Skip to main content

REVIEW article

Front. Mar. Sci., 08 August 2019
Sec. Ocean Observation
Volume 6 - 2019 | https://doi.org/10.3389/fmars.2019.00423

Global Perspectives on Observing Ocean Boundary Current Systems

  • 1Woods Hole Oceanographic Institution, Woods Hole, MA, United States
  • 2Monterey Bay Aquarium Research Institute, Moss Landing, CA, United States
  • 3Department of Ocean, Earth, and Atmospheric Sciences, Old Dominion University, Norfolk, VA, United States
  • 4LEGOS, IRD, CNES, CNRS, UPS, Universite de Toulouse, Toulouse, France
  • 5NOAA’s Atlantic Oceanographic and Meteorological Laboratory, Miami, FL, United States
  • 6Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, FL, United States
  • 7Instituto del Mar del Peru, Lima, Peru
  • 8Qingdao National Laboratory for Marine Science and Technology, Ocean University of China, Qingdao, China
  • 9Scripps Institution of Oceanography, University of California, San Diego, La Jolla, CA, United States
  • 10Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA, United States
  • 11Instituto de Oceanografía y Cambio Global, Universidad de Las Palmas de Gran Canaria, Las Palmas, Spain
  • 12European Centre for Medium-Range Weather Forecasts, Reading, United Kingdom
  • 13Department of Marine Sciences, The University of North Carolina at Chapel Hill, Chapel Hill, NC, United States
  • 14College of Earth, Ocean, and Atmospheric Sciences, Oregon State University, Corvallis, OR, United States
  • 15GEOMAR Helmholtz Centre for Ocean Research Kiel, Kiel, Germany
  • 16Faculty of Mathematics and Natural Sciences, Kiel University, Kiel, Germany
  • 17Institute of Coastal Research, Helmholtz-Zentrum Geesthacht, Geesthacht, Germany
  • 18Departamento de Oceanografia Física, Química e Geológica University of São Paulo, São Paulo, Brazil
  • 19Consejo Nacional de Investigaciones Científicas y Técnicas, Servicio de Hidrografía Naval, Buenos Aires, Argentina
  • 20Department of Meteorology, Institute of Geosciences, Federal University of Rio de Janeiro (UFRJ), Rio de Janeiro, Brazil
  • 21NOAA Pacific Marine Environmental Laboratory, Seattle, WA, United States
  • 22Deptartment of Environmental Science, Rutgers University, New Brunswick, NJ, United States
  • 23Department of Physics and Physical Oceanography, Memorial University of Newfoundland, St. John’s, NL, Canada
  • 24Instituto Milenio de Oceanografía, Universidad de Concepción, Concepción, Chile
  • 25Department of Oceanography, University of Cape Town, Cape Town, South Africa
  • 26CSIRO Oceans and Atmosphere, Crawley, WA, Australia
  • 27School of Oceanography, University of Washington, Seattle, WA, United States
  • 28Fisheries and Oceans Canada, Dartmouth, NS, Canada
  • 29Atmosphere and Ocean Research Institute, The University of Tokyo, Kashiwa, Japan
  • 30Council for Scientific and Industrial Research, Cape Town, South Africa
  • 31LEGOS, CNRS, CNES, UPS, University of Toulouse, Toulouse, France
  • 32LOCEAN-IPSL, Sorbonne Université, Paris, France
  • 33Applied Physics Laboratory, University of Washington, Seattle, WA, United States
  • 34Department of Biological Sciences, University of Southern California, Los Angeles, CA, United States
  • 35South Australian Research and Development Institute (Aquatic Sciences), West Beach, SA, Australia
  • 36Instituto Geofísico del Perú, Lima, Peru
  • 37UNC Coastal Studies Institute, Wanchese, NC, United States
  • 38Tokyo University of Marine Science and Technology, Tokyo, Japan
  • 39Wellesley College, Wellesley, MA, United States
  • 40Graduate School of Oceanography, University of Rhode Island, Narragansett, RI, United States
  • 41Institute for Marine and Antarctic Studies, University of Tasmania, Hobart, TAS, Australia
  • 42Departamento de Ciencias de la Atmósfera y los Oceanos, University of Buenos Aires, Buenos Aires, Argentina
  • 43School of Ocean and Earth Science and Technology, University of Hawai‘i at Mānoa, Honolulu, HI, United States
  • 44Departamento de Oceanografia, Universidade Federal de Santa Catarina, Florianópolis, Brazil
  • 45School of Mathematics and Statistics, University of New South Wales, Sydney, NSW, Australia
  • 46Department of Biological Sciences, University of South Carolina, Columbia, SC, United States
  • 47Consejo Nacional de Investigaciones Científicas y Técnicas, Buenos Aires, Argentina
  • 48Department of Biological Sciences, University of Cape Town, Cape Town, South Africa
  • 49CSIRO Oceans and Atmosphere, Hobart, TAS, Australia
  • 50Ministry of Fisheries and Marine Resources, Swakopmund, Namibia
  • 51Department of Marine and Coastal Sciences, Rutgers University, New Brunswick, NJ, United States
  • 52Joint Institute for the Study of the Atmosphere and Ocean, University of Washington, Seattle, WA, United States
  • 53Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China

Ocean boundary current systems are key components of the climate system, are home to highly productive ecosystems, and have numerous societal impacts. Establishment of a global network of boundary current observing systems is a critical part of ongoing development of the Global Ocean Observing System. The characteristics of boundary current systems are reviewed, focusing on scientific and societal motivations for sustained observing. Techniques currently used to observe boundary current systems are reviewed, followed by a census of the current state of boundary current observing systems globally. The next steps in the development of boundary current observing systems are considered, leading to several specific recommendations.

Introduction

Ocean boundary current systems are where society most frequently interacts with the ocean through fisheries, maritime transportation, oil and gas extraction, and recreation. These systems are home to intense and highly variable oceanic currents that redistribute mass, heat, salt, biogeochemical constituents, plankton, and pollution. Circulation patterns also influence the life history, foraging behavior, and abundance of many marine species (e.g., Mansfield et al., 2017). The coastal and open oceans are linked through boundary current systems where events such as coastal upwelling, sea level anomalies, primary productivity, fisheries, and weather are propagated between domains by various processes (e.g., eddies, Rossby waves, and advection). Boundary currents may be broadly categorized as either western boundary currents (WBCs; Imawaki et al., 2013) or eastern boundary currents (EBCs) based on their governing dynamics. In each ocean basin, WBCs play a prominent role in the climate system by redistributing heat from the equator toward the poles, while EBCs are some of the most biologically productive regions in the world and respond dramatically to climate variability (Chavez et al., 2008; Chavez and Messié, 2009).

In our changing climate, shifting hydrological cycles and weather patterns are expected to strongly impact oceanic boundary current processes. Observational evidence for such shifts is beginning to appear. Wu et al. (2012) noted enhanced warming of subtropical WBCs and their extensions during the 20th century, possibly linked to their poleward shift or intensification. Changes in the stability of WBCs have also been noted, with instabilities in the Gulf Stream shifting westward (Andres, 2016), increasing influence of warm core rings on shelf circulation (Gawarkiewicz et al., 2018), and a trend toward greater instability in the East Australian and Agulhas Currents (Sloyan and O’Kane, 2015; Beal and Elipot, 2016).

Oceanic ecosystems are being exposed to increasing pressure from major stressors including warming, deoxygenation, fishing, and acidification. EBCs in particular are projected to be strongly impacted by these stressors (Bakun et al., 2015). For instance, the Peru-Chile (Humboldt) Current system (see the section Peru-Chile Current System), a highly productive EBC and a regional source of greenhouse gases, is naturally affected by upwelling of offshore waters with low oxygen and pH onto the continental shelf (Helly and Levin, 2004) and by periodic El Niño Southern Oscillation (ENSO) events that change the water mass distributions, oxygenation, and productivity (Chavez et al., 2008; Gutiérrez, 2016; Graco et al., 2017); further, stress could have significant consequences for the regional ecosystem. Similarly, changes in the Gulf Stream under global warming are predicted to negatively impact fisheries in the Gulf of Maine and on the New England Shelf (Saba et al., 2016; Claret et al., 2018).

Sustained, interdisciplinary observations in boundary current regions are required for a comprehensive ocean observing system. For OceanObs’09, Send et al. (2010) proposed a global network of sustained monitoring arrays as part of the Global Ocean Observing System (GOOS). Send et al. (2010) broadly defined the properties to be observed as (1) the transports of mass, heat, and freshwater needed for monitoring the global climate in conjunction with basin-scale measurements and (2) local boundary-specific properties including eddy activity, changes in potential vorticity, air–sea interactions (Cronin et al., 2019), ecosystem dynamics, and biogeochemistry. More recently, the 2017 GOOS workshop on “Implementation of Multi-Disciplinary Sustained Ocean Observations” (IMSOO; Palacz et al., 2017) focused, in part, on how to proceed with the development of a truly multidisciplinary boundary current observing system, building upon the more physical and climate-focused plans of Send et al. (2010). In particular, it was noted that observations that resolve along-boundary variability are needed in order to understand climate impacts on various societally relevant uses of boundary current systems (e.g., fisheries). The need to maintain a global perspective that targets all boundary current systems has been repeatedly recognized (Send et al., 2010; Palacz et al., 2017), particularly in developing nations where fisheries can be centrally important (Palacz et al., 2017). To that end, IMSOO planned to review established observing systems in the California Current System and East Australian Current in order to develop a blueprint for an adaptive, multidisciplinary observing system with relocatable subsystems to capture finer scales (Palacz et al., 2017).

Oceanic boundaries present a variety of challenges for sustained observing systems (Send et al., 2010). With strong flows in relatively shallow areas, spatial scales of O(1)–O(10) km, and temporal scales often shorter than a few days (e.g., He et al., 2010; Todd et al., 2013; Rudnick et al., 2017), the broad-scale (i.e., Argo and gridded satellite altimetry) and long-duration (e.g., HOTS, BATS, Station P, and CARIACO) measurements that constitute the observing system for the ocean interior are insufficient for boundary current systems. Multiple observing strategies are needed to measure the essential ocean variables (EOVs; Table 1) that can be used to understand and track the physical and biogeochemical processes of interest within boundary currents (Lindstrom et al., 2012). The optimal combination of observing methods will depend upon characteristics unique to each region. Send et al. (2010) noted that an additional challenge in observing boundary current systems is that there is no well-defined offshore “end” of a boundary current but rather a temporally and spatially variable transition to the interior. At the same time, oceanic boundaries generally lie within exclusive economic zones (EEZs), so the implementation of observing systems requires significant international cooperation.

TABLE 1
www.frontiersin.org

Table 1. List of essential ocean variables from www.goosocean.org/eov with indications of which observing platforms are able to sample each variable.

The overarching purpose of this review is to examine the current state of the boundary current system component of GOOS, updating and building upon the OceanObs’09 review of Send et al. (2010). The section Scientific and Societal Needs considers the scientific and societal needs that comprehensive boundary current observing systems must fulfill. The section Observing Techniques reviews how various observing techniques are employed in boundary currents, highlighting key scientific advances from each platform. The section Current Status of Regional Boundary Current Observing Systems surveys the current state of boundary current observing systems globally. Table 2 provides a comprehensive collection of publications and datasets from the past decade, organized by region and platform. The section Future Outlook then considers the future development of boundary current observing systems. The section Summary Recommendations concludes with specific recommendations to promote development of a comprehensive global network of boundary current observing systems.

TABLE 2
www.frontiersin.org

Table 2. Examples of sustained boundary current observing efforts since 2009.

Scientific and Societal Needs

The Framework for Ocean Observing (Lindstrom et al., 2012), developed after OceanObs’09, recommended that ocean observing systems (1) be “fit for purpose” and driven by “scientific inquiry and societal needs”; (2) include physical, biogeochemical, and biological observations; (3) operate collaboratively based on established best practices; (4) balance innovation with stability; (5) promote alignment of independent user groups; (6) build on existing infrastructure as much as possible; and (7) provide maximum benefit to all users from each observation. Here, we present the scientific and societal needs that that should be met by comprehensive observing of oceanic boundary current systems, focusing on three broad categories: ecosystems and biogeochemistry (see the section Ecosystems and Biogeochemistry), weather and climate (see the section Climate and Weather), and connections between the shelves and deep ocean (see the section Shelf-Deep Ocean Connections).

Ecosystems and Biogeochemistry

Boundary current systems play an important role in carbon cycling through the physical and biological carbon pumps. WBCs are major sites of air–sea CO2 exchange (e.g., Rodgers et al., 2008; Gorgues et al., 2010; Nakano et al., 2011) and have been shown to exhibit enhanced contemporary carbon uptake from the atmosphere (Takahashi et al., 2009; Landschützer et al., 2014). WBC CO2 uptake is driven by a large pCO2 disequilibrium with the overlying mid-latitude atmosphere, which is due to the rapid cooling of low Revelle factor waters advected from the tropics to midlatitudes. Since thick subtropical mode waters form during wintertime convection on the equatorward edges of the WBC extensions, the mode waters are key carbon sinks (e.g., Bates et al., 2002; Gruber et al., 2002; Ito and Follows, 2003; Levine et al., 2011; DeVries, 2014; Iudicone et al., 2016) and have been the target of detailed observational carbon studies (Andersson et al., 2013; Palevsky and Quay, 2017). However, it is still unclear how variability in the rate of mode water formation might impact ocean carbon uptake in these regions and what impacts these changes might have on the biological pump and higher trophic levels (e.g., fisheries). In the Kuroshio Extension region, there is evidence that the majority of carbon exported from the surface ocean during the spring and summer productive season is subsequently respired in the seasonal thermocline and ventilated back to the atmosphere during wintertime mode water formation (Palevsky et al., 2016; Fassbender et al., 2017b; Palevsky and Quay, 2017; Bushinsky and Emerson, 2018). The Southern Hemisphere WBCs are chronically undersampled, particularly during winter, leading to significant uncertainty in their contribution to the global ocean carbon sink.

Boundary current systems are highly productive regions (Chavez et al., 2008). The mechanisms of nutrient supply to surface waters that drive increased primary productivity differ among EBC and WBC systems, but their global contributions are similar (Chavez and Toggweiler, 1995). In EBC systems, the dominant source of nutrients is coastal upwelling (Chavez and Messié, 2009), while in WBC systems, geostrophic- and eddy-driven upwelling predominates (Pelegrí and Csanady, 1991). Nutrient streams are important in the Gulf Stream (Pelegrí and Csanady, 1991; Pelegrí et al., 1996; Williams et al., 2006, 2011; Palter and Lozier, 2008) and the Kuroshio (Guo et al., 2012, 2013), transporting subsurface positive nitrate anomalies, which are delivered to the photic zone primarily by mesoscale and submesoscale processes (Nagai and Clayton, 2017; Honda et al., 2018; Yamamoto et al., 2018; Zhang et al., 2018). Nutrient cycles and drivers have not yet been studied in WBC systems of the Southern Hemisphere.

Western boundary currents are also enriched in micro-nutrients (e.g., Fe, Zn, Cd, Co, and Ni) from land–sea exchanges. They ultimately feed open ocean surface waters and, at lower latitudes, the equatorial undercurrent, where these micro-nutrients are critical in maintaining high levels of productivity. For instance, iron transported by boundary currents in the western Pacific feeds into the Pacific Equatorial Undercurrent, which then supplies iron to the eastern equatorial Pacific (e.g., Mackey et al., 2002; Ryan et al., 2006). In the North Atlantic, Gulf Stream rings supply iron to the subtropical gyre (e.g., Conway et al., 2018). Subpolar WBCs such as the Oyashio and Malvinas Currents are also likely to transport waters enriched in nutrients; wind-driven and shelfbreak upwelling then supplies nutrients to the euphotic layers, enhancing biological productivity (Matano and Palma, 2008; Ito et al., 2010; Valla and Piola, 2015). Locations at which subtropical and subpolar WBCs meet provide ideal environments for biological production, as warm subtropical waters converge with nutrient-rich subpolar waters (Brandini et al., 2000).

The upwelling of deep, poorly ventilated water masses rich in inorganic nutrients and CO2 and low in O2 make EBCs areas of high air–sea fluxes, and the sensitivity of the upwelling process to climate variability contributes to large interannual and decadal scale changes in the magnitude of these fluxes (Friederich et al., 2002; Brady et al., 2019). EBCs also exhibit strong cross-shore gradients in fluxes; narrow strips of the nearshore ocean act as intense sources of CO2 to the atmosphere, while the abundance of nutrients in these upwelled waters facilitates primary production that results in net uptake of CO2 (Hales et al., 2005). The supply of poorly ventilated waters combined with high levels of organic-matter remineralization resulting from intense primary production in surface waters can trigger periods of anoxia and low pH in shelf waters (Feely et al., 2008; Zhang et al., 2010) with severe consequences for demersal and pelagic ecosystems (Chan et al., 2008; Monteiro et al., 2008; Bertrand et al., 2011).

Boundary currents play an important role in ocean ecosystems across all trophic levels. The intense levels of primary production associated with EBCs support rich ecosystems with relatively short food chains, and these systems provide at least 20% of the world’s wild-caught fish despite covering less than 1% of the global ocean (Chavez and Messié, 2009). WBCs and EBCs are also oceanic regions where coastal and open ocean ecosystems are brought together and interact. Modeling studies have suggested that boundary currents are hotspots of microbial biodiversity (Barton et al., 2010; Clayton et al., 2013). This has been supported in the Kuroshio Extension by some in situ surveys (Clayton et al., 2014, 2017). At the other end of the trophic spectrum, recent work combining tag data and satellite altimetry data has shown that white sharks (Carcharodon carcharias) actively occupy warm-core anticyclonic eddies in the Gulf Stream (Gaube et al., 2018). The warmer waters in these mesoscale features allow the sharks to reduce the physiological costs of thermoregulation in cold water, thereby making prey more accessible and energetically more profitable. Similarly, the location of the Kuroshio axis and associated changes in water temperature have been shown to influence the behavior of juvenile Pacific bluefin tuna (Thunnus orientalis; Fujioka et al., 2018). In the Southern Benguela EBC upwelling system, the coastal, wind-driven upwelling along the southwest African coast supports planktonic food supplies for young pelagic fish, while the temperate Agulhas Bank shelf region provides suitable spawning habitat for large communities of fish including in particular anchovy and sardine (Hutchings et al., 2009a). Likewise, southern elephant seals feed along the intense fronts and eddies in the Brazil/Malvinas Confluence (Campagna et al., 2006). WBCs are also known to play an important role in the migration of other coastal and pelagic organisms, such as eels (Shinoda et al., 2011; Rypina et al., 2014) and salmon (Wagawa et al., 2016).

Marine heat waves (MHWs) are strongly linked with boundary current systems. For instance, the exceptional and devastating MHW event off Western Australia during summer of 2010/2011 was caused by a strengthening of the Leeuwin Current associated with La Niña conditions (Pearce and Feng, 2013; Feng et al., 2015), a 2014–2015 MHW had unprecedented impacts on the California Current System (Di Lorenzo and Mantua, 2016; Zaba and Rudnick, 2016), and an MHW in 2015–2016 impacted the Tasman Sea (Oliver et al., 2017). These discrete, prolonged periods of anomalously warm waters at particular locations (Hobday et al., 2016) can stress ecosystems, leading to increased mortality of marine species, closing of commercial and recreational fisheries, and coral bleaching (Cavole et al., 2016; Stuart-Smith et al., 2018). The addition of other stressors such as ocean acidification and deoxygenation, which are projected to increase in future warming scenarios, could amplify the ecosystem impacts of MHWs. Sustained physical and biogeochemical observations are necessary to improve forecasts of the frequency and magnitude of MHWs, as well as to assess the risk and vulnerability of marine ecosystems to extreme climate events (Frölicher and Laufkötter, 2018).

Climate and Weather

Boundary currents are an integral part of the global climate system as they redistribute heat and facilitate carbon uptake from the atmosphere (see the section Ecosystems and Biogeochemistry). In the Atlantic, boundary currents are key components of the Atlantic Meridional Overturning Circulation (AMOC; Frajka-Williams et al., 2019). Low-latitude WBCs that connect the subtropics to the equator at thermocline and intermediate levels are important contributors to the mass and heat budgets of the equatorial oceans, which influence climate modes such as ENSO (Lengaigne et al., 2012). Low-latitude WBCs are also suspected to contribute to the decadal modulation of the equatorial thermocline background state (e.g., Lee and Fukumori, 2003). Sustained monitoring of WBC transports would be particularly useful for climate and seasonal-to-decadal forecast centers (see Smith et al., 2019).

As climate change progresses, boundary current systems are likely to undergo further significant changes. Subtropical WBCs and their extensions are the fastest warming regions of the world ocean (Wu et al., 2012; Yang et al., 2016). Climate model simulations have suggested that western boundary current extensions may move poleward under climate change (Saba et al., 2016). This poleward expansion of energetic WBCs may impact extreme temperatures and marine species migration (Johnson et al., 2011), as well as enhance eddy activity regionally (e.g., Oliver et al., 2015). While low-resolution climate models suggest strengthening and poleward migration of several of these currents under climate change, particularly in the Southern Hemisphere (Sen Gupta et al., 2012; Hu et al., 2015; Pontes et al., 2016), studies leveraging in situ velocity and satellite data suggest no significant increase in their transports since the early 1990s (Rossby et al., 2014; Beal and Elipot, 2016). This discrepancy motivates the collection of long-term measurements of baroclinic changes in boundary currents (i.e., subsurface temperature and salinity properties), as well as the vertical structure of the velocity, in order to understand and predict future changes.

In addition, ocean warming and a magnified hydrological cycle could drive significant changes in shelf ocean stratification, while changes to wind forcing will directly alter rates of upwelling. These ocean circulation processes, and meteorological forcing at the scales that impact upwelling, are poorly represented in climate models (Richter, 2015; Zuidema et al., 2016). Thus, we have little capability to predict how upwelling, winds, and other physical drivers of ocean property exchanges at the coastal/open ocean boundary will change in the future. The impact these changes will have on coastal ecosystems is simply unknown.

Detection and attribution of global sea level variability has improved considerably in the last decade (Cazenave et al., 2014; Marzeion et al., 2014). The location and strength of WBCs considerably influence the mean local sea level (Domingues et al., 2016; Archer et al., 2017b), possibly accounting for part of the mismatch between forecasts and observations of sea level at the coast (Ezer, 2015). Relationships between large-scale wind anomalies, basin-wide sea surface height (SSH), and WBCs (e.g., Boening et al., 2012; Volkov et al., 2019) suggest that observations of current strength and oceanic teleconnections can be used to improve seasonal to decadal coastal sea level forecasts, leading to improved assessments of impacts on infrastructure and groundwater quality (Slangen et al., 2014; Park and Sweet, 2015).

Boundary current systems influence synoptic and longer scale weather patterns. Large upper ocean heat content within WBCs can fuel development and intensification of tropical cyclones (Bright et al., 2002; Wu et al., 2008; Nguyen and Molinari, 2012; Galarneau et al., 2013). Strong sea surface temperature (SST) gradients across WBCs, particularly during winter months, destabilize the atmospheric boundary layer, fueling the mid-latitude storm tracks and atmospheric blocking frequency, which in turn impact regional climate (Kelly et al., 2010; Nakamura, 2012; O’Reilly and Czaja, 2015; O’Reilly et al., 2016; Révelard et al., 2016; Ma et al., 2017). For instance, a weaker Gulf Stream SST front leads to a decrease in cold and dry spells over Europe (O’Reilly et al., 2016), while a sharper SST front in the Kuroshio Extension increases cyclogenesis and shifts the storm track northward, causing warming over eastern Asia and the western United States that can reduce snow cover by 4–6% (O’Reilly and Czaja, 2015; Révelard et al., 2016). Variability in the warm waters of the Agulhas influences summer rainfall over parts of South Africa (Jury et al., 1993; Nkwinkwa Njouodo et al., 2018). In EBC systems, SST minima are collocated with maxima in sea level pressure that are in turn associated with alongshore wind stress, wind stress curl, and cloud cover along the boundary (Sun et al., 2018), suggesting coupling with the full Hadley–Walker tropical atmospheric circulation, though the details of such coupling remain an open question.

Accurate weather and climate forecasting thus requires accurate representation of boundary current systems. However, most of the current ocean reanalyses used to initialize the monthly, seasonal, and decadal forecasts exhibit large errors in the boundary currents (Rouault et al., 2003; Valdivieso et al., 2017), hampering forecast performance. Coupled climate models, such as those used in the Intergovernmental Panel on Climate Change reports, also exhibit large deficiencies in boundary current regions (e.g., Siqueira and Kirtman, 2016; Zuidema et al., 2016), including warm SST biases in EBCs (e.g., Large and Danabasoglu, 2006). Current modeling and data assimilation capabilities are insufficient to fully represent boundary currents at the small spatial scales needed for forecasting. Subramanian et al. (2019) further consider how observing efforts, including within boundary currents, can contribute to improved subseasonal-to-seasonal forecasting.

Shelf-Deep Ocean Connections

The coastal ocean and nearshore zones support a broad range of human activities in maritime industries and resource extraction, and the environmental health and productivity of these regions deliver important ecosystem services. As already noted, the proximity of energetic boundary currents in deep water adjacent to continental shelves mediates shelf-sea/deep-ocean exchange of properties. Along many coasts, this forcing can match or exceed local drivers of circulation such as tides, wind, and river inflows. Coastal ocean and shelf edge dynamics have immediate impacts on ecosystem function and productivity on weekly to seasonal timescales but can also drive multi-decadal changes in ecosystem structure through effects on habitat ranges and biodiversity, not only in coastal zones but also at basin scales.

While we have a broad understanding of the dynamics of upwelling in both WBC and EBC regimes, quantitative estimates of net shelf-sea/deep-ocean exchanges of freshwater and tracers integrated over extended along-shelf distances are few. Quantifying these exchanges is challenging where shelf-edge flow–bathymetry interactions foster variability at short length and timescales. Similarly, exchange flows are not always readily observable at the sea surface from satellite or shore-based remote sensing technologies (see the section Remote Sensing) because they are associated with bottom boundary layer flow driven by the boundary current encountering the seafloor or subduction at the sea surface due to boundary current detachment and mixing. Two efforts along the U.S. East Coast are striving to make such measurements using multi-platform observing arrays: the Processes driving Exchange At Cape Hatteras (PEACH) program and the Ocean Observatories Initiative (OOI; Smith et al., 2018; Trowbridge et al., 2019) Pioneer Array (see the section Northwestern Atlantic). Similarly, in situ and satellite remote sensing observations combined with high-resolution numerical simulations have provided insights into the shelf-sea/deep-ocean exchanges near the confluence of the Brazil and Malvinas Currents (Guerrero et al., 2014; Matano et al., 2014; Strub et al., 2015).

On narrow continental shelves adjacent to intense boundary currents, the impact of deep-ocean circulation on the shelf system is immediate, driving significant fluxes across the continental shelf edge through mesoscale and boundary layer dynamics. For example, mesoscale and submesoscale meandering of the Agulhas jet leads to strong episodic exchanges with shelf waters (Krug et al., 2017; Leber et al., 2017) that support high productivity over the eastern Agulhas Bank (Probyn et al., 1994) and may influence the well-known sardine run (Fréon et al., 2010). On broad continental shelves, bathymetric constraints on cross-isobath flow can hamper exchange at the shelf edge, trapping terrestrial inflows and establishing appreciable cross-shelf buoyancy gradients that in turn sustain shelf-edge fronts (Fratantoni and Pickart, 2007; Howatt et al., 2018).

With changing climate, ocean warming and changes to the hydrological cycle could drive changes in vertical thermal stratification and across-shelf salinity stratification, altering ocean conditions at the inshore edge of boundary current systems (e.g., Gawarkiewicz et al., 2018) and potentially impacting across-shelf fluxes of nutrients and micro-nutrients that are important to sustaining coastal productivity (Fennel et al., 2006). Changes in watershed land use and global weather will alter the volume and characteristics of river flows discharged into the coastal zone. At continental shelf scales, key areas of uncertainty in the oceanographic response to climate variability and change include submesoscale processes and open ocean–shelf exchange. Sustained observing efforts are needed that more fully capture the influence of boundary currents on exchanges with the coastal zone. Designing and deploying boundary current observing systems capable of operating across shelf and deep ocean regimes to deliver coherent views of the shelf-edge exchange is challenging.

Observing Techniques

The highly variable and multi-scale characteristics of boundary currents necessitate an integrated observing system approach, in which high-resolution observations are nested within a backbone of observations over a broad area. Under the Framework for Ocean Observing (Lindstrom et al., 2012), design and implementation of ocean observing systems are focused around a set of EOVs that include physical, biogeochemical, and ecosystem parameters (Table 1)1. Design of an observing system for a particular region (e.g., a specific boundary current system) should proceed through a series of “readiness levels.” In the concept phase, initial feasibility studies and peer review of proposed plans take place. Then, in the pilot phase, small-scale deployments are used to test and validate the proposed approach. Once the observing system reaches the mature phase, it is part of the sustained Global Ocean Observing System. No single observing platform can provide all of the necessary measurements (Table 1), so an optimal mix of observing platforms is needed. Determination of this mix of platforms will be specific to a particular boundary current system, taking into consideration the unique processes and scales at play in that system. Here, we briefly review how various observing platforms are currently being used in boundary current systems; Table 2 refers to many other examples of these observing techniques being applied to boundary current systems.

Time Series

Time series measured from platforms fixed to the seafloor have long been and continue to be central to observing system design and implementation since they uniquely enable collection of long-term measurements at high temporal resolution (minutes to hours) at key locations. Traditional tall moorings (e.g., Johns et al., 2005) typically carry instruments on the mooring wire, within subsurface floats, and on surface buoys, if present; instruments are available to measure most physical EOVs and a growing number of biogeochemical and ecosystem EOVs (Table 1). Moored surface buoys additionally carry suites of meteorological sensors on the buoy tower and sensors for biogeochemical and physical EOVs on the buoy bridle and mooring line just below the sea surface; these air- and sea-surface measurements can be combined to estimate the air–sea exchanges of heat, moisture, CO2, and momentum (Cronin et al., 2019). Inverted echo sounders (IESs) measure the time for sound pulses to travel from the bottom-mounted IES to the surface and back, which, in regions with good databases of hydrographic measurements, can provide full water column estimates of temperature, salinity, and density using the gravest empirical mode technique (Meinen and Watts, 2000). In the Florida Strait, a unique time series of volume transport has resulted from measuring the voltage induced in a submarine cable by seawater moving through the Earth’s magnetic field (Larsen and Sanford, 1985; Baringer and Larsen, 2001; Meinen et al., 2010).

Dense, moored arrays of instruments remain the most effective way to return volume and property transport measurements with high temporal resolution. Subsurface moorings are more typical in WBCs due to the strong surface currents, although surface moorings have also been successfully deployed in the Gulf Stream (Weller et al., 2012) and Kuroshio Extension (Cronin et al., 2013). Arrays of IESs can be used to infer geostrophic shear profiles and, with the addition of bottom pressure sensors (PIES) and near-bottom current measurements (CPIES), can provide estimates of the absolute geostrophic current (Donohue et al., 2010; Meinen et al., 2018). However, the high costs of building, deploying, and turning around such arrays makes them feasible only at a few key locations. Other observing assets are needed to provide spatially broad measurements.

Ship-Based Measurements

Measurements from both dedicated research vessels and ships of opportunity have been central to observing boundary current systems for decades. Research vessels can measure nearly every EOV (Table 1) through the full depth of the water column and are uniquely capable of collecting many types of samples (e.g., net tows and large-volume water samples). Ongoing sustained research vessel surveys of ocean boundary currents include the global GO-SHIP transects at 25- to 50-km resolution (Talley et al., 2016) and the California Cooperative Oceanic Fisheries Investigations (CalCOFI) surveys (McClatchie, 2014) in the California Current System (see the section California Current System). The servicing of boundary current mooring arrays, generally undertaken from research vessels, provides unique opportunities to undertake intensive process studies targeting key scientific questions. The primary limitations on research vessels’ contribution to sustained boundary current observing are their high costs of operation (typically tens of thousands of dollars per day, excluding science personnel) and the infrequency of cruises.

The World Meteorological Organization (WMO) Voluntary Observing Ship (VOS) Program and Ship of Opportunity Program (SOOP) both make use of non-research vessels to collect observations globally, substantially augmenting the amount of ship-based observing. Both programs collect meteorological measurements with real-time observations benefiting weather forecasting, while SOOP additionally uses commercial ships to collect oceanographic measurements along frequently occupied trade routes in the global ocean. Oceanic measurements from SOOP include temperature profiles from expendable bathythermographs (XBTs) at 10- to 25-km resolution in boundary currents (Goni et al., 2019), surface temperature, salinity, plankton, and pCO2 from flow-through systems, and, on specially equipped vessels, velocity profiles from hull-mounted ADCPs (e.g., M/V Oleander; Rossby et al., 2010). Several repeat transects across boundary currents have been maintained for multiple decades and so represent some of the longer datasets available (see the section Current Status of Regional Boundary Current Observing Systems). Fast-moving ships are often able to occupy transects directly across strong boundary currents in short periods of time, a feat not yet possible with other sampling platforms. However, subsurface measurements of variables other than temperature and velocity have remained elusive from ships of opportunity, and recovery of instruments deployed over the side is not practical on cargo vessels.

Autonomous Underwater Gliders

Autonomous underwater gliders (Rudnick, 2016b; Testor et al., 2019) routinely collect long-duration, high-resolution observations in a variety of boundary current systems globally (Todd et al., 2018b; Table 2). Gliders typically profile from the surface to 500–1,000 m, taking 3–6 h to complete a cycle from the surface to depth and back while covering 3–6 km horizontally through the water at a speed of about 0.25 m s–1. During a mission lasting 3–6 months, a glider’s survey track extends well over 2,000 km. Owing to the relatively slow speed of gliders, care must be taken when interpreting the observations, which contain both spatial and temporal variability (Rudnick and Cole, 2011). Sustained deployment of networks of gliders can provide observations with both high spatial resolution and year-round coverage (e.g., Figures 1A,B).

FIGURE 1
www.frontiersin.org

Figure 1. Examples of multi-year, glider-based sampling in (left) an eastern boundary current system and (right) a western boundary current. Trajectories of all Spray gliders surveying the California Current System along CalCOFI lines 66.7, 80.0, and 90.0 (Rudnick et al., 2017 and references therein) and the Gulf Stream along the U.S. East Coast (Todd, 2017; Todd and Locke-Wynn, 2017; Todd et al., 2018a) are shown on the background map. (A,B) Glider sampling as a function of month and cross-shore or cross-stream distance with sampling in all years in gray and calendar year 2017 in color; Gulf Stream sampling in 2017 is colored by along-stream distance from 25°N following the mean 40-cm SSH contour (black trajectory on map with dots every 250 km). (C–F) Example transects of salinity and dissolved oxygen along CalCOFI line 90.0 off Southern California in May 2017 and of potential temperature and velocity toward 50° across the Gulf Stream near Cape Hatteras in August 2017 (red transects on map).

Realizable glider-based sampling plans in boundary currents vary primarily due to the strength of currents relative to a glider’s speed. In EBCs and other boundary currents with relatively weak depth-average currents, gliders can occupy repeat survey lines. The California Underwater Glider Network (CUGN; Figure 1, left), which consists of three cross-shore transects off southern and central California that have been continuously occupied for more than a decade (Rudnick et al., 2017), exemplifies sustained glider observations in an EBC. In WBCs and other boundary currents where depth-average currents are significantly faster than a glider’s speed through the water, gliders can be navigated so as to cross the observed flow as they are advected downstream, returning oblique transects. For example, multi-year surveys of the Gulf Stream (Figure 1, right; Todd et al., 2016, 2018a; Todd, 2017; Todd and Locke-Wynn, 2017) have now returned over 150 high-resolution transects across the WBC of the North Atlantic. Testor et al. (2019) further discuss efforts associated with the OceanGliders Boundary Ocean Observing Network (BOON).

Gliders can carry a variety of sensors (e.g., Figures 1C–F). Measurements of pressure, temperature (Figure 1D), conductivity, and depth-average currents are standard, enabling estimates of absolute geostrophic transport and other physical parameters at relevant scales in boundary currents. Measurements of bio-optical (e.g., Niewiadomska et al., 2008; Henderikx Freitas et al., 2016) and bio-acoustic properties (e.g., Baumgartner and Fratantoni, 2008; Van Uffelen et al., 2017), dissolved oxygen (e.g., Figure 1E; Perry et al., 2008), nitrate, turbulent microstructure (St. Laurent and Merrifield, 2017), and velocity profiles (Figure 1F; Todd et al., 2017) are becoming increasingly common. The main constraints on sensors for gliders are the requirements for small-size, low-power consumption and multi-month stability. As sensor technology continues to mature, gliders will be well suited to carry sensors for additional EOVs, such as pH, in boundary currents.

Drifters

Surface Velocity Program (SVP) drifters drogued at 15-m depth (Niiler et al., 1995; Niiler, 2001; Centurioni, 2018) deployed as part of the Global Drifter Program (GDP) and the Global Surface Drifter Array (GSDA) are also important for understanding the structure and variability of boundary current systems. The GSDA archive dates back to February 1979 and includes over 32 million records of geographical location, 15-m depth velocity, and SST at 6-h resolution (e.g., Hansen and Poulain, 1996; Lumpkin and Pazos, 2007).

Drifter observations have been widely used in both EBCs and WBCs (see Table 2). Recently improved analysis techniques (e.g., Lumpkin, 2003; LaCasce, 2008; Koszalka and LaCasce, 2010; Laurindo et al., 2017) and expansion of the Lagrangian drifter array have allowed gridded, Eulerian statistics of near-surface velocity to be produced at higher resolution, resulting in improved estimates of near-surface flow in boundary currents (e.g., Figure 2) at seasonal to interannual timescales (e.g., Niiler et al., 2003; Lumpkin and Johnson, 2013). Drifter observations in boundary currents offer opportunities for new analyses of long-term variability and trends (e.g., Johnson, 2001; Lumpkin and Johnson, 2013) and the dispersion of tracers and marine debris in the upper ocean (Lumpkin et al., 2012; van Sebille et al., 2015), which is driven by turbulence at scales from surface waves through the submesoscale to large-scale geostrophic eddies (Lumpkin et al., 2017; Lund et al., 2018).

FIGURE 2
www.frontiersin.org

Figure 2. Trajectories and near-surface velocity estimates from Global Drifter Program drifters in the western Pacific and marginal seas. Over 1.2 million discrete measurements from 1982 to 2014 are included. Paths of various boundary currents are clearly visible, as is the rich eddy field in the region of the Subtropical Countercurrent around 18–24°N. NEC = North Equatorial Current, NECC = North Equatorial Counter Current, SEC = South Equatorial Current, SCS = South China Sea (Figure from Todd et al., 2018b).

Argo Floats

Over the past two decades, autonomous profiling Argo floats have become cost-effective and robust platforms. Over 3,700 active Argo floats provide global measurements of temperature, salinity, and pressure in the upper 2,000 m of the ocean, and some are also equipped with sensors measuring biogeochemical properties (Riser et al., 2016; Jayne et al., 2017; Roemmich et al., 2019). Though the Argo network was not designed to capture the details of boundary currents and lacks the resolution necessary to resolve narrow boundary currents, Argo data have nevertheless been used extensively in both WBCs and EBCs (see Table 2). Argo complements other boundary current observing efforts by providing collocated temperature and salinity measurements that are used to infer geostrophic shear from XBT temperature profiles, extending geostrophic shear from XBT and ocean glider data to 2,000 m, measuring reference velocities at parking depth (typically 1,000 m), and linking transport measurements of boundary currents to the ocean interior through basin-wide integration (e.g., Zilberman et al., 2018). Following recommendations at OceanObs’09 (Roemmich et al., 2010), the Argo program is currently moving to double float density in WBC regions (Jayne et al., 2017). The Kuroshio (Figure 3A) and Gulf Stream have historically been among the more densely populated sectors in the Argo array, while other boundary current regions (e.g., the Peru-Chile system, Figure 3B) lack the desired coverage.

FIGURE 3
www.frontiersin.org

Figure 3. Sampling density of Argo float (including Core Argo and BGC Argo) profiles per 1° latitude × 1° longitude bin, collected between January 2009 and September 2018, in the Kuroshio region (A), and the Peru-Chile Current region (B).

Remote Sensing

Among the many oceanic variables that are routinely measured from satellites (Table 1), SSH, SST, and ocean color have been most used to study boundary current systems. Satellite measurements typically have resolutions of O(1)–O(10) km along the satellite track, with repeated measurements on daily to 10-day timescales at a given location. Boundary currents often have strong signatures in both SSH and SST, so satellite-derived gradients in these properties can approximate the strength and/or position of these currents (e.g., Imawaki et al., 2001), including variability on longer timescales (e.g., Qiu and Chen, 2005; Qiu et al., 2014; Andres, 2016). Synergy between dynamic height derived from temperature and salinity profiles and SSH can be exploited to produce synthetic reconstructions of boundary currents (van Sebille et al., 2010; Beal and Elipot, 2016; Majumder and Schmid, 2018; Zilberman et al., 2018), although these reconstructions crucially depend on assumptions about the non-steric (barotropic and mass) variability. Weaknesses of SSH for observing boundary currents include reduced measurement quality within 40 km of the coast due to large uncertainties in the wet-tropospheric correction, unfiltered tides, and a lack of sufficient temporal and spatial resolution to capture the full spectrum of near-surface current variability observed by drifters (Poulain and Niiler, 1989; Centurioni and Niiler, 2003; Fratantoni and Richardson, 2006; Centurioni et al., 2008, 2009; Maximenko et al., 2009). Products that combine SSH and drifter measurements have improved eddy kinetic energy and dynamic topography estimates (Maximenko et al., 2009; Lumpkin and Garzoli, 2011; Rio et al., 2014; Rio and Santoleri, 2018). Estimates of chlorophyll derived from satellite ocean color measurements provide information on biological productivity in boundary current systems worldwide (e.g., Messié and Chavez, 2015; Gómez-Letona et al., 2017). Because ocean color observations have higher resolution (O(1) km) than satellite altimetry measurements, they potentially provide insight into the rich fields of submesoscale instabilities that exist within boundary current systems (Figure 4; Everett et al., 2014; Lee and Kim, 2018).

FIGURE 4
www.frontiersin.org

Figure 4. Example of combined satellite- and land-based remote sensing of the Florida Current. (A) SST from GHRSST and surface geostrophic currents from AVISO. (B) Chlorophyll from MODIS AQUA and surface currents from HF radars (HF radar data from Archer et al., 2017b).

High-frequency (HF) radars (Paduan and Washburn, 2013) have been used effectively to monitor surface current variability of boundary currents (e.g., Kim et al., 2011; Archer et al., 2018). They directly map the total surface current within O(100) km of the coast at high resolution in time (∼1 h) and space (∼1 km) during long-term deployments (∼10 years). HF radar observations have proven useful for investigating both the mean surface velocity structure of boundary currents and associated submesoscale features that develop as boundary currents meander and shed eddies (Archer et al., 2018; Soh and Kim, 2018). Combining HF radar velocity estimates with satellite-based measurements of SST and ocean color (e.g., Figure 4) can provide a multidisciplinary view of surface circulation features at O(1)-km scales (e.g., Schaeffer et al., 2017). Some radar sites have been in continuous operation for more than a decade, offering opportunities to examine interannual to decadal variability of surface circulation. New radar sites can be installed and daisy-chained with existing sites, providing measurements of the alongshore evolution of boundary currents, as has been achieved along the West Coast of the United States (Kim et al., 2011).

Current Status of Regional Boundary Current Observing Systems

Existing observing systems for particular boundary currents are in various stages of development. Here, we review the current status of the observing systems currently operating in several EBCs and WBCs globally. The California Current System (see the section California Current System) is arguably the most well sampled boundary current in the world, offering hope that a fully integrated physical and biogeochemical system is achievable. Other boundary current systems, particularly in the Southern Hemisphere, are much less sampled. As was the case a decade ago (Send et al., 2010), biogeochemical and ecosystem EOVs (Table 1) remain much less well sampled than physical EOVs. Table 2 provides a more comprehensive collection of recent scientific results for each boundary current system as well as sources of publicly available observations.

Eastern Boundary Current Systems

California Current System

The California Current System is the EBC system of the subtropical North Pacific (Checkley and Barth, 2009). The equatorward flowing California Current carries relatively cool and fresh waters of subpolar origin, while the poleward California Undercurrent (Gay and Chereskin, 2009; Todd et al., 2011b) transports warmer saltier waters from the tropics along the continental margin. The California Current System is strongly influenced by the predominantly upwelling-favorable winds along the West Coast of North America.

Owing to the need to understand the collapse of the regional sardine fishery in the 1940s, there is a well-developed multidisciplinary observing system in the California Current System with a decades-long history of routine observations by the CalCOFI program (McClatchie, 2014, and references therein). Since 1949, CalCOFI has made regular (currently quarterly) measurements of physical, biological, and chemical properties at fixed stations along survey lines oriented perpendicular to the coast from research vessels (Figure 5). The establishment of the California Current Ecosystem Long Term Ecological Research program in 2004 brought further ship-based surveys and long-term moorings (Figure 5) focused on non-linear transitions in the pelagic ecosystem in response to ENSO, the Pacific Decadal Oscillation, and secular trends. In the Northern California Current, the Newport Hydrographic Line (44°39.1’N) has been continuously occupied since 1961 (Huyer et al., 2007). Since 2005, autonomous underwater gliders have continuously surveyed along three of the CalCOFI lines as part of the CUGN (see the section Autonomous Underwater Gliders; Figures 1, 5) as well as along cross-shore transects as far north as the Washington Coast (Figure 5), returning measurements of physical properties and some biological proxies; the gliders complement the ship-based surveys by providing observations at higher spatial and temporal resolutions (e.g., Rudnick et al., 2017), albeit of a more limited set of properties. An array of PIES with end-point moorings off of Southern California monitors full-depth geostrophic transport; gliders routinely retrieve data from the PIES and transmit them to shore (Send et al., 2013). Since 2007, NOAA has led large-scale coastal surveys along the U.S. West Coast every 2–4 years to determine the spatial distributions of carbon, oxygen, nutrient, biological, and hydrographic parameters (Feely et al., 2008, 2018). Starting in 2010, moored platforms throughout the California Current System established high-frequency time series of physical and biogeochemical parameters (Nam et al., 2011; Harris et al., 2013; Sutton et al., 2016). More recently, the OOI Endurance Array (Smith et al., 2018; Trowbridge et al., 2019) has been deployed in the northern California Current System (Figure 5); moorings on the shelf and continental slope provide high-resolution time series, while gliders provide high-spatial-resolution observations between the mooring sites. A network of shore-based HF radars provides real-time surface currents within about 150 km of the coast along nearly the entire U.S. West Coast (Kim et al., 2011).

FIGURE 5
www.frontiersin.org

Figure 5. Map of observing efforts extending more than 1 year during the past decade for the California Current System (see the section California Current System). Glider trajectories are shown in orange, SOOP/XBT lines are red, moorings are red dots, and stations routinely occupied by research vessels are green. Contours are mean sea surface height over the period 2009–2017 from AVISO.

Peru-Chile Current System

The Peru-Chile Current System (or Humboldt Current System) is the EBC system of the subtropical South Pacific, extending from the equator to southern Chile (∼45°S). It is characterized by a persistent stratus cloud deck, equatorward surface currents, strong wind-driven coastal upwelling, poleward undercurrents, and filaments and eddies that develop along the coasts of Peru and Chile (see Colas et al., 2012 and references therein). A subsurface oxygen minimum zone (e.g., Paulmier and Ruiz-Pino, 2009) results in upwelled waters being nutrient rich but low in oxygen (e.g., Silva et al., 2009; Pizarro et al., 2016). Due to its proximity to the equator, the Peru-Chile Current System is strongly influenced by equatorial variability through propagation of Kelvin and coastal trapped waves (Dewitte et al., 2012; Mosquera-Vásquez et al., 2013) and anomalous advection during strong El Niños (e.g., Colas et al., 2008).

The dramatic impacts of El Niño events on both weather and fisheries have driven monitoring of oceanographic properties and fish stock assessments along the Peruvian coast since the 1960s (Figure 6; Grados et al., 2018). Over the past decade, these surveys have taken place monthly along the northern Peruvian coast and at least twice per year along the entire Peruvian coast; shipboard ADCP surveys are conducted at least seasonally. Biweekly time series along the 100-m isobath between Paita (5°S) and Ilo (17°S), coastal tide-gauge stations, daily SST measurements at coastal laboratories, and a nearshore thermistor chain and bottom-mounted ADCP at 4°30’S (Figure 6) allow monitoring of temperature and sea level anomalies and associated fluctuations in thermocline, oxycline, and nutricline depths. Measurements from the TAO/TRITON moored array and the Argo program (Figure 3B) provide key broad-scale context to these coastal observations. Efforts are underway to improve monitoring of the Peru-Chile Current System. For instance, sustained glider surveys across the frontal region off northern Peru, where El Niño impacts are large, are planned to begin by 2020.

FIGURE 6
www.frontiersin.org

Figure 6. Map of the boundary current observing effort for the Peru-Chile Current System (see the section Peru-Chile Current System) with details as in Figure 5.

Leeuwin Current System and South Australian Current System

The boundary currents along the western and southern coasts of Australia have some unique features. The Leeuwin Current, which is the subtropical EBC of the southeastern Indian Ocean, is unusual in that it flows poleward along an eastern boundary, transporting warm, fresh tropical waters southward due to forcing by the Indonesian Throughflow and ocean atmosphere interactions in the Indian Ocean (Godfrey and Weaver, 1991); it is important for the upper ocean heat balance in the southeast Indian Ocean (Domingues et al., 2006). The Leeuwin Current hosts broad-scale downwelling (Furue et al., 2017; Liang et al., 2017) where eastward surface currents merge with the Leeuwin Current and then downwell into the Leeuwin Undercurrent at depths of 200–1,000 m. The equatorward Leeuwin Undercurrent carries waters of subantarctic origin along the western Australian coast (Woo and Pattiaratchi, 2008), leaving the coast near 22°S to contribute to the lower limb of a zonal overturning (Furue et al., 2017) and the subtropical gyre (Schott et al., 2009). In winter, the Leeuwin Current merges with the southwestward-flowing Holloway Current off the northwest coast of Australia, the eastward-flowing South Australian Current off the south coast, and the southward-flowing Zeehan Current off the west coast of Tasmania to form the longest shelfbreak boundary current system in the world (Ridgway and Condie, 2004; D’Adamo et al., 2009; Ridgway and Godfrey, 2015). Along the continental slope south of Australia, the westward flowing Flinders Current results from the collision of the equatorward deep ocean Sverdrup transport with the deep shelf slope of the Great Australian Bight (Middleton and Cirano, 2002; Middleton and Bye, 2007) and is a unique northern boundary current.

Coastal sea level observations at Fremantle have long been used as a proxy for the strength of the Leeuwin Current (Feng et al., 2003). Since 2008, the Australian Integrated Marine Observing System (IMOS; Hill et al., 2010) has been monitoring the shelf component of the Leeuwin Current near 32°S using shelf moorings (Feng et al., 2013), gliders, and HF radars (Figure 7). Short-term deployments (2012–2014) have also been carried out off the northwest coast of Australia (Ridgway and Godfrey, 2015). XBT surveys from ships of opportunity in and out of Fremantle, though not targeted for the Leeuwin Current, have taken place since the mid-1980s (Wijffels et al., 2008). IMOS makes ongoing observations of the South Australian Current system with dedicated moorings and glider missions monitoring the Flinders Current (Figure 7).

FIGURE 7
www.frontiersin.org

Figure 7. Map of the boundary current observing efforts for the Leeuwin and South Australian Current Systems (see the section Leeuwin Current System and South Australian Current System) and the Southwestern Pacific (see the section Southwestern Pacific) with details as in Figure 5.

Benguela Current System

The Benguela Current Large Marine Ecosystem is the eastern boundary upwelling system of the South Atlantic. The equatorward Benguela Current is unique in that it is bounded by warm currents at its northern and southern edges, the Angola Current to the north and the Agulhas Current (see the section Agulhas Current) to the south. Coastal upwelling is linked to the seasonal position of the South Atlantic high pressure system, resulting in a number of upwelling cells along the southern African coast with divergent seasonality; the strongest year-round upwelling occurs off Lüderitz (∼26°S), effectively dividing the Benguela Current System into northern and southern subsystems. The northern Benguela upwelling system is highly productive (e.g., Louw et al., 2016) but also prone to hypoxia over the continental shelf that is modulated by a seasonal poleward undercurrent bringing low-oxygen waters to the shelf in summer and fall and an equatorward undercurrent that brings oxygenated water in winter and spring (Duncombe Rae, 2005; Mohrholz et al., 2008; Monteiro et al., 2008). The southern Benguela upwelling system experiences intense, pulsed upwelling in summer and quiescence in winter (Shannon and Nelson, 1996; Weeks et al., 2006; Hutchings et al., 2009c), although the direction of net Ekman transport appears to be offshore year-round (Carr and Kearns, 2003). This region also experiences hypoxia (and occasionally anoxia) in inshore waters, particularly in the region of St. Helena Bay (Pitcher and Probyn, 2011; Pitcher et al., 2014), but low-oxygen events are driven solely by bacterial respiration of organic matter from surface waters (Monteiro and van der Plas, 2006) and can result in mass mortalities of commercial fish stocks and rock lobster (e.g., Cockcroft et al., 2000, 2008; Van der Lingen et al., 2006).

In the southern Benguela Current System, monthly ship-based sampling of fisheries-relevant parameters took place through the 1950s and 1960s and then intermittently until 1988, after which surveys of fisheries, hydrographic, and chemical properties have been conducted two to three times per year (Figure 8; Moloney et al., 2004). Since 2012, quarterly surveys as part of the Integrated Ecosystem Program have additionally monitored the carbonate system. Various multifunctional moorings have been deployed over the years, including a buoy for oxygen and temperature and a harmful algal bloom detection system in the vicinity of St. Helena Bay (see Hutchings et al., 2009b). The Namibian Ministry of Fisheries and Marine Resources conducts regular monitoring of hydrographic conditions and commercial fish resources in Namibian waters of the northern Benguela (Figure 8); regular shipboard oceanographic monitoring began in 1999 with sampling frequency varying from two to eight occupations annually along most lines and up to twice per month off Lüderitz during the lobster fishing season. Long-term, though intermittent, moored observations have been collected at 23°S, 14’03°E, and coastal stations are maintained along the Namibian coast (Figure 8).

FIGURE 8
www.frontiersin.org

Figure 8. Map of the boundary current observing effort for the Bengula Current System (see the section Benguela Current System) and the Agulhas Current (see the section Agulhas Current) with details as in Figure 5.

Canary Current System

The Canary Current Large Marine Ecosystem extends from the northern tip of the Iberian Peninsula (43°N) to south of Senegal (12°N), corresponding to the extent of the northeasterly trade winds in the northeastern Atlantic. Upwelling occurs year-round with meridional shifts in the trade winds leading to seasonality in the latitudinal range of upwelling, particularly in the south (Benazzouz et al., 2014; Faye et al., 2015), where strong intraseasonal to longer timescale variability is driven by internal or remotely forced pulsations of the trade winds, passages of African easterly waves, and oceanic coastally trapped waves (Polo et al., 2008; Diakhaté et al., 2016; Oettli et al., 2016). The ecosystem is broadly divided by the Strait of Gibraltar into the Iberian and the Northwest African areas, though strong subregional differences are observed due to variability in factors including coastal configuration, oxygen concentration, nutrient fertilization, and productivity (Arístegui et al., 2009). The continental shelf in the Canary Current System is the most extensive of any EBC, and persistent circulation features are associated with the topography of the shelf. Large filaments of coastal upwelled water stretch offshore from the numerous capes and promontories (e.g., Cape Guir and Cape Blanc), transporting waters rich in organic matter into the oligotrophic subtropical gyre (Álvarez-Salgado et al., 2007; Lovechio et al., 2018). The Canary Archipelago interrupts the equatorward flow of the Canary Current, leading to strong mesoscale variability downstream of the islands (Arístegui et al., 1994). Island eddies and upwelling filaments interact to exchange water properties, resulting in an efficient route for transporting organic matter to the open ocean (Arístegui et al., 1997; Barton et al., 1998). As a major upwelling area, the Canary Current System is highly productive and the focus of intensive fisheries. Interannual and decadal variability in fisheries landings and distributions of small pelagic fishes has been related to environmental changes associated with the North Atlantic Oscillation and, to a lesser extent, ENSO in the southern part of the region (see reviews in Arístegui et al., 2006; Arístegui et al., 2009).

There have been numerous process-oriented field programs in Canary Current System in the past 20 years, including the Coastal Transition Zone (CTZ; Barton and Arístegui, 2004) and Canary Islands Azores Gibraltar Observations (CANIGO; Parrilla et al., 2002) programs. However, physical, biogeochemical, and ecosystem monitoring efforts have been less routine than have other EBCs (Figure 9). The European Station for Time series in the Ocean Canary Islands (ESTOC) has completed over 20 years of continuous meteorological and surface and mid-water physical and biogeochemical monitoring. The Cape Verde Ocean and Atmospheric Observatories (CVOO/CVAO) are a deep water mooring and an atmospheric station that have been deployed since 2006 in a region critical for climate and greenhouse gas studies and for investigating dust impacts on marine ecosystems. Both ESTOC and CVOO/CVAO are part of the European open ocean fixed point observatories (FixO3). An additional mooring has been recording oceanographic properties and particle fluxes with sediment traps off Cape Blanc continuously since 2003 (Nowald et al., 2015). Long-term measurements of coastal oceanic and atmospheric properties from buoys off Morocco and Senegal have begun during the last 4 years. Ship-based hydrographic and biogeochemical sampling has taken place twice per year since 2006 at the latitude of the Canary Archipelago as part of the RAPROCAN program (Figure 9), which aims to monitor the Canary Current and maintain the ESTOC mooring. Gliders have periodically surveyed between the African coast and the Cape Verde Islands (Figure 9; Karstensen et al., 2017; Kolodziejczyk et al., 2018).

FIGURE 9
www.frontiersin.org

Figure 9. Map of the boundary current observing effort for the Canary Current System (see the section Canary Current System) with details as in Figure 5.

Western Boundary Current Systems

Northwestern Atlantic

The Gulf Stream comprises the upper limb of the AMOC in the North Atlantic subtropical gyre, carrying warm, saline waters from the tropics to higher latitudes. It flows along the eastern seaboard of the United States before separating from the continental margin near Cape Hatteras. The Labrador Current is the WBC of the subpolar gyre. The North Atlantic Deep Western Boundary Current is a deep limb of the AMOC that carries cold water masses from the tail of the Grand Banks of Newfoundland equatorward (Pickart and Watts, 1990). It encounters the Gulf Stream at the tail of the Grand Banks and again at Cape Hatteras, where a portion is entrained into the abyssal interior (Pickart and Smethie, 1993; Bower and Hunt, 2000a,b) while the rest continues to flow equatorward along the western boundary and into the Southern Hemisphere (Southwestern Atlantic). The strength of the Deep Western Boundary Current may influence the latitude at which the Gulf Stream detaches from the continental margin (Thompson and Schmitz, 1989). Along the edge of the adjacent Middle Atlantic Bight shelf, a persistent shelfbreak front and associated shelfbreak jet (Linder and Gawarkiewicz, 1998) transport waters equatorward with secondary frontal circulation leading to upwelling and elevated primary productivity (Marra et al., 1990). The shelfbreak jet continues southward until just north of Cape Hatteras, where it turns offshore as it encounters the much stronger Gulf Stream (Gawarkiewicz and Linder, 2006).

The boundary current observing system for the subtropical northwest Atlantic (Figure 10) is anchored by decades-long measurements at several fixed locations along the boundary. In the Florida Strait near 27°N, cable-based measurements of Gulf Stream transport and quarterly to bi-monthly ship-based sampling have been ongoing since 1982 as part of the Western Boundary Time Series (WBTS; Baringer and Larsen, 2001; Meinen et al., 2010). Far to the northeast, where the Gulf Stream has separated from the continental margin, XBT, shipboard ADCP, and surface temperature and salinity measurements are obtained twice weekly from M/V Oleander, a cargo ship running between New Jersey and Bermuda (Rossby et al., 2010, 2014; Wang et al., 2010). The AX10 XBT line between New York and Puerto Rico crosses the Gulf Stream just upstream of the Oleander line and conducts high-resolution sampling within the boundary current (e.g., Domingues et al., 2018). Since 2015, gliders have been used to routinely survey across the Gulf Stream between Florida and Massachusetts (Figure 1; Todd, 2017; Todd and Locke-Wynn, 2017; Todd et al., 2018b), providing subsurface observations that fill the gap between the WBTS and Oleander and AX10 lines. Two moored arrays with complementary repeat hydrographic sampling have focused on the Deep Western Boundary Current for a decade or more. The RAPID-MOCHA array of subsurface moorings and PIES near 26.5°N has been in place since 2004 with hydrographic stations reoccupied about every 9 months (Meinen et al., 2013). Farther north, the Line W array of subsurface moorings was in place from 2004 to 2014 with repeat ship-based sampling every 6–12 months (Toole et al., 2017). The OOI Pioneer Array south of New England (Smith et al., 2018; Trowbridge et al., 2019) and the PEACH array near Cape Hatteras use a mixture of moorings, gliders (e.g., Gawarkiewicz et al., 2018), and land-based remote sensing (e.g., Haines et al., 2017) to characterize the dynamics of the shelfbreak jet and exchange between the shelf and deep ocean in the vicinity of the Gulf Stream and its eddies. In the subpolar northwestern Atlantic at 53°N, transport of the Labrador Current has been monitored since 1997 using a combination of moored and shipboard observations (Zantopp et al., 2017).

FIGURE 10
www.frontiersin.org

Figure 10. Map of the boundary current observing efforts for the Northwestern Atlantic (see the section Northwestern Atlantic) with details as in Figure 5 and the addition of the submarine cable location in the Florida Strait.

Northwestern Pacific

In the Northwestern Pacific, bifurcation of the westward North Equatorial current between 11° and 13°N along the Philippine coast (Qiu and Chen, 2010; Rudnick et al., 2015b) forms the poleward Kuroshio and the equatorward Mindanao Current. The Kuroshio becomes a more coherent jet as it flows along the Taiwanese coast (e.g., Centurioni et al., 2004), into the East China Sea, and along the southern Japanese coast before separating from the continental margin near 35°N to form the Kuroshio Extension, an eastward, meandering jet in the open North Pacific. The Mindanao carries waters from the North Pacific southward to feed tropical circulation in both the Pacific and Indian Oceans (Schönau et al., 2015). The Oyashio is the western boundary current of the North Pacific subpolar gyre and converges with the Kuroshio to the east of Japan. This convergence region has rich frontal structure as various water masses meet and are modified and is a key area for fisheries (Yasuda, 2003).

The Japan Meteorological Agency (JMA) has carried out repeat hydrographic surveys two to five times annually at the PN line in the East China Sea since 1971 (Aoyama et al., 2008; Figure 11) and at the TK line south of Kyushu since 1987 (Oka and Kawabe, 2003) to monitor physical and biogeochemical EOVs in the Kuroshio. JMA has also monitored the Ryukyu Current system (Ichikawa et al., 2004) flowing south of the Ryukyu Islands at the OK line southeast of Okinawa, which is connected to a zonal section along 24°N. Furthermore, the JMA has maintained physical and biogeochemical surveys along 137°E across the western North Pacific to monitor major currents of the subtropical and tropical gyres including the Kuroshio (Nakano et al., 2015; Oka et al., 2018). Monthly fisheries surveys and hydrographic stations along the A-line off Hokkaido have been occupied since 1987 (Kuroda et al., 2015) with collocated long-term moorings (Kono and Kawasaki, 1997). JAMSTEC has sampled hydrographic stations K2 (47°N, 160°E) and KNOT (44°N, 155°E) in the subpolar north Pacific at least annually since 1997 (Wakita et al., 2010). The Kuroshio Extension Observatory (KEO; Cronin et al., 2015) is a surface mooring that has been located in the subtropical recirculation gyre south of the Kuroshio Extension at 32.3°N, 144.6°E (Figure 11) since 2004. KEO monitors air–sea exchanges of heat, moisture, momentum, and CO2; sea surface temperature, salinity, and ocean acidification; and upper ocean temperature, salinity, and currents below the surface buoy. Since 2014, a sediment trap mooring has been located at KEO (Honda et al., 2018). More recently, the CLIVAR Northwestern Pacific Ocean Circulation and Climate Experiment (NPOCE) has deployed an array of subsurface moorings, some with real-time data transmission, across the western Pacific, South China Sea, and Indonesian seas (Figure 11) that cover the major currents in these regions (e.g., Hu et al., 2013, 2015, 2016; Zhang et al., 2014; Chen et al., 2015; Wang et al., 2017). Multiple XBT transects cross boundary currents within the region (see Goni et al., 2019). Gliders have been deployed for extended periods in the Kuroshio and Mindanao (Figure 11), generally sampling obliquely across the boundary currents as they were advected downstream (e.g., Rainville et al., 2013; Schönau and Rudnick, 2017).

FIGURE 11
www.frontiersin.org

Figure 11. Map of the boundary current observing effort for the Northwestern Pacific (see the section Northwestern Pacific) with details as in Figure 5.

Southwestern Pacific

The East Australian Current is the subtropical WBC of the South Pacific; it is a strong, meandering current with large poleward heat transport (Sloyan et al., 2016) that separates from the continental margin between 30°S and 32°S to join a dynamic eddy field (Cetina Heredia et al., 2014) in the Tasman Sea. The low-latitude WBC system of the South Pacific originates as the equatorward Gulf of Papua Current along the northeast coast of Australia, which then flows through the Solomon Sea as the New Guinea Coastal Undercurrent before feeding into the equatorial current system. This is a major contributor to the mass and heat budget of the tropical Pacific, acting as a conveyor belt for micro-nutrients from the western continental margins to the eastern Equatorial Pacific upwelling region. These low-latitude WBCs split into numerous branches around topographic obstacles and flow through narrow passages, presenting challenges for sustained observing.

The sustained observing system for the East Australian Current and its extension (Figure 7) currently consists of high-density XBT transects (PX05, PX06, PX30, and PX34; Goni et al., 2019); Argo floats; a deep moored array at approximately 27°S; HF radar sites near 32°S and 30°S; a regional array of shelf moorings (including biogeochemical and biological sensors) at 30°S, 34°S, and 36°S; and numerous glider deployments from northern Australia (11°S) to the Tasman Sea (42°S) (Roughan and Morris, 2011; Roughan et al., 2013, 2015). These observational platforms complement each other well, providing a distributed boundary current observational system for the East Australian Current that has been shown to constrain ocean models well (Kerry et al., 2018). Additional sustained measurements are needed to characterize the seasonal changes in the transports of mass, heat, and freshwater in the East Australian Current and its eddy field. Effective monitoring strategies would be to deploy moored arrays in key regions; to increase Argo float and drifter density in the WBC region; and to implement glider sampling along existing high-density XBT lines within the East Australian Current, its eddy field, and recirculation.

In the low-latitude WBC system, long-term, sustained observations of the heat and mass transport through the southern entrance of the Solomon Sea have been provided by gliders since 2007 (Davis et al., 2012) and an array of PIES since 2012 (Figure 7). Concurrent, short-term process studies including mooring deployments have been conducted as part of the CLIVAR-SPICE program (Ganachaud et al., 2014). Future monitoring efforts should integrate measurements across platforms, with the existing measurements in the southern entrance complemented by observations at the northern exits of the Solomon Sea (e.g., moorings, HF radars, and glider transects) to resolve the partitioning of the flow joining the equator (see Smith et al., 2019).

Agulhas Current

The Agulhas Current is the poleward WBC of the subtropical South Indian Ocean (Lutjeharms, 2006). It flows as a fast (>1.5 m s–1), deep-reaching (>3,000 m) jet along the continental slope of southeast Africa (Beal and Bryden, 1999; Beal et al., 2015). Near 40°S, the Agulhas flows into the open ocean, where it retroflects under the influence of the large-scale wind stress curl to flow eastward into the Indian Ocean as the Agulhas Return Current (de Ruijter et al., 1999). Leakage of warm, salty Agulhas waters into the South Atlantic by rings, eddies, and filaments (Boebel et al., 2003; Richardson, 2007) is thought to influence the AMOC on timescales from decades to millennia (Beal et al., 2011).

In 2010, the Agulhas Current Time-series experiment (ACT) established a moored array to measure the volume transport of the Agulhas Current along a satellite altimeter ground-track (#96) near 34°S (Figure 8) for a period of 3 years. The array consisted of seven full-depth current meter moorings and four CPIES that captured the breadth and depth of the Agulhas jet, including during offshore meander events (Beal et al., 2015). Following ACT, a consortium of South African, U.S., and Dutch scientists deployed the Agulhas System Climate Array (ASCA) in 2016 for long-term monitoring of the Agulhas Current as part of GOOS. ASCA augmented the original ACT array design with conductivity–temperature recorders to measure the heat and freshwater fluxes. The long-term success of ASCA was dependent on an ambitious plan of capacity building and resource sharing among nations, and, owing to a number of challenges, this plan was not fulfilled, and the array was pulled out of the water in 2018, following a 2-year deployment. In 2015, the Shelf Agulhas Glider Experiment (SAGE) demonstrated the feasibility of operating autonomous robotic platforms to sample the shelf regions of the Agulhas Current (Krug et al., 2017). Since SAGE, growing regional interest in monitoring with autonomous platforms led to formation of a South African multi-institutional scientific consortium named Gliders in the Agulhas (GINA). GINA conducted two additional glider missions in 2017 and 2018 and is working toward the development of a sustained glider observing system for the Agulhas Current coastal and shelf regions. The influence of the Agulhas leakage on the AMOC has been monitored since 2013 by an array of CPIES and tall moorings as part of the SAMBA line at 34.5°S (Figure 8; Ansorge et al., 2014). Thus far, no sustained ecological or biogeochemical measurements have been made in the Agulhas, though the addition of oxygen sensors to SAMBA moorings is planned.

Southwestern Atlantic

In the South Atlantic, the WBC system consists of the poleward Brazil Current and the equatorward North Brazil Undercurrent, both of which originate from the bifurcation of the South Equatorial Current between 10°S and 20°S (e.g., da Silveira et al., 1994; Rodrigues et al., 2007), and the equatorward Malvinas current in the subpolar gyre. The Brazil Current and Malvinas Current both separate from the South American continental margin between 35°S and 40°S to flow eastward at the Brazil–Malvinas confluence (Olson et al., 1988). The North Brazil Undercurrent constitutes a bottleneck for the interhemispheric mean flow of the upper limb of the AMOC as it transports warm waters of South Atlantic origin across the equator (e.g., Schott et al., 1998; Zhang et al., 2011; Rühs et al., 2015). The Deep Western Boundary Current carries much of the lower limb of the AMOC off the coast of South America (Schott et al., 2005; Meinen et al., 2013).

For more than a decade, high-density XBT transects (Goni et al., 2019) have been occupied near 22°S and 34°S (AX97 and AX18) across South Atlantic WBCs (Figure 12; Dong et al., 2015; Lima et al., 2016). Near 11°S, an array of four tall moorings and two PIES has measured transport of the North Brazil Current since 2013 (Figure 12; Hummels et al., 2015). At 34.5°S (Figure 12), an array of PIES, CPIES, and a bottom-mounted ADCP has monitored the Brazil Current and Deep Western Boundary Current (Meinen et al., 2013, 2017, 2018) in conjunction with periodic hydrographic surveys (Valla et al., 2018). A series of yearlong deployments of current meter arrays along 41°S since 1993 (Figure 12; Vivier and Provost, 1999; Spadone and Provost, 2009; Paniagua et al., 2018), in conjunction with satellite altimetry, has allowed for production of a 24-year transport time series for the Malvinas Current (Artana et al., 2018a).

FIGURE 12
www.frontiersin.org

Figure 12. Map of the boundary current observing system for the Southwestern Atlantic (see the section Southwestern Atlantic) with details as in Figure 5.

Future Outlook

We recommend establishing and maintaining a global network of boundary current observing systems. Each distinct observing system will need to be tailored to the unique aspects of that particular boundary current system and also follow best practices established by the international community. Such a network of regional boundary current observing systems is a crucial element of GOOS. To date, boundary current observing systems in different regions and countries have developed largely independently. Development and maintenance of a global network of boundary current observing systems that is fit for purpose would benefit from the standards outlined in the Framework for Ocean Observing (Lindstrom et al., 2012). In particular, application of the Framework across boundary current observing systems should foster communication and data sharing; contribute to capacity building, particularly in developing countries; encourage confidence and support from funding agencies; and promote international collaboration and scientific and technological innovation.

Boundary currents typically lie within the EEZs of coastal states, so development and maintenance of boundary current observing systems will require the cooperation and support of appropriate governing authorities. Considering the difficulty of obtaining international funding for observations in national waters, there is a need for a community of regional boundary observers. Moreover, many boundary currents span multiple countries, so the observing system for a single boundary current system is likely to require collaboration and coordination between several countries. The advective nature of boundary currents may even require that mobile or drifting assets deployed within one country’s EEZ be recovered within another EEZ. Sharing of measurements taken within EEZs, particularly those that have economic impacts such as some biogeochemical measurements, remains a challenge. By moving toward international collaboration in the design and implementation of boundary current systems as suggested by the Framework for Ocean Observing, there is hope for building the high-level governance structure needed to surmount the challenges posed by boundary currents falling within EEZs. The Large Marine Ecosystems effort has identified distinct boundary regions that cross international borders and has gained international traction through the Global Environment Facility and the International Union for Conservation of Nature; leveraging this effort to facilitate international cooperation and governance for sustained boundary current observations may be fruitful.

For any particular boundary current system, a complete observing system will require a combination of currently available observing platforms (see the section Observing Techniques), as well as future platforms, to optimally measure EOVs at necessary spatial and temporal resolutions to address relevant scientific and societal needs. Through the Framework process, specific observing platforms, sampling choices, and instruments would be matched to the relevant questions. Drifting and mobile assets that provide spatially resolved measurements at the expense of temporal resolution will need to be combined with moored assets that provide high-frequency measurements at key locations and land- or satellite-based remote sensing that provides spatially broad measurements of sea surface properties. Such integrated arrays, as are already in place in the California Current System, at the Ocean Observatories Initiative (OOI; Smith et al., 2018; Trowbridge et al., 2019) Endurance and Pioneer Arrays, and along the Australian coasts as part of the Integrated Marine Observing System, offer critical opportunities for intercalibration between instruments on fixed and mobile assets; such intercalibration is particularly important for biogeochemical sensors (e.g., Palevsky and Nicholson, 2018). Since similar needs arise in most boundary current systems, the Framework process should provide a means for determining the extent to which the same observing strategies should be applied to address similar needs in different systems. Additional studies that compare different sampling techniques in a given boundary current system could provide guidance on the strengths and limitations of each technique and how to better exploit their complementarity.

While the discussion of observing platforms in the section Observing Techniques focused on mature observing platforms with proven records of sustained operation in boundary currents, there is no doubt that recently developed observing platforms and sensing technology will become integral parts of future boundary current observing systems. For instance, more fast-moving autonomous underwater vehicles (AUVs) and autonomous surface vehicles (ASVs) will be deployed to conduct adaptive and targeted sampling in response to real-time needs. Propeller-driven AUVs have thus far seen limited use in boundary currents. Though able to carry large instrument payloads and move much faster (1–2 m s–1) than gliders, propeller-driven AUVs have been limited by battery endurance to missions typically lasting hours to days; improvements in battery technology and autonomous charging are expected to make propeller-driven vehicles capable of long-duration sampling in the near future. Fast-moving, long-endurance ASVs (e.g., Saildrones and WaveGliders) are poised to become key platforms for making measurements near the air–sea interface, including meteorological measurements, pCO2, subsurface currents, and plankton biomass. Due to the use of renewable energy, these ASVs generally carry a larger number of sensors and have longer duration than other autonomous platforms (e.g., Zhang et al., 2017). Planned high-resolution, satellite-based altimetry measurements (e.g., SWOT), smaller and dramatically cheaper satellites (e.g., Cubesats), and, potentially, geostationary satellites positioned over boundary regions offer the prospect of dramatically increased spatial and temporal resolution of surface properties.

At some locations, boundary currents have been continuously observed for many years using various techniques. For instance, the CalCOFI program has maintained quarterly ship-based stations for more than 65 years (McClatchie, 2014), the WBTS has made cable- and ship-based measurements in the Florida Strait for more than 35 years (see the section Northwestern Atlantic), and hydrographic sampling has occurred monthly along the inside edge of the East Australian Current since the 1940s (Lynch et al., 2014) and is now an integral part of the East Australian Current observing system (Roughan and Morris, 2011). Long-term measurements like these are invaluable for capturing decadal variability and secular trends. Sites at which decades-long measurements exist should be maintained and serve as anchors for comprehensive boundary current observing systems. These long-term measurement sites at the boundaries also serve as points at which the boundary current observing systems are linked to the basin-scale ocean observing system. Since 2004, the WBTS has been integrated with the U.K.–U.S. RAPID-MOCHA program that measures meridional transport at 26.5°N in the North Atlantic, while several long-standing, cross-Pacific XBT transects intersect the U.S. West Coast within the CalCOFI domain (Goni et al., 2019).

Existing boundary current observing systems are largely focused on measuring physical processes, with biogeochemical and ecosystem processes only beginning to gain traction, largely due to the advent of new sensors. The California Current System (see the section California Current System) and Benguela (see the section Benguela Current System) are exceptions, having had sustained observations of EOVs relevant to physics, biogeochemistry, and biology and ecosystems for over 65 and 30 years, respectively. However, these ship-intensive models are unlikely to be suited to all boundary current systems due to a wide range of factors (e.g., cost, proximity to the coast, existing infrastructure, and available manpower). Although the methods for measuring many of the EOVs needed to monitor biogeochemical and, to a greater extent, ecosystem processes are time-intensive and require a platform for collecting water, new sensors are being developed to reliably measure a range of biogeochemical and biological EOVs. Many of these sensors have been successfully deployed on BGC-Argo floats as part of the SOCCOM project (Johnson et al., 2017). Increasing the measurements of biological and ecological EOVs should be prioritized if we are to understand, monitor, and predict the physical–biological connections and processes that support marine-based industries and activities and, importantly, seafood security.

Providing publicly available data in a timely manner is a key attribute of any ocean observing system. These observations should be provided in formats that are discoverable, accessible, and readily subset, following conventions agreed upon by the community (see Wilkinson et al. (2016) for a set of general principles for management of scientific data). Many platforms already provide observations in near real-time through a variety of services. Transmission of data through the Global Telecommunications System is particularly important if those observations are to be used in operational numerical modeling. Advances in real-time data collection from sub-surface moorings (e.g., Send et al., 2013) will be critical to providing boundary current observations in a timely manner for forecasting and prediction. Widespread dissemination of comprehensive boundary current observations can foster synergies with other disciplines, including the geophysics (tsunamis and earthquakes), physics, meteorological (e.g., tropical and extratropical cyclone forecasting; Domingues et al., 2019), and fisheries communities.

In addition to providing raw observations, there is a need for providing synthesized data products that are tailored to user needs. Integration of complementary data types can yield useful metrics. Further advances in data analysis techniques and statistical methods should aid in using multi-platform data to increase temporal and/or spatial resolution of metrics. The Southern California Temperature Index (Rudnick et al., 2017) is an example of such a data product.

Boundary current observations play a key role in constraining ocean models (e.g., Todd and Locke-Wynn, 2017), while models play a complementary role by filling gaps between sparse observations in a dynamically consistent manner (e.g., Todd et al., 2011b, 2012; Gopalakrishnan et al., 2013). Increased availability of boundary current observations, particularly in regions that are currently poorly sampled, should lead to continued improvements in regional models and predictive tools. At the same time, higher resolution climate models that can resolve boundary currents are becoming more plentiful and should begin to rely on high-resolution boundary current observations as constraints. One specific goal would be to reduce climate models’ warm SST biases within EBCs; continuation and expansion of long-term measurements in EBCs as well as focused process studies to study upper ocean and atmospheric dynamics in EBCs would contribute to this goal. Observation impact studies derived from data assimilating models provide guidance on the value of a range of observation types in resolving boundary current transport, as well as for constraining the eddy field in ocean reanalyses (e.g., Kerry et al., 2016, 2018). It remains an open question how best to integrate models with interdisciplinary (e.g., biogeochemical) observations to study ecosystem dynamics, though advances are being made in the assimilation of biological parameters (e.g., Song et al., 2012). Observing System Simulation Experiments tailored for boundary current systems can also provide insight to the type, spatial distribution, and frequency of observations required to improve numerical simulations of boundary current dynamics (Hoffman and Atlas, 2016). Targeted observations can reduce biases in the initialization of models used to forecast extreme weather events and support local decision making (Halliwell et al., 2017).

Downscaling coarse resolution climate model predictions through the application of higher resolution regional and coastal models is now common and has shown promise but still faces research challenges. Furthermore, a significant amount of physical, biogeochemical, and biological response on the continental shelf is due to episodic oceanic and atmospheric events at timescales of variability that are absent from coarse models and cannot be recovered locally. To be valid globally, the veracity of downscaled models needs to be appraised by supporting observations of shelf edge fluxes in a diversity of circulation regimes.

Funding sustained observing efforts is a significant challenge. Portions of the observing system that have proven their readiness for long-term deployment have been discontinued after one or more short-term funding cycles. For instance, it is currently not clear how ship-time and funding challenges will be met for a re-establishment of ASCA (see the section Agulhas Current) in the future. In the typical 3-to-4-year cycles of scientific funding, early years (e.g., pilot phases) of observing efforts are readily fundable based on the promise of quick scientific results. Observing efforts that have endured for a decade or longer can leverage their long histories and clear relevance to decadal variability or secular trends to secure continued funding. The middle years, roughly years 4 through 10 as programs transition from pilot to mature components of the GOOS, are particularly difficult to fund.

The provision of robust three-dimensional and time-varying ocean circulation estimates in boundary current systems, resolving scales of a few kilometers, is seemingly within reach through advances in data-assimilative ocean models and rapid developments in observations platforms and sensors. However, the development of integrated observing systems that deliver the scope of observations required and the models capable of fully utilizing those observations is challenging. Success will require coordinated international collaborations, bringing together the expertise of the ocean modeling and observational communities. Establishment of an Ocean Boundary Task Team would provide a mechanism for the exchange of information regarding observing and model strategies, sensor developments, analysis techniques to combine data from the various observing platforms, and model development and application. The Task Team would also enable capacity building, encourage timely and appropriate transfer of knowledge, and provide a mechanism to instigate multinational observing systems with shared goals amongst participating nations. Endorsement of the Task Team by IOC/WMO or similar international organization is critical due to interests of multiple coastal state EEZs and the resulting complex governance needs.

Summary Recommendations

The following actions are recommended to promote development of a comprehensive global network of boundary current observing systems in the next decade:

(1) Maintain existing long-term (i.e., multi-year) observational records;

(2) Expand the use of mobile, autonomous platforms (e.g., gliders, AUVs, and ASVs) to provide continuous, high-resolution, broad-scale monitoring of EOVs;

(3) Deploy moored platforms at key locations to measure high-frequency variability;

(4) Continue and expand the provisioning of real-time observations and encourage post-processed data to be made publicly available as quickly as possible; data should be provided in readily discoverable formats that can easily be subset;

(5) Continue development and expand deployment of sensors for ecological and biogeochemical EOVs;

(6) Establish an Ocean Boundary Task Team to foster international community development and end-user engagement and to guide evolution of observing systems as user requirements change;

(7) Expand collaborations between observational efforts, modeling efforts, and societal users to meet stakeholder and end-user needs; and

(8) Increase focus on exchange between continental shelves and the deep ocean boundary currents to develop observing systems that span the continuum from the land to the deep ocean.

Author Contributions

RT led the manuscript. Other lead authors (FC, SCl, SCr, MGo, MGr, XL, JS, and NZ) helped to conceive the manuscript and participated in all the stages of development. All authors provided input and/or edited the text.

Funding

RT was supported by The Andrew W. Mellon Foundation Endowed Fund for Innovative Research at WHOI. FC was supported by the David and Lucile Packard Foundation. MGo was funded by NSF and NOAA/AOML. XL was funded by China’s National Key Research and Development Projects (2016YFA0601803), the National Natural Science Foundation of China (41490641, 41521091, and U1606402), and the Qingdao National Laboratory for Marine Science and Technology (2017ASKJ01). JS was supported by NOAA’s Global Ocean Monitoring and Observing Program (Award NA15OAR4320071). DZ was partially funded by the Joint Institute for the Study of the Atmosphere and Ocean (JISAO) under NOAA Cooperative Agreement NA15OAR4320063. BS was supported by IMOS and CSIRO’s Decadal Climate Forecasting Project. We gratefully acknowledge the wide range of funding sources from many nations that have enabled the observations and analyses reviewed here.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We gratefully acknowledge the efforts of all parties involved in collecting, analyzing, and disseminating the wide-ranging observations discussed here. We thank the OCB Program, US CLIVAR, MBARI, and OMIX for supporting the Ocean Carbon Hot Spots Workshop, which facilitated the discussion of carbon cycle research in Western Boundary Current systems. We also gratefully acknowledge J. Hildebrandt (WHOI) for the assistance in editing the final manuscript. This is PMEL contribution # 4890.

Footnotes

  1. ^ http://www.goosocean.org/eov

References

Adams, K., Barth, J. A., and Shearman, R. K. (2016). Intraseasonal cross-shelf variability of hypoxia along the Newport, Oregon, hydrographic line. J. Phys. Oceanogr. 46, 2219–2238. doi: 10.1175/JPO-D-15-0119.1

CrossRef Full Text | Google Scholar

Alberty, M. S. (2018). Water Mass Transport and Transformation in the Tropics and Arctic. San Diego, CA: University of California. Available at: https://escholarship.org/uc/item/9rf158pq

Google Scholar

Alberty, M., Sprintall, J., MacKinnon, J., Germineaud, C., Cravatte, S., and Ganachaud, A. (2019). “Data from: moored observations of transport in the Solomon Sea,” in Solomon Sea SPICE Mooring Data, (La Jolla, CA: UC San Diego Library Digital Collections), doi: 10.6075/J0639N12

CrossRef Full Text | Google Scholar

Alford, M. H., Sloyan, B. M., and Simmons, H. L. (2017). Internal waves in the East Australian Current. Geophys. Res. Lett. 44, 12280–12288. doi: 10.1002/2017GL075246

CrossRef Full Text | Google Scholar

Alin, S. R., Feely, R. A., Dickson, A. G., Hernández-Ayón, J. M., Juranek, L. W., Ohman, M. D., et al. (2012). Robust empirical relationships for estimating the carbonate system in the southern California Current System and application to CalCOFI hydrographic cruise data (2005–2011). J. Geophys. Res. Oceans 117:C05033. doi: 10.1029/2011JC007511

CrossRef Full Text | Google Scholar

Álvarez-Salgado, X. A., Arístegui, J., Barton, E. D., and Hansell, D. A. (2007). Contribution of upwelling filaments to offshore carbon export in the subtropical Northeast Atlantic Ocean. Limnol. Oceanogr. 52, 1287–1292. doi: 10.4319/lo.2007.52.3.1287

CrossRef Full Text | Google Scholar

Andersson, A. J., Krug, L. A., Bates, N. R., and Doney, S. C. (2013). Sea–air CO2 flux in the North Atlantic subtropical gyre: role and influence of sub-tropical mode water formation. Deep Sea Res. Part II Top. Stud. Oceanogr. 91, 57–70. doi: 10.1016/j.dsr2.2013.02.022

CrossRef Full Text | Google Scholar

Andres, M. (2016). On the recent destabilization of the Gulf Stream path downstream of Cape Hatteras. Geophys. Res. Lett. 43, 9836–9842. doi: 10.1002/2016GL069966

CrossRef Full Text | Google Scholar

Andres, M., Jan, S., Sanford, T. B., Mensah, V., Centurioni, L. R., and Book, J. W. (2015). Mean structure and variability of the Kuroshio from northeastern Taiwan to southwestern Japan. Oceanography 28, 84–95. doi: 10.5670/oceanog.2015.84

CrossRef Full Text | Google Scholar

Ansorge, I. J., Baringer, M. O., Campos, E. J. D., Dong, S., Fine, R. A., Garzoli, S. L., et al. (2014). Basin-wide oceanographic array bridges the South Atlantic. EOS 95, 53–54. doi: 10.1002/2014eo060001

CrossRef Full Text | Google Scholar

Aoyama, M., Goto, H., Kamiya, H., Kaneko, I., Kawae, S., Kodama, H., et al. (2008). Marine biogeochemical response to a rapid warming in the main stream of the Kuroshio in the western North Pacific. Fish. Oceanogr. 17, 206–218. doi: 10.1111/j.1365-2419.2008.00473.x

CrossRef Full Text | Google Scholar

Archer, M. R., Keating, S. R., Roughan, M., Johns, W. E., Lumkpin, R., Beron-Vera, F., et al. (2018). The kinematic similarity of two western boundary currents revealed by sustained high-resolution observations. Geophys. Res. Lett. 45, 6176–6185. doi: 10.1029/2018GL078429

CrossRef Full Text | Google Scholar

Archer, M. R., Roughan, M., Keating, S. R., and Schaeffer, A. (2017a). On the variability of the East Australian Current: jet structure, meandering, and influence on shelf circulation. J. Geophys. Res. Oceans 122, 8464–8481. doi: 10.1002/2017JC013097

CrossRef Full Text | Google Scholar

Archer, M. R., Shay, L. K., and Johns, W. E. (2017b). The surface velocity structure of the Florida Current in a jet coordinate frame. J. Geophys. Res. Oceans 122, 9189–9208. doi: 10.1002/2017JC013286

CrossRef Full Text | Google Scholar

Archer, M. R., Shay, L. K., Jaimes, B., and Martinez-Pedraja, J. (2015). “Observing frontal instabilities of the Florida Current using high frequency radar,” in Coastal Ocean Observing Systems, eds Y. Liu, H. Kerkering, and R. H. Weisberg (Amsterdam: Elsevier, Inc.), 179–208. doi: 10.1016/B978-0-12-802022-7.00011-0

CrossRef Full Text | Google Scholar

Arístegui, J., Alvarez-Salgado, X. A., Barton, E. D., Figueiras, F. G., Hernández-León, S., Roy, C., et al. (2006). “Oceanography and fisheries of the Canary Current/Iberian region of the eastern North Atlantic,” in The Sea, Vol. 14, eds A. R. Robinson and K. H. Brink (Cambridge, MA: Harvard University Press), 877–931.

Google Scholar

Arístegui, J., Barton, E. D., Alvarez Salgado, X. A., Santos, A. M. P., Figueiras, F. G., Kifani, S., et al. (2009). Sub-regional ecosystem variability in the Canary Current upwelling. Prog. Oceanogr. 83, 33–48. doi: 10.1016/j.pocean.2009.07.031

CrossRef Full Text | Google Scholar

Arístegui, J., Sangrà, P., Hernández-León, S., Cantón, M., Hernández Guerra, A., and Kerling, J. L. (1994). Island-induced eddies in the Canary Islands. Deep Sea Res. I 41, 1509–1525. doi: 10.1016/0967-0637(94)90058-2

CrossRef Full Text | Google Scholar

Arístegui, J., Tett, P., Hernández-Guerra, A., Basterretxea, G., Montero, M. F., Wild, K., et al. (1997). The influence of island-generated eddies on chlorophyll distribution: a study of mesoscale variation around Gran Canaria. Deep Sea Res. 44, 71–96. doi: 10.1016/s0967-0637(96)00093-3

CrossRef Full Text | Google Scholar

Artana, C., Ferrari, R., Koenig, Z., Saraceno, M., Piola, A. R., and Provost, C. (2016). Malvinas Current variability from argo floats and satellite altimetry. J. Geophys. Res. Oceans 121, 4854–4872. doi: 10.1002/2016JC011889

CrossRef Full Text | Google Scholar

Artana, C., Ferrari, R., Koenig, Z., Sennechael, N., Saraceno, M., Piola, A. R., et al. (2018a). Malvinas Current volume transport at 41°S: a 24 yearlong time series consistent with mooring data from 3 decades and satellite altimetry. J. Geophys. Res. Oceans 123, 378–398. doi: 10.1002/2017JC013600

CrossRef Full Text | Google Scholar

Artana, C., Lellouche, J., Park, Y., Garric, G., Koenig, Z., Sennéchael, N., et al. (2018b). Fronts of the Malvinas Current System: surface and subsurface expressions revealed by satellite altimetry, Argo floats, and Mercator operational model outputs. J. Geophys. Res. Oceans 123, 5261–5285. doi: 10.1029/2018JC013887

CrossRef Full Text | Google Scholar

Auad, G., Roemmich, D., and Gilson, J. (2011). The California Current System in relation to the northeast Pacific ocean circulation. Prog. Oceanogr. 91, 576–592. doi: 10.1016/j.pocean.2011.09.004

CrossRef Full Text | Google Scholar

Bacon, S., and Saunders, P. M. (2010). The deep western boundary current at Cape Farewell: results from a moored current meter array. J. Phys. Oceanogr. 40, 815–829. doi: 10.1175/2009jpo4091.1

CrossRef Full Text | Google Scholar

Bakun, A., Black, B. A., Bograd, S. J., Garcia-Reyes, M., Miller, A. J., Rykaczewski, R. R., et al. (2015). Anticipated effects of climate change on coastal upwelling ecosystems. Curr. Clim. Change Rep. 1, 85–93. doi: 10.1007/s40641-015-0008-4

CrossRef Full Text | Google Scholar

Bane, J. M., He, R., Muglia, M., Lowcher, C. F., Gong, Y., and Haines, S. M. (2017). Marine hydrokinetic energy from western boundary currents. Annu. Rev. Mar. Sci. 9, 105–123. doi: 10.1146/annurev-marine-010816-060423

PubMed Abstract | CrossRef Full Text | Google Scholar

Baringer, M. O., and Larsen, J. (2001). Sixteen years of Florida Current transport at 27N. Geophys. Res. Lett. 28, 3179–3182. doi: 10.1029/2001GL013246

CrossRef Full Text | Google Scholar

Barton, A. D., Dutkiewicz, S., Flierl, G., Bragg, J., and Follows, M. J. (2010). Patterns of diversity in marine phytoplankton. Science 327, 1509–1511. doi: 10.1126/science.1184961

PubMed Abstract | CrossRef Full Text | Google Scholar

Barton, E. D., and Arístegui, J. (2004). The Canary Islands transition—upwelling, eddies and filaments. Prog. Oceanogr. 62, 67–69. doi: 10.1016/j.pocean.2004.08.003

CrossRef Full Text | Google Scholar

Barton, E. D., Arístegui, J., Tett, P., Cantón, M., García-Braun, J., Hernández-León, S., et al. (1998). The transition zone of the Canary Current upwelling region. Prog. Oceanogr. 41, 455–504. doi: 10.1016/s0079-6611(98)00023-8

CrossRef Full Text | Google Scholar

Bates, N., Pequignet, A. C., Johnson, R. J., and Gruber, N. (2002). A variable sink for atmospheric CO2 in subtropical mode water of the North Atlantic Ocean. Nature 420, 489–493. doi: 10.1038/nature01253

PubMed Abstract | CrossRef Full Text | Google Scholar

Baumgartner, M. F., and Fratantoni, D. M. (2008). Diel periodicity in both sei whale vocalization rates and the vertical migration of their copepod prey observed from ocean gliders. Limnol. Oceanogr. 53, 2197–2209. doi: 10.4319/lo.2008.53.5_part_2.2197

CrossRef Full Text | Google Scholar

Beal, L. M., and Bryden, H. L. (1999). The velocity and vorticity structure of the Agulhas Current at 32°S. J. Geophys. Res. 104, 5151–5176. doi: 10.1029/1998JC900056

CrossRef Full Text | Google Scholar

Beal, L. M., and Elipot, S. (2016). Broadening not strengthening of the Agulhas Current since the early 1990s. Nature 540, 570–573. doi: 10.1038/nature19853

PubMed Abstract | CrossRef Full Text | Google Scholar

Beal, L. M., De Ruijter, W. P. M., Biastoch, A., Zahn, R., Cronin, M., Hermes, J., et al. (2011). On the role of the Agulhas system in ocean circulation and climate. Nature 472, 429–436. doi: 10.1038/nature09983

PubMed Abstract | CrossRef Full Text | Google Scholar

Beal, L. M., Elipot, S., Houk, A., and Leber, G. M. (2015). Capturing the transport variability of a western boundary jet: results from the Agulhas Current Time-Series Experiment (ACT). J. Phys. Oceanogr. 45, 1302–1324. doi: 10.1175/JPO-D-14-0119.1

CrossRef Full Text | Google Scholar

Beal, L. M., Hormann, V., Lumpkin, R., and Foltz, G. R. (2013). The response of the surface circulation of the Arabian Sea to monsoonal forcing. J. Phys. Oceanogr. 43, 2008–2022. doi: 10.1175/JPO-D-13-033.1

CrossRef Full Text | Google Scholar

Bednaršek, N., Feely, R. A., Beck, M. W., Glippa, O., Kanerva, M., and Engström-Öst, J. (2018). El Niño-related thermal stress coupled with upwelling-related ocean acidification negatively impacts cellular to population-level responses in pteropods along the California Current System with implications for increased bioenergetic costs. Front. Mar. Sci. 4:486. doi: 10.3389/fmars.2018.00486

CrossRef Full Text | Google Scholar

Bednaršek, N., Feely, R. A., Reum, J. C. P., Peterson, B., Menkel, J., Alin, S. R., et al. (2014). Limacina helicina shell dissolution as an indicator of declining habitat suitability owing to ocean acidification in the California Current Ecosystem. Proc. Biol. Sci. 281:20140123. doi: 10.1098/rspb.2014.0123

PubMed Abstract | CrossRef Full Text | Google Scholar

Bednaršek, N., Feely, R. A., Tolimieri, N., Hermann, A. J., Siedlecki, S. A., Waldbusser, G. G., et al. (2017). Exposure history determines pteropod vulnerability to ocean acidification along the US West Coast. Sci. Rep. 7:4526. doi: 10.1038/s41598-017-03934-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Benazzouz, A., Mordane, S., Orbi, A., Chagdali, M., Hilmi, K., Atillah, A., et al. (2014). An improved coastal upwelling index from sea surface temperature using satellite-based approach—the case of the Canary Current upwelling system. Cont. Shelf Res. 81, 38–54. doi: 10.1016/j.csr.2014.03.012

CrossRef Full Text | Google Scholar

Bertrand, A., Chaigneau, A., Peraltilla, S., Ledesma, J., Graco, M., Monetti, F., et al. (2011). Oxygen: a fundamental property regulating pelagic ecosystem structure in the coastal southeastern tropical Pacific. PLoS One 6:e29558. doi: 10.1371/journal.pone.0029558

PubMed Abstract | CrossRef Full Text | Google Scholar

Bigorre, S. P., Weller, R. A., Edson, J. B., and Ware, J. D. (2013). A surface mooring for air–sea interaction research in the Gulf Stream. Part II: analysis of the observations and their accuracies. J. Atmos. Ocean. Technol. 30, 450–469. doi: 10.1175/JTECH-D-12-00078.1

CrossRef Full Text | Google Scholar

Boebel, O., Lutjeharms, J., Schmid, C., Zenk, W., Rossby, T., and Barron, C. (2003). The cape cauldron: a regime of turbulent inter-ocean exchange. Deep Sea Res. Part II 50, 57–86. doi: 10.1016/S0967-0645(02)00379-X

CrossRef Full Text | Google Scholar

Boening, C., Willis, J. K., Landerer, F. W., Nerem, R. S., and Fasullo, J. (2012). The 2011 La Niña: so strong, the oceans fell. Geophys. Res. Lett. 39:L19602. doi: 10.1029/2012GL053055

CrossRef Full Text | Google Scholar

Bond, N. A., Cronin, M. F., Sabine, C., Kawai, Y., Ichikawa, H., Freitag, P., et al. (2011). Upper ocean response to Typhoon Choi-wan as measured by the Kuroshio Extension Observatory mooring. J. Geophys. Res. 116:C02031. doi: 10.1029/2010JC006548

CrossRef Full Text | Google Scholar

Bowen, M., Markham, J., Sutton, P., Zhang, X., Wu, Q., Shears, N. T., et al. (2017). Interannual variability of sea surface temperature in the southwest Pacific and the role of ocean dynamics. J. Clim. 30, 7481–7492. doi: 10.1175/JCLI-D-16-0852.1

CrossRef Full Text | Google Scholar

Bower, A. S., and Hunt, H. D. (2000a). Lagrangian observations of the Deep Western Boundary Current in the North Atlantic Ocean. Part I: large-scale pathways and spreading rates. J. Phys. Oceanogr. 30, 764–783. doi: 10.1175/1520-0485(2000)030<0764:lootdw>2.0.co;2

CrossRef Full Text | Google Scholar

Bower, A. S., and Hunt, H. D. (2000b). Lagrangian observations of the Deep Western Boundary Current in the North Atlantic Ocean. Part II: the Gulf Stream–Deep Western Boundary Current crossover. J. Phys. Oceanogr. 30, 784–804. doi: 10.1175/1520-0485(2000)030<0784:lootdw>2.0.co;2

CrossRef Full Text | Google Scholar

Brady, R. X., Lovenduski, N. S., Alexander, M. A., Jacox, M., and Gruber, N. (2019). On the role of climate modes in modulating the air–sea CO2 fluxes in eastern boundary upwelling systems. Biogeosciences 16, 329–346. doi: 10.5194/bg-16-329-2019

CrossRef Full Text | Google Scholar

Brandini, F. P., Boltovskoy, D., Piola, A., Kocmur, S., Röttgers, R., Abreu, P. C., et al. (2000). Multiannual trends in fronts and distribution of nutrients and chlorophyll in the southwestern Atlantic (30–62 S). Deep Sea Res. Part I 47, 1015–1033. doi: 10.1016/s0967-0637(99)00075-8

CrossRef Full Text | Google Scholar

Brassington, G. B. (2010). Estimating surface divergence of ocean eddies using observed trajectories from a surface drifting buoy. J. Atmos. Ocean. Technol. 27, 705–720. doi: 10.1175/2009jtecho651.1

CrossRef Full Text | Google Scholar

Brassington, G. B., Summons, N., and Lumpkin, R. (2011). Observed and simulated Lagrangian and eddy characteristics of the East Australian Current and the Tasman Sea. Deep Sea Res. 58, 559–573. doi: 10.1016/j.dsr2.2010.10.001

CrossRef Full Text | Google Scholar

Bright, R. J., Xie, L., and Pietrafesa, L. J. (2002). Evidence of the Gulf Stream’s influence on tropical cyclone intensity. Geophys. Res. Lett. 29:1801. doi: 10.1029/2002GL014920

CrossRef Full Text | Google Scholar

Bushinsky, S. M., and Emerson, S. R. (2018). Biological and physical controls on the oxygen cycle in the Kuroshio Extension from an array of profiling floats. Deep Sea Res. I 141, 51–70. doi: 10.1016/j.dsr.2018.09.005

CrossRef Full Text | Google Scholar

Bushinsky, S. M., Emerson, S. R., Riser, S. C., and Swift, D. D. (2016). Accurate oxygen measurements on modified Argo floats using in situ air calibrations. Limnol. Oceanogr. Methods 14, 491–505. doi: 10.1002/lom3.10107

CrossRef Full Text | Google Scholar

Campagna, C., Piola, A. R., Marin, M. R., Lewis, M., and Fernández, T. (2006). Southern elephant seal trajectories, ocean fronts and eddies in the Brazil/Malvinas Confluence. Deep Sea Res. I 53, 1907–1924. doi: 10.1016/j.dsr.2006.08.015

CrossRef Full Text | Google Scholar

Capet, X., Estrade, P., Machu, É., Ndoye, S., Grelet, J., Lazar, A., et al. (2017). On the dynamics of the southern Senegal upwelling center: observed variability from synoptic to superinertial scales. J. Phys. Oceanogr. 47, 155–180. doi: 10.1175/JPO-D-15-0247.1

CrossRef Full Text | Google Scholar

Carr, M.-E., and Kearns, E. J. (2003). Production regimes in four eastern boundary current systems. Deep Sea Res. II 50, 3199–3221. doi: 10.1016/j.dsr2.2003.07.015

CrossRef Full Text | Google Scholar

Castelao, R., Glenn, S., and Schofield, O. (2010). Temperature, salinity, and density variability in the central Middle Atlantic Bight. J. Geophys. Res. 115:C10005. doi: 10.1029/2009JC006082

CrossRef Full Text | Google Scholar

Cavole, L. M., Demko, A. M., Diner, R. E., Giddings, A., Koester, I., Pagniello, C. M. L. S., et al. (2016). Biological impacts of the 2013–2015 warm-water anomaly in the Northeast Pacific: winners, losers, and the future. Oceanography 29, 273–285. doi: 10.5670/oceanog.2016.32

CrossRef Full Text | Google Scholar

Cazenave, A., Dieng, H.-B., Meyssignac, B., von Schuckmann, K., Decharme, B., and Berthier, E. (2014). The rate of sea-level rise. Nat. Clim. Change 4, 358–361. doi: 10.1038/nclimate2159

CrossRef Full Text | Google Scholar

Centurioni, L. R. (2018). “Drifter technology and impacts for sea surface temperature, sea-level pressure, and ocean circulation studies,” in Observing the Oceans in Real Time, eds R. Venkatesan, A. Tandon, E. D’Asaro, and M. A. Atmanand (Cham: Springer International Publishing), 37–57. doi: 10.1007/978-3-319-66493-4_3

CrossRef Full Text | Google Scholar

Centurioni, L. R., and Niiler, P. P. (2003). On the surface currents of the Caribbean Sea. Geophys. Res. Lett. 30:1279. doi: 10.1029/2002GL016231

CrossRef Full Text | Google Scholar

Centurioni, L. R., Hormann, V., Talley, L. D., Arzeno, I., Beal, L., Caruso, M., et al. (2017). Northern Arabian Sea Circulation–autonomous research (NASCar). A research initiative based on autonomous sensors. Oceanography 30, 74–87. doi: 10.5670/oceanog.2017.224

CrossRef Full Text | Google Scholar

Centurioni, L. R., Niiler, P. N., and Lee, D.-K. (2009). Near-surface circulation in the South China Sea during the winter monsoon. Geophys. Res. Lett. 36:L06605. doi: 10.1029/2008GL037076

CrossRef Full Text | Google Scholar

Centurioni, L. R., Niiler, P. P., and Lee, D.-K. (2004). Observations of inflow of Philippine sea surface water into the South China Sea through the Luzon Strait. J. Phys. Oceanogr. 34, 113–121. doi: 10.1175/1520-0485(2004)034<0113:ooiops>2.0.co;2

CrossRef Full Text | Google Scholar

Centurioni, L. R., Ohlmann, J. C., and Niiler, P. P. (2008). Permanent meanders in the California Current System. J. Phys. Oceanogr. 38, 1690–1710. doi: 10.1175/2008JPO3746.1

CrossRef Full Text | Google Scholar

Cetina Heredia, P., Roughan, M., Van Sebille, E., and Coleman, M. A. (2014). Long-term trends in the East Australian Current separation latitude and eddy driven transport. J. Geophys. Res. 119, 4351–4366. doi: 10.1002/2014JC010071

CrossRef Full Text | Google Scholar

Chan, H. Y., Xu, W. Z., Shin, P. K. S., and Cheung, S. G. (2008). Prolonged exposure to low dissolved oxygen affects early development and swimming behaviour in the gastropod Nassarius festivus (Nassariidae). Mar. Biol. 153, 735–743. doi: 10.1007/s00227-007-0850-6

CrossRef Full Text | Google Scholar

Chavez, F. P., and Messié, M. (2009). A comparison of eastern boundary upwelling ecosystems. Prog. Oceanogr. 83, 80–96. doi: 10.1016/j.pocean.2009.07.032

PubMed Abstract | CrossRef Full Text | Google Scholar

Chavez, F. P., and Toggweiler, J. R. (1995). “Physical estimates of global new production: the upwelling contribution,” in Upwelling in the Ocean: Modern Processes and Ancient Records, eds C. P. Summerhayes, et al. (Hoboken, NJ: John Wiley), 313–320.

Google Scholar

Chavez, F. P., Bertrand, A., Guevara, R., Soler, P., and Csirke, J. (2008). The northern Humboldt Current System: brief history, present status and a view towards the future. Prog. Oceanogr. 79, 95–105. doi: 10.1016/j.pocean.2008.10.012

CrossRef Full Text | Google Scholar

Checkley, D., and Barth, J. A. (2009). Patterns and processes in the California Current System. Prog. Oceanogr. 83, 49–64. doi: 10.1016/j.pocean.2009.07.028

CrossRef Full Text | Google Scholar

Chen, K., Gawarkiewicz, G., and Plueddemann, A. J. (2018). Atmospheric and offshore forcing of temperature variability at the shelf break: observations from the OOI Pioneer Array. Oceanography 31, 72–79. doi: 10.5670/oceanog.2018.112

CrossRef Full Text | Google Scholar

Chen, Z., Wu, L., Qiu, B., Li, L., Hu, D., Liu, C., et al. (2015). Strengthening Kuroshio observed at its origin during November 2010 to October 2012. Oceans 120, 2460–2470. doi: 10.1002/2014JC010590

CrossRef Full Text | Google Scholar

Claret, M., Galbraith, E. D., Palter, J. B., Bianchi, D., Fennel, K., Gilbert, D., et al. (2018). Rapid coastal deoxygenation due to ocean circulation shift in the northwest Atlantic. Nat. Clim. Chang. 8, 868–872. doi: 10.1038/s41558-018-0263-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Clayton, S., Dutkiewicz, S., Jahn, O., and Follows, M. J. (2013). Dispersal, eddies, and the diversity of marine phytoplankton. Limnol. Oceanogr. 3, 182–197. doi: 10.1215/21573689-2373515

CrossRef Full Text | Google Scholar

Clayton, S., Lin, Y. C., Follows, M. J., and Worden, A. Z. (2017). Co-existence of distinct Ostreococcus ecotypes at an oceanic front. Limnol. Oceanogr. 62, 75–88. doi: 10.1002/lno.10373

CrossRef Full Text | Google Scholar

Clayton, S., Nagai, T., and Follows, M. J. (2014). Fine scale phytoplankton community structure across the Kuroshio Front. J. Plankton Res. 36, 1017–1030. doi: 10.1093/plankt/fbu020

CrossRef Full Text | Google Scholar

Cockcroft, A. C., Schoeman, D. S., Pitcher, G. C., Bailey, G. W., and van Zyl, D. L. (2000). “A mass standing, or “walkout,” of west coast rock lobster Jasus lalandii in Elands Bay, South Africa: causes, results and implications,” in The Biodiversity Crisis and Crustacea, Vol. 12, eds J. C. von Kaupel Klein and F. R. Schram (Rotterdam: A. A. Balkema), 673–688.

Google Scholar

Cockcroft, A. C., van Zyl, D., and Hutchings, L. (2008). Large-scale changes in the spatial distribution of South African West Coast rock lobsters: an overview. Afr. J. Mar. Sci. 30, 149–159. doi: 10.2989/ajms.2008.30.1.15.465

CrossRef Full Text | Google Scholar

Colas, F., Capet, X., McWilliams, J. C., and Shchepetkin, A. (2008). 1997–1998 El Niño off Peru: a numerical study. Prog. Oceanogr. 79, 138–155.

Google Scholar

Colas, F., McWilliams, J. C., Capet, X., and Kurian, J. (2012). Heat balance and eddies in the Peru-Chile current system. Clim. Dyn. 39, 509–529. doi: 10.1007/s00382-011-1170-6

CrossRef Full Text | Google Scholar

Conway, T. M., Palter, J. B., and de Souza, G. F. (2018). Gulf Stream rings as a source of iron to the North Atlantic subtropical gyre. Nat. Geosci. 11, 594–598. doi: 10.1038/s41561-018-0162-0

CrossRef Full Text | Google Scholar

Cravatte, S., Ganachaud, A., Sprintall, J., Alberty, M., Germineaud, C., Brachet, C., et al. (2019). Solomon Sea SPICE Mooring Data. La Jolla, CA: UC San Diego Library Digital Collections, doi: 10.6075/J09W0CS2

CrossRef Full Text | Google Scholar

Cronin, M. F., Bond, N. A., Farrar, J. T., Ichikawa, H., Jayne, S. R., Kawai, Y., et al. (2013). Formation and erosion of the seasonal thermocline in the Kuroshio Extension recirculation gyre. Deep Sea Res. 85, 62–74. doi: 10.1016/j.dsr2.2012.07.018

CrossRef Full Text | Google Scholar

Cronin, M. F., Gentemann, C. L., Edson, J., Ueki, I., Bourassa, M., Brown, S., et al. (2019). Air–sea fluxes with a focus on heat and momentum. Front. Mar. Sci. 6:430. doi: 10.3389/fmars.2019.00430

CrossRef Full Text | Google Scholar

Cronin, M. F., Pelland, N. A., Emerson, S. R., and Crawford, W. R. (2015). Estimating diffusivity from the mixed layer heat and salt balances in the North Pacific. J. Geophys. Res. Oceans 120, 7346–7362. doi: 10.1002/2015JC011010

CrossRef Full Text | Google Scholar

D’Adamo, N., Fandry, C., Bubarton, S., and Domingues, C. (2009). Northern sources of the Leeuwin Current and the “Holloway Current” on the North West Shelf. J. R. Soc. West Aust. 92, 53–66.

Google Scholar

da Silveira, I. C. A., deMiranda, L. B., and Brown, W. S. (1994). On the origins of the North Brazil Current. J. Geophys. Res. 99, 22501–22512. doi: 10.1029/94JC01776

CrossRef Full Text | Google Scholar

Davis, R. (2016). Data From: Solomon Sea Ocean Transport From Gliders. La Jolla, CA: Scripps Institution of Oceanography, Instrument Development Group, doi: 10.21238/s8SPRAY2718

CrossRef Full Text | Google Scholar

Davis, R. E., Kessler, W. S., and Sherman, J. T. (2012). Gliders measure western boundary current transport from the South Pacific to the Equator. J. Phys. Oceanogr. 42, 2001–2013. doi: 10.1175/JPO-D-12-022.1

CrossRef Full Text | Google Scholar

de Ruijter, W. P. M., Biastoch, A., Drijfhout, S. S., Lutjeharms, J. R. E., Matano, R. P., Pichevin, T., et al. (1999). Indian-Atlantic interocean exchange: dynamics, estimation and impact. J. Geophys. Res. 104, 20885–20910. doi: 10.1029/1998JC900099

CrossRef Full Text | Google Scholar

Dengler, M., Fischer, J., Schott, F. A., and Zantopp, R. (2006). Deep Labrador Current and its variability in 1996–2005. Geophys. Res. Lett. 33:L21S06. doi: 10.1029/2006GL026702

CrossRef Full Text | Google Scholar

Dever, M., Hebert, D., Greenan, B. J. W., Sheng, J., and Smith, P. C. (2016). Hydrography and coastal circulation along the Halifax Line and the connections with the Gulf of St. Lawrence. Atmos. Ocean 54, 199–217. doi: 10.1080/07055900.2016.1189397

CrossRef Full Text | Google Scholar

DeVries, T. (2014). The oceanic anthropogenic CO2 sink: storage, air–sea fluxes, and transports over the industrial era. Glob. Biogeochem. Cycles 28, 631–647. doi: 10.1002/2013GB004739

CrossRef Full Text | Google Scholar

Dewitte, B., Vazquez-Cuervo, J., Goubanova, K., Illig, S., Takahashi, K., Cambon, G., et al. (2012). Change in El Niño flavours over 1958–2008: implications for the long-term trend of the upwelling off Peru. Deep Sea Res. Part II 77, 143–156.

Google Scholar

deYoung, B., von Oppeln-Bronikowski, N., Matthews, J. B. R., and Bachmayer, R. (2018). Glider operations in the Labrador Sea. J. Ocean Tech. 13, 108–120.

Google Scholar

Di Lorenzo, E., and Mantua, N. (2016). Multi-year persistence of the 2014/15 North Pacific marine heatwave. Nat. Clim. Change 6, 1042–1047. doi: 10.1038/nclimate3082

CrossRef Full Text | Google Scholar

Diakhaté, M., de Coëtlogon, G., Lazar, A., Wade, M., and Gaye, A. T. (2016). Intraseasonal variability of tropical Atlantic sea-surface temperature: air–sea interaction over upwelling fronts. Q. J. R. Meteorol. Soc. 142, 372–386. doi: 10.1002/qj.2657

CrossRef Full Text | Google Scholar

Domingues, C. M., Wijffels, S. E., Maltrud, M. E., Church, J. A., and Tomczak, M. (2006). Role of eddies in cooling the Leeuwin Current. Geophys. Res. Lett. 33:L05603. doi: 10.1029/2005GL025216

CrossRef Full Text | Google Scholar

Domingues, R., Baringer, M., and Goni, G. (2016). Remote sources for year-to-year changes in the seasonality of the Florida Current transport. J. Geophys. Res. 121, 7547–7559. doi: 10.1002/2016JC012070

CrossRef Full Text | Google Scholar

Domingues, R., Goni, G., Baringer, M., and Volkov, D. L. (2018). What caused the accelerated sea level changes along the United States East Coast during 2010–2015? Geophys. Res. Lett. 45, 13367–13376. doi: 10.1029/2018GL081183

CrossRef Full Text | Google Scholar

Domingues, R., Kuwano-Yoshida, A., Chardon-Maldonado, P., Todd, R. E., Halliwell, G. R., Kim, H.-S., et al. (2019). Ocean observations in support of studies and forecasts of tropical and extratropical cyclones. Front. Mar. Sci. 6:446. doi: 10.3389/fmars.2019.00446

CrossRef Full Text | Google Scholar

Dong, S., Goni, G., and Bringas, F. (2015). Temporal variability of the south Atlantic Meridional Overturning Circulation between 20°S and 35°S. Geophys. Res. Lett. 42, 7655–7662. doi: 10.1002/2015GL065603

CrossRef Full Text | Google Scholar

Donohue, K. A., Watts, D. R., Tracey, K. L., Greene, A. D., and Kennelly, M. (2010). Mapping circulation in the Kuroshio Extension with an array of current and pressure recording inverted echo sounders. J. Atmos. Ocean. Technol. 27, 507–527. doi: 10.1175/2009JTECHO686.1

CrossRef Full Text | Google Scholar

Douglass, E., Roemmich, D., and Stammer, D. (2010). Interannual variability in North Pacific heat and freshwater budgets. Deep Sea Res. Part II 57, 1127–1140. doi: 10.1016/j.dsr2.2010.01.001

CrossRef Full Text | Google Scholar

Duncombe Rae, C. M. (2005). A demonstration of the hydrographic partition of the Benguela upwelling ecosystem at 26’40°S. Afr. J. Mar. Sci. 27, 617–628. doi: 10.2989/18142320509504122

CrossRef Full Text | Google Scholar

Elipot, S., and Beal, L. M. (2015). Characteristics, energetics, and origins of Agulhas Current meanders and their limited influence on ring shedding. J. Phys. Oceanogr. 45, 2294–2314. doi: 10.1175/JPO-D-14-0254.1

CrossRef Full Text | Google Scholar

Elipot, S., and Beal, L. M. (2018). Observed Agulhas Current sensitivity to interannual and long-term trend atmospheric forcings. J. Clim. 31, 3077–3098. doi: 10.1175/JCLI-D-17-0597.1

CrossRef Full Text | Google Scholar

Escribano, R., and Morales, C. E. (2012). Spatial and temporal scales of variability in the coastal upwelling and coastal transition zone off central-southern Chile (35–40°S). Progr. Oceanogr. 92, 1–7. doi: 10.1016/j.pocean.2011.07.019

CrossRef Full Text | Google Scholar

Espinoza, P., Lorrain, A., Ménard, F., Cherel, Y., Tremblay-Boyer, L., Argüelles, J., et al. (2017). Trophic structure in the northern Humboldt Current System: new perspectives from stable isotope analysis. Mar. Biol. 164:86. doi: 10.1007/s00227-017-3119-8

CrossRef Full Text | Google Scholar

Everett, J. D., Baird, M. E., Roughan, M., Suthers, I. M., and Doblin, M. A. (2014). Relative impact of seasonal and oceanographic drivers on surface chlorophyll a along a Western Boundary Current. Prog. Oceanogr. 120, 340–351. doi: 10.1016/j.pocean.2013.10.016

CrossRef Full Text | Google Scholar

Ezer, T. (2015). Detecting changes in the transport of the Gulf Stream and the Atlantic overturning circulation from coastal sea level data: the extreme decline in 2009–2010 and estimated variations for 1935–2012, Global Planet. Change 129, 23–36. doi: 10.1016/j.gloplacha.2015.03.002

CrossRef Full Text | Google Scholar

Fassbender, A. J., Alin, S. R., Feely, R. A., Sutton, A. J., Newton, J. A., Krembs, C., et al. (2018). Seasonal carbonate chemistry variability in marine surface waters of the US Pacific Northwest. Earth Syst. Sci. Data 10, 1367–1401. doi: 10.5194/essd-10-1367-2018

CrossRef Full Text | Google Scholar

Fassbender, A. J., Alin, S. R., Feely, R. A., Sutton, A. J., Newton, J. A., and Byrne, R. H. (2017a). Estimating total alkalinity in the Washington State coastal zone: complexities and surprising utility for ocean acidification research. Estuar. Coast. 40, 404–418. doi: 10.1007/s12237-016-0168-z

CrossRef Full Text | Google Scholar

Fassbender, A. J., Sabine, C. L., Cronin, M. F., and Sutton, A. J. (2017b). Mixed-layer carbon cycling at the Kuroshio extension observatory. Glob. Biogeochem. Cycles 31, 272–288. doi: 10.1002/2016GB005547

CrossRef Full Text | Google Scholar

Fassbender, A. J., Sabine, C. L., and Feifel, K. M. (2016). Consideration of coastal carbonate chemistry in understanding biological calcification. Geophys. Res. Lett. 43, 4467–4476. doi: 10.1002/2016gl068860

CrossRef Full Text | Google Scholar

Fassbender, A. J., Sabine, C. L., Feely, R. A., Langdon, C., and Mordy, C. W. (2011). Inorganic carbon dynamics during northern California coastal upwelling. Cont. Shelf Res. 31, 1180–1192. doi: 10.1016/j.csr.2011.04.006

CrossRef Full Text | Google Scholar

Faye, S., Lazar, A., Sow, B. A., and Gaye, A. T. (2015). A model study of the seasonality of sea surface temperature and circulation in the Atlantic north-eastern tropical upwelling system. Front. Phys. 3:76. doi: 10.3389/fphy.2015.00076

CrossRef Full Text | Google Scholar

Feely, R. A., Alin, S., Carter, B., Bednaršek, N., Hales, B., Chan, F., et al. (2016). Chemical and biological impacts of ocean acidification along the west coast of North America. Estuar. Coast. Shelf Sci. 183, 260–270. doi: 10.1016/j.ecss.2016.08.043

CrossRef Full Text | Google Scholar

Feely, R. A., Okazaki, R. R., Cai, W.-J., Bednaršek, N., Alin, S. R., Byrne, R. H., et al. (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. Cont. Shelf Res. 152, 50–60. doi: 10.1016/j.csr.2017.11.002

CrossRef Full Text | 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, 1490–1492. doi: 10.1126/science.1155676

PubMed Abstract | CrossRef Full Text | Google Scholar

Feng, M., Hendon, H. H., Xie, S. P., Marshall, A. G., Schiller, A., Kosaka, Y., et al. (2015). Decadal increase in Ningaloo Niño since the late 1990s. Geophys. Res. Lett. 42, 104–112. doi: 10.1002/2014GL062509

CrossRef Full Text | Google Scholar

Feng, M., McPhaden, M. J., Xie, S. P., and Hafner, J. (2013). La Niña forces unprecedented Leeuwin Current warming in 2011. Sci. Rep. 3:1277. doi: 10.1038/srep01277

PubMed Abstract | CrossRef Full Text | Google Scholar

Feng, M., Meyer, G., Pearce, A., and Wijffels, S. (2003). Annual and interannual variations of the Leeuwin Current at 32S. J. Geophys. Res. 108:3355. doi: 10.1029/2002JC001763

CrossRef Full Text | Google Scholar

Fennel, K., Wilkin, J., Levin, J., Moisan, J., O’Reilly, J., and Haidvogel, D. (2006). Nitrogen cycling in the middle atlantic bight: results from a three-dimensional model and implications for the North Atlantic nitrogen budget. Glob. Biogeochem. Cycles 20:GB3007. doi: 10.1029/2005GB002456

CrossRef Full Text | Google Scholar

Fernandez, D., Bowen, M., and Sutton, P. (2018). Variability, coherence and forcing mechanisms in the New Zealand ocean boundary currents. Prog. Oceanogr. 165, 1680–1688. doi: 10.1016/j.pocean.2018.06.002

CrossRef Full Text | Google Scholar

Ferrari, R., Artana, C., Saraceno, M., Piola, A. R., and Provost, C. (2017). Satellite altimetry and current-meter velocities in the Malvinas Current at 41°S: comparisons and modes of variations. J. Geophys. Res. 122, 9572–9590. doi: 10.1002/2017JC013340

CrossRef Full Text | Google Scholar

Fischer, J., Schott, F., and Dengler, M. (2004). Boundary circulation at the exit of the Labrador Sea. J. Phys. Oceanogr. 34, 1548–1570. doi: 10.1175/1520-0485(2004)034<1548:bcateo>2.0.co;2

CrossRef Full Text | Google Scholar

Fischer, J., Visbeck, M., Zantopp, R., and Nunes, N. (2010). Interannual to decadal variability of outflow from the Labrador Sea. Geophys. Res. Lett. 37:L24610. doi: 10.1029/2010GL045321

CrossRef Full Text | Google Scholar

Frajka-Williams, E., Ansorge, I. J., Baehr, J., Bryden, H. L., Chidichimo, M. P., Cunningham, S. A., et al. (2019). Atlantic Meridional Overturning Circulation: observed transport and variability. Front. Mar. Sci. 6:260. doi: 10.3389/fmars.2019.00260

CrossRef Full Text | Google Scholar

Fratantoni, D. M., and Richardson, P. L. (2006). The evolution and demise of North Brazil current rings. J. Phys. Oceanogr. 36, 1241–1264. doi: 10.1175/JPO2907.1

CrossRef Full Text | Google Scholar

Fratantoni, P. S., and Pickart, R. S. (2007). The western North Atlantic shelfbreak current system in summer. J. Phys. Oceanogr. 37, 2509–2533. doi: 10.1175/JPO3123

CrossRef Full Text | Google Scholar

Fréon, P., Coetzee, J. C., Van der Lingen, C. D., Connell, A. D., O’Donoghue, S. H., Roberts, M. J., et al. (2010). A review and tests of hypotheses about causes of the KwaZulu-Natal sardine run. Afr. J. Mar. Sci. 32, 449–479. doi: 10.2989/1814232X.2010.519451

CrossRef Full Text | Google Scholar

Friederich, G. E., Walz, P. M., Burczynski, M. G., and Chavez, F. P. (2002). Inorganic carbon in the central California upwelling system during the 1997–1999 El Niño–La Niña event. Prog. Oceanogr. 54, 185–203. doi: 10.1016/S0079-6611(02)00049-6

CrossRef Full Text | Google Scholar

Frölicher, T. L., and Laufkötter, C. (2018). Emerging risks from marine heat waves. Nat. Commun. 9:650. doi: 10.1038/s41467-018-03163-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Fujioka, K., Fukuda, H., Furukawa, S., Tei, Y., Okamoto, S., and Ohshimo, S. (2018). Habitat use and movement patterns of small (age-0) juvenile Pacific bluefin tuna (Thunnus orientalis) relative to the Kuroshio. Fish. Oceanogr. 27, 185–198. doi: 10.1111/fog.12244

CrossRef Full Text | Google Scholar

Furue, R., Guerreiro, K., Phillips, H. E., McCreary, J. P., and Bindoff, N. (2017). On the Leeuwin Current system and its linkage to zonal flows in the South Indian ocean as inferred from a gridded hydrography. J. Phys. Oceanogr. 47, 583–602. doi: 10.1175/JPO-D-16-0170.1

CrossRef Full Text | Google Scholar

Galarneau, T. J. Jr., Davis, C. A., and Shapiro, M. A. (2013). Intensification of Hurricane Sandy (2012) through extratropical warm core seclusion. Mon. Weather Rev. 141, 4296–4321. doi: 10.1175/MWR-D-13-00181.1

CrossRef Full Text | Google Scholar

Ganachaud, A., Cravatte, S., Melet, A., Schiller, A., Holbrook, N. J., Sloyan, B. M., et al. (2014). The Southwest Pacific Ocean Circulation and Climate Experiment (SPICE). J. Geophys. Res. Oceans 119, 7660–7686. doi: 10.1002/2013JC00967

CrossRef Full Text | Google Scholar

Ganachaud, A., Cravatte, S., Sprintall, J., Germineaud, C., Alberty, M., Jeandel, C., et al. (2017). The Solomon Sea: its circulation, chemistry, geochemistry and biology explored during two oceanographic cruises. Elem. Sci. Anth. 5:33. doi: 10.1525/elementa.221

CrossRef Full Text | Google Scholar

Garzoli, S. L., Baringer, M. O., Dong, S., Perez, R. C., and Yao, Q. (2012). South Atlantic meridional fluxes. Deep Sea Res. Part I 71, 21–32. doi: 10.1016/j.dsr.2012.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Gaube, P., Braun, C. D., Lawson, G. L., McGillicuddy, D. J., Della Penna, A., Skomal, G., et al. (2018). Mesoscale eddies influence the movements of mature female white sharks in the Gulf Stream and Sargasso Sea. Sci. Rep. 8:7363. doi: 10.1038/s41598-018-25565-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Gawarkiewicz, G. G., and Linder, C. A. (2006). Lagrangian flow patterns north of Cape Hatteras using near-surface drifters. Prog. Oceanogr. 70, 181–195. doi: 10.1016/j.pocean.2006.03.020

CrossRef Full Text | Google Scholar

Gawarkiewicz, G., Todd, R. E., Zhang, W., Partida, J., Gangopadhyay, A., Monim, M.-U.-H., et al. (2018). The changing nature of shelf–break exchange revealed by the OOI Pioneer Array. Oceanography 31, 60–70. doi: 10.5670/oceanog.2018.110

CrossRef Full Text | Google Scholar

Gay, P. S., and Chereskin, T. K. (2009). Mean structure and seasonal variability of the poleward undercurrent off southern California. J. Geophys. Res. 114:C02007. doi: 10.1029/2008JC004886

CrossRef Full Text | Google Scholar

Godfrey, J. S., and Weaver, A. J. (1991). Is the Leeuwin current driven by Pacific heating and winds? Prog. Oceanogr. 27, 225–272. doi: 10.1016/0079-6611(91)90026-I

CrossRef Full Text | Google Scholar

Gómez-Letona, M., Ramos, A. G., Coca, J., and Arístegui, J. (2017). Trends in primary production in the Canary Current upwelling system—a regional perspective comparing remote sensing models. Front. Mar. Sci. 4:370. doi: 10.3389/fmars.2017.00370

CrossRef Full Text | Google Scholar

Goni, G. J., Sprintall, J., Bringas, F., Cheng, L., Cirano, M., Dong, S., et al. (2019). More than 50 years of continuous temperature section measurements by the Global Expendable Bathythermograph (XBT) Network, its integrability, societal benefits, and future. Front. Mar. Sci. 6:452. doi: 10.3389/fmars.2019.00452

CrossRef Full Text | Google Scholar

Gopalakrishnan, G., Cornuelle, B. D., Hoteit, I., Rudnick, D. L., and Owens, W. B. (2013). State estimates and forecasts of the Loop Current in the Gulf of Mexico using the MITgcm and its adjoint. J. Geophys. Res. 118, 3292–3314. doi: 10.1002/jgrc.20239

CrossRef Full Text | Google Scholar

Gordon, A. L., Flament, P., Villanoy, C., and Centurioni, L. (2014). The nascent Kuroshio of Lamon Bay. J. Geophys. Res. 119, 4251–4263. doi: 10.1002/2014JC009882

CrossRef Full Text | Google Scholar

Gorgues, T., Aumont, O., and Rodgers, K. B. (2010). A mechanistic account of increasing seasonal variations in the rate of ocean uptake of anthropogenic carbon. Biogeosciences 7, 2581–2589. doi: 10.5194/bg-7-2581-2010

CrossRef Full Text | Google Scholar

Graco, M., Purca, S., Dewitte, B., Castro, C., Morón, O., Ledesma, J., et al. (2017). The OMZ and nutrient features as a signature of interannual and low-frequency variability in the Peruvian upwelling system. Biogeosciences 14, 4601–4617. doi: 10.5194/bg-14-1-2017

CrossRef Full Text | Google Scholar

Grados, C., Chaigneau, A., Echevin, V., and Dominguez, N. (2018). Upper ocean hydrology of the northern Humboldt Current System at seasonal, interannual and interdecadal scales. Prog. Oceanogr. 165, 123–144. doi: 10.1016/j.pocean.2018.05.005

CrossRef Full Text | Google Scholar

Gruber, N., Keeling, C. D., and Bates, N. R. (2002). Interannual variability in the North Atlantic Ocean carbon sink. Science 298, 2374–2378. doi: 10.1126/science.1077077

PubMed Abstract | CrossRef Full Text | Google Scholar

Guerrero, R. A., Piola, A. R., Fenco, H., Matano, R. P., Combes, V., Chao, Y., et al. (2014). The salinity signature of the cross-shelf exchanges in the Southwestern Atlantic Ocean: satellite observations. J. Geophys. Res. Oceans 119, 7794–7810. doi: 10.1002/2014JC010113

PubMed Abstract | CrossRef Full Text | Google Scholar

Gula, J., Blacic, T. M., and Todd, R. E. (2019). Submesoscale coherent vortices in the Gulf Stream. Geophys. Res. Lett. 46, 2704–2714. doi: 10.1029/2019GL081919

CrossRef Full Text | Google Scholar

Guo, X. Y., Zhu, X. H., Long, Y., and Huang, D. J. (2013). Spatial variations in the Kuroshio nutrient transport from the East China Sea to south of Japan. Biogeosciences 10, 6403–6417. doi: 10.5194/bg-10-6403-2013

CrossRef Full Text | Google Scholar

Guo, X., Zhu, X. H., Wu, Q. S., and Huang, D. (2012). The Kuroshio nutrient stream and its temporal variation in the East China Sea. J. Geophys. Res. Oceans 117:C01026. doi: 10.1029/2011JC007292

CrossRef Full Text | Google Scholar

Gutiérrez, D. (2016). Variabilidad climática, procesos oceanográficos y producción primaria frente al Perú. Boletín Técnico “Generación de información y monitoreo del Fenómeno El Niño.”. Instituto Geofísico del Perú 3, 4–8.

Google Scholar

Haines, S., Seim, H., and Muglia, M. (2017). Implementing quality control of high-frequency radar estimates and application to Gulf Stream surface currents. J. Atmos. Ocean. Technol. 34, 1207–1224. doi: 10.1175/JTECH-D-16-0203.1

CrossRef Full Text | Google Scholar

Hales, B., Takahashi, T., and Bandstra, L. (2005). Atmospheric CO2 uptake by a coastal upwelling system. Glob. Biogeochem. Cycles 19:GB1009. doi: 10.1029/2004GB002295

CrossRef Full Text | Google Scholar

Halliwell, G. R., Mehari, M., Shay, L. K., Kourafalou, V. H., Kang, H., Kim, J.-S., et al. (2017). OSSE quantitative assessment of rapid-response prestorm ocean surveys to improve coupled tropical cyclone prediction. J. Geophys. Res. Oceans 122, 5729–5748. doi: 10.1002/2017JC012760

CrossRef Full Text | Google Scholar

Hansen, D. V., and Poulain, P.-M. (1996). Quality control and interpolation of WOCE/TOGA drifter data. J. Atmos. Ocean. Technol. 13, 900–909. doi: 10.1175/1520-0426(1996)013<0900:qcaiow>2.0.co;2

CrossRef Full Text | Google Scholar

Harris, K. E., DeGrandpre, M. D., and Hales, B. (2013). Aragonite saturation states in a coastal upwelling zone. Geophys. Res. Lett. 40, 2720–2725. doi: 10.1002/grl.50460

CrossRef Full Text | Google Scholar

He, R., Chen, K., Moore, T., and Li, M. (2010). Mesoscale variations of sea surface temperature and ocean color patterns at the Mid-Atlantic Bight shelfbreak. Geophys. Res. Lett. 37:L09607. doi: 10.1029/2010GL042658

CrossRef Full Text | Google Scholar

Helly, J. J., and Levin, L. A. (2004). Global distribution of naturally occurring marine hypoxia on continental margins. Deep Sea Res. I 51, 1159–1168. doi: 10.1016/j.dsr.2004.03.009

CrossRef Full Text | Google Scholar

Henderikx Freitas, F., Saldiìas, G. S., Goñi, M., Shearman, R. K., and White, A. E. (2018). Temporal and spatial dynamics of physical and biological properties along the Endurance Array of the California Current Ecosystem. Oceanography 31, 80–89. doi: 10.5670/oceanog.2018.113

CrossRef Full Text | Google Scholar

Henderikx Freitas, F., Siegel, D. A., Washburn, L., Halewood, S., and Stassinos, E. (2016). Assessing controls on cross-shelf phytoplankton and suspended particle distributions using repeated bio-optical glider surveys. Oceans 121, 7776–7794. doi: 10.1002/2016JC011781

CrossRef Full Text | Google Scholar

Heslop, E. E., Ruiz, S., Allen, J., López-Jurado, J. L., Renault, L., and Tintoré, J. (2012). Autonomous underwater gliders monitoring variability at “choke points” in our ocean system: a case study in the Western Mediterranean Sea. Geophys. Res. Lett. 39:L20604.

Google Scholar

Hill, K., Moltmann, T., Meyers, G., and Proctor, R. (2010). “The Australian integrated marine observing system (IMOS),” in Proceedings of OceanObs’09: Sustained Ocean Observations and Information for Society, eds J. Hall, D. E. Harrison, and D. Stammer (Venice: ESA Publication WPP-306), doi: 10.5270/OceanObs09

CrossRef Full Text | Google Scholar

Hill, K., Rintoul, S. R., Ridgway, K. R., and Oke, P. R. (2011). Decadal changes in the South Pacific western boundary current system revealed in observations and ocean state estimates. J. Geophys. Res. 116:C01009. doi: 10.1029/2009JC005926

CrossRef Full Text | Google Scholar

Hobday, A. J., Alexander, L. V., Perkins, S. E., Smale, D. A., Straub, S. C., Oliver, E. C. J., et al. (2016). A hierarchical approach to defining marine heatwaves. Prog. Oceanogr. 141, 227–238. doi: 10.1016/j.pocean.2015.12.014

CrossRef Full Text | Google Scholar

Hoffman, R. N., and Atlas, R. (2016). Future observing system simulation experiments. Bull. Am. Meteor. Soc. 97, 1601–1616. doi: 10.1175/BAMS-D-15-00200.1

CrossRef Full Text | Google Scholar

Honda, M. C., Sasai, Y., Siswanto, E., Kuwano-Yoshida, A., Aiki, H., and Cronin, M. F. (2018). Impact of cyclonic eddies and typhoons on biogeochemistry in the oligotrophic ocean based on biogeochemical/physical/meteorological time-series at station KEO. Prog. Earth Planet. Sci. 5:42. doi: 10.1186/s40645-018-0196-3

CrossRef Full Text | Google Scholar

Houpert, L., Inall, M. E., Dumont, E., Gary, S., Johnson, C., Porter, M., et al. (2018). Structure and transport of the North Atlantic Current in the eastern subpolar gyre from sustained glider observations. J. Geophys. Res. Oceans 123, 6019–6038. doi: 10.1029/2018JC014162

CrossRef Full Text | Google Scholar

Howatt, T., Palter, J. B., Matthews, J. B., deYoung, B., Bachmayer, R., and Claus, B. (2018). Ekman and eddy exchange of freshwater and oxygen across the Labrador Shelf Break. J. Phys. Oceanogr. 48, 1015–1031. doi: 10.1175/JPO-D-17-0148.1

CrossRef Full Text | Google Scholar

Hu, D., Hu, S., Wu, L., Li, L., Zhang, L., Diao, X., et al. (2013). Direct measurements of the Luzon undercurrent. J. Phys. Oceanogr. 43, 1417–1425. doi: 10.1175/JPO-D-12-0165.1

CrossRef Full Text | Google Scholar

Hu, D., Wu, L., Cai, W., Sen Gupta, A., Ganachaud, A., Qiu, B., et al. (2015). Pacific western boundary currents and their roles in climate. Nature 522, 299–308. doi: 10.1038/nature14504

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, S., Hu, D., Guan, C., Wang, F., Zhang, L., Wang, F., et al. (2016). Interannual variability of the Mindanao Current/Undercurrent in direct observations and numerical simulations. J. Phys. Oceanogr. 46, 483–499. doi: 10.1175/JPO-D-15-0092.1

CrossRef Full Text | Google Scholar

Hummels, R., Brandt, P., Dengler, M., Fischer, J., Araujo, M., Veleda, D., et al. (2015). Interannual to decadal changes in the western boundary circulation in the Atlantic at 11 degrees S. Geophys. Res. Lett. 42, 7615–7622. doi: 10.1002/2015gl065254

CrossRef Full Text | Google Scholar

Hutchings, L., Augustyn, C. J., Cockcroft, A., Van der Lingen, C., Coetzee, J., Leslie, R. W., et al. (2009a). Marine fisheries monitoring programmes in South Africa. S. Afr. J. Sci. 105, 182–192.

PubMed Abstract | Google Scholar

Hutchings, L., Roberts, M. J., and Verheye, H. M. (2009b). Marine environmental monitoring programmes in South Africa: a review. S. Afr. J. Sci. 105, 94–102.

Google Scholar

Hutchings, L., van der Lingen, C., Shannon, L. J., Cawford, R. J. M., Verheye, H. M., Bartholomae, C. H., et al. (2009c). The Benguela Current: an ecosystem of four components. Progr. Oceanogr. 83, 15–32. doi: 10.1016/j.pocean.2009.07.046

CrossRef Full Text | Google Scholar

Huyer, A., Wheeler, P. A., Strub, P. T., Smith, R. L., Letelier, R., and Kosro, P. M. (2007). The Newport Line off Oregon—studies in the northeast Pacific. Prog. Oceanogr. 75, 126–160. doi: 10.1016/j.pocean.2007.08.003

CrossRef Full Text | Google Scholar

Ichikawa, H., Nakamura, H., Nishina, A., and Higashi, M. (2004). Variability of northeastward current southeast of Northern Ryukyu Islands. J. Oceanogr. 60, 351–363. doi: 10.1023/B:JOCE.0000038341.27622.73

CrossRef Full Text | Google Scholar

Imawaki, S., Bower, A., Beal, L., and Qiu, B. (2013). “Western boundary currents,” in Ocean Circulation and Climate – a 21st Century Perspective, eds G. Siedler, S. M. Griffies, J. Gould, and J. A. Church (Amsterdam: Elsevier Academic Press), 305–338. doi: 10.1016/B978-0-12-391851-2.00013-1

CrossRef Full Text | Google Scholar

Imawaki, S., Uchida, H., Ichikawa, H., Fuksawa, M., Umatani, S., and The Asuka group. (2001). Satellite altimeter monitoring the Kuroshio transport south of Japan. Geophys. Res. Lett. 28, 17–20. doi: 10.1029/2000GL011796

CrossRef Full Text | Google Scholar

Inoue, R., Honda, M., Fujiki, T., Matsumoto, K., Kouketsu, S., Suga, T., et al. (2016a). Western North Pacific integrated physical-biogeochemical ocean observation experiment (INBOX): Part 2. Biogeochemical responses to eddies and typhoons revealed from the S1 mooring and shipboard measurements. J. Mar. Res. 74, 71–99. doi: 10.1357/002224016819257335

CrossRef Full Text | Google Scholar

Inoue, R., Suga, T., Kouketsu, S., Kita, T., Hosoda, S., Kobayashi, T., et al. (2016b). Western north Pacific integrated physical–biogeochemical ocean observation experiment (INBOX): Part 1. Specifications and chronology of the S1-INBOX floats. J. Mar. Res. 74, 43–69. doi: 10.1357/002224016819257344

CrossRef Full Text | Google Scholar

Ito, S., Yoshie, N., Okunishi, T., Ono, T., Okazaki, Y., Kuwataet, A., et al. (2010). Application of an automatic approach to calibrate the NEMURO nutrient-phytoplankton-zooplankton food web model in the Oyashio region. Progr. Oceanogr. 87, 186–200. doi: 10.1016/j.pocean.2010.08.004

CrossRef Full Text | Google Scholar

Ito, T., and Follows, M. J. (2003). Upper ocean control on the solubility pump of CO2. J. Mar. Res. 61, 465–489. doi: 10.1357/002224003322384898

CrossRef Full Text | Google Scholar

Iudicone, D., Rodgers, K. B., Plancherel, Y., Aumont, O., Ito, T., Key, R. M., et al. (2016). The formation of the ocean’s anthropogenic carbon reservoir. Sci. Rep. 6:35473. doi: 10.1038/srep35473

PubMed Abstract | CrossRef Full Text | Google Scholar

Jayne, S. R., Roemmich, D., Zilberman, N., Riser, S. C., Johnson, K. S., Johnson, G. C., et al. (2017). The Argo program: present and future. Oceanography 30, 18–28. doi: 10.5670/oceanog.2017.213

CrossRef Full Text | Google Scholar

Johns, W. E., Baringer, M. O., Beal, L. M., Cunningham, S. A., Kanzow, T., Bryden, H. L., et al. (2011). Continuous array-based estimates of Atlantic Ocean heat transport at 26.5°N. J. Clim. 24, 2429–2449. doi: 10.1175/2010JCLI3997.1

CrossRef Full Text | Google Scholar

Johns, W. E., Beal, L. M., Baringer, M. O., Molina, J. R., Cunningham, S. A., Kanzow, T., et al. (2008). Variability of shallow and deep western boundary currents off the Bahamas during 2004–2005: results from the 26°N RAPID-MOC array. J. Phys. Oceanogr. 38, 605–623. doi: 10.1175/2007JPO3791.1

CrossRef Full Text | Google Scholar

Johns, W. E., Kanzow, T., and Zantopp, R. (2005). Estimating ocean transports with dynamic height moorings: an application in the Atlantic Deep Western Boundary Current at 26°N. Deep Sea Res. Part I 52, 1542–1567. doi: 10.1016/j.dsr.2005.02.002

CrossRef Full Text | Google Scholar

Johnson, C. R., Banks, S. C., Barrett, N. S., Cazassus, F., Dunstan, P. K., Edgar, G. J., et al. (2011). Climate change cascades: shifts in oceanography, species’ ranges and subtidal marine community dynamics in eastern Tasmania. J. Exp. Mar. Biol. Ecol. 400, 17–32. doi: 10.1016/j.jembe.2011.02.032

CrossRef Full Text | Google Scholar

Johnson, G. C. (2001). The Pacific Ocean subtropical cell surface limb. Geophys. Res. Lett. 28, 1771–1774. doi: 10.1029/2000gl012723

CrossRef Full Text | Google Scholar

Johnson, K. S., Plant, J. N., Coletti, L. J., Jannasch, H. W., Sakamoto, C. M., Riser, S. C., et al. (2017). Biogeochemical sensor performance in the SOCCOM profiling float array. J. Geophys. Res. 122, 6416–6436. doi: 10.1002/2017JC012838

CrossRef Full Text | Google Scholar

Johnston, T. M. S., and Rudnick, D. L. (2015). Mixing estimates in the California Current System from sustained observations by underwater gliders. Deep Sea Res. II 112, 61–78. doi: 10.1016/j.dsr2.2014.03.009

CrossRef Full Text | Google Scholar

Johnston, T. M., Rudnick, D. L., Alford, M. H., Pickering, A., and Simmons, H. L. (2013). Internal tidal energy fluxes in the South China Sea from density and velocity measurements by gliders. J. Geophys. Res. Oceans 118, 3939–3949. doi: 10.1002/jgrc.20311

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Schmidt, M., Siegfried, L., and van der Plas, A. (2019). Coastal trapped wave propagation along the southwest African shelf as revealed by moored observations. J. Phys. Oceanogr. 49, 851–866. doi: 10.1175/JPO-D-18-0046

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Siegfried, L., and van der Plas, A. (2017a). Seasonal to interannual variability of water mass characteristics and currents on the Namibian shelf. J. Mar. Syst. 165, 36–46. doi: 10.1016/j.jmarsys.2016.09.003

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Siegfried, L., van der Plas, A., Heene, T., and Beier, S. (2017b). Daily Means of Bottom Temperature and Bottom Oxygen Measurements on the Namibian Shelf at 18°S. Warnemünde: Leibniz Institute for Baltic Sea Research, doi: 10.1594/PANGAEA.871251

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Siegfried, L., van der Plas, A., Heene, T., and Beier, S. (2017c). Daily Means of Horizontal Currents Measurements on the Namibian Shelf at 18°S. Warnemünde: Leibniz Institute for Baltic Sea Research, doi: 10.1594/PANGAEA.871253

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Siegfried, L., van der Plas, A., Heene, T., Beier, S., et al. (2017d). Daily Means of Horizontal Currents Measurements on the Namibian Shelf at 20°S. Warnemünde: Leibniz Institute for Baltic Sea Research, doi: 10.1594/PANGAEA.872098

CrossRef Full Text | Google Scholar

Junker, T., Mohrholz, V., Siegfried, L., van der Plas, A., Heene, T., and Breier, S. (2017e). Daily Means of Bottom Temperature and Bottom Oxygen Measurements on the Namibian Shelf at 20°S. Warnemünde: Leibniz Institute for Baltic Sea Research, doi: 10.1594/PANGAEA.872099

CrossRef Full Text | Google Scholar

Juranek, L. W., Feely, R. A., Peterson, W. T., Alin, S. R., Hales, B., Lee, K., et al. (2009). A novel method for determination of aragonite saturation state on the continental shelf of central Oregon using multi-parameter relationships with hydrographic data. Geophys. Res. Lett. 36:L24601. doi: 10.1029/2009GL040778

CrossRef Full Text | Google Scholar

Jury, M. R., Valentine, H. R., and Lutjeharms, J. R. (1993). Influence of the Agulhas Current on summer rainfall along the southeast coast of South Africa. J. Appl. Meteorol. 32, 1282–1287. doi: 10.1175/1520-0450(1993)032<1282:iotaco>2.0.co;2

CrossRef Full Text | Google Scholar

Karstensen, J., Schütte, F., Pietri, A., Krahmann, G., Fiedler, B., Grundle, D., et al. (2017). Upwelling and isolation in oxygen-depleted anticyclonic modewater eddies and implications for nitrate cycling. Biogeosciences 14, 2167–2181. doi: 10.5194/bg-14-2167-2017

CrossRef Full Text | Google Scholar

Kelly, K. A., Small, R. J., Samelson, R. M., Qiu, B., Joyce, T. M., Kwon, Y. O., et al. (2010). Western boundary currents and frontal air–sea interaction: Gulf Stream and Kuroshio Extension. J. Clim. 23, 5644–5667. doi: 10.1175/2010JCLI3346.1

CrossRef Full Text | Google Scholar

Kerry, C., Powell, B., Roughan, M., and Oke, P. (2016). Development and evaluation of a high-resolution reanalysis of the East Australian Current region using the regional ocean modelling system (ROMS 3.4) and incremental strong-constraint 4-Dimensional variational (IS4D-Var) data assimilation. Geosci. Model. Dev. 9, 3779–3801. doi: 10.5194/gmd-9-3779-2016

CrossRef Full Text | Google Scholar

Kerry, C., Roughan, M., and Powell, B. (2018). Observation impact in a regional reanalysis of the East Australian Current System. J. Geophys. Res. 123, 7511–7528. doi: 10.1029/2017JC013685

CrossRef Full Text | Google Scholar

Kersalé, M., Lamont, T., Speich, S., Terre, T., Laxenaire, R., Roberts, M. J., et al. (2018). Moored observations of mesoscale features in the Cape Basin: characteristics and local impacts on water mass distributions. Ocean Sci. 14, 923–945. doi: 10.5194/os-14-923-2018

CrossRef Full Text | Google Scholar

Kim, S. Y. (2010). Observations of submesoscale eddies using high-frequency radar-derived kinematic and dynamic quantities. Cont. Shelf Res. 30, 1639–1655. doi: 10.1016/j.csr.2010.06.011

CrossRef Full Text | Google Scholar

Kim, S. Y., and Kosro, P. M. (2013). Observations of near-inertial surface currents off Oregon: decorrelation time and length scales. Oceans 118, 3723–3736. doi: 10.1002/jgrc.20235

CrossRef Full Text | Google Scholar

Kim, S. Y., Terrill, E. J., Cornuelle, B. D., Jones, B., Washburn, L., Moline, M. A., et al. (2011). Mapping the U.S. West Coast surface circulation: a multiyear analysis of high-frequency radar observations. J. Geophys. Res. 116:C03011. doi: 10.1029/2010JC006669

CrossRef Full Text | Google Scholar

Klenz, T., Dengler, M., and Brandt, P. (2018). Seasonal variability of the Mauritania Current and hydrography at 18°N. J. Geophys. Res. 123, 8122–8137. doi: 10.1029/2018JC014264

CrossRef Full Text | Google Scholar

Kolodziejczyk, N., Testor, P., Lazar, A., Echevin, V., Krahmann, G., Chaigneau, A., et al. (2018). Subsurface fine-scale patterns in an anticyclonic eddy off Cap-Vert peninsula observed from glider measurements. J. Geophys. Res. Oceans 123, 6312–6329. doi: 10.1029/2018JC014135

CrossRef Full Text | Google Scholar

Kono, T., and Kawasaki, Y. (1997). Results of CTD and mooring observations southeast of Hokkaido. 1. Annual verocity and transport variations in the Oyashio. Bull. Hokkaido Natl. Fish. Res. Inst. 61, 65–81.

Google Scholar

Kopte, R., Brandt, P., Claus, M., Greatbatch, R. J., and Dengler, M. (2018). Role of equatorial basin-mode resonance for the seasonal variability of the Angola Current at 11°S. J. Phys. Oceanogr. 48, 261–281. doi: 10.1175/JPO-D-17-0111.1

CrossRef Full Text | Google Scholar

Kopte, R., Brandt, P., Dengler, M., Tchipalanga, P. C. M., Macuéria, M., and Ostrowski, M. (2017). The Angola Current: flow and hydrographic characteristics as observed at 11°S. J. Geophys. Res. Oceans 122, 1177–1189. doi: 10.1002/2016JC012374

CrossRef Full Text | Google Scholar

Koszalka, I., and LaCasce, J. H. (2010). Lagrangian analysis by clustering. Ocean Dyn. 60, 957–972. doi: 10.1007/s10236-010-0306-2

CrossRef Full Text | Google Scholar

Krug, M., Swart, S., and Gula, J. (2017). Submesoscale cyclones in the Agulhas Current. Geophys. Res. Lett. 44, 346–354. doi: 10.1002/2016GL071006

CrossRef Full Text | Google Scholar

Kuroda, H., Wagawa, T., Kakehi, S., Shimizu, Y., Kusaka, A., Okunishi, T., et al. (2017). Long-term mean and seasonal variation of altimetry-derived Oyashio transport across the A-line off the southeastern coast of Hokkaido. Jpn. Deep Sea Res. Part I 121, 95–109. doi: 10.1016/j.dsr.2016.12.006

CrossRef Full Text | Google Scholar

Kuroda, H., Wagawa, T., Shimizu, Y., Ito, S., Kakehi, S., Okunishi, T., et al. (2015). Interdecadal decrease of the Oyashio transport on the continental slope off the southeastern coast of Hokkaido, Japan. J. Geophys. Res. Oceans 120, 2504–2522. doi: 10.1002/2014JC010402

CrossRef Full Text | Google Scholar

LaCasce, J. H. (2008). Statistics from Lagrangian observations. Prog. Oceanogr. 77, 1–29. doi: 10.1016/j.pocean.2008.02.002

CrossRef Full Text | Google Scholar

Landschützer, P., Gruber, N., Bakker, D. C. E., and Schuster, U. (2014). Recent variability of the global ocean carbon sink. Glob. Biogeochem. Cycles 28, 927–949. doi: 10.1002/2014GB004853

CrossRef Full Text | Google Scholar

Large, W. G., and Danabasoglu, G. (2006). Attribution and impacts of upper-ocean biases in CCSM3. J. Clim. 19, 2325–2346. doi: 10.1175/JCLI3740.1

CrossRef Full Text | Google Scholar

Larsen, J. C., and Sanford, T. B. (1985). Florida Current volume transports from voltage measurements. Science 227, 302–304. doi: 10.1126/science.227.4684.302

PubMed Abstract | CrossRef Full Text | Google Scholar

Laurindo, L. C., Mariano, A., and Lumpkin, R. (2017). An improved near-surface velocity climatology for the global ocean from drifter observations. Deep Sea Res. I 124, 73–92. doi: 10.1016/j.dsr.2017.04.009

CrossRef Full Text | Google Scholar

Leber, G. M., Beal, L. M., and Elipot, S. (2017). Wind and current forcing combine to drive strong upwelling in the Agulhas Current. J. Phys. Oceanogr. 47, 123–134. doi: 10.1175/JPO-D-16-0079.1

CrossRef Full Text | Google Scholar

Lee, E. A., and Kim, S. Y. (2018). Regional variability and turbulent characteristics of the satellite-sensed submesoscale surface chlorophyll concentrations. J. Geopyhs. Res. 123, 4250–4279. doi: 10.1029/2017JC013732

CrossRef Full Text | Google Scholar

Lee, T., and Fukumori, I. (2003). Interannual-to-decadal variations of tropical–subtropical exchange in the Pacific Ocean: boundary versus interior pycnocline transports. J. Clim. 16, 4022–4042. doi: 10.1175/1520-0442(2003)016<4022:ivotei>2.0.co;2

CrossRef Full Text | Google Scholar

Lengaigne, M., Hausmann, U., Madec, G., Menkes, C. E., Vialard, J., and Molines, J. M. (2012). Mechanisms controlling warm water volume interannual variations in the equatorial Pacific: diabatic versus adiabatic processes. Clim. Dyn. 38, 1031–1046. doi: 10.1007/s00382-011-1051-z

CrossRef Full Text | Google Scholar

Levine, N. M., Doney, S. C., Lima, I., Wanninkhof, R., Bates, N. R., and Feely, R. A. (2011). The impact of the North Atlantic Oscillation on the uptake and accumulation of anthropogenic CO2 by North Atlantic Ocean mode waters. Glob. Biogeochem. Cycles 25:GB3022. doi: 10.1029/2010GB003892

CrossRef Full Text | Google Scholar

Liang, X., Spall, M., and Wunsch, C. (2017). Global ocean vertical velocity from a dynamically consistent ocean state estimate. J. Geophys. Res. 122, 8208–8224. doi: 10.1002/2017JC012985

CrossRef Full Text | Google Scholar

Lien, R.-C., Ma, B., Cheng, Y.-H., Ho, C.-R., Qiu, B., Lee, C. M., et al. (2014). Modulation of Kuroshio transport by mesoscale eddies at the Luzon Strait entrance. J. Geophys. Res. Oceans 119, 2129–2142. doi: 10.1002/2013JC009548

CrossRef Full Text | Google Scholar

Lien, R.-C., Ma, B., Lee, C. M., Sanford, T. B., Mensah, V., Centurioni, L. R., et al. (2015). The Kuroshio and Luzon Undercurrent east of Luzon Island. Oceanography 28, 54–63. doi: 10.5670/oceanog.2015.81

CrossRef Full Text | Google Scholar

Lima, M. O., Cirano, M., Mata, M., Goes, M., Goni, G., and Baringer, M. O. (2016). An assessment of the Brazil Current baroclinic structure and variability near 22°S in distinct ocean forecasting and analysis systems. Ocean Dyn. 66, 893–916. doi: 10.1007/s10236-016-0959-6

CrossRef Full Text | Google Scholar

Linder, C. A., and Gawarkiewicz, G. G. (1998). A climatology of the shelfbreak front in the Middle Atlantic Bight. J. Geophys. Res. 103, 18405–18423. doi: 10.1029/98JC01438

CrossRef Full Text | Google Scholar

Lindstrom, E., Gunn, J., Fischer, A., McCurdy, A., and Glover, L. K. (2012). A Framework for Ocean Observing. By the Task Team for an Integrated Framework for Sustained Ocean Observing. Paris: UNESCO, doi: 10.5270/OceanObs09-FOO

CrossRef Full Text | Google Scholar

Louw, D., van der Plas, A., Mohrholz, V., Wasmund, N., Junker, T., and Eggert, A. (2016). Seasonal and interannual phytoplankton dynamics and forcing mechanisms in the northern Benguela upwelling system. J. Mar. Syst. 157, 124–134. doi: 10.1016/j.jmarsys.2016.01.009

CrossRef Full Text | Google Scholar

Lovechio, E., Gruber, N., and Münnich, M. (2018). Mesoscale contribution to the long-range offshore transport of organic carbon from the Canary Upwelling system to the open North Atlantic. Biogeosciences 15, 5061–5091. doi: 10.5194/bg-15-5061-2018

CrossRef Full Text | Google Scholar

Lowcher, C. F., Muglia, M., Bane, J. M., He, R., Gong, Y., Haines, S. M., et al. (2017). “Marine hydrokinetic energy in the Gulf Stream off North Carolina: an assessment using observations and ocean circulation models,” in Marine Renewable Energy, eds Z. Yang and A. Copping (Cham: Springer), doi: 10.1007/978-3-319-53536-4_10

CrossRef Full Text | Google Scholar

Lumpkin, R. (2003). Decomposition of surface drifter observations in the Atlantic Ocean. Geophys. Res. Lett. 30:1753. doi: 10.1029/2003GL017519

CrossRef Full Text | Google Scholar

Lumpkin, R., and Garzoli, S. L. (2011). Interannual to decadal changes in the western South Atlantic’s surface circulation. J. Geophys. Res. 116:C01014. doi: 10.1029/2010JC006285

CrossRef Full Text | Google Scholar

Lumpkin, R., and Johnson, G. (2013). Global ocean surface velocities from drifters: mean, variance, El Niño–Southern Oscillation response, and seasonal cycle. J. Geophys. Res. 118, 2992–3006. doi: 10.1002/jgrc.20210

CrossRef Full Text | Google Scholar

Lumpkin, R., and Pazos, M. C. (2007). “Measuring surface currents with surface velocity program drifters: the instrument, its data, and some recent results,” in Lagrangian Analysis and Prediction of Coastal and Ocean Dynamics, eds A. Griffa, A. D. Kirwan, A. J. Mariano, T. Ozgokmen, and H. T. Rossby (Cambridge: Cambridge University Press), 39–67. doi: 10.1017/cbo9780511535901.003

CrossRef Full Text | Google Scholar

Lumpkin, R., Maximenko, N., and Pazos, M. (2012). Evaluating where and why drifters die. J. Atmos. Ocean. Technol. 29, 300–308. doi: 10.1175/jtech-d-11-00100.1

CrossRef Full Text | Google Scholar

Lumpkin, R., Özgökmen, T., and Centurioni, L. (2017). Advances in the application of surface drifters. Annu. Rev. Mar. Sci. 9, 59–81. doi: 10.1146/annurev-marine-010816-060641

PubMed Abstract | CrossRef Full Text | Google Scholar

Lund, B., Haus, B. K., Horstmann, J., Graber, H. C., Carrasco, R., Laxague, N. J., et al. (2018). Near-surface current mapping by shipboard marine X-band radar: a validation. J. Atmos. Oceanic Technol. 35, 1077–1090. doi: 10.1175/JTECH-D-17-0154.1

CrossRef Full Text | Google Scholar

Lutjeharms, J. R. (2006). The Agulhas Current. Berlin: Springer.

Google Scholar

Lynch, T. P., Morello, E. B., Evans, K., Richardson, A., Rochester, W., Steinberg, C. R., et al. (2014). IMOS national reference stations: a continental scale physical, chemical, biological coastal observing system. PLoS One 9:e113652. doi: 10.1371/journal.pone.0113652

PubMed Abstract | CrossRef Full Text | Google Scholar

Ma, X., Chang, P., Saravanan, R., Montuoro, R., Nakamura, H., Wu, D., et al. (2017). Importance of resolving Kuroshio front and eddy influence in simulating the North Pacific storm track. J. Clim. 30, 1861–1880. doi: 10.1175/JCLI-D-16-0154.1

CrossRef Full Text | Google Scholar

Machu, E., Capet, X., Estrade, P. A., Ndoye, S., Brajard, J., Baurand, F., et al. (2019). First evidence of anoxia and nitrogen loss in the southern Canary upwelling system. Geophys. Res. Lett. 46, 2619–2627. doi: 10.1029/2018GL079622

CrossRef Full Text | Google Scholar

Mackey, D. J., O’Sullivan, J. E., and Watson, R. J. (2002). Iron in the western Pacific: a riverine or hydrothermal source for iron in the equatorial undercurrent. Deep Sea. Res. I 49, 877–893. doi: 10.1016/S0967-0637(01)00075-9

CrossRef Full Text | Google Scholar

Majumder, S., and Schmid, C. (2018). A study of the variability in the Benguela Current volume transport. Ocean Sci. 14, 273–283. doi: 10.5194/os-14-273-2018

CrossRef Full Text | Google Scholar

Majumder, S., Goes, M., Polito, P. S., Lumpkin, R., Schmid, C., and Lopez, H. (2019). Propagating modes of variability and their impact on the western boundary current in the South Atlantic. J. Geophys. Res. Oceans 124, 3168–3185. doi: 10.1029/2018JC014812

CrossRef Full Text | Google Scholar

Mansfield, K. L., Mendilaharsu, M. L., Putman, N. F., dei Marcovaldi, M. A., Sacco, A. E., Lopez, G., et al. (2017). First satellite tracks of South Atlantic sea turtle ‘lost years’: seasonal variation in trans-equatorial movement. Proc. Biol. Sci. 284:20171730. doi: 10.1098/rspb.2017.1730

PubMed Abstract | CrossRef Full Text | Google Scholar

Mantovanelli, A., Keating, S., Wyatt, L., Roughan, M., and Schaeffer, A. (2017). Lagrangian and Eulerian characterization of two counter rotating submesoscale eddies in a western boundary current. J. Geophys. Res. Oceans 122, 4902–4921. doi: 10.1002/2016JC011968

CrossRef Full Text | Google Scholar

Marra, J., Houghton, R. W., and Garside, C. (1990). Phytoplankton growth at the shelf-break front in the Middle Atlantic Bight. J. Mar. Res. 48, 851–868. doi: 10.1357/002224090784988665

CrossRef Full Text | Google Scholar

Marzeion, B., Cogley, J. G., Richter, K., and Parkes, D. (2014). Attribution of global glacier mass loss to anthropogenic and natural causes. Science 345, 919–921. doi: 10.1126/science.1254702

PubMed Abstract | CrossRef Full Text | Google Scholar

Mata, M. M., Cirano, M., van Caspel, M. R., Fonteles, C. S., Goni, G., and Baringer, M. (2012). Observations of Brazil Current baroclinic transport near 22°S: variability from the AX97 XBT transect. CLIVAR Exchanges 58, 5–10. doi: 10.1007/s10236-016-0959-6

CrossRef Full Text | Google Scholar

Matano, R. P., and Palma, E. D. (2008). On the upwelling of downwelling currents. J. Phys. Oceanogr. 38, 2482–2500. doi: 10.1175/2008jpo3783.1

CrossRef Full Text | Google Scholar

Matano, R. P., Combes, V., Piola, A. R., Guerrero, R. A., Palma, E. D., Strub, P. T., et al. (2014). The salinity signature of the cross-shelf exchanges in the Southwestern Atlantic Ocean: numerical simulations. J. Geophys. Res. Oceans 119, 7949–7968. doi: 10.1002/2014JC010116

PubMed Abstract | CrossRef Full Text | Google Scholar

Maximenko, N., Niiler, P., Centurioni, L., Rio, M.-H., Melnichenko, O., Chambers, D., et al. (2009). Mean dynamic topography of the ocean derived from satellite and drifting buoy data using three different techniques. J. Atmos. Ocean. Technol. 26, 1910–1919. doi: 10.1175/2009JTECHO672.1

CrossRef Full Text | Google Scholar

Mazzini, P. L. F., Barth, J. A., Shearman, R. K., and Erofeev, A. (2014). Buoyancy-driven coastal currents off the Oregon coast during fall and winter. J. Phys. Oceanogr. 44, 2854–2876. doi: 10.1175/JPO-D-14-0012.1

CrossRef Full Text | Google Scholar

McClatchie, S. (2014). Regional Fisheries Oceanography of the California Current System: The CalCOFI Program. Berlin: Springer, doi: 10.1007/978-94-007-7223-6

CrossRef Full Text | Google Scholar

McClatchie, S., Thompson, A. R., Alin, S. R., Siedlecki, S., Watson, W., and Bograd, S. J. (2016). The influence of Pacific Equatorial Water on fish diversity in the southern California Current System. J. Geophys. Res. Oceans 121, 6121–6136. doi: 10.1002/2016JC011672

CrossRef Full Text | Google Scholar

Meinen, C. S., and Watts, D. R. (2000). Vertical structure and transport on a transect across the North Atlantic Current near 42°N: time series and mean. J. Geophys. Res. 105, 21869–21891. doi: 10.1029/2000JC900097

CrossRef Full Text | Google Scholar

Meinen, C. S., Baringer, M. O., and Garcia, R. F. (2010). Florida Current transport variability: an analysis of annual and longer-period signals. Deep Sea Res. I 57, 835–846. doi: 10.1016/j.dsr.2010.04.001

CrossRef Full Text | Google Scholar

Meinen, C. S., Garzoli, S. L., Perez, R. C., Campos, E., Piola, A. R., Chidichimo, M. P., et al. (2017). Characteristics and causes of Deep Western Boundary Current transport variability at 34.5°S during 2009–2014. Ocean Sci. 13, 175–194. doi: 10.5194/os-13-175-2017

CrossRef Full Text | Google Scholar

Meinen, C. S., Johns, W. E., Garzoli, S. L., van Sebille, E., Rayner, D., Kanzow, T., et al. (2013). Variability of the Deep Western Boundary Current at 26.5N during 2004–2009. Deep Sea Res. II 85, 154–168. doi: 10.1016/j.dsr2.2012.07.036

CrossRef Full Text | Google Scholar

Meinen, C. S., Speich, S., Piola, A. R., Ansorge, I., Campos, E., Kersale, M., et al. (2018). Meridional overturning circulation transport variability at 34.5°S during 2009–2017: baroclinic and barotropic flows and the dueling influence of the boundaries. Geophys. Res. Lett. 45, 4180–4188. doi: 10.1029/2018GL077408

CrossRef Full Text | Google Scholar

Menna, M., Faye, S., Poulain, P.-M., Centurioni, L., Lazar, A., Gaye, A., et al. (2016). Upwelling features of the coast of north-western Africa in 2009–2013. Boll. Geofis. Teor. Appl. 57, 71–86. doi: 10.4430/bgta0164

CrossRef Full Text | Google Scholar

Messié, M., and Chavez, F. P. (2015). Seasonal regulation of primary production in eastern boundary upwelling systems. Prog. Oceanogr. 134, 1–18. doi: 10.1016/j.pocean.2014.10.011

CrossRef Full Text | Google Scholar

Middleton, J. F., and Bye, J. T. (2007). A review of the shelf-slope circulation along Australia’s southern shelves: Cape Leeuwin to Portland. Prog. Oceanogr. 75, 1–41. doi: 10.1016/j.pocean.2007.07.001

CrossRef Full Text | Google Scholar

Middleton, J. F., and Cirano, M. (2002). A boundary current along Australia’s southern shelves: the Flinders Current. J. Geophys. Res. 107:3129. doi: 10.1029/2000JC000701

CrossRef Full Text | Google Scholar

Mihanoviæ, H., Pattiaratchi, C., and Verspecht, F. (2016). Diurnal sea breezes force near-inertial waves along Rottnest continental shelf, southwestern Australia. J. Phys. Oceanogr. 46, 3487–3508. doi: 10.1175/JPO-D-16-0022.1

CrossRef Full Text | Google Scholar

Mohrholz, V., Bartholomae, C. H., van der Plas, A. K., and Lass, H. U. (2008). The seasonal variability of the northern Benguela undercurrent and its relation to the oxygen budget on the shelf. Cont. Shelf Res. 28, 424–441. doi: 10.1016/j.csr.2007.10.001

CrossRef Full Text | Google Scholar

Moloney, C. L., Van Der Lingen, C. D., Hutchings, L., and Field, J. G. (2004). Contributions of the Benguela ecology programme to pelagic fisheries management in South Africa. Afr. J. Mar. Sci. 26, 37–51. doi: 10.2989/18142320409504048

CrossRef Full Text | Google Scholar

Monteiro, P. M. S., and van der Plas, A. K. (2006). “Low oxygen water (LOW) variability in the Benguela system: key processes and forcing scales relevant to forecasting,” in Benguela: Predicting a Large Marine Ecosystem, eds V. Shannon, et al. (Netherlands: ElsevierAmsterdam), 71–90. doi: 10.1016/s1570-0461(06)80010-8

CrossRef Full Text | Google Scholar

Monteiro, P. M. S., van der Plas, A. K., Mélice, J.-L., and Florenchie, P. (2008). Interannual hypoxia variability in a coastal upwelling system: ocean–shelf exchange, climate and ecosystem-state implications. Deep Sea Res. I 55, 435–450. doi: 10.1016/j.dsr.2007.12.010

CrossRef Full Text | Google Scholar

Mosquera-Vásquez, K., Dewitte, B., Illig, S., Takahashi, K., and Garric, G. (2013). The 2002/2003 El Niño: equatorial waves sequence and their impact on sea surface temperature. J. Geophys. Res. Oceans 118, 346–357.

Google Scholar

Nagai, T., and Clayton, S. (2017). Nutrient interleaving below the mixed layer of the Kuroshio Extension front. Ocean Dyn. 67, 1027–1046. doi: 10.1007/s10236-017-1070-3

CrossRef Full Text | Google Scholar

Nagano, A., Kizu, S., Hanawa, K., and Roemmich, D. (2016). Heat transport variation due to change of North Pacific subtropical gyre interior flow during 1993–2012. Ocean Dyn. 66, 1637–1649. doi: 10.1007/s10236-016-1007-2

CrossRef Full Text | Google Scholar

Nakamura, M. (2012). Impacts of SST anomalies in the Agulhas Current System on the regional climate variability. J. Clim. 25, 1213–1229. doi: 10.1175/JCLI-D-11-00088.1

CrossRef Full Text | Google Scholar

Nakano, H., Tsujino, H., Yasuda, M., Hirabara, T., Motoi, T., Ishii, M., et al. (2011). Uptake mechanisms of anthropogenic CO2 in the Kuroshio Extension region in an ocean general circulation model. J. Oceanogr. 67, 765–783. doi: 10.1007/s10872-011-0075-7

CrossRef Full Text | Google Scholar

Nakano, T., Kitamura, T., Sugimoto, S., Suga, T., and Kamachi, M. (2015). Long-term variations of North Pacific Tropical Water along the 137°E repeat hydrographic section. J. Oceanogr. 71, 229–238. doi: 10.1007/s10872-015-0279-3

CrossRef Full Text | Google Scholar

Nam, S., Kim, H. J., and Send, U. (2011). Amplification of hypoxic and acidic events by La Niña conditions on the continental shelf off California. Geophys. Res. Lett. 38:L22602. doi: 10.1029/2011gl049549

CrossRef Full Text | Google Scholar

Nguyen, L. T., and Molinari, J. (2012). Rapid intensification of a sheared, fast-moving hurricane over the Gulf Stream. Mon. Weather Rev. 140, 3361–3378. doi: 10.1175/MWR-D-11-00293.1

CrossRef Full Text | Google Scholar

Niewiadomska, K., Claustre, H., Prieur, L., and d’Ortenzio, F. (2008). Submesoscale physical-biogeochemical coupling across the Ligurian Current (northwestern Mediterranean) using a bio-optical glider. Limnol. Oceanogr. 53, 2210–2225. doi: 10.4319/lo.2008.53.5_part_2.2210

CrossRef Full Text | Google Scholar

Niiler, P. P. (2001). “The world ocean surface circulation,” in Ocean Circulation and Climate, eds G. Siedler, J. Church, and J. Gould (Cambridge, MA: Academic Press), 193–204. doi: 10.1016/s0074-6142(01)80119-4

CrossRef Full Text | Google Scholar

Niiler, P. P., Maximenko, N. A., Panteleev, G. G., Yamagata, T., and Olson, D. B. (2003). Near-surface dynamical structure of the Kuroshio Extension. J. Geophys. Res. 108:3193. doi: 10.1029/2002JC001461

CrossRef Full Text | Google Scholar

Niiler, P. P., Sybrandy, A., Bi, K., Poulain, P. M., and Bitterman, D. (1995). Measurements of the water-following capability of holey-sock and TRISTAR drifters. Deep Sea Res. I 42, 1951–1964. doi: 10.1016/0967-0637(95)00076-3

CrossRef Full Text | Google Scholar

Nkwinkwa Njouodo, A. S., Koseki, S., Keenlyside, N., and Rouault, M. (2018). Atmospheric signature of the Agulhas Current. Geophys. Res. Lett. 45, 5185–5193. doi: 10.1029/2018GL077042

CrossRef Full Text | Google Scholar

Nowald, N., Iversen, M. H., Fischer, G., Ratmeyer, V., and Wefer, G. (2015). Time series of in-situ particle properties and sediment trap fluxes in the coastal upwelling filament off Cape Blanc, Mauritania. Prog. Oceanogr. 137, 1–11. doi: 10.1016/j.pocean.2014.12.015

CrossRef Full Text | Google Scholar

O’Reilly, C. H., and Czaja, A. (2015). The response of the Pacific storm track and atmospheric circulation to Kuroshio Extension variability. Q. J. R. Meteorol. Soc. 141, 52–66. doi: 10.1002/qj.2334

CrossRef Full Text | Google Scholar

O’Reilly, C. H., Minobe, S., and Kuwano-Yoshida, A. (2016). The influence of the Gulf Stream on wintertime European blocking. Clim. Dyn. 47, 1545–1567. doi: 10.1007/s00382-015-2919-0

CrossRef Full Text | Google Scholar

Oettli, P., Morioka, Y., and Yamagata, T. (2016). A regional climate mode discovered in the North Atlantic: Dakar Niño/Niña. Sci. Rep. 6:18782. doi: 10.1038/srep18782

PubMed Abstract | CrossRef Full Text | Google Scholar

Ohman, M. D., Rudnick, D. L., Chekalyuk, A., Davis, R. E., Feely, R. A., Kahru, M., et al. (2013). Autonomous ocean measurements in the California Current Ecosystem. Oceanography 26, 18–25. doi: 10.5670/oceanog.2013.41

PubMed Abstract | CrossRef Full Text | Google Scholar

Oka, E., and Kawabe, M. (2003). Dynamic structure of the Kuroshio south of Kyushu in relation to the kuroshio path variations. J. Oceanogr. 59, 595–608. doi: 10.1023/B:JOCE.0000009589.28241.93

CrossRef Full Text | Google Scholar

Oka, E., Ishii, M., Nakano, T., Suga, T., Kouketsu, S., Miyamoto, M., et al. (2018). Fifty years of the 137E repeat hydrographic section in the western North Pacific Ocean. J. Oceanogr. 74, 115–145. doi: 10.1007/s10872-017-0461-x

CrossRef Full Text | Google Scholar

Oka, E., Qiu, B., Takatani, Y., Enyo, K., Sasano, D., Kosugi, N., et al. (2015). Decadal variability of Subtropical Mode Water subduction and its impact on biogeochemistry. J. Oceanogr. 71, 389–400. doi: 10.1007/s10872-015-0300-x

CrossRef Full Text | Google Scholar

Oliveira, L. R., Piola, A. R., Mata, M. M., and Soares, I. D. (2009). Brazil Current surface circulation and energetics observed from drifting buoys. J. Geophys. Res. 114:C10006. doi: 10.1029/2008JC004900

CrossRef Full Text | Google Scholar

Oliver, E. C. J., O’Kane, T. J., and Holbrook, N. J. (2015). Projected changes to Tasman Sea eddies in a future climate. Oceans 120, 7150–7165. doi: 10.1002/2015JC010993

CrossRef Full Text | Google Scholar

Oliver, E. C., Benthuysen, J. A., Bindoff, N. L., Hobday, A. J., Holbrook, N. J., Mundy, C. N., et al. (2017). The unprecedented 2015/16 Tasman Sea marine heatwave. Nat. Commun. 8:16101. doi: 10.1038/ncomms16101

PubMed Abstract | CrossRef Full Text | Google Scholar

Olson, D., Podesta, G. P., Evans, R. H., and Brown, O. (1988). Temporal variations in the separation of Brazil and Malvinas Currents. Deep Sea Res. 35, 1971–1990. doi: 10.1016/0198-0149(88)90120-3

CrossRef Full Text | Google Scholar

Paduan, J. D., and Washburn, L. (2013). High-frequency radar observations of ocean surface currents. Annu. Rev. Mar. Sci. 5, 115–136. doi: 10.1146/annurev-marine-121211-172315

PubMed Abstract | CrossRef Full Text | Google Scholar

Palacz, A. P., Pearlman, J., Simmons, S., Hill, K., Miloslavich, P., Telszewski, M., et al. (2017). Report of the Workshop on the Implementation of Multi-disciplinary Sustained Ocean Observations (IMSOO). Global Ocean Observing System (GOOS) Report No. 223. Available at: http://www.goosocean.org/imsoo-report (accessed July 23, 2019).

Google Scholar

Palevsky, H. I., and Nicholson, D. P. (2018). The North Atlantic biological pump: insights from the Ocean Observatories Initiative Irminger Sea array. Oceanography 31, 42–49. doi: 10.5670/oceanog.2018.108

CrossRef Full Text | Google Scholar

Palevsky, H. I., and Quay, P. D. (2017). Influence of biological carbon export on ocean carbon uptake over the annual cycle across the North Pacific Ocean. Glob. Biogeochem. Cycles 31, 1–15. doi: 10.1002/2016GB005527

CrossRef Full Text | Google Scholar

Palevsky, H. I., Quay, P. D., Lockwood, D. E., and Nicholson, D. P. (2016). The annual cycle of gross primary production, net community production, and export efficiency across the North Pacific Ocean. Glob. Biogeochem. Cycles 30, 361–380. doi: 10.1002/2015GB005318

CrossRef Full Text | Google Scholar

Palter, J. B., and Lozier, M. S. (2008). On the source of Gulf Stream nutrients. J. Geophys. Res. 113:C06018. doi: 10.1029/2007jc004611

CrossRef Full Text | Google Scholar

Paniagua, G. F., Saraceno, M., Piola, A. R., Guerrero, R., Provost, C., Ferrari, R., et al. (2018). Malvinas Current at 40°S–41°S: first assessment of temperature and salinity temporal variability. J. Geophys. Res. Oceans 123, 5323–5340. doi: 10.1029/2017JC013666

CrossRef Full Text | Google Scholar

Park, J., and Sweet, W. (2015). Accelerated sea level rise and Florida Current transport. Ocean Sci. 11, 607–615. doi: 10.5194/os-11-607-2015

CrossRef Full Text | Google Scholar

Parks, A. B., Shay, L. K., Johns, W. E., Martinez-Pedraja, J., and Gurgel, K. W. (2009). HF radar observations of small-scale surface current variability in the Straits of Florida. J. Geophys. Res. 114:C08002. doi: 10.1029/2008JC005025

CrossRef Full Text | Google Scholar

Parrilla, G., Neuer, S., and Le Traon, P.-Y. (2002). Topical studies in oceanography: Canary Islands Azores Gibraltar Observations (CANIGO). Volume 1: studies in the northern Canary Islands basin. Deep Sea Res. II 49, 3409–3413. doi: 10.1016/s0967-0645(02)00104-2

CrossRef Full Text | Google Scholar

Pattiaratchi, C., Hollings, B., Woo, M., and Welhena, T. (2011). Dense shelf water formation along the south-west Australian inner shelf. Geophys. Res. Lett. 38:L10609. doi: 10.1029/2011GL046816

CrossRef Full Text | Google Scholar

Paulmier, A., and Ruiz-Pino, D. (2009). Oxygen minimum zones (OMZs) in the modern ocean. Prog. Oceanogr. 80, 113–128. doi: 10.1016/j.pocean.2008.08.001

CrossRef Full Text | Google Scholar

Pearce, A., and Feng, M. (2013). The rise and fall of the “marine heat wave” off western Australia during the summer of 2010/11. J. Mar. Syst. 111-112, 139–156. doi: 10.1016/j.jmarsys.2012.10.009

CrossRef Full Text | Google Scholar

Pegliasco, C., Chaigneau, A., and Morrow, R. (2015). Main eddy vertical structures observed in the four major eastern boundary upwelling systems. Oceans 120, 6008–6033. doi: 10.1002/2015JC010950

CrossRef Full Text | Google Scholar

Pelegrí, J. L., and Csanady, G. T. (1991). Nutrient transport and mixing in the Gulf Stream. J. Geophys. Res. 96, 2577–2583. doi: 10.1029/90JC02535

CrossRef Full Text | Google Scholar

Pelegrí, J. L., Csanady, G. T., and Martins, A. (1996). The North Atlantic nutrient stream. J. Oceanogr. 52, 275–299. doi: 10.1007/BF02235924

CrossRef Full Text | Google Scholar

Pelland, N. A., Eriksen, C. C., and Lee, C. M. (2013). Subthermocline eddies over the Washington continental slope as observed by Seagliders, 2003–09. J. Phys. Oceanogr. 43, 2025–2053. doi: 10.1175/jpo-d-12-086.1

CrossRef Full Text | Google Scholar

Perry, M. J., Sackmann, B. S., Eriksen, C. C., and Lee, C. M. (2008). Seaglider observations of blooms and subsurface chlorophyll maxima off the Washington coast. Limnol. Oceanogr. 53, 2169–2179. doi: 10.4319/lo.2008.53.5_part_2.2169

CrossRef Full Text | Google Scholar

Pickart, R. S., and Smethie, W. M. (1993). How does the Deep Western Boundary Current cross the Gulf Stream? J. Phys. Oceanogr. 23, 2602–2616. doi: 10.1038/s41598-018-22758-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Pickart, R. S., and Watts, D. R. (1990). Deep western boundary current variability at Cape Hatteras. J. Mar. Res. 48, 765–791. doi: 10.1357/002224090784988674

CrossRef Full Text | Google Scholar

Pietri, A., Echevin, V., Testor, P., Chaigneau, A., Mortier, L., Grados, C., et al. (2014). Impact of a coastal-trapped wave on the near-coastal circulation of the Peru upwelling system from glider data. Oceans 119, 2109–2120. doi: 10.1002/2013JC009270

CrossRef Full Text | Google Scholar

Pietri, A., Testor, P., Echevin, V., Chaigneau, A., Mortier, L., Eldin, G., et al. (2013). Finescale vertical structure of the upwelling system off southern Peru as observed from glider data. J. Phys. Oceanogr. 43, 631–646. doi: 10.1175/JPO-D-12-035.1

CrossRef Full Text | Google Scholar

Pitcher, G. C., and Probyn, T. A. (2011). Anoxia in southern Benguela during the autumn of 2009 and its linkage to a bloom of the dinoflagellate Ceratium balechii. Harmful Algae 11, 23–32. doi: 10.1016/j.hal.2011.07.001

CrossRef Full Text | Google Scholar

Pitcher, G. C., Probyn, T. A., du Randt, A., Lucas, A. J., Bernard, S., Evers-King, H., et al. (2014). Dynamics of oxygen depletion in the nearshore of a coastal embayment of the southern Benguela upwelling system. J. Geophys. Res. Oceans 119, 2183–2200. doi: 10.1002/2013JC009443

CrossRef Full Text | Google Scholar

Pizarro, O., Ramírez, N., Castillo, M. I., Cifuentes, U., Rojas, W., and Pizarro-Koch, M. (2016). Underwater glider observations in the oxygen minimum zone off central Chile. Bull. Am. Meteor. Soc. 97, 1783–1789. doi: 10.1175/BAMS-D-14-00040.1

CrossRef Full Text | Google Scholar

Polo, I., Lazar, A., Rodriguez-Fonseca, B., and Arnault, S. (2008). Oceanic Kelvin waves and tropical Atlantic intraseasonal variability: 1. Kelvin wave characterization. J. Geophys. Res. 113:C07009. doi: 10.1029/2007JC004495

CrossRef Full Text | Google Scholar

Pontes, G. M., Sen Gupta, A., and Taschetto, A. S. (2016). Projected changes to South Atlantic boundary currents and confluence region in the CMIP5 models: the role of wind and deep ocean changes. Environ. Res. Lett. 11:094013. doi: 10.1088/1748-9326/11/9/094013

CrossRef Full Text | Google Scholar

Poulain, P. M., and Niiler, P. P. (1989). Statistical-analysis of the surface circulation in the California Current System using satellite-tracked drifters. J. Phys. Oceanogr. 19, 1588–1603. doi: 10.1175/1520-0485(1989)019<1588:saotsc>2.0.co;2

CrossRef Full Text | Google Scholar

Probyn, T. A., Mitchellinnes, B. A., Brown, P. C., Hutchings, L., and Carter, R. A. (1994). Review of primary production and related processes on the Agulhas Bank. S. Afr. J. Sci. 90, 166–173.

Google Scholar

Qiu, B., and Chen, S. (2005). Variability of the kuroshio extension jet, recirculation gyre, and mesoscale eddies on decadal time scales. J. Phys. Oceanogr. 35, 2090–2103. doi: 10.1175/JPO2807.1

CrossRef Full Text | Google Scholar

Qiu, B., and Chen, S. (2010). Interannual-to-decadal variability in the bifurcation of the North Equatorial Current off the Philippines. J. Phys. Oceanogr. 40, 2525–2538. doi: 10.1175/2010JPO4462.1

CrossRef Full Text | Google Scholar

Qiu, B., Chen, S., Schneider, N., and Taguchi, B. (2014). A coupled decadal prediction of the dynamic state of the Kuroshio Extension system. J. Clim. 27, 1751–1764. doi: 10.1175/JCLI-D-13-00318.1

CrossRef Full Text | Google Scholar

Rainville, L., Lee, C. M., Rudnick, D. L., and Yang, K.-C. (2013). Propagation of internal tides generated near Luzon Strait: observations from autonomous gliders. J. Geophys. Res. 118, 4125–4138. doi: 10.1002/jgrc.20293

CrossRef Full Text | Google Scholar

Reum, J. C. P., Alin, S. R., Harvey, C. J., Bednaršek, N., Evans, W., Feely, R. A., et al. (2016). Interpretation and design of ocean acidification experiments in upwelling systems in the context of carbonate chemistry co-variation with temperature and oxygen. ICES J. Mar. Sci. 73, 582–595. doi: 10.1093/icesjms/fsu231

CrossRef Full Text | Google Scholar

Reum, J. C., Alin, S. R., Feely, R. A., Newton, J., Warner, M., and McElhany, P. (2014). Seasonal carbonate chemistry covariation with temperature, oxygen, and salinity in a fjord estuary: implications for the design of ocean acidification experiments. PLoS One 9:e89619. doi: 10.1371/journal.pone.0089619

PubMed Abstract | CrossRef Full Text | Google Scholar

Révelard, A., Frankignoul, C., Sennéchael, N., Kwon, Y. O., and Qiu, B. (2016). Influence of the decadal variability of the kuroshio extension on the atmospheric circulation in the cold season. J. Clim. 29, 2123–2144. doi: 10.1175/JCLI-D-15-0511.1

CrossRef Full Text | Google Scholar

Richardson, P. L. (2007). Agulhas leakage into the Atlantic estimated with subsurface floats and surface drifters. Deep Sea Res. I 54, 1361–1389. doi: 10.1016/j.dsr.2007.04.010

CrossRef Full Text | Google Scholar

Richter, I. (2015). Climate model biases in the eastern tropical oceans: causes, impacts and ways forward. WIREs Clim. Chang 6, 345–358. doi: 10.1002/wcc.338

CrossRef Full Text | Google Scholar

Ridgway, K. R., and Condie, S. A. (2004). The 5500-km-long boundary flow off western and southern Australia. J. Geophys. Res. 109:C04017. doi: 10.1029/2003JC001921

CrossRef Full Text | Google Scholar

Ridgway, K. R., and Godfrey, J. S. (2015). The source of the Leeuwin Current seasonality. J. Geophys. Res. 120, 6843–6864. doi: 10.1002/2015JC011049

CrossRef Full Text | Google Scholar

Rio, M. H., and Santoleri, R. (2018). Improved global surface currents from the merging of altimetry and Sea surface temperature data. Remote Sens. Environ. 216, 770–785. doi: 10.1016/j.rse.2018.06.003

CrossRef Full Text | Google Scholar

Rio, M.-H., Mulet, S., and Picot, N. (2014). Beyond GOCE for the ocean circulation estimate: synergetic use of altimetry, gravimetry, and in situ data provides new insight into geostrophic and Ekman currents. Geophys. Res. Lett. 41, 8918–8925. doi: 10.1002/2014GL061773

CrossRef Full Text | Google Scholar

Riser, S. C., Freeland, H. J., Roemmich, D., Wijffels, S., Troisi, A., Belbéoch, M., et al. (2016). Fifteen years of ocean observations with the global Argo array. Nat. Clim. Change 6, 145–153. doi: 10.1038/nclimate2872

CrossRef Full Text | Google Scholar

Rodgers, K. B., Sarmiento, J. L., Crevoisier, C., de Boyer Montegut, C., Metzl, N., and Aumont, O. (2008). A wintertime uptake window for anthropogenic CO2 in the North Pacific. Glob. Biogeochem. Cycles 22:GB2020. doi: 10.1029/2006GB002920

CrossRef Full Text | Google Scholar

Rodrigues, R. R., Rothstein, L. M., and Wimbush, M. (2007). Seasonal variability of the South Equatorial Current bifurcation in the Atlantic Ocean: a numerical study. J. Phys. Oceanogr. 37, 16–30. doi: 10.1175/JPO2983.1

CrossRef Full Text | Google Scholar

Roemmich, D., Alford, M. H., Claustre, H., Johnson, K. S, King, B., Moum, J., et al. (2019). On the future of Argo: A global, full-depth, multi-disciplinary array. Front. Mar. Sci. 6:439. doi: 10.3389/fmars.2019.00439

CrossRef Full Text | Google Scholar

Roemmich, D., Boehme, L., Claustre, H., Freeland, H., Fukasawa, M., Goni, G., et al. (2010). “Integrating the ocean observing system: mobile platforms,” in Proceedings of OceanObs’09: Sustained Ocean Observations and Information for Society, eds J. Hall, D. E. Harrison, and D. Stammer (Venice: ESA Publication WPP-306), 21–25. doi: 10.5270/OceanObs09

CrossRef Full Text | Google Scholar

Rossby, H. T., Flagg, C. N., and Donohue, K. (2010). On the variability of Gulf Stream transport from seasonal to decadal timescales. J. Mar. Res. 68, 503–522. doi: 10.1357/002224010794657128

CrossRef Full Text | Google Scholar

Rossby, T., Flagg, C. N., Donohue, K., Sanchez-Franks, A., and Lillibridge, J. (2014). On the long-term stability of Gulf Stream transport based on 20 years of direct measurements. Geophys. Res. Lett. 41, 114–120. doi: 10.1002/2013GL058636

CrossRef Full Text | Google Scholar

Rouault, M., Reason, C. J. C., Lutjeharms, J. R. E., and Beljaars, A. C. M. (2003). Underestimation of latent and sensible heat fluxes above the Agulhas Current in NCEP and ECMWF analyses. J. Clim. 16, 776–782. doi: 10.1175/1520-0442(2003)016<0776:uolash>2.0.co;2

CrossRef Full Text | Google Scholar

Roughan, M., and Morris, B. D. (2011). “Using high-resolution ocean timeseries data to give context to long term hydrographic sampling off Port Hacking, NSW, Australia,” in Proceedings of the Oceans’11 MTS/IEEE Kona, Waikoloa, HI, doi: 10.23919/OCEANS.2011.6107032

CrossRef Full Text | Google Scholar

Roughan, M., Schaeffer, A., and Kioroglou, S. (2013). “Assessing the design of the NSW-IMOS Moored Observation Array from 2008–2013: recommendations for the future,” in Proceedinds of the 2013 OCEANS—San Diego, San Diego, CA, doi: 10.23919/OCEANS.2013.6741092

CrossRef Full Text | Google Scholar

Roughan, M., Schaeffer, A., and Suthers, I. M. (2015). Sustained ocean observing along the coast of Southeastern Australia: NSW-IMOS 2007–2014. Coast. Ocean Observing Syst. 2015, 76–98. doi: 10.1016/B978-0-12-802022-7.00006-7

CrossRef Full Text | Google Scholar

Rudnick, D. L. (2017). Data from: Spray Underwater Glider Campaign in Gulf of Mexico. La Jolla, CA: Instrument Development Group, Scripps Institution of Oceanography, doi: 10.21238/s8SPRAY0420

CrossRef Full Text | Google Scholar

Rudnick, D. L. (2016a). Data from: California Underwater Glider Network. La Jolla, CA: Instrument Development Group, Scripps Institution of Oceanography, doi: 10.21238/s8SPRAY1618

CrossRef Full Text | Google Scholar

Rudnick, D. L. (2016b). Ocean research enabled by underwater gliders. Annu. Rev. Mar. Sci. 8, 519–541. doi: 10.1146/annurev-marine-122414-033913

PubMed Abstract | CrossRef Full Text | Google Scholar

Rudnick, D. L., and Cole, S. T. (2011). On sampling the ocean using underwater gliders. J. Geophys. Res. 116:C08010. doi: 10.1029/2010JC006849

CrossRef Full Text | Google Scholar

Rudnick, D. L., Gopalakrishnan, G., and Cornuelle, B. D. (2015a). Cyclonic eddies in the Gulf of Mexico: observations by underwater gliders and simulations by numerical model. J. Phys. Oceanogr. 45, 313–326. doi: 10.1175/JPO-D-14-0138.1

CrossRef Full Text | Google Scholar

Rudnick, D. L., Jan, S., and Lee, C. M. (2015b). A new look at circulation in the western North Pacific. Oceanography 28, 16–23. doi: 10.5670/oceanog.2015.77

CrossRef Full Text | Google Scholar

Rudnick, D. L., Jan, S., Centurioni, L., Lee, C., Lien, R.-C., Wang, J., et al. (2011). Seasonal and mesoscale variability of the Kuroshio near its origin. Oceanography 24, 52–63. doi: 10.5670/oceanog.2011.94

CrossRef Full Text | Google Scholar

Rudnick, D. L., Johnston, T. M., and Sherman, J. T. (2013). High-frequency internal waves near the Luzon Strait observed by underwater gliders. J. Geophys. Res. Oceans 118, 774–784. doi: 10.1002/jgrc.20083

CrossRef Full Text | Google Scholar

Rudnick, D. L., Zaba, K. D., Todd, R. E., and Davis, R. E. (2017). A climatology of the California Current System from a network of underwater gliders. Prog. Oceanogr. 154, 64–106. doi: 10.1016/j.pocean.2017.03.002

CrossRef Full Text | Google Scholar

Rühs, S., Getzlaff, K., Durgadoo, J. V., Biastoch, A., and Böning, C. W. (2015). On the suitability of North Brazil current transport estimates for monitoring basin-scale AMOC changes. Geophys. Res. Lett. 42, 8072–8080. doi: 10.1002/2015GL065695

CrossRef Full Text | Google Scholar

Ryan, J. P., Ueki, I., Chao, Y., Zhang, H., Polito, P. S., and Chavez, F. P. (2006). Western Pacific modulation of large phytoplankton blooms in the central and eastern equatorial Pacific. J. Geophys. Res. 111:G02013. doi: 10.1029/2005JG000084

CrossRef Full Text | Google Scholar

Rypina, I. I., Llopiz, J. K., Pratt, L. M., and Lozier, M. S. (2014). Dispersal pathways of American eel larvae from the Sargasso Sea. Limnol. Oceanogr. 59, 1704–1714. doi: 10.4319/lo.2014.59.5.1704

CrossRef Full Text | Google Scholar

Saba, V. S., Griffies, S. M., Anderson, W. G., Winton, M., Alexander, M. A., Delworth, T. L., et al. (2016). Enhanced warming of the Northwest Atlantic Ocean under climate change. J. Geophys. Res. 121, 118–132. doi: 10.1002/2015JC011346

PubMed Abstract | CrossRef Full Text | Google Scholar

Saraceno, M., Guerrero, R., Piola, A., Provost, C., Perault, F., Ferrari, R., et al. (2017). Malvinas Current 2014–2015: Mooring Velocities. France: SEANOE, doi: 10.17882/51492

CrossRef Full Text | Google Scholar

Schaeffer, A., Roughan, M., Jones, E. M., and White, D. (2016). Physical and biogeochemical spatial scales of variability in the East Australian Current separation from shelf glider measurements. Biogeosciences 13, 1967–1975. doi: 10.5194/bg-13-1967-2016

CrossRef Full Text | Google Scholar

Schaeffer, A. M., Roughan, T., Austin, J. D., Everett, D., Griffin, B., Hollings, E., et al. (2016). Mean hydrography on the continental shelf from 26 repeat glider deployments along Southeastern Australia. Sci. Data 3:160070. doi: 10.1038/sdata.2016.70

PubMed Abstract | CrossRef Full Text | Google Scholar

Schaeffer, A., and Roughan, M. (2015). Influence of a western boundary current on shelf dynamics and upwelling from repeat glider deployments. Geophys. Res. Lett. 42, 121–128. doi: 10.1002/2014GL062260

CrossRef Full Text | Google Scholar

Schaeffer, A., and Roughan, M. (2017). Subsurface intensification of marine heatwaves off southeastern Australia: the role of stratification and local winds. Geophys. Res. Lett. 44, 5025–5033. doi: 10.1002/2017GL073714

CrossRef Full Text | Google Scholar

Schaeffer, A., Gramoulle, A., Roughan, M., and Mantovanelli, A. (2017). Characterizing frontal eddies along the East Australian Current from HF radar observations. J. Geophys. Res. 122, 3964–3980. doi: 10.1002/2016JC012171

CrossRef Full Text | Google Scholar

Schaeffer, A., Roughan, M., and Morris, B. (2013). Cross-shelf dynamics in a western boundary current. Implications for upwelling. J. Phys. Oceanogr. 43, 1042–1059. doi: 10.1175/JPO-D-12-0177.1

CrossRef Full Text | Google Scholar

Schaeffer, A., Roughan, M., and Morris, B. (2014). Corrigendum. J. Phys. Oceanogr. 44, 2812–2813. doi: 10.1175/JPO-D-14-0091.1

CrossRef Full Text | Google Scholar

Schmid, C., and Majumder, S. (2018). Transport variability of the Brazil Current from observations and a data assimilation model. Ocean Sci. 14, 417–436. doi: 10.5194/os014-417-2018

CrossRef Full Text | Google Scholar

Schneider, W., Donoso, D., Garcés-Vargas, J., and Escribano, R. (2016). Water-column cooling and sea surface salinity increase in the upwelling region off central-south Chile driven by a poleward displacement of the South Pacific. Prog. Oceanogr. 141, 38–58. doi: 10.1016/j.pocean.2016.11.004

CrossRef Full Text | Google Scholar

Schönau, M. C., and Rudnick, D. L. (2017). Mindanao Current and Undercurrent: thermohaline structure and transport from repeat glider observations. J. Phys. Oceanogr. 47, 2055–2075. doi: 10.1175/JPO-D-16-0274.1

CrossRef Full Text | Google Scholar

Schönau, M. C., Rudnick, D. L., Cerovecki, I., Gopalakrishnan, G., Cornuelle, B. D., McClean, J. L., et al. (2015). The Mindanao Current: mean structure and connectivity. Oceanography 28, 34–45. doi: 10.5670/oceanog.2015.79

CrossRef Full Text | Google Scholar

Schott, F. A., Fischer, J., and Stramma, L. (1998). Transports and pathways of the upper-layer circulation in the western tropical Atlantic. J. Phys. Oceanogr. 28, 1904–1928. doi: 10.1175/1520-0485(1998)028<1904:tapotu>2.0.co;2

CrossRef Full Text | Google Scholar

Schott, F., Dengler, M., Zantopp, R. J., Stramma, L., Fischer, J., and Brandt, P. (2005). The shallow and deep western boundary circulation of the South Atlantic at 5°–11°S. J. Phys. Oceanogr. 35, 2031–2053. doi: 10.1175/JPO2813.1

CrossRef Full Text | Google Scholar

Schott, F., Xie, S.-P., and McCreary, J. P. (2009). Indian Ocean circulation and climate variability. Rev. Geophys. 47:RG1002. doi: 10.1029/2007RG000245

CrossRef Full Text | Google Scholar

Sen Gupta, A., Ganachaud, A., McGregor, S., Brown, J. N., and Muir, L. (2012). Drivers of the projected changes to the Pacific Ocean equatorial circulation. Geophys. Res. Lett. 39:L09605. doi: 10.1029/2012GL051447

CrossRef Full Text | Google Scholar

Send, U. (2018). Data from: Spray Glider Data in Support of Mooring Observations from the CORC Project in the California Current Starting 2007. La Jolla, CA: Instrument Development Group, Scripps Institution of Oceanography, doi: 10.21238/s8SPRAY8857

CrossRef Full Text | Google Scholar

Send, U., Davis, R. E., Fischer, J., Imawaki, S., Kessler, W., Meinen, C., et al. (2010). “A global boundary current circulation observing network,” in Proceedings of OceanObs’09: Sustained Ocean Observations and Information for Society, Vol. 2, eds J. Hall, D. E. Harrison, and D. Stammer (Venice: ESA Publication WPP-306), doi: 10.5270/OceanObs09.cwp.78

CrossRef Full Text | Google Scholar

Send, U., Regier, L., and Jones, B. (2013). Use of underwater gliders for acoustic data retrieval from subsurface oceanographic instrumentation and bidirectional communication in the deep ocean. J. Atmos. Ocean Technol. 30, 984–998. doi: 10.1175/JTECH-D-11-00169.1

CrossRef Full Text | Google Scholar

Shannon, L. V., and Nelson, G. (1996). “The Benguela: large-scale features and processes and system variability,” in The South Atlantic: Present and Past Circulation, eds G. Wefer, W. H. Berger, G. Seidler, and D. J. Webb (Berlin: Springer-Verlag), 163–210. doi: 10.1007/978-3-642-80353-6_9

CrossRef Full Text | Google Scholar

Shinoda, A., Aoyama, J., Miller, M. J., Otake, T., Mochioka, N., Watanabe, S., et al. (2011). Evaluation of the larval distribution and migration of the Japanese eel in the western North Pacific. Rev. Fish. Biol. Fisher. 21, 591–611. doi: 10.1007/s11160-010-9195-1

CrossRef Full Text | Google Scholar

Siedlecki, S. A., Kaplan, I. C., Hermann, A. J., Nguyen, T. T., Bond, N. A., Newton, J. A., et al. (2016). Experiments with seasonal forecasts of ocean conditions for the Northern region of the California Current upwelling system. Sci. Rep. 6:27203. doi: 10.1038/srep27203

PubMed Abstract | CrossRef Full Text | Google Scholar

Silva, N., Rojas, N., and Fedele, A. (2009). Water masses in the Humboldt Current System: properties, distribution, and the nitrate deficit as a chemical water mass tracer for Equatorial Subsurface Water off Chile. Deep Sea Res. Part II 56, 1004–1020. doi: 10.1016/j.dsr2.2008.12.013

CrossRef Full Text | Google Scholar

Siqueira, L., and Kirtman, B. P. (2016). Atlantic near-term climate variability and the role of a resolved Gulf Stream. Geophys. Res. Lett. 43, 3964–3972. doi: 10.1002/2016GL068694

CrossRef Full Text | Google Scholar

Slangen, A. B. A., Carson, M., Katsman, C. A., van de Wal, R. S. W., Köhl, A., Vermeersen, L. L. A., et al. (2014). Projecting twenty-first century regional sea-level changes. Clim. Chang. 124, 317–332. doi: 10.1007/s10584-014-1080-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Sloyan, B. M., and O’Kane, T. J. (2015). Drivers of decadal variability in the Tasman Sea. J. Geophys. Res. 120, 3193–3210. doi: 10.1002/2014JC010550

CrossRef Full Text | Google Scholar

Sloyan, B. M., Ridgway, K., and Cowley, R. (2016). The East Australian Current and property transport at 27°S from 2012-2013. J. Phys. Oceanogr. 46, 993–1008. doi: 10.1175/JPO-D-15-0052.1

CrossRef Full Text | Google Scholar

Smith, L. M., Barth, J. A., Kelley, D. S., Plueddemann, A., Rodero, I., Ulses, G. A., et al. (2018). The Ocean Observatories Initiative. Oceanography 31, 16–35. doi: 10.5670/oceanog.2018.105

CrossRef Full Text | Google Scholar

Smith, N., Kessler, W. S., Cravatte, S. E., Sprintall, J., Wijffels, S. E., Cronin, M. F., et al. (2019). Tropical Pacific Observing System. Front. Mar. Sci. 6:31. doi: 10.3389/fmars.2019.00031

CrossRef Full Text | Google Scholar

Soh, H. S., and Kim, S. Y. (2018). Diagnostic characteristics of submesoscale coastal surface currents. J. Geophys. Res. 123, 1838–1859. doi: 10.1002/2017JC013428

CrossRef Full Text | Google Scholar

Song, H., Edwards, C. A., Moore, A. M., and Fiechter, J. (2012). Incremental four-dimensional variational data assimilation of positive-definite oceanic variables using a logarithm transformation. Ocean Model. 54–55, 1–17. doi: 10.1016/j.ocemod.2012.06.001

CrossRef Full Text | Google Scholar

Spadone, A., and Provost, C. (2009). Variations in the Malvinas Current volume transport since October 1992. J. Geophys. Res. 114:C02002. doi: 10.1029/2008JC004882

CrossRef Full Text | Google Scholar

St. Laurent, L., and Merrifield, S. (2017). Measurements of near-surface turbulence and mixing from autonomous ocean gliders. Oceanography 30, 116–125. doi: 10.5670/oceanog.2017.231

CrossRef Full Text | Google Scholar

Steinfeldt, R., Sültenfuß, J., Dengler, M., Fischer, T., and Rhein, M. (2015). Coastal upwelling off Peru and Mauritania inferred from helium isotope disequilibrium. Biogeosciences 12, 7519–7533. doi: 10.5194/bg-12-7519-2015

CrossRef Full Text | Google Scholar

Strub, P. T., James, C., Combes, V., Matano, R., Piola, A., Palma, E., et al. (2015). Altimeter-derived seasonal circulation on the SW Atlantic Shelf: 27°–43°S. J. Geophys. Res. Oceans 120, 3391–3418. doi: 10.1002/2015JC010769

PubMed Abstract | CrossRef Full Text | Google Scholar

Stuart-Smith, R. D., Brown, C. J., Ceccarelli, D. M., and Edgar, G. J. (2018). Ecosystem restructuring along the Great Barrier Reef following mass coral bleaching. Nature 560, 92–96. doi: 10.1038/s41586-018-0359-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Subramanian, A., Balmaseda, M. A., Chattopadhyay, R., Centurioni, L. R., Cornuelle, B. D., DeMott, C., et al. (2019). Ocean observations to improve our understanding, modeling, and forecasting of subseasonal-to-seasonal variability. Front. Mar. Sci. 6:427. doi: 10.3389/fmars.2019.00427

CrossRef Full Text | Google Scholar

Sugimoto, S., and Hanawa, K. (2014). Influence of Kuroshio path variation south of Japan on formation of subtropical mode water. J. Phys. Oceanogr. 44, 1065–1077. doi: 10.1175/jpo-d-13-0114.1

CrossRef Full Text | Google Scholar

Sun, X., Vizy, E. K., and Cook, K. H. (2018). Land–atmosphere–ocean interactions in the southeastern Atlantic: interannual variability. Clim. Dyn. 52:539. doi: 10.1007/s00382-018-4155-x

CrossRef Full Text | Google Scholar

Suthers, I. M., Young, J. W., Baird, M. E., Roughan, M., Everett, J. E., Brassington, G. B., et al. (2011). The strengthening East Australian Current, its eddies and biological effects—an introduction and overview. Deep Sea Res. 58, 538–546. doi: 10.1016/j.dsr2.2010.09.029

CrossRef Full Text | Google Scholar

Sutton, A. J., Sabine, C. L., Feely, R. A., Cai, W.-J., Cronin, M. F., McPhaden, M. J., et al. (2016). Using present-day observations to detect when anthropogenic change forces surface ocean carbonate chemistry outside preindustrial bounds. Biogeosciences 13, 5065–5083. doi: 10.5194/bg-2016-104

CrossRef Full Text | Google Scholar

Sutton, A. J., Sabine, C. L., Maenner-Jones, S., Lawrence-Slavas, N., Meinig, C., Feely, R. A., et al. (2014). A high-frequency atmospheric and seawater pCO2 data set from 14 open-ocean sites using a moored autonomous system. Earth Syst. Sci. Data 6, 353–366. doi: 10.5194/essd-6-353-2014

CrossRef Full Text | Google Scholar

Sutton, A. J., Wanninkhof, R., Sabine, C. L., Feely, R. A., Cronin, M. F., and Weller, R. A. (2017). Variability and trends in surface seawater pCO2 and CO2 flux in the Pacific Ocean. Geophys. Res. Lett. 44, 5627–5636. doi: 10.1002/2017GL073814

CrossRef Full Text | Google Scholar

Takahashi, T., Sutherland, S. C., Wanninkhof, R., Sweeney, C., Feely, R. A., Chipman, D. W., et al. (2009). Climatological mean and decadal change in surface ocean pCO2, and net sea–air CO2 flux over the global oceans. Deep Sea Res. II 56, 554–577. doi: 10.1016/j.dsr2.2008.12.009

CrossRef Full Text | Google Scholar

Talley, L. D., Feely, R. A., Sloyan, B. M., Wanninkhof, R., Baringer, M. O., Bulliser, J. L., et al. (2016). Changes in ocean heat, carbon content, and ventilation: a review of the first decade of GO-SHIP global repeat hydrography. Ann. Rev. Mar. Sci. 8, 185–215. doi: 10.1146/annurev-marine-052915-100829

PubMed Abstract | CrossRef Full Text | Google Scholar

Tchipalanga, P., Dengler, M., Brandt, P., Kopte, R., Macuéria, M., Coelho, P., et al. (2018). Eastern boundary circulation and hydrography off Angola—building Angolan oceanographic capacities. Bull. Am. Meteor. Soc. 99, 1589–1605. doi: 10.1175/BAMS-D-17-0197.1

CrossRef Full Text | Google Scholar

Testor, P., DeYoung, B., Rudnick, D. L., Glenn, S., Hayes, D., Lee, C., et al. (2019). OceanGliders: a component of the integrated GOOS. Front. Mar. Sci. 6:422. doi: 10.3389/fmars.2019.00422

CrossRef Full Text | Google Scholar

Thompson, J. D., and Schmitz, W. J. Jr. (1989). A limited-area model of the Gulf Stream: design, initial experiments, and model-data intercomparison. J. Phys. Oceanogr. 19, 791–814. doi: 10.1175/1520-0485(1989)019<0791:alamot>2.0.co;2

CrossRef Full Text | Google Scholar

Thomsen, S., Karstensen, J., Kiko, R., Krahmann, G., Dengler, M., and Engel, A. (2019). Remote and local drivers of oxygen and nitrate variability in the shallow oxygen minimum zone off Mauritania in June 2014. Biogeosciences 16, 979–998. doi: 10.5194/bg-16-979-2019

CrossRef Full Text | Google Scholar

Todd, R. E. (2017). High-frequency internal waves and thick bottom mixed layers observed by gliders in the Gulf Stream. Geophys. Res. Lett. 44, 6316–6325. doi: 10.1002/2017GL072580

CrossRef Full Text | Google Scholar

Todd, R. E., and Locke-Wynn, L. (2017). Underwater glider observations and the representation of western boundary currents in numerical models. Oceanography 30, 88–89. doi: 10.5670/oceanog.2017.225

CrossRef Full Text | Google Scholar

Todd, R. E., and Owens, B. (2016). Data from: Gliders in the Gulf Stream. La Jolla, CA: Instrument Development Group, Scripps Institution of Oceanography, doi: 10.21238/s8SPRAY2675

CrossRef Full Text | Google Scholar

Todd, R. E., Asher, T. G., Heiderich, J., Bane, J. M., and Luettich, R. A. (2018a). Transient response of the Gulf Stream to multiple hurricanes in 2017. Geophys. Res. Lett. 45, 10509–10519. doi: 10.1029/2018GL079180

CrossRef Full Text | Google Scholar

Todd, R. E., Rudnick, D. L., Centurioni, L. R., Jayne, S. R., and Lee, C. M. (2018b). “Boundary current observations with ALPS,” in ALPS II—Autonomous and Lagrangian Platforms and Sensors. A Report of the ALPS II Workshop, eds D. Rudnick, D. Costa, C. Lee, and M.-L. Timmermans (Washington, DC: National Oceanographic Partnership Program).

Google Scholar

Todd, R. E., Gawarkiewicz, G. G., and Owens, W. B. (2013). Horizontal scales of variability over the Middle Atlantic Bight shelf break and continental rise from finescale observations. J. Phys. Oceanogr. 43, 222–230. doi: 10.1175/JPO-D-12-099.1

CrossRef Full Text | Google Scholar

Todd, R. E., Owens, W. B., and Rudnick, D. L. (2016). Potential vorticity structure in the North Atlantic western boundary current from underwater glider observations. J. Phys. Oceanogr. 46, 327–348. doi: 10.1175/JPO-D-15-0112.1

CrossRef Full Text | Google Scholar

Todd, R. E., Rudnick, D. L., Davis, R. E., and Ohman, M. D. (2011a). Underwater gliders reveal rapid arrival of El Niño effects off California’s coast. Geophys. Res. Lett. 38:L03609. doi: 10.1029/2010GL046376

CrossRef Full Text | Google Scholar

Todd, R. E., Rudnick, D. L., Mazloff, M. R., Davis, R. E., and Cornuelle, B. D. (2011b). Poleward flows in the southern California Current System: Glider observations and numerical simulation. J. Geophys. Res. 116:C02026. doi: 10.1029/2010JC006536

CrossRef Full Text | Google Scholar

Todd, R. E., Rudnick, D. L., Mazloff, M. R., Cornuelle, B. D., and Davis, R. E. (2012). Thermohaline structure in the California Current System: Observations and modeling of spice variance. J. Geophys. Res. 117:C02008. doi: 10.1029/2011JC007589

CrossRef Full Text | Google Scholar

Todd, R. E., Rudnick, D. L., Sherman, J. T., Owens, W. B., and George, L. (2017). Absolute velocity estimates from autonomous underwater gliders equipped with Doppler current profilers. J. Atmos. Ocean. Technol. 34, 309–333. doi: 10.1175/JTECH-D-16-0156.1

CrossRef Full Text | Google Scholar

Toole, J. M., Andres, M., Le Bras, I. A., Joyce, T. M., and McCartney, M. S. (2017). Moored observations of the deep western boundary current in the NW Atlantic: 2004–2014. Oceans 122, 7488–7505. doi: 10.1002/2017JC012984

CrossRef Full Text | Google Scholar

Trowbridge, J., Weller, R., Kelley, D., Dever, E., Plueddemann, A., Barth, J. A., et al. (2019). The Ocean Observatories Initiative. Front. Mar. Sci. 6:74. doi: 10.3389/fmars.2019.00074

CrossRef Full Text | Google Scholar

Valdivieso, M., Haines, K., Balmaseda, M., Chang, Y.-S., Drevillon, M., Ferry, N., et al. (2017). An assessment of air–sea heat fluxes from ocean and coupled reanalyses. Clim. Dyn. 49, 983–1008. doi: 10.1007/s00382-015-2843-3

CrossRef Full Text | Google Scholar

Valla, D., and Piola, A. R. (2015). Evidence of upwelling events at the northern Patagonian shelf break. J. Geophys. Res. Oceans 120, 7635–7656. doi: 10.1002/2015jc011002

CrossRef Full Text | Google Scholar

Valla, D., Piola, A. R., Meinen, C. S., and Campos, E. (2018). Strong mixing and recirculation in the northwestern Argentine Basin. J. Geophys. Res. Oceans 123, 4624–4648. doi: 10.1029/2018JC013907

CrossRef Full Text | Google Scholar

Van der Lingen, C. D., Fréon, P., Hutchings, L., Roy, C., Bailey, G. W., Bartholomae, C., et al. (2006). “Forecasting shelf processes of relevance to living marine resources in the BCLME,” in Benguela: Predicting a Large Marine Ecosystem, Large Mar. Ecosyst. Ser, Vol. 14, ed. V. Shannon (Amsterdam: Elsevier), 309–347. doi: 10.1016/s1570-0461(06)80019-4

CrossRef Full Text | Google Scholar

van Sebille, E., Baringer, M. O., Johns, W. E., Beal, L. M., de Jong, M. F., van Aken, H. M., et al. (2011). Propagation pathways of classical Labrador Sea Water from its source region to 26°N. J. Geophys. Res. 116:C12027. doi: 10.1029/2011JC007171

CrossRef Full Text | Google Scholar

van Sebille, E., Beal, L. M., and Biastoch, A. (2010). Sea surface slope as a proxy for Agulhas Current strength. Geophys. Res. Lett 37:L09610.

Google Scholar

van Sebille, E., Wilcox, C., Lebreton, L., Maximenko, N., Hardesty, B. D., Van Franekar, J. A., et al. (2015). A global inventory of small floating plastic debris. Environ. Res. Lett 10:124006. doi: 10.1088/1748-9326/10/12/124006

CrossRef Full Text | Google Scholar

Van Uffelen, L. J., Roth, E. H., Howe, B. M., Oleson, E. M., and Barkley, Y. (2017). A Seaglider-integrated digital monitor for bioacoustic sensing. IEEE J. Ocean. Eng. 42, 800–807. doi: 10.1109/JOE.2016.2637199

CrossRef Full Text | Google Scholar

Vélez-Belchí, P., Centurioni, L. R., Lee, D.-K., Jan, S., and Niiler, P. P. (2013). Eddy induced Kuroshio intrusions onto the continental shelf of the East China Sea. J. Mar. Res. 71, 83–107. doi: 10.1357/002224013807343470

PubMed Abstract | CrossRef Full Text | Google Scholar

Vivier, F., and Provost, C. (1999). Direct velocity measurements in the malvinas current. J. Geophys. Res. 104, 21083–21103. doi: 10.1029/1999JC900163

CrossRef Full Text | Google Scholar

Volkov, D. L., Baringer, M., Smeed, D., Johns, W., and Landerer, F. (2019). Teleconnection between the Atlantic meridional overturning circulation and sea level in the Mediterranean Sea. J. Clim. 32, 935–955. doi: 10.1175/JCLI-D-18-0474.1

CrossRef Full Text | Google Scholar

Wada, A., Cronin, M. F., Sutton, A. J., Kawai, Y., and Ishii, M. (2013). Numerical simulations of oceanic pCO2 variations and interactions between typhoon Choi-wan (0914) and the ocean. J. Geophys. Res. Oceans 118, 2667–2684. doi: 10.1002/jgrc.20203

CrossRef Full Text | Google Scholar

Wagawa, T., Tamate, T., Kuroda, H., Ito, S., Kakehi, S., Yamanome, T., et al. (2016). Relationship between coastal water properties and adult return of chum salmon (Oncorhynchus keta) along the Sanriku coast, Japan. Fish. Oceangr. 25, 598–609. doi: 10.1111/fog.12175

CrossRef Full Text | Google Scholar

Wakita, M., Watanabe, S., Murata, A., Tsurushima, N., and Honda, M. (2010). Decadal change of dissolved inorganic carbon in the subarctic western North Pacific Ocean. Tellus B 62, 608–620. doi: 10.1111/j.1600-0889.2010.00476.x

CrossRef Full Text | Google Scholar

Wang, D., Flagg, C. N., Donohue, K., and Rossby, H. T. (2010). Wavenumber spectrum in the Gulf Stream from shipboard ADCP observations and comparison with altimetry measurements. J. Phys. Oceanogr. 40, 840–844. doi: 10.1175/2009JPO4330.1

CrossRef Full Text | Google Scholar

Wang, F., Zhang, L., Hu, D., Wang, Q., Zhai, F., and Hu, S. (2017). The vertical structure and variability of the western boundary currents east of the Philippines from direct observations. J. Oceanogr. 73, 743–758. doi: 10.1007/s10872-017-0429-x

CrossRef Full Text | Google Scholar

Weeks, S. J., Barlow, R., Roy, C., and Shillington, F. A. (2006). Remotely sensed variability of temperature and chlorophyll in the southern Benguela: upwelling frequency and phytoplankton response. Afr. J. Mar. Sci. 28, 493–509. doi: 10.2989/18142320609504201

CrossRef Full Text | Google Scholar

Weller, R. A., Bigorre, S. P., Lord, J., Ware, J. D., and Edson, J. B. (2012). A surface mooring for air–sea interaction research in the Gulf Stream. Part I: mooring design and instrumentation. J. Atmos. Ocean. Technol. 29, 1363–1376. doi: 10.1175/JTECH-D-12-00060.1

CrossRef Full Text | Google Scholar

Wijffels, S. E., Meyers, G., and Godfrey, J. S. (2008). A 20-yr average of the Indonesian Throughflow: regional currents and the interbasin exchange. J. Phys. Oceanogr. 38, 1965–1978. doi: 10.1175/2008JPO3987.1

CrossRef Full Text | Google Scholar

Wilkinson, M. D., Dumontier, M., Aalbersberg, I. J., Appleton, G., Axton, M., Baak, A., et al. (2016). The FAIR guiding principles for scientific data management and stewardship. Sci. Data 3:160018. doi: 10.1038/sdata.2016.18

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, R. G., McDonagh, E. L., Roussenov, V. M., Torres-Valdes, S., King, B., Sanders, R., et al. (2011). Nutrient streams in the North Atlantic: advective pathways of inorganic and organic nutrients. Glob. Biogeochem. Cycles 25:GB4008. doi: 10.1029/2010GB003853

CrossRef Full Text | Google Scholar

Williams, R. G., Roussenov, V., and Follows, M. J. (2006). Nutrient streams and their induction into the mixed layer. Glob. Biogeochem. Cycles 20:GB1016. doi: 10.1029/2005gb002586

CrossRef Full Text | Google Scholar

Woo, L. M., and Pattiaratchi, C. B. (2008). Hydrography and water masses off the western Australian coast. Deep Sea Res. I 55, 1090–1104. doi: 10.1016/j.dsr.2008.05.005

CrossRef Full Text | Google Scholar

Wu, C.-R., Chang, Y.-L., Oey, L.-Y., Chang, C.-W. J., and Hsin, Y.-C. (2008). Air–sea interaction between tropical cyclone Nari and Kuroshio. Geophys. Res. Lett. 35:L12605. doi: 10.1029/2008GL033942

CrossRef Full Text | Google Scholar

Wu, L., Cai, W., Zhang, L., Nakamura, H., Timmermann, A., Joyce, T., et al. (2012). Enhanced warming over the global subtropical western boundary currents. Nat. Clim. Chang. 2, 161–166. doi: 10.1038/nclimate1353

CrossRef Full Text | Google Scholar

Wyatt, L. R., Mantovanelli, A., Heron, M. L., Roughan, M., and Steinberg, C. R. (2018). Assessment of surface currents measured with high-frequency phased-array radars in two regions of complex circulation IEEE. J. Ocean. Eng. 43, 484–505. doi: 10.1109/JOE.2017.2704165

CrossRef Full Text | Google Scholar

Yamamoto, A., Palter, J. B., Dufour, C. O., Griffies, S. M., Bianchi, D., Claret, M., et al. (2018). Roles of the ocean mesoscale in the horizontal supply of mass, heat, carbon and nutrients to the Northern Hemisphere subtropical gyres. J. Geophys. Res. 123, 7016–7036. doi: 10.1029/2018JC013969

CrossRef Full Text | Google Scholar

Yang, H., Lohmann, G., Wei, W., Dima, M., Ionita, M., and Liu, J. (2016). Intensification and poleward shift of subtropical western boundary currents in a warming climate. Oceans 121, 4928–4945. doi: 10.1002/2015JC011513

CrossRef Full Text | Google Scholar

Yang, Y. J., Jan, S., Chang, M.-H., Wang, J., Mensah, V., Kuo, T.-H., et al. (2015). Mean structure and fluctuations of the Kuroshio east of Taiwan from in situ and remote observations. Oceanography 28, 74–83. doi: 10.5670/oceanog.2015.83

CrossRef Full Text | Google Scholar

Yasuda, I. (2003). Hydrographic structure and variability in the Kuroshio–Oyashio transition area. J. Oceanogr. 59, 389–402. doi: 10.1023/A:1025580313836

CrossRef Full Text | Google Scholar

Yasunaka, S., Nojiri, Y., Nakaoka, S., Ono, T., Mukai, H., and Usui, N. (2013). Monthly maps of sea surface dissolved inorganic carbon in the North Pacific: basin-wide distribution and seasonal variation. J. Geophys. Res. Oceans 118, 3843–3850. doi: 10.1002/jgrc.20279

CrossRef Full Text | Google Scholar

Yasunaka, S., Nojiri, Y., Nakaoka, S.-I., Ono, T., Whitney, F. A., and Telszewski, M. (2014). Mapping of sea surface nutrients in the North Pacific: basin-wide distribution and seasonal to interannual variability. J. Geophys. Res. Oceans 119, 7756–7771. doi: 10.1002/2014jc010318

CrossRef Full Text | 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. Geophys. Res. Lett. 43, 1241–1248. doi: 10.1002/2015GL067550

CrossRef Full Text | Google Scholar

Zantopp, R., Fischer, J., Visbeck, M., and Karstensen, J. (2017). From interannual to decadal: 17 years of boundary current transports at the exit of the Labrador Sea. J. Geophys. Res. Oceans 122, 1724–1748. doi: 10.1002/2016JC012271

CrossRef Full Text | Google Scholar

Zhang, D., Cronin, M. F., Lin, X., Inoue, R., Fassbender, A. J., Bishop, S. P., et al. (2017). Observing air–sea interaction in western boundary currents and their extension regions: considerations for OceanObs’19. CLIVAR Variations 15, 23–30. doi: 10.5065/D6SJ1JB2

CrossRef Full Text | Google Scholar

Zhang, D., Msadek, R., McPhaden, M. J., and Delworth, T. (2011). Multidecadal variability of the North Brazil Current and its connection to the atlantic meridional overturning circulation. J. Geophys. Res. 116:C04012. doi: 10.1029/2010JC006812

CrossRef Full Text | Google Scholar

Zhang, J., Gilbert, D., Gooday, A. J., Levin, L., Naqvi, S. W. A., Middelburg, J. J., et al. (2010). Natural and human-induced hypoxia and consequences for coastal areas: synthesis and future development. Biogeosciences 7, 1443–1467. doi: 10.5194/bg-7-1443-2010

CrossRef Full Text | Google Scholar

Zhang, L., Hu, D., Hu, S., Wang, F., Wang, F., and Yuan, D. (2014). Mindanao current/undercurrent measured by a subsurface mooring. J. Geophys. Res. 119, 3617–3628. doi: 10.1002/2013JC009693

CrossRef Full Text | Google Scholar

Zhang, S., Curchitser, E. N., Kang, D., Stock, C. A., and Dussin, R. (2018). Impacts of mesoscale eddies on the vertical nitrate flux in the Gulf stream region. J. Geophys. Res. 123, 497–513. doi: 10.1002/2017JC013402

CrossRef Full Text | Google Scholar

Zhang, W. G., and Gawarkiewicz, G. G. (2015). Dynamics of the direct intrusion of gulf stream ring water onto the Mid-Atlantic Bight shelf. Geophys. Res. Lett. 42, 7687–7695. doi: 10.1002/2015GL065530

CrossRef Full Text | Google Scholar

Zhang, W. G., and Partida, J. (2018). Frontal subduction of the Mid-Atlantic Bight shelf water at the onshore edge of a warm-core ring. J. Geophys. Res. 123, 7795–7818. doi: 10.1029/2018JC013794

CrossRef Full Text | Google Scholar

Zhang, Z., Zhao, W., Tian, J., Yang, Q., and Qu, T. (2015). Spatial structure and temporal variability of the zonal flow in the Luzon Strait. J. Geophys. Res. 120, 759–776. doi: 10.1002/2014JC010308

CrossRef Full Text | Google Scholar

Zhou, C., Zhao, W., Tian, J., Yang, Q., and Qu, T. (2014). Variability of the deep-water overflow in the luzon strait. J. Phys. Oceanogr. 44, 2972–2986. doi: 10.1175/JPO-D-14-0113.1

CrossRef Full Text | Google Scholar

Zilberman, N. V., Roemmich, D. H., Gille, S. T., and Gilson, J. (2018). Estimating the velocity and transport of western boundary current systems: a case study of the East Australian Current near Brisbane. J. Atmos. Ocean. Technol. 35, 1313–1329. doi: 10.1175/JTECH-D-17-0153.1

CrossRef Full Text | Google Scholar

Zilberman, N. V., Roemmich, D., and Gille, S. (2013). The mean and the time variability of the shallow meridional overturning circulation in the tropical South Pacific Ocean. J. Clim. 26, 4069–4087. doi: 10.1175/JCLI-D-12-00120.1

CrossRef Full Text | Google Scholar

Zilberman, N. V., Roemmich, D., and Gille, S. (2014). Meridional volume transport in the South Pacific: mean and SAM-related variability. J. Geophys. Res. Oceans. 119, 2658–2678. doi: 10.1002/2013JC009688

CrossRef Full Text | Google Scholar

Zuidema, P., Chang, P., Medeiros, B., Kirtman, B. P., Mechoso, R., Schneider, E. K., et al. (2016). Challenges and prospects for reducing coupled climate model SST biases in the eastern tropical Atlantic and Pacific Oceans: the U.S. CLIVAR eastern tropical oceans synthesis working group. Bull. Am. Meteor. Soc. 97, 2305–2328. doi: 10.1175/BAMS-D-15-00274.1

CrossRef Full Text | Google Scholar

Keywords: western boundary current systems, eastern boundary current systems, ocean observing systems, time series, autonomous underwater gliders, drifters, remote sensing, moorings

Citation: Todd RE, Chavez FP, Clayton S, Cravatte S, Goes M, Graco M, Lin X, Sprintall J, Zilberman NV, Archer M, Arístegui J, Balmaseda M, Bane JM, Baringer MO, Barth JA, Beal LM, Brandt P, Calil PHR, Campos E, Centurioni LR, Chidichimo MP, Cirano M, Cronin MF, Curchitser EN, Davis RE, Dengler M, deYoung B, Dong S, Escribano R, Fassbender AJ, Fawcett SE, Feng M, Goni GJ, Gray AR, Gutiérrez D, Hebert D, Hummels R, Ito S-i, Krug M, Lacan F, Laurindo L, Lazar A, Lee CM, Lengaigne M, Levine NM, Middleton J, Montes I, Muglia M, Nagai T, Palevsky HI, Palter JB, Phillips HE, Piola A, Plueddemann AJ, Qiu B, Rodrigues RR, Roughan M, Rudnick DL, Rykaczewski RR, Saraceno M, Seim H, Sen Gupta A, Shannon L, Sloyan BM, Sutton AJ, Thompson L, van der Plas AK, Volkov D, Wilkin J, Zhang D and Zhang L (2019) Global Perspectives on Observing Ocean Boundary Current Systems. Front. Mar. Sci. 6:423. doi: 10.3389/fmars.2019.00423

Received: 31 October 2018; Accepted: 05 July 2019;
Published: 08 August 2019.

Edited by:

Sabrina Speich, École Normale Supérieure, France

Reviewed by:

Moacyr Cunha de Araujo Filho, Federal Rural University of Pernambuco, Brazil
Eitarou Oka, The University of Tokyo, Japan

Copyright © 2019 Todd, Chavez, Clayton, Cravatte, Goes, Graco, Lin, Sprintall, Zilberman, Archer, Arístegui, Balmaseda, Bane, Baringer, Barth, Beal, Brandt, Calil, Campos, Centurioni, Chidichimo, Cirano, Cronin, Curchitser, Davis, Dengler, deYoung, Dong, Escribano, Fassbender, Fawcett, Feng, Goni, Gray, Gutiérrez, Hebert, Hummels, Ito, Krug, Lacan, Laurindo, Lazar, Lee, Lengaigne, Levine, Middleton, Montes, Muglia, Nagai, Palevsky, Palter, Phillips, Piola, Plueddemann, Qiu, Rodrigues, Roughan, Rudnick, Rykaczewski, Saraceno, Seim, Sen Gupta, Shannon, Sloyan, Sutton, Thompson, van der Plas, Volkov, Wilkin, Zhang and Zhang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Robert E. Todd, rtodd@whoi.edu

Download