1. Introduction
Along eastern boundary upwelling systems (EBUS) equatorward winds are responsible for a thermally indirect Ekman circulation that brings heavy, deep, and generally nutrient-rich water toward the surface. Nutrient enrichment and new primary production have traditionally been thought to linearly depend on the intensity of the alongshore wind stress component (Bakun et al. 2010). This is what hundreds of studies concerned with EBUS ecosystem functioning have implicitly assumed by relying on the so-called Bakun index (derived from alongshore coastal winds, sometimes complemented by wind stress curl information [i.e., the coastal upwelling transport index by Jacox et al. (2018)], as a proxy for planktonic food availability. However, the rate at which water is being transferred from the ocean interior into the mixed layer may not be simply related to the Ekman transport.
Quite recently, the contribution of alongshore pressure gradients have increasingly been accounted for. In many upwelling sectors they can yield onshore velocities that counteract, or more infrequently reinforce, the Ekman flow (Colas et al. 2008; Marchesiello et al. 2010; Ndoye et al. 2017; Jacox et al. 2018). In this study, we shed light on another potentially important effect: quasi-balanced mesoscale and submesoscale turbulence is responsible for rectified eddy transport which also tends to counteract the Ekman upwelling cell. Several studies have addressed this topic (e.g., Lathuilière et al. 2010; Gruber et al. 2011; Colas et al. 2013) but the knowledge of the eddies’ role in EBUS remains fragmented. In an attempt to remedy this the present study borrows more directly from and draw connections with the Southern Ocean (SO) literature, in which eddy effects have been more extensively investigated (Marshall and Radko 2003; Hallberg and Gnanadesikan 2006; Morrison et al. 2011). In the SO, transport in the direction transverse to the Antarctic Circumpolar Current (ACC) axis computed in isopycnal coordinates is a small residual between the mean Ekman-driven circulation due to wind pumping/suction (the so-called Deacon cell) and eddy-induced transfers of mass driven by baroclinic instability processes (Marshall and Radko 2006; Marshall and Speer 2012). The eddy-induced circulation attempts to flatten isopycnal surfaces, reduce available potential energy, and therefore largely opposes the thermally direct Deacon cell. As a result, it is widely accepted that the sole knowledge of the wind forcing does not provide useful insight into the mean cross-frontal circulation of the SO. There air–sea buoyancy fluxes are, somewhat counterintuitively, much more informative than the wind forcing with respect to the meridional tracer transport (Marshall 1997; Gent 2016).
Indeed, they constrain rates of fluid transfer across isopycnals in the mixed layer, hence also subduction and obduction rates at the base of the mixed layer in steady state [subduction (obduction) being related to a convergence (divergence) of diapycnal flux of fluid, Fig. 1]. Neglecting mixing in the ocean interior the time-averaged residual meridional transport is thus entirely determined by the knowledge of air–sea buoyancy fluxes (Walin 1982; Marshall 1997). In particular, with zero air–sea buoyancy fluxes the residual circulation in the surface layer (and in the interior) must vanish, which means that eddies exactly cancel the Eulerian wind-driven circulation (Fig. 1c).

(a)–(c) Schematic showing the effect of atmospheric buoyancy forcing
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(a)–(c) Schematic showing the effect of atmospheric buoyancy forcing
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
(a)–(c) Schematic showing the effect of atmospheric buoyancy forcing
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(a) Annual mean net buoyancy flux for 2009–13 period from TropFlux (Praveen Kumar et al. 2012). The four eastern boundary upwelling systems (EBUS): California, West Africa, Humboldt, and Benguela are marked with black, red, blue, and green squares. (b) Annual mean (squares) heat fluxes, standard deviation (crosses), and maximum and minimum (dots) for the four EBUS Systems for the 2009–13 period. (c) Daily net buoyancy fluxes for each EBUS system for 2012. The TropFlux data are produced under a collaboration between Laboratoire d’Océanographie: Expérimentation et Approches Numériques (LOCEAN) from Institut Pierre Simon Laplace (IPSL, Paris, France) and National Institute of Oceanography/CSIR (NIO, Goa, India), and supported by Institut de Recherche pour le Développement (IRD, France). TropFlux relies on data provided by the ECMWF interim reanalysis (ERA-I) and ISCCP projects. The vertical dashed black lines in (b) and (c) indicate the parameter space (0–80 W m−2) explored in this idealized study.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(a) Annual mean net buoyancy flux for 2009–13 period from TropFlux (Praveen Kumar et al. 2012). The four eastern boundary upwelling systems (EBUS): California, West Africa, Humboldt, and Benguela are marked with black, red, blue, and green squares. (b) Annual mean (squares) heat fluxes, standard deviation (crosses), and maximum and minimum (dots) for the four EBUS Systems for the 2009–13 period. (c) Daily net buoyancy fluxes for each EBUS system for 2012. The TropFlux data are produced under a collaboration between Laboratoire d’Océanographie: Expérimentation et Approches Numériques (LOCEAN) from Institut Pierre Simon Laplace (IPSL, Paris, France) and National Institute of Oceanography/CSIR (NIO, Goa, India), and supported by Institut de Recherche pour le Développement (IRD, France). TropFlux relies on data provided by the ECMWF interim reanalysis (ERA-I) and ISCCP projects. The vertical dashed black lines in (b) and (c) indicate the parameter space (0–80 W m−2) explored in this idealized study.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
(a) Annual mean net buoyancy flux for 2009–13 period from TropFlux (Praveen Kumar et al. 2012). The four eastern boundary upwelling systems (EBUS): California, West Africa, Humboldt, and Benguela are marked with black, red, blue, and green squares. (b) Annual mean (squares) heat fluxes, standard deviation (crosses), and maximum and minimum (dots) for the four EBUS Systems for the 2009–13 period. (c) Daily net buoyancy fluxes for each EBUS system for 2012. The TropFlux data are produced under a collaboration between Laboratoire d’Océanographie: Expérimentation et Approches Numériques (LOCEAN) from Institut Pierre Simon Laplace (IPSL, Paris, France) and National Institute of Oceanography/CSIR (NIO, Goa, India), and supported by Institut de Recherche pour le Développement (IRD, France). TropFlux relies on data provided by the ECMWF interim reanalysis (ERA-I) and ISCCP projects. The vertical dashed black lines in (b) and (c) indicate the parameter space (0–80 W m−2) explored in this idealized study.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
By analogy with the SO, we hypothesize and will demonstrate that (i) in a 3D upwelling system eddy fluxes are responsible for a large contribution to the time and/or alongshore average buoyancy equation [ignored in Eq. (1), see Eq. (A3)] that cancels a fraction of the buoyancy advection by the Eulerian Ekman circulation and (ii) the structure and intensity of the eddy and residual circulation is modulated by the buoyancy input
One practical difficulty is that upwelling systems lack the periodicity attribute, so the ACC framework and theory are not directly applicable. Net buoyancy input can arise from lateral fluxes at the northern and southern edges of the upwelling region, and this is an important source of complication. In addition, most upwelling systems have intense seasonal fluctuations, so time averaging over one or several years and widely distinct ocean states is needed to ensure that the neglect of the buoyancy tendency term in Eq. (1) is valid (see the appendix).
To make progress in the overall understanding of EBUS functioning, we thus consider an idealized upwelling configuration with periodic boundary conditions following Lathuilière et al. (2010). The setting and methods are described in section 2. The results are presented in section 3. How these results can be interpreted in the context of real ocean upwelling systems is extensively discussed in section 4. The paper finishes with our conclusions in section 5.
2. Model setup and methods
a. Idealized upwelling model configuration

Schematic of idealized eastern boundary upwelling model configuration. Model setup and schematic were inspired by Lathuilière et al. (2010). The model grid consists of a shallow shelf, a continental slope, and a flat bottom. The Coriolis parameter f is constant and corresponds to a latitude of 14.5°S. The surface ocean is forced with a constant wind stress of 0.075 N m−2 and varying heat fluxes (details in section 2b). A sponge layer exists offshore with enhanced diffusivity and viscosity coefficients. There full depth temperatures are restored to the initial temperature profile To(z) [see Eq. (3)]. An alongshore pressure gradient is prescribed in the upper 200 m which drives an onshore flow balancing the offshore Ekman transport.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

Schematic of idealized eastern boundary upwelling model configuration. Model setup and schematic were inspired by Lathuilière et al. (2010). The model grid consists of a shallow shelf, a continental slope, and a flat bottom. The Coriolis parameter f is constant and corresponds to a latitude of 14.5°S. The surface ocean is forced with a constant wind stress of 0.075 N m−2 and varying heat fluxes (details in section 2b). A sponge layer exists offshore with enhanced diffusivity and viscosity coefficients. There full depth temperatures are restored to the initial temperature profile To(z) [see Eq. (3)]. An alongshore pressure gradient is prescribed in the upper 200 m which drives an onshore flow balancing the offshore Ekman transport.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Schematic of idealized eastern boundary upwelling model configuration. Model setup and schematic were inspired by Lathuilière et al. (2010). The model grid consists of a shallow shelf, a continental slope, and a flat bottom. The Coriolis parameter f is constant and corresponds to a latitude of 14.5°S. The surface ocean is forced with a constant wind stress of 0.075 N m−2 and varying heat fluxes (details in section 2b). A sponge layer exists offshore with enhanced diffusivity and viscosity coefficients. There full depth temperatures are restored to the initial temperature profile To(z) [see Eq. (3)]. An alongshore pressure gradient is prescribed in the upper 200 m which drives an onshore flow balancing the offshore Ekman transport.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
We use periodic boundary conditions in the alongshore direction. At the coastline the domain is closed. For numerical simplicity a closed boundary is also present along the offshore side of the domain. To limit the spurious effects of this western wall a sponge layer of width equal to 96 km is included in which momentum and temperature are subjected to harmonic diffusion and temperature is restored toward a background temperature profile To(z) [see Eq. (3)]. In all simulations diffusion and viscosity coefficients increase smoothly via a cosine function from 0 m2 s−1 at y = 506 km to a maximum value of 600 m2 s−1 toward the closed boundary at y = 600 km (Fig. 3). For lateral momentum advection the default CROCO third-order upstream biased advection scheme with implicit diffusion is used. For temperature we use the Z version of the fifth-order Weighted Essentially Non-Oscillatory (WENO) scheme to prevent temperature under/overshoots near sharp horizontal fronts.1 Spurious diapycnal mixing associated with this scheme cannot be entirely avoided but shall have negligible effects in our study because 1) sigma coordinates are quasi-horizontal over most of the domain where the bottom is flat, 2) we are concerned with the upper 100–150 m of the water column in which (physical) diapycnal mixing rates are moderately high, and 3) we use kilometer-scale horizontal resolution [except for some sensitivity runs; see Marchesiello et al. (2009) and Lemarié et al. (2012) for relevant details on spurious diapycnal mixing]. Free slip boundary conditions are used at the onshore and offshore walls. Linear friction with a drag coefficient
The Coriolis parameter (f = −2π/Tf) is set constant (no β effect) and corresponds to a f-plane latitude of about 14.5°S resulting in an inertial period (Tf) of 2 days. Due to the absence of any β effect, there is no westward propagation of Rossby waves. This is not of major importance as we restrict the duration of our experiment to 120 days (including a spinup of 80 days) but is presumably responsible for eddy kinetic energy (EKE) levels above typical values found in the real ocean for EBUS sectors (Haney et al. 2001; Marchesiello et al. 2003; Gruber et al. 2011). The first baroclinic Rossby Radius is around 57 km within our idealized setting. This choice of parameters makes our results applicable to the Peruvian, West African, and northern Benguelan upwelling systems, but the main findings should also be valid at higher latitude.
To resolve sharp vertical velocity and temperature gradients in the upper ocean, we use 100 vertical sigma levels for all configurations. The vertical coordinate by Shchepetkin and McWilliams (2009) are used with the vertical grid parameters θs = 6, θb = 0, and hc = 10.2 Upwelling systems typically have water depth between 4000 and 6000 m offshore of the continental slope. Here we use a reduced bottom depth of 1960 m in combination with the high number of vertical levels results in a relatively high vertical resolution especially near the surface. Over the upper shelf the vertical resolution ranges from 0.1 m near the surface to 1.2 m above the bottom. Offshore the vertical resolution ranges from 0.7 m at the surface to 6 m at 100-m depth and 112 m near the bottom.

Domain averaged (a) wind stress and (b) heat flux forcing during the model spinup (day 0–80) and the three different heat flux experiments (day 80–120). The NF, MF, and HF experiments are shown in black/gray, dark/light blue, and dark/light red for the submesoscale (Δx = 800 m) and mesoscale (Δx = 8 km) runs, respectively. Single runs and ensemble mean are shown in thin and thick lines, respectively.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

Domain averaged (a) wind stress and (b) heat flux forcing during the model spinup (day 0–80) and the three different heat flux experiments (day 80–120). The NF, MF, and HF experiments are shown in black/gray, dark/light blue, and dark/light red for the submesoscale (Δx = 800 m) and mesoscale (Δx = 8 km) runs, respectively. Single runs and ensemble mean are shown in thin and thick lines, respectively.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Domain averaged (a) wind stress and (b) heat flux forcing during the model spinup (day 0–80) and the three different heat flux experiments (day 80–120). The NF, MF, and HF experiments are shown in black/gray, dark/light blue, and dark/light red for the submesoscale (Δx = 800 m) and mesoscale (Δx = 8 km) runs, respectively. Single runs and ensemble mean are shown in thin and thick lines, respectively.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
b. Heat flux sensitivity experiments and ensemble runs

Temporal-mean alongshore-averaged (a) heat flux, (b) zeta, (c) alongshore, and (d) cross-shore circulation in reference simulation (MF) averaged over the time of the experiments (day 81–120). Isotherms are contoured in white in (c) and (d). Eulerian mean streamfunction is shown in black contours in (d).
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

Temporal-mean alongshore-averaged (a) heat flux, (b) zeta, (c) alongshore, and (d) cross-shore circulation in reference simulation (MF) averaged over the time of the experiments (day 81–120). Isotherms are contoured in white in (c) and (d). Eulerian mean streamfunction is shown in black contours in (d).
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Temporal-mean alongshore-averaged (a) heat flux, (b) zeta, (c) alongshore, and (d) cross-shore circulation in reference simulation (MF) averaged over the time of the experiments (day 81–120). Isotherms are contoured in white in (c) and (d). Eulerian mean streamfunction is shown in black contours in (d).
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Model run acronyms, horizontal resolution, number of grid cells, time steps, duration, ensemble number, heat flux parameter dQ/dT, and wind stress.


c. Quantification of eddy effects and overturning streamfunction calculation
Baroclinic instability (hereafter BCI) processes are known to be essential in upwelling systems as a source of mesoscale and submesoscale turbulence (Marchesiello et al. 2003). The strength of BCI will be classically evaluated by computing available potential energy (APE) release as the averaged covariance of the vertical velocity w and buoyancy b perturbations:
Our focus is on the description of the eddy role on the transport of buoyancy. In practice, we will use three different approaches to determine variants of residual overturning (transverse) streamfunctions: 1) the transformed Eulerian mean formulation of Held and Schneider (1999); isopycnal averaging of water volume flux at 2) constant across-shore distance (equivalent to averaging at constant latitude in the SO, Döös and Webb 1994) or at 3) constant depth (Nurser and Lee 2004a,b). Each of these three approaches has its own advantages and limitations as detailed below. Thus, we apply all three of them to ensure that our main findings are robust, as previously done for the SO or in the atmosphere.
The ocean sector where upwelling takes place is in close proximity to the eastern boundary and not situated very deep below the surface. Indeed we are typically interested in the depth range from below the mixed layer down to 150–200-m depth or less, i.e., the source region of the upwelled water. The interpretation of the transport streamfunctions
3. Results
a. General hydrography and circulation in the idealized upwelling system
The general thermal structures and circulation features of the idealized upwelling configuration resemble those typical of real upwelling systems (Figs. 5 and 6). The wind forcing results in an offshore Ekman transport in the upper 30–40 m of the water column (Fig. 5d). The Ekman cell (ΨMean) is closed by a return flow reaching down to 200-m depth and has a maximum strength of 2.3 m2 s−1 around 50-m depth (Fig. 5d). The Eulerian circulation is broadly consistent with isopycnal doming in the upper 150 m of the water column especially within 80 km from shore (Figs. 5c,d). Mean surface temperatures as low as 14°C are simulated at the shelf break (Figs. 5c,d). A frontal zone separates the upwelling waters from the open ocean where surface temperatures increase to 20°C at 200 km offshore (Figs. 5c,d). An alongshore surface jet of about 0.4 m s−1 in same direction as wind forcing develops over the shelf and upper slope (Fig. 5c). A subsurface undercurrent in opposite direction hugs the continental slope with maximal velocities of 0.15 m s−1 in 150–200-m depth (Fig. 5c).

(a),(b) Sea surface temperature (SST; °C) and (c),(d) vertical velocities (m day−1) in 20-m depth at day 110 for medium heat flux forcing for (left) submesoscale (Δx = 800 m) and (right) mesoscale (Δx = 8 km) horizontal resolution, respectively. Sea level anomaly (zeta) in cm is shown in black contours.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(a),(b) Sea surface temperature (SST; °C) and (c),(d) vertical velocities (m day−1) in 20-m depth at day 110 for medium heat flux forcing for (left) submesoscale (Δx = 800 m) and (right) mesoscale (Δx = 8 km) horizontal resolution, respectively. Sea level anomaly (zeta) in cm is shown in black contours.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
(a),(b) Sea surface temperature (SST; °C) and (c),(d) vertical velocities (m day−1) in 20-m depth at day 110 for medium heat flux forcing for (left) submesoscale (Δx = 800 m) and (right) mesoscale (Δx = 8 km) horizontal resolution, respectively. Sea level anomaly (zeta) in cm is shown in black contours.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
The idealized upwelling solution also produces turbulent structures as found in upwelling systems: mesoscale eddies, filaments and sharp submesoscale temperature fronts particularly at the finest resolution (Fig. 6a). Typically an anticyclonic mesoscale warm core eddy forms at some point near the shelf break with diameter of roughly 150 km and a sea level anomaly of about 10 cm (Figs. 6a,b). Several smaller-scale cyclonic vortices are formed around the anticyclone with cold filaments at the edges of the mesoscale structures (Figs. 6a,b). Within the filaments the SST can drop down to about 18°C. Submesoscale frontal dynamics are obviously richer at the higher resolution. Downward velocities of up to 100 m day−1 are found at the cold side of these fronts (e.g., at x = 150–200 km and y = 150 km in Fig. 6c). These vertical velocities are a crucial part of the eddy flow that counteracts the Ekman transport of buoyancy as we investigate in detail in the following chapter through three different approaches.
b. Eddy effects under varying heat flux forcing
The effect of varying heat flux forcing on the cross-shore circulation is investigated in this section. Based on the SO literature our hypothesis is that changing heat fluxes affects the eddy field with implications on the eddy heat fluxes, hence also on the eddy-induced and residual transverse circulations. To test this, we compute the three different variants of eddy and residual streamfunctions defined in section 2c. This is done for the submesoscale permitting simulations which represent the turbulent processes more accurately. A detailed comparison between the simulations with 8-km and 800-m horizontal resolution is carried out in the next section 3c.
We first quantify the APE release by BCI to start looking into the eddy effect sensitivity to heat fluxes. The mean cross-shore distribution of

(a),(d) Ensemble mean vertical eddy buoyancy fluxes and (b),(e) eddy streamfunctions based on transformed Eulerian mean theory (
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(a),(d) Ensemble mean vertical eddy buoyancy fluxes and (b),(e) eddy streamfunctions based on transformed Eulerian mean theory (
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
(a),(d) Ensemble mean vertical eddy buoyancy fluxes and (b),(e) eddy streamfunctions based on transformed Eulerian mean theory (
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
We now turn to the effects of turbulence on buoyancy advection and start with the most dramatic NF case. The

(top) Eulerian
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

(top) Eulerian
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
(top) Eulerian
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

As in Fig. 8, but for
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

As in Fig. 8, but for
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
As in Fig. 8, but for
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Mean–eddy cancellation is manifest in Figs. 8i and 9i. In the upper 50 m at 200 km and even down to 100 m closer to the shelf break the intensity of the residual circulation is smaller than that of the mean Eulerian circulation by a factor of 4 or more. Most importantly, advective feeding of the surface layer with upwelling water vanishes. At greater depth (125 m) residual streamfunction values above 1 m2 s−1 are found away from the coast (Fig. 8i) where the tilt of the isopycnals and BCI strength (Fig. 7a) is reduced. Note that the residual circulation below the surface layer is not particularly better aligned with the isopycnals than the mean Eulerian circulation, contrary to what is typically found in the SO (Karsten and Marshall 2002). There are two reasons for this. First, our simulations are not in statistical equilibrium so that transient adjustments might be responsible for apparent “mean” diapycnal flow over the period of analysis. The fact that our simulations for the different air–sea flux cases exhibit similar residual circulations indicates that this reason is secondary. Second, and more importantly, our entire region of interest is in close proximity to the mean ML base, and it is subjected to intense intermittent mixing. To substantiate this statement, we show the maximum mixed layer depths reached for every simulation in Fig. 7. More specifically, mixed layer depths between 60 and 100 m are reached in 5% of the time. Although our simulations do not resolve all relevant processes, a transition layer with relatively elevated mixing levels is found below the mixed layer (Large et al. 1994; Johnston and Rudnick 2009). This provides the required diabatic forcing to accommodate a slow but nonzero residual circulation consistent with upwelling to about 80–120-m depth which bifurcates offshore above this depth range (Figs. 8i and 9i). This circulation pattern is very similar for
How these results are affected by the air–sea buoyancy forcings is described in the remainder of this section. In contrast to ΨMean, the sensitivity of the eddy-induced and residual circulation to air–sea heat forcing is noticeable well below the mixed layer base (Figs. 9h,l). For instance, the lower part of the eddy cell weakens by over 50% (from −1.3 m2 s−1 in NF to −0.55 m2 s−1 in HF) at 70 km offshore and 75-m depth. Although a large degree of eddy-mean cancellation is still present at HF its residual circulation has recovered a structure that more closely resembles the mean Ekman circulation, with some streamlines unambiguously connecting the offshore-subsurface to the nearshore-surface sectors. Again no major difference is found between the sensitivities exhibited by
Subtle differences between simulations with different heat fluxes are more readily apparent in Fig. 10 where streamfunction values are shown at a particular location chosen because it lies on the (Eulerian) mean upwelling pathway (see “x” symbol in Figs. 8 and 9). Figure 10 also gives a sense of the dispersion among realizations from the ensemble runs. The behaviors of

Strength of (a) Eulerian, (b) eddy, and (c) residual streamfunction under varying heat flux forcing in 70-m depth at 64 km offshore (black cross in Figs. 7b, 8, and 9). Streamfunction estimates based on TEM theory (red, orange; formulation 1), isopycnal integration of horizontal (black, dark blue; formulation 2), and vertical (gray, light blue; formulation 3) velocities are shown for mesoscale and submesoscale horizontal resolution, respectively. Details on the calculation are in section 2c. Values found for each run of the ensembles are represented with a cross. Solid lines connect the mean ensemble values found for the three different air–sea heat flux choices.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

Strength of (a) Eulerian, (b) eddy, and (c) residual streamfunction under varying heat flux forcing in 70-m depth at 64 km offshore (black cross in Figs. 7b, 8, and 9). Streamfunction estimates based on TEM theory (red, orange; formulation 1), isopycnal integration of horizontal (black, dark blue; formulation 2), and vertical (gray, light blue; formulation 3) velocities are shown for mesoscale and submesoscale horizontal resolution, respectively. Details on the calculation are in section 2c. Values found for each run of the ensembles are represented with a cross. Solid lines connect the mean ensemble values found for the three different air–sea heat flux choices.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Strength of (a) Eulerian, (b) eddy, and (c) residual streamfunction under varying heat flux forcing in 70-m depth at 64 km offshore (black cross in Figs. 7b, 8, and 9). Streamfunction estimates based on TEM theory (red, orange; formulation 1), isopycnal integration of horizontal (black, dark blue; formulation 2), and vertical (gray, light blue; formulation 3) velocities are shown for mesoscale and submesoscale horizontal resolution, respectively. Details on the calculation are in section 2c. Values found for each run of the ensembles are represented with a cross. Solid lines connect the mean ensemble values found for the three different air–sea heat flux choices.
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
c. Resolution sensitivity
Submesoscale frontal processes are well known to be instrumental in the dynamics of the mixed layer. Our focus is on the dynamics in a layer of upper-ocean fluid situated below the mixed layer, where the role of the submesoscale has not been clearly established and may vary depending on the regime under consideration (Capet et al. 2016). The influence of resolution on our findings is thus investigated, by comparing analyses for simulations at Δx = 800 m with analogs at Δx = 8 km.
Starting with
The mean cross-shore buoyancy gradient
In short, the eddy cancellation mechanism and its sensitivity to air–sea buoyancy forcing highlighted in the previous section for submesoscale-permitting simulations results mainly from mesoscale turbulence effects. Although this may seem surprising given the shallow nature of the problem under consideration, this is consistent with the limited vertical extension of submesoscale turbulence, which is typically confined into the mixed layer in most oceanic regimes [see Capet et al. (2016) for a counterexample where Charney instability is present].
4. Discussion
The overarching objective of this research is to contribute to the understanding of the factors influencing upper-ocean enrichment in nutrients, primary production, and the fate of organic matter in EBUS, including considerations on their temporal variability and spatial heterogeneity (e.g., differences between upwelling systems or upwelling sectors within a given EBUS). The results suggest that (i) the eddy-induced circulation can counteract (or cancel in the SO terminology) mean advection by the Ekman circulation to a large degree, and (ii) the level of cancellation can be subjected to spatiotemporal modulations due to air–sea buoyancy flux variability, which can lead to changes in the degree to which eddy fluxes counteract the mean advection by the Ekman flow that is central in coastal upwelling dynamics. This being said, the cancellation mechanism has been investigated in a simplified numerical configuration which has three important limitations in terms of realism.
First and foremost, the simulations we analyze have EKE levels that significantly exceed those found in the real ocean, by a factor of 2–10 depending on which EBUS sector is considered. Elevated EKE is an inherent consequence of our periodic channel simplification with which the β effect is not compatible. The absence of beta precludes the westward radiation of energy, for instance through Rossby wave propagation or beta drift of vortices (Carton 2010). To make progress despite this important caveat we assume that the intensity of the eddy overturning streamfunctions scales linearly with EKE. Marshall et al. (2012) and Mak et al. (2017) provide some support to do so although obviously not in the specific context where the aim is to account for a missing β drift. Figure 11 is produced based on this assumption, on the EKE degree of cancellation pairs obtained for each of our simulations, and on the EBUS EKE values reported by Gruber et al. (2011). See caption of Fig. 11 for more details. It provides estimates for the amount of cancellation in the four EBUS, going from 3% to 6% in the northern Benguela, to 10%–30% in the California Current System. Upper range values correspond to situations with ≈0 net air–sea buoyancy fluxes, which are not very common in this latter system when upwelling conditions prevail. These results are consistent with the well-accepted view that the role of eddies in EBUS is not of leading order (as it is in the SO). But they point to the possible importance of eddy-induced circulations during periods when air–sea buoyancy forcings are weak and perhaps more so when they are temporarily negative (see Fig. 2). Note that this situation was not investigated. Eddy fluxes may also have an appreciable time-averaged effect on tracer advection in sectors of the California Current System where standing meanders yield intensified mesoscale and submesoscale activity (Centurioni et al. 2008; Colas et al. 2013). An important caveat regarding this rescaling approach concerns the possibly subtle relationship between the cross-shore EKE distribution and the cancellation of the Ekman cell by eddies. Take for instance the southern Benguela where offshore Agulhas rings lead EKE levels in excess of 400 m2 s−2 just offshore of the continental shelf (Capet et al. 2008). Mesoscale structures produced remotely by the Agulhas retroflection or by the baroclinically unstable offshore flowing California Current contribute to EKE in the vicinity of the coastal upwelling sector. However, these structures may not be an important local source of APE to EKE conversion, which is more naturally related to cancellation (see the formulation of ΨTEM above). Most generally, the local relationships between surface EKE, upper ocean

Simulated eddy kinetic energy (EKE; cm2 s−2) vs eddy cancellation (%) from submesoscale simulations. The EKE values colored dots are derived from geostrophic velocity estimates of spatially filtered (rectangular centered 24-km running mean both in x and y direction) sea surface height anomalies of the dx = 800 m simulations (orange = HF, blue = MF, and gray = NF case). The diagonal lines represent the best fit through the 12 ensembles of the different heat flux cases (same color code). The eddy cancellation (%) for each individual ensemble run is estimated via
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1

Simulated eddy kinetic energy (EKE; cm2 s−2) vs eddy cancellation (%) from submesoscale simulations. The EKE values colored dots are derived from geostrophic velocity estimates of spatially filtered (rectangular centered 24-km running mean both in x and y direction) sea surface height anomalies of the dx = 800 m simulations (orange = HF, blue = MF, and gray = NF case). The diagonal lines represent the best fit through the 12 ensembles of the different heat flux cases (same color code). The eddy cancellation (%) for each individual ensemble run is estimated via
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Simulated eddy kinetic energy (EKE; cm2 s−2) vs eddy cancellation (%) from submesoscale simulations. The EKE values colored dots are derived from geostrophic velocity estimates of spatially filtered (rectangular centered 24-km running mean both in x and y direction) sea surface height anomalies of the dx = 800 m simulations (orange = HF, blue = MF, and gray = NF case). The diagonal lines represent the best fit through the 12 ensembles of the different heat flux cases (same color code). The eddy cancellation (%) for each individual ensemble run is estimated via
Citation: Journal of Physical Oceanography 51, 11; 10.1175/JPO-D-20-0294.1
Second, our study was based on numerical simulations that were analyzed only over specific time periods of 40 days. Our findings on the cancellation process itself do not specifically depend on this particular time scale but the modulation of the cancellation in response to air–sea buoyancy fluctuations may. To explore this potential issue, we have computed the temporal evolution of the APE release rate
Third, for simplicity, wind variability has been ignored throughout the study whereas it is an important aspect of upwelling dynamics. In the real ocean wind and air–sea buoyancy flux variability may be correlated and combine to produce results distinct from those found in this study. Such correlations vary from place to place (Send et al. 1987; Beardsley et al. 1998; Flynn et al. 2017; Lübbecke et al. 2019). For instance low winds can yield large (resp. weak) air–sea net heating into the surface ocean because cooling through latent heat release is reduced (resp. because in some regions like central California upwelling relaxations are associated with increased nebulosity and reduced incoming solar radiation). By analogy with the SO functioning and in agreement with the general understanding of baroclinic processes the intensity of the eddy cancellation mechanisms shall roughly scale with that of the Ekman flow, i.e., we expect compensation (Marshall and Radko 2003) to occur whereby increasing upwelling winds steepens the isopycnals which increases the counteracting effect of eddies. In Fig. 11 the intensity of the eddy-induced circulation is expressed as a fraction of the Ekman cell transport, for the particular wind strength that we chose. How that fraction actually varies depending on the wind conditions and wind spatial structure (Capet et al. 2004; Small et al. 2015; Bonino et al. 2019) would need to be investigated. Furthermore, the use of a pure flux versus restoring condition for the surface buoyancy boundary condition, as well as the time scale in the restoring case (Zhai and Munday 2014) may have an influence on the sensitivity of the residual overturning to wind stress changes (Abernathey et al. 2011). This needs to be investigated in future EBUS studies.
Finally, note that alongshore pressure gradient variability is another factor that can modulate upwelling intensity on synoptic, seasonal, and longer time scales (Werner and Hickey 1983; Huyer et al. 1987; Colas et al. 2008; Marchesiello et al. 2010; McCabe et al. 2015; Jacox et al. 2018), frequently in the sense of an upwelling reduction. In the northern Benguela and southern Canary current sectors where the eddy cancellation is particularly weak this effect and its temporal variability may thus be difficult to discern. In the northern Humboldt the eddy cancellation is presumably stronger but still modest in magnitude (~10%–20%, Fig. 11). Diagnostics of the eddy-induced circulation in realistic simulations for this upwelling sector tend to confirm our estimates. Maximum eddy-induced streamfunction values for summer reported in Colas et al. (2013, see their Fig. 8) reach about 0.2 m2 s−1, i.e., about 15% of the Ekman transport for that season. The CCS is the system where eddy cancellation is expected to be strongest and possibly cancel a large fraction of the Ekman circulation in some circumstances (up to 30%–40%, Fig. 11). It is also the system where the manifestation of eddy fluxes has received the most observational attention (Shearman et al. 1999; Pallàs-Sanz et al. 2010b), including on biogeochemical tracer dynamics (Bograd and Mantyla 2005; Huyer et al. 2005; Pallàs-Sanz et al. 2010a). Two estimations of the eddy cancellation strength in the California current system can be drawn from Nagai et al. (2015) and Colas et al. (2013). Note that the two studies use very similar numerical configurations. Counting streamfunction contours in Figs. 5a and 5b of Nagai et al. (2015) gives an annual mean eddy cancellation intensity ~2/8 = 25% of the Ekman flow at 50-m depth and 100 km from shore for central California. Slightly weaker but comparable values of 10%–15% are obtained in Colas et al. (2013) for summer, when upwelling winds and air–sea heat fluxes are most positive. Both estimates are within the range of values inferred from the present study.
5. Summary and conclusions
An idealized numerical model is used to study the effect of eddies on the (alongshore) mean transport of buoyancy in a coastal upwelling. The eddy contribution to buoyancy advection tends to counteract the advection by the Ekman transport, so as to limit the slope of the isopycnals. The efficiency of this eddy cancellation process varies with the strength of the air–sea buoyancy flux forcing: eddies are most effective at impeding the transport of buoyancy by Ekman currents in situations where air–sea buoyancy fluxes provide no or limited warming of upwelled surface waters drifting offshore, which allows more intense and deeper-reaching frontal conditions to be produced. Limitations imposed by our idealized framework do not allow us to work in steady state and limit the duration of our experiments to periods of tens of days. However, drifts in the thermohaline structure of our simulations remain small and our results can be interpreted using a steady-state Walin type reasoning (Marshall 1997): given the mean frontal thermohaline structure of an upwelling system vanishing (or negative) air–sea buoyancy fluxes would imply that the surface Ekman flow produces diapycnal transport of mass unless it is counteracted by eddy transport, resulting in partial or total cancellation of the mean buoyancy advection by the Ekman flow. Two limit cases and an intermediate situation are represented in Fig. 1. Our simulations fall in between the intermediate case and the total cancellation case. Concerned by the methodological limitations inherent to eddy flux estimations and descriptions we used three different standard methods and found good agreement between them in our idealized setting. Identifying and quantifying eddy cancellation in realistic model simulations is far more difficult, mainly because the alongshore periodicity of our numerical configuration offers a much simpler framework for analyses. It also limits the time-averaging and/or ensemble run size requirements in a context where stochastic variability is important and can blur the role of eddies and its sensitivity to forcings. However, the simplifications we take advantage of have important implications in terms of model realism. As thoroughly discussed in the previous section, the real ocean behavior is expected to differ, with much smaller eddy effects in terms of Ekman flow cancellation than the ones we reported. Despite this important caveat, we think that the idealized posing on which the present work is based is useful to develop intuition on the role of eddies in upwelling systems. In the same spirit, a follow-up study attempting to gain insight into biogeochemical tracer dynamics is in progress. More realistic EBUS studies on eddy cancellation would be useful to further clarify the eddy role on tracer transport and distribution in EBUS, as pursued in the context of subtropical gyres by Doddridge et al. (2016) and Doddridge and Marshall (2018).
Acknowledgments
ST received funding be the European Commission (Horizon 2020, MSCA-IF-2016, WACO 749699: Fine-scale Physics, Biogeochemistry and Climate Change in the West African Coastal Ocean). ST further acknowledges support by the DFG project SFB 754 and the Excellence Cluster: Future Ocean Kiel IMAP PostDoc network, respectively, for two research visits at LOCEAN in Paris in 2016 and 2017. Model simulations were performed on the CINES Occigen HPC under DARI projects Dynamique et Couplage de l’Océan de surface A0050101140 and A0060101140. We thank F. Colas for providing detailed information on streamfunction estimates and fruitful discussions which helped to improve the manuscript.
Data availability statement
The overall size of the model output of all ensemble runs used for this publication is 2.7 TB. This large size makes it impossible for us to provide a constant online data access. Instead the simulations are stored safely by the first author and can be made available on request.
APPENDIX
Eddy–Mean Flow Decomposition, Bolus Velocities, and Buoyancy Advection
REFERENCES
Abernathey, R., J. Marshall, and D. Ferreira, 2011: The dependence of Southern Ocean meridional overturning on wind stress. J. Phys. Oceanogr., 41, 2261–2278, https://doi.org/10.1175/JPO-D-11-023.1.
Bakun, A., D. B. Field, A. Redondo-Rodriguez, and S. J. Weeks, 2010: Greenhouse gas, upwelling-favorable winds, and the future of coastal ocean upwelling ecosystems. Global Change Biol., 16, 1213–1228, https://doi.org/10.1111/j.1365-2486.2009.02094.x.
Barnier, B., L. Siefridt, and P. Marchesiello, 1995: Thermal forcing for a global ocean circulation model using a three-year climatology of ECMWF analyses. J. Mar. Syst., 6, 363–380, https://doi.org/10.1016/0924-7963(94)00034-9.
Beardsley, R. C., E. P. Dever, S. J. Lentz, and J. P. Dean, 1998: Surface heat flux variability over the northern California shelf. J. Geophys. Res., 103, 21 553–21 586, https://doi.org/10.1029/98JC01458.
Bograd, S. J., and A. W. Mantyla, 2005: On the subduction of upwelled waters in the California Current. J. Mar. Res., 63, 863–885, https://doi.org/10.1357/002224005774464229.
Bonino, G., E. Di Lorenzo, S. Masina, and D. Iovino, 2019: Interannual to decadal variability within and across the major eastern boundary upwelling systems. Sci. Rep., 9, 19949, https://doi.org/10.1038/s41598-019-56514-8.
Capet, X., P. Marchesiello, and J. C. McWilliams, 2004: Upwelling response to coastal wind profiles. Geophys. Res. Lett., 31, L13311, https://doi.org/10.1029/2004GL020123.
Capet, X., F. Colas, P. Penven, P. Marchesiello, and J. C. McWilliams, 2008: Eddies in eastern-boundary subtropical upwelling systems. Ocean Modeling in an Eddying Regime, Geophys. Monogr., Vol. 177, Amer. Geophys. Union, 131–147, https://doi.org/10.1029/177GM10.
Capet, X., G. Roullet, P. Klein, and G. Maze, 2016: Intensification of upper-ocean submesoscale turbulence through Charney baroclinic instability. J. Phys. Oceanogr., 46, 3365–3384, https://doi.org/10.1175/JPO-D-16-0050.1.
Carton, X., 2010: Oceanic vortices. Fronts, Waves and Vortices in Geophysical Flows, J.-B. Flor, Ed., Springer, 61–108, https://doi.org/10.1007/978-3-642-11587-5_3.
Centurioni, L. R., J. C. Ohlmann, and P. P. Niiler, 2008: Permanent meanders in the California Current System. J. Phys. Oceanogr., 38, 1690–1710, https://doi.org/10.1175/2008JPO3746.1.
Chelton, D. B., P. Gaube, M. G. Schlax, J. J. Early, and R. M. Samelson, 2011: The influence of nonlinear mesoscale eddies on near-surface oceanic chlorophyll. Science, 334, 328–332, https://doi.org/10.1126/science.1208897.
Colas, F., X. Capet, J. McWilliams, and A. Shchepetkin, 2008: 1997–1998 El Niño off Peru: A numerical study. Prog. Oceanogr., 79, 138–155, https://doi.org/10.1016/j.pocean.2008.10.015.
Colas, F., J. C. McWilliams, X. Capet, and J. Kurian, 2012: Heat balance and eddies in the Peru-Chile current system. Climate Dyn., 39, 509–529, https://doi.org/10.1007/s00382-011-1170-6.
Colas, F., X. Capet, J. C. McWilliams, and Z. Li, 2013: Mesoscale eddy buoyancy flux and eddy-induced circulation in eastern boundary currents. J. Phys. Oceanogr., 43, 1073–1095, https://doi.org/10.1175/JPO-D-11-0241.1.
Condie, S., and J. R. Dunn, 2006: Seasonal characteristics of the surface mixed layer in the Australasian region: Implications for primary production regimes and biogeography. Mar. Freshwater Res., 57, 569–590, https://doi.org/10.1071/MF06009.
Doddridge, E. W., and D. P. Marshall, 2018: Implications of eddy cancellation for nutrient distribution within subtropical gyres. J. Geophys. Res. Oceans, 123, 6720–6735, https://doi.org/10.1029/2018JC013842.
Doddridge, E. W., D. P. Marshall, and A. M. Hogg, 2016: Eddy cancellation of the Ekman cell in subtropical gyres. J. Phys. Oceanogr., 46, 2995–3010, https://doi.org/10.1175/JPO-D-16-0097.1.
Döös, K., and D. J. Webb, 1994: The deacon cell and the other meridional cells of the southern ocean. J. Phys. Oceanogr., 24, 429–442, https://doi.org/10.1175/1520-0485(1994)024<0429:TDCATO>2.0.CO;2.
Flynn, K. R., M. R. Fewings, C. Gotschalk, and K. Lombardo, 2017: Large-scale anomalies in sea-surface temperature and air-sea fluxes during wind relaxation events off the United States west coast in summer. J. Geophys. Res. Oceans, 122, 2574–2594, https://doi.org/10.1002/2016JC012613.
Fox-Kemper, B., R. Ferrari, and R. Hallberg, 2008: Parameterization of mixed layer eddies. Part I: Theory and diagnosis. J. Phys. Oceanogr., 38, 1145–1165, https://doi.org/10.1175/2007JPO3792.1.
Garabato, A. C. N., D. P. Stevens, A. J. Watson, and W. Roether, 2007: Short-circuiting of the overturning circulation in the Antarctic circumpolar current. Nature, 447, 194–197, https://doi.org/10.1038/nature05832.
Gent, P. R., 2016: Effects of southern hemisphere wind changes on the meridional overturning circulation in ocean models. Annu. Rev. Mar. Sci., 8, 79–94, https://doi.org/10.1146/annurev-marine-122414-033929.
Gruber, N., Z. Lachkar, H. Frenzel, P. Marchesiello, M. Münnich, J. C. McWilliams, T. Nagai, and G.-K. Plattner, 2011: Eddy-induced reduction of biological production in eastern boundary upwelling systems. Nat. Geosci., 4, 787–792, https://doi.org/10.1038/ngeo1273.
Hallberg, R., and A. Gnanadesikan, 2006: The role of eddies in determining the structure and response of the wind-driven southern hemisphere overturning: Results from the modeling eddies in the southern ocean (MESO) project. J. Phys. Oceanogr., 36, 2232–2252, https://doi.org/10.1175/JPO2980.1.
Haney, R. L., R. A. Hale, and D. E. Dietrich, 2001: Offshore propagation of eddy kinetic energy in the California current. J. Geophys. Res., 106, 11 709–11 717, https://doi.org/10.1029/2000JC000433.
Held, I. M., and T. Schneider, 1999: The surface branch of the zonally averaged mass transport circulation in the troposphere. J. Atmos. Sci., 56, 1688–1697, https://doi.org/10.1175/1520-0469(1999)056<1688:TSBOTZ>2.0.CO;2.
Huyer, A., R. L. Smith, and T. Paluszkiewicz, 1987: Coastal upwelling off Peru during normal and El Nino times, 1981–1984. J. Geophys. Res., 92, 14 297–14 307, https://doi.org/10.1029/JC092iC13p14297.
Huyer, A., J. H. Fleischbein, J. Keister, P. M. Kosro, N. Perlin, R. L. Smith, and P. A. Wheeler, 2005: Two coastal upwelling domains in the northern California current system. J. Mar. Res., 63, 901–929, https://doi.org/10.1357/002224005774464238.
Jacox, M. G., C. A. Edwards, E. L. Hazen, and S. J. Bograd, 2018: Coastal upwelling revisited: Ekman, Bakun, and improved upwelling indices for the U.S. West Coast. J. Geophys. Res. Oceans, 123, 7332–7350, https://doi.org/10.1029/2018JC014187.
Johnston, T. S., and D. L. Rudnick, 2009: Observations of the transition layer. J. Phys. Oceanogr., 39, 780–797, https://doi.org/10.1175/2008JPO3824.1.
Karsten, R. H., and J. Marshall, 2002: Constructing the residual circulation of the ACC from observations. J. Phys. Oceanogr., 32, 3315–3327, https://doi.org/10.1175/1520-0485(2002)032<3315:CTRCOT>2.0.CO;2.
Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Lathuilière, C., V. Echevin, M. Lévy, and G. Madec, 2010: On the role of the mesoscale circulation on an idealized coastal upwelling ecosystem. J. Geophys. Res., 115, C09018, https://doi.org/10.1029/2009JC005827.
Lee, M.-M., and A. J. G. Nurser, 2012: Eddy subduction and the vertical transport streamfunction. J. Phys. Oceanogr., 42, 1762–1780, https://doi.org/10.1175/JPO-D-11-0219.1.
Lemarié, F., J. Kurian, A. F. Shchepetkin, M. Jeroen Molemaker, F. Colas, and J. C. McWilliams, 2012: Are there inescapable issues prohibiting the use of terrain-following coordinates in climate models? Ocean Modell., 42, 57–79, https://doi.org/10.1016/j.ocemod.2011.11.007.
Lentz, S. J., and D. C. Chapman, 2004: The importance of nonlinear cross-shelf momentum flux during wind-driven coastal upwelling. J. Phys. Oceanogr., 34, 2444–2457, https://doi.org/10.1175/JPO2644.1.
Lübbecke, J. F., P. Brandt, M. Dengler, R. Kopte, J. Lüdke, I. Richter, M. S. Martins, and P. C. Tchipalanga, 2019: Causes and evolution of the southeastern tropical Atlantic warm event in early 2016. Climate Dyn., 53, 261–274, https://doi.org/10.1007/s00382-018-4582-8.
Mak, J., D. P. Marshall, J. R. Maddison, and S. D. Bachman, 2017: Emergent eddy saturation from an energy constrained eddy parameterisation. Ocean Modell., 112, 125–138, https://doi.org/10.1016/j.ocemod.2017.02.007.
Marchesiello, P., J. C. McWilliams, and A. Shchepetkin, 2003: Equilibrium structure and dynamics of the California Current System. J. Phys. Oceanogr., 33, 753–783, https://doi.org/10.1175/1520-0485(2003)33<753:ESADOT>2.0.CO;2.
Marchesiello, P., L. Debreu, and X. Couvelard, 2009: Spurious diapycnal mixing in terrain-following coordinate models: The problem and a solution. Ocean Modell., 26, 156–169, https://doi.org/10.1016/j.ocemod.2008.09.004.
Marchesiello, P., J. Lefèvre, A. Vega, X. Couvelard, and C. Menkes, 2010: Coastal upwelling, circulation and heat balance around New Caledonia’s barrier reef. Mar. Pollut. Bull., 61, 432–448, https://doi.org/10.1016/j.marpolbul.2010.06.043.
Marshall, D. P., 1997: Subduction of water masses in an eddying ocean. J. Mar. Res., 55, 201–222, https://doi.org/10.1357/0022240973224373.
Marshall, D. P., J. R. Maddison, and P. S. Berloff, 2012: A framework for parameterizing eddy potential vorticity fluxes. J. Phys. Oceanogr., 42, 539–557, https://doi.org/10.1175/JPO-D-11-048.1.
Marshall, J., and T. Radko, 2003: Residual-mean solutions for the Antarctic circumpolar current and its associated overturning circulation. J. Phys. Oceanogr., 33, 2341–2354, https://doi.org/10.1175/1520-0485(2003)033<2341:RSFTAC>2.0.CO;2.
Marshall, J., and T. Radko, 2006: A model of the upper branch of the meridional overturning of the southern ocean. Prog. Oceanogr., 70, 331–345, https://doi.org/10.1016/j.pocean.2006.07.004.
Marshall, J., and K. Speer, 2012: Closure of the meridional overturning circulation through southern ocean upwelling. Nat. Geosci., 5, 171–180, https://doi.org/10.1038/ngeo1391.
McCabe, R. M., B. M. Hickey, E. P. Dever, and P. MacCready, 2015: Seasonal cross-shelf flow structure, upwelling relaxation, and the alongshelf pressure gradient in the northern California Current System. J. Phys. Oceanogr., 45, 209–227, https://doi.org/10.1175/JPO-D-14-0025.1.
Morrison, A. K., A. M. Hogg, and M. L. Ward, 2011: Sensitivity of the southern ocean overturning circulation to surface buoyancy forcing. Geophys. Res. Lett., 38, L14602, https://doi.org/10.1029/2011GL048031.
Nagai, T., N. Gruber, H. Frenzel, Z. Lachkar, J. C. McWilliams, and G.-K. Plattner, 2015: Dominant role of eddies and filaments in the offshore transport of carbon and nutrients in the California current system. J. Geophys. Res. Oceans, 120, 5318–5341, https://doi.org/10.1002/2015JC010889.
Ndoye, S., X. Capet, P. Estrade, B. Sow, E. Machu, T. Brochier, J. Döring, and P. Brehmer, 2017: Dynamics of a “low-enrichment high-retention” upwelling center over the southern Senegal shelf. Geophys. Res. Lett., 44, 5034–5043, https://doi.org/10.1002/2017GL072789.
Nurser, A. J. G., and M.-M. Lee, 2004a: Isopycnal averaging at constant height. Part I: The formulation and a case study. J. Phys. Oceanogr., 34, 2721–2739, https://doi.org/10.1175/JPO2649.1.
Nurser, A. J. G., and M.-M. Lee, 2004b: Isopycnal averaging at constant height. Part II: Relating to the residual streamfunction in Eulerian space. J. Phys. Oceanogr., 34, 2740–2755, https://doi.org/10.1175/JPO2650.1.
Pallàs-Sanz, E., T. Johnston, and D. Rudnick, 2010a: Frontal dynamics in a California Current System shallow front: 1. Frontal processes and tracer structure. J. Geophys. Res., 115, C12067, https://doi.org/10.1029/2009JC006032.
Pallàs-Sanz, E., T. Johnston, and D. Rudnick, 2010b: Frontal dynamics in a California Current System shallow front: 2. Mesoscale vertical velocity. J. Geophys. Res., 115, C12068, https://doi.org/10.1029/2010JC006474.
Praveen Kumar, B., J. Vialard, M. Lengaigne, V. S. N. Murty, and M. J. McPhaden, 2012: Tropflux: Air-sea fluxes for the global tropical oceans description and evaluation. Climate Dyn., 38, 1521–1543, https://doi.org/10.1007/s00382-011-1115-0.
Ridgway, K., J. R. Dunn, and J. Wilkin, 2002: Ocean interpolation by four-dimensional weighted least squares—Application to the waters around Australasia. J. Atmos. Oceanic Technol., 19, 1357–1375, https://doi.org/10.1175/1520-0426(2002)019<1357:OIBFDW>2.0.CO;2.
Send, U., R. C. Beardsley, and C. D. Winant, 1987: Relaxation from upwelling in the coastal ocean dynamics experiment. J. Geophys. Res., 92, 1683–1698, https://doi.org/10.1029/JC092iC02p01683.
Shchepetkin, A. F., and J. C. McWilliams, 2009: Correction and commentary for “ocean forecasting in terrain-following coordinates: Formulation and skill assessment of the regional ocean modeling system” by Haidvogel et al., J. Comp. Phys. 227, pp. 3595–3624. J. Comput. Phys., 228, 8985–9000, https://doi.org/10.1016/j.jcp.2009.09.002.
Shearman, R. K., J. A. Barth, and P. M. Kosro, 1999: Diagnosis of the three-dimensional circulation associated with mesoscale motion in the California Current. J. Phys. Oceanogr., 29, 651–670, https://doi.org/10.1175/1520-0485(1999)029<0651:DOTTDC>2.0.CO;2.
Silvester, J. M., Y.-D. Lenn, J. A. Polton, T. P. Rippeth, and M. M. Maqueda, 2014: Observations of a diapycnal shortcut to adiabatic upwelling of Antarctic Circumpolar Deep Water. Geophys. Res. Lett., 41, 7950–7956, https://doi.org/10.1002/2014GL061538.
Small, R. J., E. Curchitser, K. Hedstrom, B. Kauffman, and W. G. Large, 2015: The Benguela upwelling system: Quantifying the sensitivity to resolution and coastal wind representation in a global climate model. J. Climate, 28, 9409–9432, https://doi.org/10.1175/JCLI-D-15-0192.1.
Walin, G., 1982: On the relation between sea-surface heat flow and thermal circulation in the ocean. Tellus, 34, 187–195, https://doi.org/10.3402/tellusa.v34i2.10801.
Werner, F. E., and B. M. Hickey, 1983: The role of a longshore pressure gradient in Pacific Northwest coastal dynamics. J. Phys. Oceanogr., 13, 395–410, https://doi.org/10.1175/1520-0485(1983)013<0395:TROALP>2.0.CO;2.
Zhai, X., and D. R. Munday, 2014: Sensitivity of southern ocean overturning to wind stress changes: Role of surface restoring time scales. Ocean Modell., 84, 12–25, https://doi.org/10.1016/j.ocemod.2014.09.004.
This result is seemingly at odds with previous studies including Colas et al. (2012, their Fig. 10) where a 15-fold increase in horizontal resolution strongly increases APE to EKE conversion near the surface. In reality, note that the absolute change in mixed layer