal ocean off Oregon and northern California during the 1997–8 El Niño". Progress in Oceanography. 54 (1–4): 311–341. doi:10.1016/S0079-6611(02)00056-3.
  29. ^ a b Jacox, Michael G.; Hazen, Elliott L.; Zaba, Katherine D.; Rudnick, Daniel L.; Edwards, Christopher A.; Moore, Andrew M.; Bograd, Steven J. (2016). "Impacts of the 2015-2016 El Niño on the California Current System: Early assessment and comparison to past events: 2015-2016 El Niño Impact in the CCS". Geophysical Research Letters. 43 (13): 7072–7080. doi:10.1002/2016GL069716.
  30. ^ Bograd, Steven J.; Lynn, Ronald J. (2001-01-15). "Physical-biological coupling in the California Current during the 1997-99 El Niño-La Niña Cycle". Geophysical Research Letters. 28 (2): 275–278. doi:10.1029/2000GL012047.
  31. ^ Lynn, R.J; Bograd, S.J (2002). "Dynamic evolution of the 1997–1999 El Niño–La Niña cycle in the southern California Current System". Progress in Oceanography. 54 (1–4): 59–75. doi:10.1016/S0079-6611(02)00043-5.
  32. ^ Schwing, F.B.; Murphree, T.; deWitt, L.; Green, P.M. (2002). "The evolution of oceanic and atmospheric anomalies in the northeast Pacific during the El Niño and La Niña events of 1995–2001". Progress in Oceanography. 54 (1–4): 459–491. doi:10.1016/S0079-6611(02)00064-2.
  33. ^ Di Lorenzo, Emanuele; Combes, Vincent; Keister, Julie E; Strub, P. Ted; Thomas, Andrew C; Franks, Peter J.S.; Ohman, Mark D; Furtado, Jason C.; Bracco, Annalisa; Bograd, Steven J.; Peterson, William T. (2013). "Synthesis of Pacific Ocean Climate and Ecosystem Dynamics". Oceanography. 26 (4): 68–81. ISSN 1042-8275.
  34. ^ Lorenzo, E. Di; Schneider, N.; Cobb, K. M.; Franks, P. J. S.; Chhak, K.; Miller, A. J.; McWilliams, J. C.; Bograd, S. J.; Arango, H.; Curchitser, E.; Powell, T. M. (2008). "North Pacific Gyre Oscillation links ocean climate and ecosystem change". Geophysical Research Letters. 35 (8). doi:10.1029/2007GL032838. ISSN 1944-8007.
  35. ^ a b Capone, Douglas G.; Hutchins, David A. (2013). "Microbial biogeochemistry of coastal upwelling regimes in a changing ocean". Nature Geoscience. 6 (9): 711–717. doi:10.1038/ngeo1916. ISSN 1752-0908.
  36. ^ a b c Chan, Francis; Barth, John; Kroeker, Kristy; Lubchenco, Jane; Menge, Bruce (2019). "The Dynamics and Impact of Ocean Acidification and Hypoxia: Insights from Sustained Investigations in the Northern California Current Large Marine Ecosystem". Oceanography. 32 (3): 62–71. doi:10.5670/oceanog.2019.312. ISSN 1042-8275.
  37. ^ a b Connolly, T. P.; Hickey, B. M.; Geier, S. L.; Cochlan, W. P. (2010). "Processes influencing seasonal hypoxia in the northern California Current System". Journal of Geophysical Research. 115 (C3): C03021. doi:10.1029/2009JC005283. ISSN 0148-0227. PMC 2867361. PMID 20463844.{{cite journal}}: CS1 maint: PMC format (link)
  38. ^ a b Ryan, John P.; Fischer, Andrew M.; Kudela, Raphael M.; Gower, James F. R.; King, Stephanie A.; Marin, Roman; Chavez, Francisco P. (2009). "Influences of upwelling and downwelling winds on red tide bloom dynamics in Monterey Bay, California". Continental Shelf Research. 29 (5): 785–795. doi:10.1016/j.csr.2008.11.006. ISSN 0278-4343.
  39. ^ a b Mackey, Katherine R. M.; Chien, Chia-Te; Paytan, Adina (2014). "Microbial and biogeochemical responses to projected future nitrate enrichment in the California upwelling system". Frontiers in Microbiology. 5. doi:10.3389/fmicb.2014.00632. ISSN 1664-302X. PMC 4238378. PMID 25477873.{{cite journal}}: CS1 maint: PMC format (link) CS1 maint: unflagged free DOI (link)
  40. ^ a b Hickey, Barbara; Banas, Neil (2008-12-01). "Why is the Northern End of the California Current System So Productive?". Oceanography. 21 (4): 90–107. doi:10.5670/oceanog.2008.07. ISSN 1042-8275.
  41. ^ USGS (2020). "USGS 14144700 Columbia River at Vancouver, WA". National water Information System.{{cite web}}: CS1 maint: url-status (link)
  42. ^ Simenstad, Charles A.; Small, Lawrence F.; David McIntire, C.; Jay, David A.; Sherwood, Christopher (1990). "Columbia river estuary studies: An introduction to the estuary, a brief history, and prior studies". Progress in Oceanography. 25 (1): 1–13. doi:10.1016/0079-6611(90)90002-J. ISSN 0079-6611.
  43. ^ Pennington, Timothy J.; Chavez, Francisco P. (2000). "Seasonal fluctuations of temperature, salinity, nitrate, chlorophyll and primary production at station H3/M1 over 1989–1996 in Monterey Bay, California". Deep Sea Research Part II: Topical Studies in Oceanography. 47 (5–6): 947–973. doi:10.1016/S0967-0645(99)00132-0.
  44. ^ Landry, M.R.; Postel, J.R.; Peterson, W.K.; Newman, J. (1989), "Broad-scale distributional patterns of hydrographic variables on the Washington/Oregon shelf", Coastal Oceanography of Washington and Oregon, vol. 47, Elsevier, pp. 1–40, ISBN 0-444-87308-2
  45. ^ Karstensen, Johannes; Stramma, Lothar; Visbeck, Martin (2008). "Oxygen minimum zones in the eastern tropical Atlantic and Pacific oceans". Progress in Oceanography. 77 (4): 331–350. doi:10.1016/j.pocean.2007.05.009.
  46. ^ Testa, J. M.; Kemp, W. M. (2011), Wolanski, Eric; McLusky, Donald (eds.), "5.05 - Oxygen – Dynamics and Biogeochemical Consequences", Treatise on Estuarine and Coastal Science, Waltham: Academic Press, pp. 163–199, doi:10.1016/b978-0-12-374711-2.00505-2, ISBN 978-0-08-087885-0, retrieved 2020-11-04
  47. ^ Keller, Aa; Ciannelli, L; Wakefield, Ww; Simon, V; Barth, Ja; Pierce, Sd (2017-03-24). "Species-specific responses of demersal fishes to near-bottom oxygen levels within the California Current large marine ecosystem". Marine Ecology Progress Series. 568: 151–173. doi:10.3354/meps12066. ISSN 0171-8630.
  48. ^ Pozo Buil, Mercedes; Di Lorenzo, Emanuele (2017). "Decadal dynamics and predictability of oxygen and subsurface tracers in the California Current System". Geophysical Research Letters. 44 (9): 4204–4213. doi:10.1002/2017GL072931.
  49. ^ a b c Gruber, N.; Hauri, C.; Lachkar, Z.; Loher, D.; Frolicher, T. L.; Plattner, G.-K. (2012). "Rapid Progression of Ocean Acidification in the California Current System". Science. 337 (6091): 220–223. doi:10.1126/science.1216773. ISSN 0036-8075.
  50. ^ Feely, Richard A.; Okazaki, Remy R.; Cai, Wei-Jun; Bednaršek, Nina; Alin, Simone R.; Byrne, Robert H.; Fassbender, Andrea (2018). "The combined effects of acidification and hypoxia on pH and aragonite saturation in the coastal waters of the California current ecosystem and the northern Gulf of Mexico". Continental Shelf Research. 152: 50–60. doi:10.1016/j.csr.2017.11.002. ISSN 0278-4343.
  51. ^ Waldbusser, George G.; Salisbury, Joseph E. (2014-01-03). "Ocean Acidification in the Coastal Zone from an Organism's Perspective: Multiple System Parameters, Frequency Domains, and Habitats". Annual Review of Marine Science. 6 (1): 221–247. doi:10.1146/annurev-marine-121211-172238. ISSN 1941-1405.
  52. ^ Hales, Burke; Takahashi, Taro; Bandstra, Leah (2005). "Atmospheric CO2 uptake by a coastal upwelling system". Global Biogeochemical Cycles. 19 (1). doi:10.1029/2004GB002295. ISSN 1944-9224.
  53. ^ Harvey, Christopher James; Garfield, Newell; Williams, Gregory D. (Gregory Dean), 1967-; Tolimieri, Nick; Schroeder, Isaac; Andrews, Kelly S.; Barnas, Katie; Bjorkstedt, Eric Peter; Bograd, Steven J., 1963-; Brodeur, Richard D.; Burke, Brian J. (Brian Joseph) (2019). "Ecosystem Status Report of the California Current for 2019: A Summary of Ecosystem Indicators Compiled by the California Current Integrated Ecosystem Assessment Team (CCIEA)". doi:10.25923/p0ed-ke21. {{cite journal}}: Cite journal requires |journal= (help)CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  54. ^ "Subduction of the Juan de Fuca Plate beneath the North American Pla..." www.usgs.gov. Retrieved 2020-11-16.
  55. ^ Bray, N. A.; Keyes, A.; Morawitz, W. M. L. (1999). "The California Current system in the Southern California Bight and the Santa Barbara Channel". Journal of Geophysical Research: Oceans. 104 (C4): 7695–7714. doi:10.1029/1998jc900038. ISSN 0148-0227.
  56. ^ Atwater, T. (1998). "Plate Tectonic History of Southern California with Emphasis on the Western Transverse Ranges and Northern Channel Islands". doi:10.32375/1998-mp45.1. {{cite journal}}: Cite journal requires |journal= (help)
  57. ^ a b Leonard, L. J.; Currie, C. A.; Mazzotti, S.; Hyndman, R. D. (2010). "Rupture area and displacement of past Cascadia great earthquakes from coastal coseismic subsidence". Geological Society of America Bulletin. 122 (11–12): 2079–2096. doi:10.1130/B30108.1. ISSN 0016-7606.
  58. ^ Muhs, Daniel R.; Simmons, Kathleen R.; Schumann, R. Randall; Groves, Lindsey T.; DeVogel, Stephen B.; Minor, Scott A.; Laurel, DeAnna (2014). "Coastal tectonics on the eastern margin of the Pacific Rim: late Quaternary sea-level history and uplift rates, Channel Islands National Park, California, USA". Quaternary Science Reviews. 105: 209–238. doi:10.1016/j.quascirev.2014.09.017. ISSN 0277-3791.
  59. ^ Barth, John A; Pierce, Stephen D; Smith, Robert L (2000-05-01). "A separating coastal upwelling jet at Cape Blanco, Oregon and its connection to the California Current System". Deep Sea Research Part II: Topical Studies in Oceanography. 47 (5): 783–810. doi:10.1016/S0967-0645(99)00127-7. ISSN 0967-0645.
  60. ^ Barth, J. A.; Pierce, S. D.; Castelao, R. M. (2005). "Time-dependent, wind-driven flow over a shallow midshelf submarine bank". Journal of Geophysical Research. 110 (C10): C10S05. doi:10.1029/2004JC002761. ISSN 0148-0227.
  61. ^ Inman, Douglas L. (2005), Schwartz, Maurice L. (ed.), "Littoral Cells", Encyclopedia of Coastal Science, Dordrecht: Springer Netherlands, pp. 594–599, doi:10.1007/1-4020-3880-1_196, ISBN 978-1-4020-3880-8, retrieved 2020-11-16
  62. ^ Gardner, James V.; Dean, Walter E.; Dartnell, Peter (1997). "Biogenic sedimentation beneath the California Current System for the past 30 kyr and its paleoceanographic significance". Paleoceanography. 12 (2): 207–225. doi:10.1029/96pa03567. ISSN 0883-8305.
  63. ^ Gorsline, D. S. (1981). "Fine Sediment Transport and Deposition in Active Margin Basins": 39–59. {{cite journal}}: Cite journal requires |journal= (help)
  64. ^ George, D.A.; Largier, J.L.; Storlazzi, C.D.; Barnard, P.L. (2015). "Classification of rocky headlands in California with relevance to littoral cell boundary delineation". Marine Geology. 369: 137–152. doi:10.1016/j.margeo.2015.08.010. ISSN 0025-3227.
  65. ^ Drake, David E.; Cacchione, David A. (1985). "Seasonal variation in sediment transport on the Russian River shelf, California". Continental Shelf Research. 4 (5): 495–514. doi:10.1016/0278-4343(85)90007-X. ISSN 0278-4343.
  66. ^ Wolf, S.C. (1970). "Coastal currents and mass transport of surface sediments over the shelf regions of Monterey Bay, California". Marine Geology. 8 (5): 16.
  67. ^ Gardner, James V.; Dean, Walter E.; Dartnell, Peter (1997). "Biogenic sedimentation beneath the California Current System for the past 30 kyr and its paleoceanographic significance". Paleoceanography. 12 (2): 207–225. doi:10.1029/96PA03567.
  68. ^ Ryther, John H. (1969). "Photosynthesis and Fish Production in the Sea". Science. 166 (3901): 72–76. ISSN 0036-8075.
  69. ^ a b Kolody, B. C.; McCrow, J. P.; Allen, L. Zeigler; Aylward, F. O.; Fontanez, K. M.; Moustafa, A.; Moniruzzaman, M.; Chavez, F. P.; Scholin, C. A.; Allen, E. E.; Worden, A. Z. (2019). "Diel transcriptional response of a California Current plankton microbiome to light, low iron, and enduring viral infection". The ISME Journal. 13 (11): 2817–2833. doi:10.1038/s41396-019-0472-2. ISSN 1751-7370.
  70. ^ Fisheries, NOAA (2020-11-09). "Blue Whale | NOAA Fisheries". NOAA. Retrieved 2020-11-30.
  71. ^ a b McGowan, John A; Bograd, Steven J; Lynn, Ronald J; Miller, Arthur J (2003). "The biological response to the 1977 regime shift in the California Current". Deep Sea Research Part II: Topical Studies in Oceanography. CalCOFI: A Half Century of Physical, Chemical and Biological Research in the California Current System. 50 (14): 2567–2582. doi:10.1016/S0967-0645(03)00135-8. ISSN 0967-0645.
  72. ^ a b Harvey, Chris; Garfield, Newell (Toby); Williams, Gregory; Tolimieri, Nick; Andrews, Kelly; Barnas, Katie; Bjorkstedt, Eric; Bograd, Steven; Borchert, Jerry; Braby, Caren; Brodeur, Richard (2020). "Ecosystem Status Report of the California Current for 2019-20: A Summary of Ecosystem Indicators Compiled by the California Current Integrated Ecosystem Assessment Team (CCIEA)". doi:10.25923/e5rb-9f55. {{cite journal}}: Cite journal requires |journal= (help)
  73. ^ Chavez, F. P. (2003-01-10). "From Anchovies to Sardines and Back: Multidecadal Change in the Pacific Ocean". Science. 299 (5604): 217–221. doi:10.1126/science.1075880.
  74. ^ a b c d e Heine, J., & Checkley, D. (Eds.). (2014). CalCOFI (No. 55; Fisheries Review). CalCOFI.https://www.calcofi.org/publications/calcofireports/v55/Vol_55_CalCOFIReport.pdf
  75. ^ Acevedo-Trejos, Esteban; Brandt, Gunnar; Bruggeman, Jorn; Merico, Agostino (2015). "Mechanisms shaping size structure and functional diversity of phytoplankton communities in the ocean". Scientific Reports. 5 (1): 8918. doi:10.1038/srep08918. ISSN 2045-2322.
  76. ^ Santora, Jarrod A.; Hazen, Elliott L.; Schroeder, Isaac D.; Bograd, Steven J.; Sakuma, Keith M.; Field, John C. (2017). "Impacts of ocean climate variability on biodiversity of pelagic forage species in an upwelling ecosystem". Marine Ecology Progress Series. 580: 205–220. doi:10.3354/meps12278. ISSN 0171-8630.
  77. ^ Hewitt, R. (1981). Eddies and Speciation in the California Current. CalCOFI Report, 12, 3.
  78. ^ Drake, Patrick T.; Edwards, Christopher A.; Morgan, Steven G.; Dever, Edward P. (2013-07-01). "Influence of larval behavior on transport and population connectivity in a realistic simulation of the California Current System". Journal of Marine Research. 71 (4): 317–350. doi:10.1357/002224013808877099.
  79. ^ Morgan, Steven G. (2014). "Behaviorally Mediated Larval Transport in Upwelling Systems". Advances in Oceanography. 2014: 1–17. doi:10.1155/2014/364214. ISSN 2356-6809.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  80. ^ Miner, C. Melissa; Burnaford, Jennifer L.; Ambrose, Richard F.; Antrim, Liam; Bohlmann, Heath; Blanchette, Carol A.; Engle, John M.; Fradkin, Steven C.; Gaddam, Rani; Harley, Christopher D. G.; Miner, Benjamin G. (2018). "Large-scale impacts of sea star wasting disease (SSWD) on intertidal sea stars and implications for recovery". PLOS ONE. 13 (3): e0192870. doi:10.1371/journal.pone.0192870. ISSN 1932-6203. PMC 5860697. PMID 29558484.{{cite journal}}: CS1 maint: PMC format (link) CS1 maint: unflagged free DOI (link)
  81. ^ "Pacific Shell Institute".{{cite web}}: CS1 maint: url-status (link)
  82. ^ Pounds and Values of Commercially Caught Fish and Shellfish Landed in Oregon (FTR 205). Oregon Department of Fish and Wildlife. https://www.dfw.state.or.us/fish/commercial/landing_stats/2019/10%20year%20(2019).pdf
  83. ^ a b Peterson, B., Emmett, R., Ralston, S., Forney, K. A., Road, S., Cruz, S., Benson, S., Road, S., & Landing, M. (2006). The State of the California Current, 2005–2006: Warm in the North, Cool in the South. CalCOFI Report, 47, 46.
  84. ^ Koehn, Laura E.; Essington, Timothy E.; Marshall, Kristin N.; Kaplan, Isaac C.; Sydeman, William J.; Szoboszlai, Amber I.; Thayer, Julie A. (2016). "Developing a high taxonomic resolution food web model to assess the functional role of forage fish in the California Current ecosystem". Ecological Modelling. 335: 87–100. doi:10.1016/j.ecolmodel.2016.05.010. ISSN 0304-3800.
  85. ^ Carlisle, Aaron B.; Litvin, Steven Y.; Hazen, Elliott L.; Madigan, Daniel J.; Goldman, Kenneth J.; Lea, Robert N.; Block, Barbara A. (2015). "Reconstructing habitat use by juvenile salmon sharks links upwelling to strandings in the California Current". Marine Ecology Progress Series. 525: 217–228. doi:10.3354/meps11183. ISSN 0171-8630.
  86. ^ Shaffer, S. A.; Tremblay, Y.; Weimerskirch, H.; Scott, D.; Thompson, D. R.; Sagar, P. M.; Moller, H.; Taylor, G. A.; Foley, D. G.; Block, B. A.; Costa, D. P. (2006-08-14). "Migratory shearwaters integrate oceanic resources across the Pacific Ocean in an endless summer". Proceedings of the National Academy of Sciences. 103 (34): 12799–12802. doi:10.1073/pnas.0603715103. ISSN 0027-8424.
  87. ^ Barlow, Jay; Forney, Karin (2007). "Abundance and density of cetaceans in the California Current ecosystem". NOAA Southwest Fisheries Science Center Bulletin. 105: 509–526.
  88. ^ Benson, Scott R; Croll, Donald A; Marinovic, Baldo B; Chavez, Francisco P; Harvey, James T (2002-07-01). "Changes in the cetacean assemblage of a coastal upwelling ecosystem during El Niño 1997–98 and La Niña 1999". Progress in Oceanography. Observations of the 1997-98 El Nino along the West Coast of North America. 54 (1): 279–291. doi:10.1016/S0079-6611(02)00054-X. ISSN 0079-6611.
  89. ^ NOAA Fisheries (2020). "Marine Mammals on the West Coast | NOAA Fisheries". NOAA. Retrieved 2020-11-16.{{cite web}}: CS1 maint: url-status (link)
  90. ^ Fisheries, NOAA (2020-11-18). "Humpback Whale | NOAA Fisheries". NOAA. Retrieved 2020-11-30.
  91. ^ Antonelis, G.; Fiscus, C. (1980). "Pinnipeds of the California Current". California Cooperative Oceanic Fisheries Investigations.
  92. ^ Veit, Richard R.; Harrison, Nancy M. (2017). "Positive Interactions among Foraging Seabirds, Marine Mammals and Fishes and Implications for Their Conservation". Frontiers in Ecology and Evolution. 5. doi:10.3389/fevo.2017.00121. ISSN 2296-701X.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  93. ^ Holen, Steven R.; Deméré, Thomas A.; Fisher, Daniel C.; Fullagar, Richard; Paces, James B.; Jefferson, George T.; Beeton, Jared M.; Cerutti, Richard A.; Rountrey, Adam N.; Vescera, Lawrence; Holen, Kathleen A. (2017). "A 130,000-year-old archaeological site in southern California, USA". Nature. 544 (7651): 479–483. doi:10.1038/nature22065. ISSN 1476-4687.
  94. ^ Gobalet, Kenneth W. (1992). "Inland Utilization of Marine Fishes by Native Americans along the Central California Coast". Journal of California and Great Basin Anthropology. 14 (1): 72–84. ISSN 0191-3557.
  95. ^ Fisheries, NOAA (2020-11-25). "West Coast | NOAA Fisheries". NOAA. Retrieved 2020-11-28.
  96. ^ a b McClatchie, Sam (2013), "Oceanography of the Southern California Current System Relevant to Fisheries", Regional Fisheries Oceanography of the California Current System, Dordrecht: Springer Netherlands, pp. 13–60, ISBN 978-94-007-7222-9, retrieved 2020-11-16
  97. ^ Pacific Fisheries Management Council. "Fishery management plans". Pacific Fishery Management Council. Retrieved 2020-11-16.{{cite web}}: CS1 maint: url-status (link)
  98. ^ Field, John C.; Francis, Robert C. (2006). "Considering ecosystem-based fisheries management in the California Current". Marine Policy. 30 (5): 552–569. doi:10.1016/j.marpol.2005.07.004. ISSN 0308-597X.
  99. ^ Fisheries, NOAA (2020-11-03). "California Market Squid | NOAA Fisheries". NOAA. Retrieved 2020-11-16.
  100. ^ a b Fisheries, NOAA (2020). "Pacific Sardine | NOAA Fisheries". NOAA. Retrieved 2020-11-16.{{cite web}}: CS1 maint: url-status (link)
  101. ^ "Pacific Sardine | California Sea Grant". caseagrant.ucsd.edu. Retrieved 2020-11-16.
  102. ^ "California Current - Salmon | Integrated Ecosystem Assessment". www.integratedecosystemassessment.noaa.gov. Retrieved 2020-11-09.
  103. ^ National Research Council (2002). "Effects of Trawling and Dredging on Seafloor Habitat". doi:10.17226/10323. {{cite journal}}: Cite journal requires |journal= (help)
  104. ^ Molnar, Jennifer L.; Gamboa, Rebecca L.; Revenga, Carmen; Spalding, Mark D. (2008). "Assessing the global threat of invasive species to marine biodiversity". Frontiers in Ecology and the Environment. 6 (9): 485–492. doi:10.1890/070064. ISSN 1540-9309.
  105. ^ Cordes, Erik E.; Jones, Daniel O. B.; Schlacher, Thomas A.; Amon, Diva J.; Bernardino, Angelo F.; Brooke, Sandra; Carney, Robert; DeLeo, Danielle M.; Dunlop, Katherine M.; Escobar-Briones, Elva G.; Gates, Andrew R. (2016). "Environmental Impacts of the Deep-Water Oil and Gas Industry: A Review to Guide Management Strategies". Frontiers in Environmental Science. 4. doi:10.3389/fenvs.2016.00058. ISSN 2296-665X.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  106. ^ Sullivan, Florence A.; Torres, Leigh G. (2018). "Assessment of vessel disturbance to gray whales to inform sustainable ecotourism". The Journal of Wildlife Management. 82 (5): 896–905. doi:10.1002/jwmg.21462. ISSN 1937-2817.
  107. ^ Cooke, T (2007). "How Will Climate Change Affect the California Current Upwelling?". Eos. Retrieved 2020-11-17.{{cite web}}: CS1 maint: url-status (link)
  108. ^ García-Reyes, Marisol; Sydeman, William J.; Schoeman, David S.; Rykaczewski, Ryan R.; Black, Bryan A.; Smit, Albertus J.; Bograd, Steven J. (2015). "Under Pressure: Climate Change, Upwelling, and Eastern Boundary Upwelling Ecosystems". Frontiers in Marine Science. 2. doi:10.3389/fmars.2015.00109. ISSN 2296-7745.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  109. ^ a b c d Doney, Scott C.; Fabry, Victoria J.; Feely, Richard A.; Kleypas, Joan A. (2009). "Ocean Acidification: The Other CO2 Problem". Annual Review of Marine Science. 1 (1): 169–192. doi:10.1146/annurev.marine.010908.163834. ISSN 1941-1405.
  110. ^ Lu, Jian; Vecchi, Gabriel A.; Reichler, Thomas (2007). "Expansion of the Hadley cell under global warming". Geophysical Research Letters. 34 (6): L06805. doi:10.1029/2006GL028443. ISSN 0094-8276.
  111. ^ a b Sydeman, W. J.; García-Reyes, M.; Schoeman, D. S.; Rykaczewski, R. R.; Thompson, S. A.; Black, B. A.; Bograd, S. J. (2014). "Climate change and wind intensification in coastal upwelling ecosystems". Science. 345 (6192): 77–80. doi:10.1126/science.1251635. ISSN 0036-8075.
  112. ^ Bakun, Arthur (1990). "Global climate change and intensification of coastal ocean upwelling". Science. 247: 198–201.
  113. ^ Palacios, Daniel M.; Bograd, Steven J.; Mendelssohn, Roy; Schwing, Franklin B. (2004). "Long-term and seasonal trends in stratification in the California Current, 1950–1993". Journal of Geophysical Research: Oceans. 109 (C10). doi:10.1029/2004JC002380. ISSN 2156-2202.
  114. ^ a b Di Lorenzo, Emanuele; Miller, Arthur (2004). "The warming of the California Current System: Dynamics and Ecosystem Implications". Journal of Physical Oceanography. 35: 336–362.
  115. ^ a b Bograd, Steven J.; Castro, Carmen G.; Di Lorenzo, Emanuele; Palacios, Daniel M.; Bailey, Helen; Gilly, William; Chavez, Francisco P. (2008). "Oxygen declines and the shoaling of the hypoxic boundary in the California Current: Hypoxia in the California Current". Geophysical Research Letters. 35 (12): n/a–n/a. doi:10.1029/2008GL034185.
  116. ^ Gruber, Nicolas (2011). "Warming up, turning sour, losing breath: ocean biogeochemistry under global change". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 369 (1943): 1980–1996. doi:10.1098/rsta.2011.0003. ISSN 1364-503X.
  117. ^ Feely, R. A.; Sabine, C. L.; Hernandez-Ayon, J. M.; Ianson, D.; Hales, B. (2008). "Evidence for Upwelling of Corrosive "Acidified" Water onto the Continental Shelf". Science. 320 (5882): 1490–1492. doi:10.1126/science.1155676. ISSN 0036-8075.
  118. ^ McCabe, Ryan M.; Hickey, Barbara M.; Kudela, Raphael M.; Lefebvre, Kathi A.; Adams, Nicolaus G.; Bill, Brian D.; Gulland, Frances M. D.; Thomson, Richard E.; Cochlan, William P.; Trainer, Vera L. (2016-10-16). "An unprecedented coastwide toxic algal bloom linked to anomalous ocean conditions". Geophysical Research Letters. 43 (19). doi:10.1002/2016GL070023. ISSN 0094-8276. PMC 5129552. PMID 27917011.{{cite journal}}: CS1 maint: PMC format (link)
  119. ^ a b Grémillet, David; Lewis, Sue; Drapeau, Laurent; van Der Lingen, Carl D.; Huggett, Jenny A.; Coetzee, Janet C.; Verheye, Hans M.; Daunt, Francis; Wanless, Sarah; Ryan, Peter G. (2008). "Spatial match-mismatch in the Benguela upwelling zone: should we expect chlorophyll and sea-surface temperature to predict marine predator distributions?". Journal of Applied Ecology. 45 (2): 610–621. doi:10.1111/j.1365-2664.2007.01447.x.
  120. ^ a b King, Jacquelynne R.; Agostini, Vera N.; Harvey, Christopher J.; McFarlane, Gordon A.; Foreman, Michael G. G.; Overland, James E.; Di Lorenzo, Emanuele; Bond, Nicholas A.; Aydin, Kerim Y. (2011-07-01). "Climate forcing and the California Current ecosystem". ICES Journal of Marine Science. 68 (6): 1199–1216. doi:10.1093/icesjms/fsr009. ISSN 1054-3139.
  121. ^ a b c Sydeman, Wj; Thompson, Sa; Kitaysky, A (2012). "Seabirds and climate change: roadmap for the future". Marine Ecology Progress Series. 454: 107–117. doi:10.3354/meps09806. ISSN 0171-8630.
  122. ^ Cury, Philippe; Roy, Claude (2011-04-11). "Optimal Environmental Window and Pelagic Fish Recruitment Success in Upwelling Areas". Canadian Journal of Fisheries and Aquatic Sciences. doi:10.1139/f89-086.
  123. ^ Suchman, Cynthia L.; Brodeur, Richard D.; Daly, Elizabeth A.; Emmett, Robert L. (2012). "Large medusae in surface waters of the Northern California Current: variability in relation to environmental conditions". Hydrobiologia. 690 (1): 113–125. doi:10.1007/s10750-012-1055-7. ISSN 0018-8158.
  124. ^ Frawley, Timothy H.; Muhling, Barbara A.; Brodie, Stephanie; Fisher, Mary C.; Tommasi, Desiree; Le Fol, Gwendal; Hazen, Elliott L.; Stohs, Stephen S.; Finkbeiner, Elena M.; Jacox, Michael G. (2020-11-18). "Changes to the structure and function of an albacore fishery reveal shifting social‐ecological realities for Pacific Northwest fishermen". Fish and Fisheries. doi:10.1111/faf.12519. ISSN 1467-2960.
  125. ^ Smith, James A.; Tommasi, Desiree; Sweeney, Jonathan; Brodie, Stephanie; Welch, Heather; Hazen, Elliott L.; Muhling, Barbara; Stohs, Steven M.; Jacox, Michael G. (2020-01-31). "Lost opportunity: Quantifying the dynamic economic impact of time‐area fishery closures". Journal of Applied Ecology. 57 (3): 502–513. doi:10.1111/1365-2664.13565. ISSN 0021-8901.
  126. ^ Impacts of climate change on fisheries and aquaculture : synthesis of current knowledge, adaptation and mitigation options. Barange, Manuel, 1961-, Bahri, Tarûb,, Beveridge, Malcolm C. M.,, Cochrane, K. L.,, Funge Smith, S. (Simon),, Poulain, Florence,. Rome. ISBN 92-5-130607-9. OCLC 1078885208.{{cite book}}: CS1 maint: extra punctuation (link) CS1 maint: others (link)