1. Introduction
North of 28°S, off the Brazilian coast, the Antarctic Intermediate Water (AAIW) flow bifurcation sets up an equatorward-flowing Intermediate Western Boundary Current (IWBC), opposing the Brazil Current (BC) direction (Boebel et al. 1997, 1999; Legeais et al. 2013). The IWBC was first predicted by Stommel (1965) in his seminal book The Gulf Stream as part of the South Atlantic meridional overturning circulation (MOC). Observational evidence of the IWBC and the AAIW transport, however, dates to the late 1990s and early 2000s (Boebel et al. 1997; Müller et al. 1998; Boebel et al. 1999; Schmid et al. 2000; da Silveira et al. 2004; Campos 2006). More recent investigations define the IWBC as an equatorward jet, spanning from ~400 m to greater than ~1600-m depth, that carries about 6 Sv (1 Sv ≡ 106 m3 s−1) of intermediate waters (da Silveira et al. 2008; Rocha et al. 2014; Biló et al. 2014). The most well-established of those water masses is the AAIW, traced oceanwide by its low-salinity core (Wüst 1935) and surface-referenced potential density between 1027.1 and 1027.4 kg m−3 (Tsuchiya et al. 1994). This northward volume flux of intermediate water closes the MOC, and thus the AAIW is an essential component of the climate system (Rintoul 1991; Schmitz 1995).
The Vitória–Trindade Ridge (VTR)—a quasi-zonal seamount chain at 20°S—is a western boundary current rendezvous point, where the BC and IWBC meet important topographic constraints and are forced to go through the banks, generating mesoscale (and most likely submesoscale) structures. For the IWBC, three main obstacles are the probable causes for the current to meander: Cape São Tomé, Tubarão Bight, and the VTR topographic features (Fig. 1).
Mesoscale variability in the IWBC has been explored in the last decade by the works of da Silveira et al. (2008), Mano et al. (2009), Legeais et al. (2013), and Costa et al. (2017). Da Silveira et al. (2008) show baroclinic instability as the main forcing mechanism for the IWBC unstable meanders. Mano et al. (2009) showed that the BC–IWBC meandering starts at the IWBC and transfers energy from intermediate to upper layers. Legeais et al. (2013) presented evidence of abundant mesoscale motions at the IWBC level north of the VTR. Costa et al. (2017) described a tight cyclonic recirculation within Tubarão Bight, which is likely to be either permanent or semipermanent. Previous results on the IWBC include current-meter mooring velocity measurements (Evans and Signorini 1985; Müller et al. 1998; Costa et al. 2017), numerical models (da Silveira et al. 2004, 2008; Costa et al. 2017), CTD-derived velocities (da Silveira et al. 2004, 2008), and Lagrangian studies (Boebel et al. 1999; Schmid and Garzoli 2009; Legeais et al. 2013; Costa et al. 2017); however, few have solely focused on the IWBC.
Given the recent findings of Costa et al. (2017) on the stationarity and supposed quasi-steadiness of the Tubarão Bight recirculation and its impact on the dynamics of the IWBC, we formulate the following questions: (i) What is the basic state of the IWBC off southeast Brazil (24°–18°S), and how is this basic state related to the temporal-mean spatial pattern depicted from observations and numerical simulations? (ii) Which mechanisms drive the observed mesoscale variability along the IWBC path?
2. The intermediate circulation between 24° and 18°S
a. Quasi-synoptic observations
We employ recent quasi-synoptic observations obtained in the “Marine Environment Characterization of the Espírito Santo Sedimentary Basin” Experiment (hereafter AMBES), conducted through a partnership of the Oceanographic Institute of the University of São Paulo (IOUSP) and Petróleo Brasileiro S.A. (Petrobras). The spring 2012 AMBES cruise obtained direct velocity measurements with a 38-kHz RDI shipboard acoustic Doppler current profiler (ADCP), which sampled the upper 650–850 m of the water column. We processed the shipboard ADCP data following the guidelines of Firing et al. (1995). We used 10-min ensemble averages and discarded data with a return signal (the so-called “percent good”) below 85%. To the best of our knowledge, these are the first quasi-synoptic velocity observations within the IWBC in the VTR region.
Figure 2a displays ADCP velocity observations at 600 m along the ship track, collected during the AMBES cruise. Figures 2b–d show cross-transect velocity vertical sections for three selected transects (I, II, and III). Transect I in Fig. 2b shows a typical BC–IWBC system pattern: the opposing flows of the southward-flowing BC and the northward-flowing IWBC on the Brazilian southeast continental slope (e.g., Boebel et al. 1999; da Silveira et al. 2008; Lima et al. 2016). The IWBC, which is apparently meandering in transect I, presents a core speed of 0.25 m s−1—the weakest measured during the cruise. (Although treated here as a meander, without additional data south of transect I we cannot rule out other possibilities, for example, the presence of an isolated eddy by the IWBC.) The meandering of the BC–IWBC jet near Cape São Tomé and Cape Frio (~23.5°S) was also investigated by da Silveira et al. (2008), Mano et al. (2009), and Rocha et al. (2014).
Figure 2c shows that the intermediate-level circulation inside Tubarão Bight resembles a cyclone structure. The lobe adjacent to the continental margin exhibits velocities up to 0.50 m s−1. This pattern in transect II confirms the description by Costa et al. (2017): an IWBC cyclonic recirculation within Tubarão Bight. These authors used data from two current-meter moorings, together with Argo float trajectories and the output of a numerical model. Their findings indicated that this recirculation weakens the northward flow that crosses the VTR, and consequently intensifies the flow downstream along the western boundary. This intensification of the IWBC inside Tubarão Bight was also described by Legeais et al. (2013). Moreover, Costa et al. (2017) observed the shoaling of the IWBC within the bight, using mooring velocity data, with a mean velocity reversal depth of 370 m, but with instantaneous reversals at 150 m. As for the BC, our observations depict a ~200-m-deep jet confined to the shelf break.
The IWBC crosses the VTR through a narrow channel between the Besnard and Vitória–Congress Banks, which is the main IWBC path out of Tubarão Bight (Legeais et al. 2013; Costa et al. 2017). Transect III in Fig. 2d captures the main branch of the intermediate current reorganized downstream of the seamounts region, where the IWBC shows strong instantaneous velocities (up to 0.48 m s−1) and the BC is detached from the slope.
The features presented in Figs. 2a–d are consistent with what has been described by da Silveira et al. (2004, 2008), Legeais et al. (2013), and Costa et al. (2017), namely, an organized IWBC off Cape São Tomé (transect I), a strengthened and recirculating IWBC within Tubarão Bight (transect II), and an IWBC branch reorganizing north of the VTR (transect III). However, the data are restricted to a few transects and occupied over a month. These quasi-synoptic observations include a number of transients, and thus cannot be used to assess the stationarity of the IWBC flow and its recirculation.
b. The Regional Ocean Modeling Experiment
To fully examine the steadiness and stability of the circulation within the AAIW layer in the study area, long time series of potential density and velocity are required. Given the paucity of such observations in the region, we opt to answer the questions posed herein with the aid of a regional circulation experiment output. We use the Regional Oceanic Modeling System (ROMS) with a configuration for the western portion of the South Atlantic Ocean (41°16′–10°01′S, 62°34′–19°49′W). The model has a horizontal resolution of 6 km and 30 vertical levels in terrain-following coordinates, which suffices to resolve mesoscale eddies in the region. The simulation was initialized on 1 January 2000 with temperature and salinity fields from the Simple Ocean Data Assimilation (SODA) project and ran for 11 years subject to SODA fields on the open boundaries, with a spinup time of 6 years. Our analysis below uses data spanning the last 7 years of the simulation, in which the dynamical fields were statistically equilibrated. The model is forced by climatological monthly surface wind and heat fluxes from QuikSCAT and COADS, respectively. We emphasize that our goal in using a model forced with monthly mean climatologies of wind and heat fluxes is to obtain a consistent dynamical simulation of the area rather than a hindcast simulation of the observed events.
Figures 2e–g present the September–October cross-transect velocity averages for transects I, II, and III over the 7 years of ROMS output. In Figs. 2h–j, we also show 7-yr monthly averages of transports of the BC and the IWBC.
The model simulated the typical BC–IWBC vertical structure in transect I (Fig. 2e). The modeled IWBC transport shows no significant seasonal variability and is consistent with the simulation in Costa et al. (2017). Our simulated BC transport displays two annual maxima (2-H), while Schmid and Majumder (2018) reported one (summer) or two (summer and spring) annual maxima.
The observed BC–IWBC structure in transect II is qualitatively well reproduced by ROMS. As in the ADCP data, the simulated IWBC is strongest inside Tubarão Bight; this enhanced transport is present in the model, although the recirculating branch is not well captured by the simulation’s transect II (cf. Figs. 2c,f). The recirculation in ROMS is more confined within Tubarão Bight than the one inferred from Argo, as well as displaced northward (more on this in the next section). The modeled BC displays large transport monthly variations in transect II (Fig. 2i), most likely associated with the BC negotiating topography while crossing the VTR (cf. Costa et al. 2017).
The ROMS 2-month average in Fig. 2g depicts a weak and shallow BC, with a transport of 2 Sv. Soutelino et al. (2011) characterized the BC north of the VTR as a shallow and eddy-dominated flow, consistent with our simulation. Just north of the VTR, the simulated IWBC is highly variable on a monthly time scale, alternating northward and eastward flows. The IWBC transport is largest in months of predominantly northward flow (March, April, and September); on the other hand, transport through transect III is nearly zero when the IWBC veers eastward (February, June, and July) (see Fig. 2j). Although apparently overestimated by our simulation, this eastward flow north of the VTR was previously inferred by Wienders et al. (2000) and observed by Schmid and Garzoli (2009), who qualitatively associated it with eastward penetrations of AAIW. Legeais et al. (2013) suggested a link between the weakening of the IWBC north of 20°S and an exchange of AAIW between the western boundary and the ocean interior.
Both BC and IWBC transports in our ROMS simulation match those simulated by Costa et al. (2017). The eastward flow north of the VTR is not present in the simulation in Costa et al. (2017); this is the main inconsistency between the two models. In a literature survey presented in Table 1, we detail a comparison between previous estimates of the BC–IWBC transports and our study.
Comparison between the AMBES cruise, ROMS outputs, and literature. ADCP transports are restricted to the depth range presented in Figs. 2b–d. ROMS transports are averaged for the simulation period (7 years) in virtually the same ADCP transect position. Positive (negative) values indicate northward (southward) volume transport. REMO is the abbreviation for Rede de Modelagem e Observação Oceanogáfica, an initiative of the Brazilian Navy and Petróleo Brasileiro S.A. based on a 1/12° Hybrid Coordinate Ocean Model run.
c. The Argo float climatology
Given that the parking depths of Argo floats lie within the IWBC domain, we can use float trajectories to characterize the time-mean velocity pattern near the VTR (26°–16°S, 43°–32°W). We obtained the Argo dataset (Argo 2019) from the AOML database (http://tds0.ifremer.fr/thredds/catalog/CORIOLIS-ARGO-GDAC-OBS/aoml/catalog.html) and selected floats that entered the domain from January 2000 through April 2019, drifting within the IWBC with parking depths between 600 and 1100 m.
We determined the location and period of an Argo cycle by the middle point of the float trajectory and its duration, respectively. (Fig. 3 shows the spatial distribution of the number of Argo observations on a 12.5 km × 12.5 km grid.) We identify each Argo cycle in the float trajectory and treat it as an independent sample. In doing so, we eliminate surface drift during data transmission and ignore the short profiling time (Park et al. 2005). Also, floats that dwelt for too short (<5 days) or too long (>15 days) in their parking depths were removed from the analysis. These criteria yield 5503 cycles from 90 floats that occupied the region, with an average cycle of 9.6 ± 0.18 days.
We estimate horizontal velocity (u and υ) at the parking depth from the position difference between the end and the beginning of each cycle divided by the average parking time (Lebedev et al. 2007). We then interpolate the parking-depth velocity onto a 12.5 km × 12.5 km grid. Figure 3 shows that the Argo array presents good coverage in the region compared to previous datasets, but the number of Argo cycles per grid point is still small, particularly north of the VTR.
The Argo-derived flow (Fig. 4a) depicts the IWBC flowing northward along the continental slope with maximum speed of ~0.30 m s−1. Argo climatological speeds are close to the Costa et al. (2017) mooring time averages. Müller et al. (1998) analyzed moorings in the region, with mean velocities ranging from 6 to 21 cm s−1, both in the same ballpark as our estimates. Near the Tubarão Bight northern limit, the flow splits into two branches, one exiting the bight flowing northward (Legeais et al. 2013; Costa et al. 2017), and the other veering southward and forming a cyclonic recirculation. This Argo climatology does not display an intermediate flow through the Besnard Passage, in contrast to the strong synoptic flow observed in the AMBES ADCP survey (Fig. 2a) and two RAFOS float trajectories that exited the bight through this passage (Legeais et al. 2013).
d. Comparison between Argo float climatology and ROMS
The horizontal mean velocity of our ROMS simulation is qualitatively consistent with the Argo intermediate circulation pattern (cf. Figs. 4a and 4b). ROMS maximum velocities are 25% larger than maximum Argo velocities, and these quantitative differences are likely due to the smoothing nature of the trajectory-based Argo velocity (an average over 9.6 days and 12.5 km × 12.5 km) compared to ROMS’s daily snapshots on a 6-km horizontal grid.
According to the Argo estimates, about 58% of the flow that enters Tubarão Bight at Cape São Tomé exits it through the main channel; the remaining 42% recirculates. In ROMS, 70% of the flow exits the bight through the main channel, and the remaining 30% recirculates. The flow that crosses the ridge through the main channel bifurcates: 58% flows north and 42% veers eastward. Similar to ROMS, Argo presents an eastward flow between 36° and 34°W. But Argo does not clearly depict a zonally elongated recirculation, which in ROMS extends east of Tubarão Bight and around the VTR for 400 km. As mentioned earlier, our simulation may be overestimating this eastward flow and, owing to paucity of observations, care must be taken in interpreting model results north of the VTR [see Fig. 3 and the discussion of data density in Schmid and Garzoli (2009)].
Two mesoscale features appear in the lee of the VTR between ~18° and 20°S: a cyclone centered at ~37°W and a larger anticyclone at ~35°W, off the Abrolhos Bank. Legeais et al. (2013) noted that some floats stalled for a long time off the Abrolhos Bank. Those floats were probably trapped by these features.
3. The IWBC steady state in the vicinity of the VTR
In the previous section, we show that mesoscale features appear in the time-mean IWBC flow. Are these time-mean features the result of strong unsteady eddies? Or are they permanent structures caused by the topographic steering of the IWBC?
If the mesoscale features are steady, then the flow should occur along potential vorticity contours (e.g., Bretherton and Haidvogel 1976). On the other hand, mean potential vorticity contours cross streamlines when mesoscales features are unsteady (e.g., Vallis 2017). To address this question, we formulate an intermediate-layer quasigeostrophic model within the AAIW layer (1027.1–1027.4 kg m−3; Tsuchiya et al. 1994). We diagnose quasigeostrophic streamfunction ψ and potential vorticity Q, and investigate the properties of the mean flow using ψ–Q scatter diagrams (Bretherton and Haidvogel 1976; Read et al. 1986).
a. Model formulation
b. The steady solution
Figure 6 shows the streamfunction
c. Stability of the steady state
Focusing on smaller regions along the IWBC path, different values of I occur according to the flow characteristics: We see values in the Cape São Tomé region higher than in the recirculation and in the main channel, which is explained by the rich and frequent meandering activity of the IWBC that adds perturbation terms and weakens the mean flow (da Silveira et al. 2008). This also implies that the flow in the vicinity of Tubarão Bight is steadier (I = 0.05, not shown).
The
So far, we have shown that the time-mean IWBC is representative of the steady state, although instabilities may be present. Hence, we shall ask: do local conversions account for all the variability? Or is the variability driven by remote forcing? We next address these questions, using a detailed analysis of the energetics of the eddy–mean flow interactions of the IWBC.
4. Energetics of the eddy–mean flow interaction
In oceanography, we must use density (buoyancy) to calculate the potential energy budget, especially in cases where salinity plays a major role in the fluid density, as in the AAIW. The potential energy above is only an exact definition for constant N2 (Huang 2005).
a. Energy conservation equations
Main terms from the eddy energy conservation equations: u = (u, υ, w) is the velocity vector, ∇ is the gradient operator, Ψ is the horizontal streamfunction, N2 is the buoyancy frequency, p is pressure, and b is buoyancy. Subscripted indices indicate derivatives.
The time-varying budgets on the left side of (15) and (16) depend on the balance between the divergence of energy at the boundaries of the domain and the local energy production on their right-hand sides. The conversion terms on the right quantify the transformations between the different forms of energy. The terms
The flux divergence terms Ge, Fe, and Pw account for the redistribution of eddy energy through advection and pressure work. Although not computed explicitly, the residual term that represents horizontal and vertical mixing, heat and freshwater fluxes, wind forcing, and bottom drag is included to close the energy balance (Chen et al. 2014).
We calculate energy budgets in the IWBC using ROMS daily outputs to estimate the conversion and redistribution terms in Table 2. We vertically average ROMS fields within the AAIW layer (1027.1–1027.4 kg m−3; Tsuchiya et al. 1994) and apply the decomposition (12) to the resulting 2D fields (terms with z dependence are calculated prior to averaging). We define the eddy component as variability with time scales shorter than 60 days.
b. Energy conversions
Existing studies of the BC–IWBC system energetics focused on regions south of the VTR, restricted either to Cape São Tomé and Cape Frio (e.g., Mano et al. 2009) or to southeast Brazil (e.g., Oliveira et al. 2009; Magalhães et al. 2017). Those studies showed that baroclinic conversions account for most of the eddy generation, though barotropic conversions may be important in some regions.
Here, both barotropic and baroclinic conversions are at play (see Fig. 9). In particular, barotropic conversion through horizontal shear production (HSP) is larger where the IWBC splits into two branches near the exit from Tubarão Bight and off the Abrolhos Bank (see Fig. 9a). Baroclinic conversion, through horizontal buoyancy production (HBP), is also enhanced north of the VTR; HSP is twice as large as HBP. Vertical conversions, through vertical shear production (VSP) and vertical buoyancy production (VBP), are small over most of the region but large near seamounts, becoming important in the energy budget of subregions such as Tubarão Bight. In the ocean, this is associated with enhanced vertical mixing (e.g., Polzin et al. 1997).
The theoretical analysis of section 3 suggests that the region south of the VTR is prone to instabilities of the IWBC mean flow, yet the energetics of the numerical model in this section show that most conversions occur north of the VTR. To solve this apparent contradiction, we next look into the details of the model output within the Tubarão Bight region.
c. Breaking the steadiness
Figure 10c shows a Hovmöller diagram of ROMS meridional eddy velocity along the path in Fig. 10a. We average
We track one of the strongest perturbation events in the simulation (Fig. 11). This perturbation is enhanced at ~36°W on 1 May 2002 (see Fig. 11a) and propagates westward at 4.47 km day−1. About 25 days after this enhancement at ~36°W, an anticyclonic structure begins to approach the eastern limit of Tubarão Bight (see Fig. 11b). Within the bight, the ring propagates to the southwest through the time-mean flow; the eddy velocity is perpendicular to the time-mean flow (see vectors and streamlines in Fig. 11c). Once the ring reaches the western boundary, it is strained by the mean flow and quickly advected downstream (see Fig. 11d). The perturbation grows significantly as it crosses the VTR, generating a strong eddy field north of the ridge (Figs. 11e,f). (This sequence of events occurs for most of the perturbations seen in the Hovmöller diagram.) The strong eddy field off the Abrolhos Bank is consistent with the swirly float trajectories reported by Legeais et al. (2013).
We select 29 events from the series and estimate a mean wavelength of 344 ± 89 km, a mean period of 36 ± 7 days, and a mean phase speed c = 0.06 ± 0.01 m s−1 (5.18 km day−1). The mean eddy speed is |u| = 0.07 ± 0.03 m s−1 (6.05 km day−1), so |u|/c ≈ 1.14 ± 0.57, suggesting nonlinearity (Chelton et al. 2011). We show that these nonlinear westward-propagating eddies have Eulerian phase speeds consistent with the nondispersive linear Rossby-wave dispersion relation (Fig. 12a), as previously remarked by Morten et al. (2017). A standard modal analysis of these subsurface-intensified eddies indicates that the first baroclinic mode accounts for most of the variance, about 40%. And a synthesis with the gravest four modes accounts for 80% of the eddy vertical structure (Figs. 12b–d).
d. Tubarão Bight energetics
We now turn to the energy budget in Tubarão Bight, defined as the ~19–23°S domain west of ~36°W. For this region, we calculate the energy conversion terms and the energy fluxes in Table 2. Figures 13a–c show in colors the total eddy kinetic energy generation,
In summary, the energetics suggest that westward-propagating features, such as the one in Fig. 11, interact with the mean flow as they are advected by the IWBC, experiencing explosive growth downstream of the VTR. The mechanism of interaction is barotropic conversions, and the process appears to be constrained by topography. A simple representation of the Tubarão Bight energy budget is shown in Fig. 16.
5. Final remarks
New direct velocity measurements of the Intermediate Western Boundary Current (IWBC) show (and detail) a circulation pattern consistent with the existing literature, namely, (i) off Cape São Tomé, the IWBC flows northward along the continental slope, underneath the southward-flowing Brazil Current; (ii) at Tubarão Bight, topography steers the IWBC, generating a cyclonic recirculation that intensifies the flow along the continental slope south of the bight; and (iii) the remaining IWBC flow exits the bight through the Vitória–Trindade Ridge main channel. Argo float trajectories reveal a larger, zonally elongated IWBC recirculation around the VTR that extends to ~35°W. A regional simulation with ROMS shows good skill in simulating these observational patterns, including the IWBC recirculation cells.
Analysis of an intermediate-layer QG model shows that the time-mean ROMS circulation is a good proxy for the IWBC steady state (>90%), with the linear inversion relation
Despite a steadiness indicated by the QG analysis, ROMS eddy–mean energy exchanges are important throughout the model domain. A detailed energy analysis around Tubarão Bight shows that steadiness is broken by nonlinear eddies that enter the domain through the eastern boundary. Despite their nonlinearity, these eddies have phase speeds consistent with nondispersive linear Rossby-wave theory. The perturbations interact with the IWBC mean flow via barotropic conversions. As they are advected downstream, the eddies grow by feeding off the mean flow through standard Reynolds-stress horizontal shear production, with topography seemly playing an important role in this growth process.
The model results highlight the complexity of the eddy–mean flow interactions off east Brazil, with both remote forcing and downstream eddy growth playing critical roles. Process-oriented observational studies are needed to test these model predictions and further characterize local and nonlocal eddies in this region and their effects on the IWBC and the Meridional Overturning Circulation.
Acknowledgments
We thank Frank O. Smith for copy editing and proofreading this manuscript. This study was financed in part by Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—CAPES, Brazil—Finance Code 001 and by Projeto REMARSUL (Processo CAPES 88882.158621/2014-01), Projeto VT-Dyn (Processo FAPESP 2015/21729-4) and Projeto SUBMESO (Processo CNPq 442926/2015-4). Rocha was supported by a WHOI Postdoctoral Scholarship.
APPENDIX A
Calculation of g′
APPENDIX B
Estimation of Scatter Cloud Relative Area
Read et al. (1986) discuss the caveats of this index, which depends on the orientation of the polygon, as well as on its shape in ψ–Q space. The authors also propose an alternative metric based on the perpendicular width of the scatter cloud relative to its length: given a scatter cloud angle ϕ ≤ π/2, the width-to-length ratio is thus tan(ϕ)/2 and a measure of the departure from free mode. For the IWBC, this metric yields ~0.15, against 0.08 of the I index.
REFERENCES
Argo, 2019: Argo float data and metadata from Global Data Assembly Centre (Argo GDAC). SEANOE, https://doi.org/10.17882/42182.
Biló, T. C., I. C. A. da Silveira, W. C. Belo, B. M. Castro, and A. R. Piola, 2014: Methods for estimating the velocities of the Brazil Current in the pre-salt reservoir area off southeast Brazil (23° S–26° S). Ocean Dyn., 64, 1431–1446, https://doi.org/10.1007/s10236-014-0761-2.
Blumen, W., 1968: On the stability of quasi-geostrophic flow. J. Atmos. Sci., 25, 929–931, https://doi.org/10.1175/1520-0469(1968)025<0929:OTSOQG>2.0.CO;2.
Boebel, O., C. Schmid, and W. Zenk, 1997: Flow and recirculation of Antarctic Intermediate Water across the Rio Grande rise. J. Geophys. Res., 102, 20 967–20 986, https://doi.org/10.1029/97JC00977.
Boebel, O., R. Davis, M. Ollitrault, R. Peterson, P. Richardson, C. Schmid, and W. Zenk, 1999: The intermediate depth circulation of the western South Atlantic. Geophys. Res. Lett., 26, 3329–3332, https://doi.org/10.1029/1999GL002355.
Bretherton, F. P., and D. B. Haidvogel, 1976: Two-dimensional turbulence above topography. J. Fluid Mech., 78, 129–154, https://doi.org/10.1017/S002211207600236X.
Campos, E. J., 2006: Equatorward translation of the Vitória Eddy in a numerical simulation. Geophys. Res. Lett., 33, L22607, https://doi.org/10.1029/2006GL026997.
Chelton, D. B., M. G. Schlax, and R. M. Samelson, 2011: Global observations of nonlinear mesoscale eddies. Prog. Oceanogr., 91, 167–216, https://doi.org/10.1016/j.pocean.2011.01.002.
Chen, R., G. R. Flierl, and C. Wunsch, 2014: A description of local and nonlocal eddy–mean flow interaction in a global eddy-permitting state estimate. J. Phys. Oceanogr., 44, 2336–2352, https://doi.org/10.1175/JPO-D-14-0009.1.
Costa, V. S., G. N. Mill, M. Gabioux, G. S. Grossmann-Matheson, and A. M. Paiva, 2017: The recirculation of the intermediate western boundary current at the Tubarão Bight–Brazil. Deep-Sea Res. I, 120, 48–60, https://doi.org/10.1016/j.dsr.2016.12.001.
da Silveira, I. C. A., and G. R. Flierl, 2002: Eddy formation in 2 1/2-layer, quasigeostrophic jets. J. Phys. Oceanogr., 32, 729–745, https://doi.org/10.1175/1520-0485(2002)032<0729:EFILQJ>2.0.CO;2.
da Silveira, I. C. A., L. Calado, B. Castro, M. Cirano, J. Lima, and A. S. Mascarenhas, 2004: On the baroclinic structure of the Brazil Current–Intermediate Western Boundary Current system at 22°–23°S. Geophys. Res. Lett., 31, L14308, https://doi.org/10.1029/2004GL020036.
da Silveira, I. C. A., J. Lima, A. Schmidt, W. Ceccopieri, A. Sartori, C. Franscisco, and R. Fontes, 2008: Is the meander growth in the Brazil Current system off Southeast Brazil due to baroclinic instability? Dyn. Atmos. Oceans, 45, 187–207, https://doi.org/10.1016/j.dynatmoce.2008.01.002.
Evans, D. L., and S. S. Signorini, 1985: Vertical structure of the Brazil Current. Nature, 315, 48–50, https://doi.org/10.1038/315048a0.
Evans, D. L., S. R. Signorini, and L. B. Miranda, 1983: A note on the transport of the Brazil Current. J. Phys. Oceanogr., 13, 1732–1738, https://doi.org/10.1175/1520-0485(1983)013<1732:ANOTTO>2.0.CO;2.
Firing, E., J. Ranada, and P. Caldwell, 1995: Processing ADCP data with the CODAS software system version 3.1. Tech. Doc., 218 pp.
Gill, A., J. Green, and A. Simmons, 1974: Energy partition in the large-scale ocean circulation and the production of mid-ocean eddies. Deep-Sea Res. Oceanogr. Abstr., 21, 499–528, https://doi.org/10.1016/0011-7471(74)90010-2.
Hart, J., 1974: On the mixed stability program for quasi-geostrophic ocean currents. J. Phys. Oceanogr., 4, 349–356, https://doi.org/10.1175/1520-0485(1974)004<0349:OTMSPF>2.0.CO;2.
Huang, R. X., 2005: Available potential energy in the world’s oceans. J. Mar. Res., 63, 141–158, https://doi.org/10.1357/0022240053693770.
Isern-Fontanet, J., B. Chapron, G. Lapeyre, and P. Klein, 2006: Potential use of microwave sea surface temperatures for the estimation of ocean currents. Geophys. Res. Lett., 33, L24608, https://doi.org/10.1029/2006GL027801.
Lebedev, K. V., H. Yoshinari, N. A. Maximenko, and P. W. Hacker, 2007: Velocity data assessed from trajectories of Argo floats at parking level and at the sea surface. IPRC Tech. Note 4(2), 16 pp., http://apdrc.soest.hawaii.edu/projects/yomaha/yomaha07/YoMaHa070612.pdf.
Legeais, J.-F., M. Ollitrault, and M. Arhan, 2013: Lagrangian observations in the intermediate western boundary current of the South Atlantic. Deep-Sea Res. II, 85, 109–126, https://doi.org/10.1016/j.dsr2.2012.07.028.
Lima, M. O., M. Cirano, M. M. Mata, M. Goes, G. Goni, and M. Baringer, 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, https://doi.org/10.1007/s10236-016-0959-6.
Magalhães, F. C., J. L. L. Azevedo, and L. R. Oliveira, 2017: Energetics of eddy-mean flow interactions in the Brazil Current between 20°S and 36°S. J. Geophys. Res. Oceans, 122, 6129–6146, https://doi.org/10.1002/2016JC012609.
Mano, M. F., A. M. Paiva, A. R. Torres Jr., and A. L. Coutinho, 2009: Energy flux to a cyclonic eddy off Cabo Frio, Brazil. J. Phys. Oceanogr., 39, 2999–3010, https://doi.org/10.1175/2009JPO4026.1.
Mata, M. M., S. E. Wijffels, J. A. Church, and M. Tomczak, 2006: Eddy shedding and energy conversions in the East Australian Current. J. Geophys. Res., 111, C09034, https://doi.org/10.1029/2006JC003592.
Mata, M. M., M. Cirano, M. R. van Caaspel, C. Fonteles, G. Goni, and M. Baringer, 2013: Observations of the Brazil Current baroclinic transportation near 22°S: Variability from the AX97 XBT transect and satellite altimetry. Proceedings of 20 years of progress in radar altimetry: 24–29 September 2012, Venice, Italy, ESA SP, Vol. 710, European Space Agency, 151.
Morten, A. J., B. K. Arbic, and G. R. Flierl, 2017: Wavenumber-frequency analysis of single-layer shallow-water beta-plane quasi-geostrophic turbulence. Phys. Fluids, 29, 106602, https://doi.org/10.1063/1.5003846.
Müller, T. J., Y. Ikeda, N. Zangenberg, and L. V. Nonato, 1998: Direct measurements of western boundary currents off Brazil between 20°S and 28°S. J. Geophys. Res., 103, 5429–5437, https://doi.org/10.1029/97JC03529.
Oliveira, L. R., A. R. Piola, M. M. Mata, and I. D. Soares, 2009: Brazil Current surface circulation and energetics observed from drifting buoys. J. Geophys. Res. Oceans, 114, C10006, https://doi.org/10.1029/2008JC004900.
Park, J. J., K. Kim, B. A. King, and S. C. Riser, 2005: An advanced method to estimate deep currents from profiling floats. J. Atmos. Oceanic Technol., 22, 1294–1304, https://doi.org/10.1175/JTECH1748.1.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. Springer, 710 pp.
Polzin, K., J. Toole, J. Ledwell, and R. Schmitt, 1997: Spatial variability of turbulent mixing in the abyssal ocean. Science, 276, 93–96, https://doi.org/10.1126/science.276.5309.93.
Read, P., P. Rhines, and A. White, 1986: Geostrophic scatter diagrams and potential vorticity dynamics. J. Atmos. Sci., 43, 3226–3240, https://doi.org/10.1175/1520-0469(1986)043<3226:GSDAPV>2.0.CO;2.
Rintoul, S. R., 1991: South Atlantic interbasin exchange. J. Geophys. Res., 96, 2675–2692, https://doi.org/10.1029/90JC02422.
Rocha, C. B., I. C. da Silveira, B. M. Castro, and J. A. M. Lima, 2014: Vertical structure, energetics, and dynamics of the Brazil Current System at 22°S–28°S. J. Geophys. Res. Oceans, 119, 52–69, https://doi.org/10.1002/2013JC009143.
Schmid, C., and S. L. Garzoli, 2009: New observations of the spreading and variability of the Antarctic Intermediate Water in the Atlantic. J. Mar. Res., 67, 815–843, https://doi.org/10.1357/002224009792006151.
Schmid, C., and S. Majumder, 2018: Transport variability of the Brazil Current from observations and a data assimilation model. Ocean Sci., 14, 417–436, https://doi.org/10.5194/os-14-417-2018.
Schmid, C., G. Siedler, and W. Zenk, 2000: Dynamics of intermediate water circulation in the subtropical South Atlantic. J. Phys. Oceanogr., 30, 3191–3211, https://doi.org/10.1175/1520-0485(2000)030<3191:DOIWCI>2.0.CO;2.
Schmitz, W. J., 1995: On the interbasin-scale thermohaline circulation. Rev. Geophys., 33, 151–173, https://doi.org/10.1029/95RG00879.
Soutelino, R. G., I. C. da Silveira, A. A. Gangopadhyay, and J. Miranda, 2011: Is the Brazil Current eddy-dominated to the north of 20°S? Geophys. Res. Lett., 38, L03607, https://doi.org/10.1029/2010GL046276.
Stommel, H. M., 1965: The Gulf Stream: A Physical and Dynamical Description. 2nd ed., University of California Press, 248 pp.
Stramma, L., Y. Ikeda, and R. G. Peterson, 1990: Geostrophic transport in the Brazil Current region north of 20 S. Deep-Sea Res., 37A, 1875–1886, https://doi.org/10.1016/0198-0149(90)90083-8.
Tsuchiya, M., L. D. Talley, and M. S. McCartney, 1994: Water-mass distributions in the western South Atlantic; A section from South Georgia Island (54S) northward across the equator. J. Mar. Res., 52, 55–81, https://doi.org/10.1357/0022240943076759.
Vallis, G. K., 2017: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. 2nd ed. Cambridge University Press, 964 pp.
Weatherall, P., and Coauthors, 2015: A new digital bathymetric model of the world’s oceans. Earth Space Sci., 2, 331–345, https://doi.org/10.1002/2015EA000107.
Wienders, N., M. Arhan, and H. Mercier, 2000: Circulation at the western boundary of the South and Equatorial Atlantic: Exchanges with the ocean interior. J. Mar. Res., 58, 1007–1039, https://doi.org/10.1357/002224000763485782.
Wüst, G., 1935: The Stratosphere of the Atlantic Ocean. Scientific Results of the German Atlantic Expedition of the Research Vessel “Meteor” 1925–27, Vol. 6, 109–288 (English translation, W. J. Emery, Ed., Amerind, 1978).