1. Introduction
The densest waters found in the ocean interior are formed at several high-latitude sites where atmospheric forcing causes surface water temperature and salinity conditions to match those of the deepest waters. These waters spill into the abyss and are carried equatorward by a system of deep boundary currents (DBCs). This view was outlined by Stommel (1957), and a model proposed by Stommel and Arons (1960) described a simple set of dynamics that link the DBCs to a weak interior circulation. Being far removed from the wind-driven Ekman layer, flow in the abyss is driven by buoyancy forcing. Lacking any information on the buoyancy forcing, Stommel and Arons calculated the circulation resulting from uniform upwelling. Under this scenario, poleward flow dominates the interior. However, early experiments by Stommel et al. (1958) demonstrated the tendency for zonal flow when the deep interior is characterized by a strong zonal gradient in the magnitude of buoyancy forcing. Specifically, the Stommel et al. experiments showed that zonal currents would transport mass between the deep western boundary current and isolated sources of upwelling to the east, with meridional motions in the interior inhibited by the absence of buoyancy forces.
The Stommel–Arons model is still central to our view of the abyssal circulation problem, and the important role of buoyancy forcing in the abyss is not questioned. Several studies have attempted to estimate the mixing rates responsible for abyssal buoyancy forcing. In particular, measurements of flow through deep passages have been used to formulate closed heat and mass budgets for semienclosed basins. Such heat–mass budgets have been the primary means of assessing abyssal mixing rates, and vertical diffusivities of kρ ≃ (1–10) × 10−4 m2 s−1 have been reported in the North and South Atlantic (Saunders 1987; Whitehead and Worthington 1982), the South Pacific (Roemmich et al. 1996), and the Indian Ocean (Barton and Hill 1989). These mixing rates are up to several orders of magnitude larger than typical thermocline values. In contrast to the mixing rates implied by abyssal budgets, direct observations of dissipation over abyssal plains suggest that diffusivities there are comparable to thermocline diffusivity levels of kρ ≃ 1 × 10−5 m2 s−1 (Toole et al. 1994; Kunze and Sanford 1996). Instead, the larger mixing rates needed to reconcile the budget estimates seem to occur near rough bathymetry, as suggested by several studies (Toole et al. 1994, 1997a; Polzin et al. 1997).
The focus of the present work is the South Atlantic, where the densest waters of the global ocean are formed in the Weddell Sea off Antarctica. These waters, so-called Antarctic Bottom Water (AABW), flow northward along the western side of the Argentine Basin, reaching the Brazil Basin via the deep passages of the Vema and Hunter Channels. The isotherm bounding the water colder than about 1°C potential temperature intersects the seafloor within the basin, such that the coldest class of AABW does not exit the Brazil Basin adiabatically. The diffusion of heat driven by turbulent mixing provides the buoyancy needed to upwell this water. The Brazil Basin budget of AABW was considered by Hogg et al. (1982) and revisited by Morris et al. (2001). These investigations indicated that a basin-average vertical diffusivity of (2–4) × 10−4 m2 s−1 was necessary to close the basin's heat and mass budgets.
However, as demonstrated by the experiments of Stommel et al. (1958), the structure of the interior flow is dependent on the specific distribution of deep sources and sinks of buoyancy, and budget studies give only spatially integrated estimates of buoyancy forcing. To investigate how spatial patterns of mixing influence flow in the abyss, a combined microstructure observation program and tracer dispersion experiment was started during 1996: the Brazil Basin Tracer Release Experiment (BBTRE).
Full depth microstructure data from the Brazil Basin were collected during January 1996 and April 1997. A free-falling, autonomous instrument platform called the high resolution profiler was used (HRP: Schmitt et al. 1988). The collective microstructure observations spanned nearly 30° longitude, covering the western basin and extending to the fracture zones (FZs) of the Mid-Atlantic Ridge (MAR). A preliminary presentation of microstructure-derived dissipation rates from the 1996 survey was given by Polzin et al. (1997), and data from the 1997 survey were discussed by Toole et al. (1997b). While these previous reports have focused on the estimation of turbulent diffusivities for the deep water of the Brazil Basin, the focus here will be to assimilate the dissipation data into an inverse model for the steady-state circulation.
The tracer dispersion experiment also began during January 1996 with an injection of 110 kg of sulfur hexafluoride at 4000-m depth, near 21.5°S, 18.5°W. The initial tracer plume was neutrally buoyant along the σ4 = 45.94 kg m−3 isopycnal. This density surface is 500 to 1000 m above the FZ bathymetry at the longitude of the injection but deepens and intersects the seafloor about 400 km to the east. Tracer inventories were conducted during April 1997 (14 months postinjection) and March 1998 (26 months postinjection). Details of the tracer's dispersion are given by Ledwell et al. (2000) and will generally not be further described here. However, the tracer results compliment the microstructure-derived estimates, and relevant comparisons will be cited throughout this report.
Previous attempts to examine mixing rates and diapycnal velocities in inverse models have been done for both box models of the advective budgets (e.g., Zhang and Hogg 1992) and beta-spiral inversions (e.g., Hautala and Riser 1993). These inversions all rely on either integrated property budgets (box inversions) or differential property budgets (beta-spiral inversions), and the mixing rates and diapycnal velocity estimates are determined through the divergences of lateral fluxes. This form of mixing-rate estimation is highly sensitive to small uncertainties in the advective terms, as discussed by Wunsch (1996). The inverse model used in this study resembles the model of Fukumori (1991) in that both integrated and differential budgets are used. In this manner, mass conservation and the thermal advection are closed in volume-integrated cells, while the advection of potential vorticity is examined in the differential form of the beta-spiral equation.
The description of the turbulent dissipation observed during BBTRE will begin in section 2. There, we will discuss spatial and temporal features of mixing. In section 3, an inverse model for estimating the circulation will be formulated. There we will describe the governing equations for the model (section 3a), the model for the time-mean dissipation (section 3b), the treatment of data (section 3c), and the details of the linear inverse (section 3d). The results of the inverse calculation are presented in section 4. There we present estimates of both shallow (section 4a) and deep (section 4b) flow, canyon flow (section 4c), and deep upwelling and mixing efficiency (section 4d). A discussion of the estimated circulation will be given in section 5.
2. Observations
In the period between January 1996 and April 1997, two microstructure surveys were made in the Brazil Basin. During the first cruise, January through February 1996, a 75 station HRP survey consisting of two section legs covered a region extending from the continental slope regions off Rio de Janeiro (24°S) and Recife (11°S) to the western edge of the MAR (16°W at 21°S) where the tracer was injected. A second survey was carried out in March and April of 1997. An irregular grid of HRP stations was occupied over the region (19°–25°S, 12°–23°W). The primary focus of the 1996 survey was the spatial variability of mixing levels, and a report on the observed variability is given by Polzin et al. (1997).
The present work will focus on dissipation data collected over the fracture zones east of 25°W. This dataset consists of 87 stations from the 1997 survey and an additional 42 stations from the 1996 survey (Fig. 1). All profiles consist of conventional hydrographic variables (e.g., Θ, S, U, V) in addition to microstructure, and generally extend from the surface mixed layer to within 20 m of the bottom. The bathymetric data shown in Fig. 1 were derived from satellite measurements of the marine gravity field (Smith and Sandwell 1997). The map clearly shows the network of FZs leading east to the MAR. This system of FZs is characterized by a series of canyons bounded latitudinally by crests that rise up to 1 km above the canyon floors. The canyon of an unnamed FZ between 21° and 22°S was heavily sampled during the 1997 survey. Observations from this canyon will receive careful attention in section 4c.
A summary of the observed deep-dissipation data was made by collapsing the latitudinal spread of the stations (Fig. 2). To simplify the presentation, only the depth range deeper than 2000 m is shown. The dissipation data were depth averaged into 100-m intervals, and individual profiles are shown as columns of binned dissipation. Since the latitudinal extent of the survey has been collapsed, no particular section of bathymetry is entirely appropriate. Instead, two classes of representative bathymetry are shown: the bathymetry of FZ crests and the bathymetry of canyon floors, as sampled from the latitude range 21°–23°S.
Within about 500 m of the bottom, there is an enhancement of dissipation by up to two orders of magnitude over values observed at middepth. Enhanced dissipation generally occurs between the level of FZ crests and canyon floors. Thus, the largest mixing rates are observed in canyons.
a. Spatial variability
The spatial patterns of enhanced dissipation apparent in Fig. 2 suggest that variability of dissipation is associated with the different sloping regions that characterize the crest–canyon system of the FZs. This also suggests that height above bottom (hab) may serve as a better coordinate than depth for examining the spatial distribution of ϵ. This approach was adopted, and the available profiles of ϵ were classified according to the meridional slope of the local bathymetry Zb(x, y). HRP stations were categorized as being FZ crests (∂yZb ≅ 0, ∂yyZb < 0), canyon floors (∂yZb ≅ 0, ∂yyZb > 0), or sloping regions (∂yZb ≠ 0). The profile data, classified in this manner, are shown as functions of hab in Fig. 3. Each ensemble profile results from data that has been vertically averaged into 100-m bins. Auto-lag-correlation analysis of ϵ profiles shows zero crossings at 20-m vertical separations, so each 100-m bin contains five degrees of freedom. The 95% confidence interval of each bin was calculated using the “boot strap” Monte Carlo method (Efron and Gong 1983). Dissipation decreases with height in each of the three classes, and all profiles reach background levels of dissipation, ϵ ≅ 3 × 10−10 W kg−1, at heights above bottom greater than 1000 m. The slope profile is notably different from the other two, in both the magnitude of the maximum dissipation and the scale of decay. The slope profile has a bottom value of ϵ ≅ (3–9) × 10−9 W kg−1 and decays with an e-fold scale of (150 ± 50) m. The crest and canyon profiles peak at ϵ ≅ (2–5) × 10−9 W kg−1 and decay over a larger scale, (500 ± 100) m.
b. Temporal variability
In addition to this spatial structure, temporal variability also exists in the dissipation data. Daily averages of dissipation data roughly modulate with the spring–neap cycle of tidal currents. This correlation is most evident when the dissipation profiles are vertically integrated, and a comparison between integrated ϵ and tidal current speed is made using estimates of barotropic tidal velocity produced by the TPXO model of Egbert et al. (1994) (Fig. 4). For the time intervals where dissipation data were available, the TPXO model was used to produce velocity estimates at the latitude and longitude appropriate for each HRP station.
In both 1996 and 1997 comparisons, profiles of ρϵ were integrated up from the local bottom to hab = 2000 m, and all profiles were time binned in 24-h intervals and averaged. The modulation of integrated dissipation follows the phase of the spring–neap tidal cycle closely, and a one day lead of the tidal signal is consistent within the statistical spread of the data. While the correlation between the two records is quite low for the 1996 data, near 0.3, the correlation coefficient for the 1997 data is 0.7. Furthermore, this comparison uses only the tidal record reconstructed along the survey track, though the dissipation is possibly influenced by tidal beams that have radiated from other sites. We regard the temporal fluctuations in the dissipation record as suggestive of tidal forcing, and explore this possibility further in section 3b.
3. Inverse model
a. Model equations
We focus here on the ocean abyss and make the a priori assumption that the diabatic terms in the governing equations are significant. All dynamical equations will be cast in terms of flow along and across isopycnals. The dynamically relevant density variable for this analysis is neutral density (γn: Jackett and McDougall 1997): the neutral density coordinates are everywhere parallel to the locally defined potential density. Preliminary examination of multiple-year hydrographic data revealed that the temporal variation of the isopycnal-height field was negligible (Ledwell et al. 2000). Therefore, we will assume that the density field can be regarded as steady over the 14-month period spanned by the observations.
In expressions (10) and (11),
b. Model for the turbulent dissipation
Spatial and temporal variations in the observed dissipation data were described in sections 2a and 2b. These variations are not random, and simple arithmetic averaging may alias modes of variability in a manner that will produce biased estimates of mean dissipation. Estimates of spatially averaged dissipation are sensitive to the layout of bathymetric crests, slopes, and canyons in that ϵ levels above slopes can exceed levels above crests and canyon floors. Similarly, estimates of time-averaged dissipation are sensitive to the phase of the spring–neap tidal cycle, as ϵ levels are maximum near spring tide and minimal near neap tide. Due to the logistics of the tracer component of the experiment, canyon sites were favored during data collection, and three of five weeks of sampling occurred during neap tides. Thus, simple averages of the observed dissipation data are biased low relative to the true temporal and spatial mean. We therefore rely on simple modeling to recover an unbiased estimate of mean dissipation.
First, we consider the fortnightly modulation of the dissipation with the barotropic tidal current. Traditionally, frictional boundary-layer dynamics are assumed as the mechanism for tidally driven dissipation. The energy dissipated in the frictional boundary layer is estimated as ϵ = ρ0cd|u|u2, where cd ≃ 0.0025 is the drag coefficient and u is the barotropic velocity of flow. Taking u = 0.02 m s−1, the tidal dissipation in the frictional boundary layer is Wf ≃ 0.02 mW m−2. This is two orders of magnitude smaller than the integrated dissipation occurring between the bottom and a height of 2000 m. It is further noted that HRP profiles seldom extend close enough to the bottom to sample the frictional boundary layer, which is typically 1 to 10 m thick.
In contrast to bottom friction, the internal waves generated by tidal flow over rough bathymetry radiate energy upward, away from the bottom. Instabilities in the internal wave field act to transfer some of this energy to the small scales where it is dissipated by viscosity. Thus, internal waves provide an additional, though indirect, mechanism by which tidal energy is dissipated. The internal wave generation problem of flow over bottom bathymetry is well known, and a general discussion of steady flow “lee wave” generation is given by Gill (1982). Internal waves produced by tidal flow are termed “internal tides,” and unlike lee waves, internal tides have intrinsic frequencies equal to those of the barotropic tides. Internal tide generation models are discussed by Baines (1982), who shows that the energy flux associated with the internal tide radiated from linear slope topography scales as Ef ≃ Kρ0Nu2h2, where N is the buoyancy frequency, u is the barotropic tidal velocity, h is the depth, and K is a parameter that depends on the height of the topography relative to the water depth and the geometry of the slope. This scaling is valid for linear slopes of arbitrary steepness, though the particular value of K depends on whether the slope is supercritical (slope steeper than the beam of radiated waves) or subcritical (slope less steep than waves). For the case of subcritical topography, Bell (1975) shows that the energy flux for topography of arbitrary shape scales as Ef ≃ kρNu2d2, where (k, d) are the wavenumber and height of a topographic obstacle. Using this scaling as a lower bound, an energy flux of Ef ≃ 1–10 mW m−2 is likely to characterize the internal tide generated across the FZs. This value for energy flux is the same scale as the observed integrated dissipation values to a height of 2000 m above the topography.
c. Treatment of data
The dynamics described in section 3a can be employed in an inverse model where the observed density field (h(x, y), N2), geopotential-anomaly field φ(x, y, p), temperature field Θ(x, y, z), and modeled dissipation
The 1997 HRP survey data provided the hydrographic fields. The temperature and salinity data were first used to calculate isobaric maps of geopotential anomaly relative to 2000 dbar. These fields were modeled using a polynomial fit over the survey domain with the form φ(X, Y) = ϕ0 + ϕxX + ϕyY + ϕxxX2 + ϕxyXY + ϕyyY2, where (X, Y) are Cartesian coordinates and ϕ( · ) are the coefficients of the fit. The second-order polynomial fit was found to filter tidal displacements in the density field (St. Laurent 1999). The gradient of the fit is used with (6) to give estimates of the geostrophic flow relative to 2000 m. This reference level was chosen based on ideas about basin-scale water masses, with 2000 m being roughly between northward flowing intermediate water and southward flowing North Atlantic Deep Water (NADW). The hydrographic data were sorted into layers of neutral density, and polynomial fits were used to give quasi-analytic expressions for Θ(x, y), Θz(x, y), h(x, y), and N2(x, y) in each layer. This layered system comprises the static stratification field, which is used for the inversion. Vertical derivatives for N2 and Θz and their associated uncertainty were calculated using the slopes from linear fits to the data. For each density layer, the linear fits were computed about the level of the bounding isopycnal over a depth interval comparable to the density layer thickness. Assignments of geostrophic shear for each layer were made through explicit reference to the isobaric surfaces used in the geopotential-anomaly calculation. Thus, the inverse domain is a layered system of hydrographic estimates on a staggered grid of nested cells. Various domain dimensions were explored, but the calculations discussed here cover 19°–25°S, 14°–22°W and six density layers, with γn = 27.6 kg m−3 (density units are suppressed hereafter) serving as the uppermost interface and the seafloor bathymetry serving as the lower boundary of the inverse domain. Each cell of the staggered grid has Mercator dimensions of (0.5°)2, and the altimetry-derived bathymetry data of Smith and Sandwell (1997) were subsampled onto this grid. The modeled dissipation data (12) were incorporated into the inverse domain by identifying each estimation point as a bathymetric crest, slope, or canyon. This was done through a classification scheme that utilizes both the first and second meridional derivatives of bathymetry, which were computed at two arc minute resolution.
Residuals between the fitted and observed density field are typically 20–50 m, assumed to be the result of tidal heaving, and an assessment of hydrographic noise was done using Monte Carlo methods. In particular, the influence of 50-m amplitude vertical displacements of the isopycnal-height field arising from tidal frequency motions was examined. To accomplish this, the observed isopycnal-height field was perturbed by a random field of Gaussian noise with 50-m standard deviation. The perturbed density field was then treated with the polynomial smoothing discussed above, and this smoothed but perturbed hydrography was used for the inversion. In this manner, 10 alternate realizations of the density field were produced, and an ensemble of inverse solutions was calculated.
Since the survey data were concentrated above several FZs along an irregular grid (Fig. 1), some regions of the inverse domain correspond to regions that were sparsely sampled. There is no survey data in the northeast and southeast corners of the 19°–25°S, 14°–22°W inverse domain, so the mapped properties in these regions are poorly constrained. The inverse estimates in these sparsely sampled regions will receive minimal attention.
d. Linear algebraic equations
The fitted hydrographic fields and the dissipation data were used to assemble a linear system of coupled equations of the form Eijxj + ni = di, where Eij is the coefficient matrix of the terms in (7), (8), and (9) involving the unknowns xj = [u0, υ0, w∗, κ]T, plus the estimates w∗ ≅ N−2Γ0∂
4. Results
The matrix model for the six-layer system consisted of about 7000 equations for 3200 unknowns. In a series of initial calculations, lateral diffusivity κ was treated as constant along each density surface used in the inversion. Using this approach, the estimates of κ were inconsistent with the a priori estimates of κ ≃ 100 m2 s−1. Spatially variations of κ on isopycnals intersecting topography may account for this discrepancy. Additionally, diagnostic scaling with κ ≃ 100 m2 s−1 indicated that the primary thermal-advection balance in (9) occurs between the terms involving advection (both lateral and vertical) and the vertical diffusion term. For these reasons, terms involving the lateral diffusivity were dropped from the thermal-advection equations, leaving the three-dimensional flow field (u, w∗) as the model unknowns. The singular value decomposition (SVD) of the scaled matrix
a. Shallow flow
Diapycnal mixing in the layers shallower than 3000 m was weak and diffusivity estimates were kρ = (Γ/N2)
Within the domain treated by the inversion, the bottom topography does not shoal less than 3000 m with the exception of a seamount in the southeast corner of the region. Thus, geostrophic dynamics were assumed valid everywhere in layers shallower than 3 km, and the combined use of the beta spiral with mass and temperature advection yielded an overdetermined system for the barotropic flow at each grid point. The estimated geostrophic flow fields for levels shallower than 3000 m, including the reference level, are shown in Fig. 5. The largest current speeds in these layers are 2 mm s−1, and the mean speeds are around 0.5 mm s−1. The flows in these layers are generally consistent with ideas about water mass circulation. Above and below a level of very weak flow (γn = 27.83, z ≅ 2200 m), the intermediate water layer (γn = 28.01, z ≅ 1400 m) flows in a northward direction, while the NADW layer (γn = 28.06, z ≅ 2900 m) flows in a southward direction. The local nature of these flow domains makes detailed comparisons with previous results difficult, as most past work has considered basin scales. Durrieu De Madron and Weatherly (1994) describe the Brazil Basin circulation of NADW and AABW water masses, but their observations do not cover the region considered in our study. Reid (1989) estimated net northward flow at z = 2000 m near 20°S, 20°W in his study of large-scale South Atlantic circulation. Also, Speer et al. (1995) describe a section at 22°S, 15°W with net eastward transport in the depth range between 1300 and 3200 m as part of the Namib Col Current, a feature not apparent in our data. However, the flows found here have qualitative similarity to flow estimates discussed by Zhang and Hogg (1992).
In Fig. 5, an ensemble of 10 velocity estimates is plotted at each grid point. The spread of vectors at each point is the result of uncertainty in the determination of the reference-level flow due to ±50 m noise in the isopycnal-height field. In many cases, the estimated variations are small and not apparent in the figures. The flow estimates are most sensitive in the northeast and southeast corners of the domain, where no stations were occupied. This sensitivity amounts to an uncertainty in flow direction not greater than π/12 radians. Statistical uncertainty in the magnitude of the flow field is typically 0.1–0.2 mm s−1, as assessed from standard error propagation techniques used with the polynomial fits to the geopotential-anomaly fields.
b. Deep flow
In the layers deeper than 3000 m, bottom bathymetry has a clear influence on the flow dynamics. In particular, proximity to the bottom is the controlling factor for the magnitude of the diabatic forcing. The vertical diffusivity characterizing these layers is kρ ≃ 1 × 10−4 m2 s−1 or greater, and this corresponds to an increased importance of the diabatic terms in the governing equations. The estimated lateral flow fields, with associated vertical diffusivities, are shown in Figs. 6, 7, and 8. Flow in these layers is generally southward and westward, with average current magnitudes of 2–4 mm s−1 and the faster currents occurring in the western portion of the region exceeding 5 mm s−1. The influence of ±50 m noise in the isopycnal-height field has also been evaluated, and the lateral flow is very stable to perturbations of this magnitude.
The neutral density range 28.12 γn < 28.18 (45.88 < σ4 < 45.96) is of particular interest, because this density layer bounds the target isopycnal used for the tracer release, γn = 28.16 (σ4 = 45.94). The tracer inventory made during the 1997 survey indicated that the tracer's center of mass was advected 130 km southwest over a period of 14 months, as reported by Ledwell et al. (2000). This is equivalent to a time average speed of 3.6 mm s−1. The average velocity in the density layer 28.12 < γn < 28.18, as estimated by the inversion for the region 22°–25°S, 18.5°–22°W, is |u| = (3 ± 0.3) mm s−1 in the direction 220°–230°. Thus, the lateral flow estimate given by the inversion is consistent with the rate of lateral advection measured by the tracer. The calculated flow in these layers is also consistent with float observations made in this region during 1994–96, as reported by Hogg and Owens (1999). These estimates of flow are not consistent with the estimates of Reid (1989), who found northward flow near 20°S, 20°W at 4000 and 4500 m.
A detailed discussion of tracer- and microstructure-derived diffusivities for the target isopycnal is given by Ledwell et al. (2000). The tracer-derived result of
Diapycnal advection estimates were made on density surfaces, and these estimates are presented in Figs. 9, 10, and 11. The γn = 28.08 (σ4 = 45.83) surface is characterized by downward advection with mean w∗ = −3±1 m yr−1; larger downward pumping occurs in the east. The deeper surfaces, γn = 28.16 (σ4 = 45.94) and γn = 28.20 (σ4 = 45.98) are characterized by larger values of advection, |w∗| ≃ 10 m yr−1, with individual estimates having associated uncertainty of roughly 20%–40%. The γn = 28.16 surface was the target isopycnal for the tracer release experiment, and downward advection is estimated over the part of the domain occupied by the tracer patch. This is consistant with the 14-month tracer inventory, and a later 28-month inventory, both of which indicate that the tracer's center of mass sank relative to the target isopycnal (Ledwell et al. 2000).
Diapycnal advection through the eastern range of the tracer surface is upward, as it is for nearly all of the surfaces deeper than the γn = 28.20 surface. Figures 10 and 11 show contoured fields of w∗ that have been highly smoothed. Where ridge topography occupies each density layer there is no flow, and the contoured maps show estimates of w∗ that have been averaged in 1° × 1° cells, which often include regions blocked by topography. Estimates of upward advection in canyons exceed w∗ ≅ 30 m yr−1 and are as large as w∗ ≅ 100 m yr−1 in some local regions. In FZ canyons, eastward flow is estimated with magnitude 0.4 ± 0.2 mm s−1. It is this eastward canyon transport that supplies fluid for the diapycnal upwelling.
c. Mixing rates and flow in an abyssal canyon
A fundamental feature of the abyssal Brazil Basin is the system of crests and canyons associated with the FZs of the MAR. Many canyons run zonally for hundreds of kilometers, and each canyon is bounded to the north and south by crests extending 500 to 1000 m above the canyon floors. The meridional wavelength of the canyon–crest system is roughly 30–50 km, and because of this the flow in these FZs was difficult to resolve in the inverse model calculations. Instead, each canyon is better modeled locally so that the special geometry of the system can be considered.
This expression was evaluated using data from the section of HRP profiles shown in Fig. 12 for the density layer with γn > 28.20. The diffusivities along the γn = 28.20 isopycnal were estimated as the ensemble average kρ = 〈Γϵ(h = hσ)N−2〉 across the canyon width, down the length of the canyon. Error analyses for the kρ estimates were done using Monte Carlo methods incorporating an uncertainty of ±100 m in the bottom depth, along with the statistical uncertainty in
Net diapycnal upwelling is required across the γn = 28.20 surface between 18° and 16.5°W. From mass conservation (15), the average diapycnal advection across γn = 28.20 in the canyon is
d. Deep upwelling and mixing efficiency
As presented above, strong diapycnal upwelling is required to close the heat and mass budgets in the canyons. Figure 13 presents a comparison of the a priori estimates of w∗ to the inverse solutions for w∗ on the γn = 28.20 neutral surface. Inverse estimates are generally consistent with the a priori estimates of w∗ at sites where the γn = 28.20 surface is more than 300 m above the bottom. However, the inverse estimates suggest diapycnal upwelling is favored at sites where the γn = 28.20 surface is within 300 m of the bottom.
Diapycnal upwelling must be occurring in regions where turbulence supports a convergent buoyancy flux: w∗N2 ≅ −∂Jb/∂z > 0. This condition is satisfied by turbulence occurring close to the bottom since the buoyancy flux must decrease with depth to meet a no-flux condition at the seafloor. Within the context of (3), the buoyancy flux is related to the dissipation rate by a mixing efficiency parameter, Jb = −Γ
5. Discussion
This work has examined the buoyancy forcing and circulation occurring in a region of the abyssal ocean. Turbulent mixing is enhanced over the rough bathymetry near the MAR and its associated FZs, and this leads to enhanced diapycnal advection in these regions. Furthermore, the vortex stretching achieved by this diapycnal advection serves a significant role in the local balance of potential vorticity.
The signal of abyssal dissipation was found to have both spatial and temporal content. The spatial distribution of dissipation along an isopycnal is clearly related to bottom proximity. In the context of the complex bottom topography associated with FZs, dissipation profiles were found to have differing character for three classes of bathymetry. While dissipation generally increases toward the bottom, levels occurring over the flanks of sloping bathymetry exceed levels occurring on FZ crests and the floors of canyons. These spatial trends accompany a temporal modulation of dissipation, correlated with the spring–neap tidal cycle. An attempt was made to model these spatial and temporal trends, and the resulting model features a dependence with hab specific to the bathymetric classes of crests, canyons, and slopes. The temporal variability of dissipation was modeled according to a u2 energy flux scaling of linear internal waves.
The modeled dissipation profiles were employed in an inverse model that combines their thermodynamic information with the dynamic constraints implied by the beta spiral and hydrographic budgets. Specifically, the dissipation rate model was used to provide both a measure of the diapycnal diffusivity kρ and an a priori estimate of the diapycnal advection w∗. This approach proved highly effective in resolving both the isopycnal and diapycnal components of flow.
a. Local consequences of buoyancy forcing
The results of the inverse calculation suggest that, while diabatic flow is weak in much of the interior, stronger levels of turbulence above rough bathymetry lead to enhanced levels of diabatic flow. Divergence of the diabatic flow acts as a mechanism of buoyancy forcing for the lateral circulation though the vortex stretching term (f/h)(w∗(zu) − w∗(zl)) in the vorticity budget. The significance of diabatic vortex stretching as a forcing mechanism must be assessed relative to other forcing agents for the flow. From the inverse estimates, we may compare the strength of buoyancy forcing to the adiabatic mechanism of vortex stretching, represented by (f/h)u · ∇h in the vorticity budget.
In the deep density layers, adiabatic vortex stretching is cyclonic, while diabatic vortex stretching is anticyclonic. In the region where mixing rates are largest, these two mechanisms are comparable in magnitude, and the vortex stretching terms in (2) nearly cancel. Since βυ ≅ 0, there is little meridional flow just above the level of FZ crests (Figs. 6 and 7). In addition to the stretching mechanisms, friction imposed by topography may also be a significant forcing mechanism near the bottom, though we have not attempted to assess this contribution. Farther west of the Mid-Atlantic Ridge where mixing rates are weaker, the adiabatic stretching exceeds the diabatic stretching, resulting in the southward flow where the largest lateral velocities occur.
Spall (2001) has examined the planetary geostrophic dynamics for a circulation driven by buoyancy forcing over sloping topography. In his model, Spall represents only the large-scale zonal slope of FZ topography, with diapycnal upwelling from FZ canyons represented as flow out of the sloping bottom. Spall finds that bottom pressure torque tends to balance the diabatic vortex stretching in the deepest density layer, keeping the flow along the bottom weak even when the buoyancy forcing is strong.
The net transport through a density surface (Tσ) can be expressed in terms of the streamfunction as Tσ = ψ(xw) − ψ(xe), where (xe, xw) are the eastern and western limits of the surface area integral. The transport calculation is simplified if the eastern limit is taken to be the longitude where the surface incrops along the bottom, so that ψ(xe) = 0. For the γn = 28.20 (σ4 = 45.98 and θ ≅ 0.8°C) surface, the net diapycnal transport between 19° and 25°S is (0.34 ± 0.14) × 106 m3 s−1. This is the integrated upwelling over 6° of latitude; the actual upwelling occurs in the canyons. Assuming that an up-canyon flow of 3.5 mm s−1 is typical of canyons that are 500 m deep and 30 km wide, then six or seven canyons would carry the required mass for upwelling. This is consistent with the number of canyons in the region.
b. Consequences for the basin-scale mass budget
Maps of seafloor bathymetry show a network of 20–30 FZs occupying the entire meridional extent of the Brazil Basin. Additionally, the character of barotropic tidal energy does not vary significantly along the axis of the MAR. Treating these as proxies for turbulent dissipation, the spatial distribution of mixing depicted in Fig. 2 may be characteristic of the FZs throughout the greater basin. This would imply that buoyancy forcing plays an important role in driving the abyssal circulation. Indeed, O'Dwyer and Williams (1997) have produced basin-scale maps of q = f/h that show that the steamfunction for adiabatic flow in the Brazil Basin is blocked by boundaries at depths greater than 3500 m. Flow in regions of blocked q contours occurs only under the action of buoyancy forcing and friction.
Some basin-scale consequences of abyssal buoyancy forcing are examined by considering the mass budget for water of temperature class θ < 0.8°C in the Brazil Basin. This budget was considered by Hogg et al. (1982) and Morris et al. (2001). Morris et al. (2001) find that (3.7 ± 1.7) × 106 m3 s−1 of net upwelling occurs across the θ = 0.8°C isotherm surface in the latitude range between 30°S and the equator. If the circulation shown in Fig. 14 is characteristic of other FZ regions, a simple scaling of our localized estimate (6° of latitude) for net upwelling to the basin scale (30° of latitude) suggests an upwelling across θ = 0.8°C isotherm of (1.7 ± 0.7) × 106 m3 s−1. This number is in the general range needed to close the basin-scale mass budget.
The various regions of bottom water circulation in the Brazil Basin are shown in Fig. 15. Bottom water enters the basin through the Vema Channel, and a boundary current carries this deep water northward along the western portion of the basin (Reid 1989; Speer and Zenk 1993; Durrieu De Madron and Weatherly 1994). However, over much of the zonal band between 35° and 25°W, FZs are absent and mixing levels and isopycnal slopes are likely to be small, so the mean flow should satisfy βυ ≅ 0. Indeed, float observations in this region suggest that the flow is dominantly zonal (Hogg and Owens 1999). Farther to the east, mixing levels and isopycnal slopes increase above FZ topography. In this region, our inverse results serve as an example, and diapycnal advection will influence the lateral flow through the vorticity relation βυ = (f/h)u · ∇h + f(∂w∗/∂z). Net diapycnal upwelling of water masses colder than θ ≅ 0.8°C likely occurs in FZ canyons. Inversion of observations from January 1996 to April 1997 indicate a net southward transport of −(0.34 ± 0.05) × 106 m3 s−1 for the water class denser than γn > 28.08 in the boxed region (Fig. 15), though recent float observations suggest that meridional flows in this region are time varying (N. Hogg 2001, personal communication).
The work described here demonstrates that direct measures of mixing can be combined with hydrographic observations to give an estimate for circulation that is consistent with other large-scale budgets. This is in contrast to many earlier attempts to reconcile local dissipation estimates with mixing rates inferred from advective–diffusive budgets. A notable difference between the current approach and early treatments of dissipation data rests with the application of an inverse technique with both microstructure and conventional hydrographic data. While direct observations of mixing tend to give detailed information on local diapycnal exchange processes, hydrographic budgets predict only the mean of diabatic forcing over large scales. Neither approach alone would give a meaningful measure of the full three-dimensional circulation.
Buoyancy forcing provided by turbulent mixing likely provides the dominant forcing for the abyssal circulation, and the highly localized observations described here constitute the only available observations of deep dissipation. The localized measurements presented here suggest that the spatial structure of abyssal buoyancy forcing is far richer than Stommel and Aron's example of uniform upwelling.
Acknowledgments
We thank Kurt Polzin and Jim Ledwell for helpful discussions during the preparation of this manuscript. The authors also thank Ellyn Montgomery, Dave Wellwood, and Tom Bolmer, and the officers and crew of the R/V Seward Johnson for support of our field program. Comments from Chris Garrett, Nelson Hogg, Trevor McDougall, Mike Spall, and two anonymous reviewers improved this presentation. This work was supported by Grant OCE94-15589 of the National Science Foundation. The support of the Office of Naval Research is also acknowledged.
REFERENCES
Baines, P. G., 1982: On internal tide generation models. Deep-Sea Res, 29 , 307–338.
Barton, E. D., and A. E. Hill, 1989: Abyssal flow through the Amirante Trench (Western Indian Ocean). Deep-Sea Res, 36 , 1121–1126.
Bell, T. H., 1975: Lee waves in stratified flows with simple harmonic time dependence. J. Fluid Mech, 67 , 705–722.
Bevington, P. R., and D. K. Robinson, 1992: Data Reduction and Error Analysis for the Physical Sciences. McGraw-Hill, 328 pp.
Bretherton, F. P., R. E. Davis, and C. B. Fandry, 1976: A technique for objective analysis and design of oceanographic experiments applied to MODE-73. Deep-Sea Res, 23 , 559–582.
Durrieu De Madron, X., and G. Weatherly, 1994: Circulation, transport and bottom boundary layers of the deep currents in the Brazil Basin. J. Mar. Res, 52 , 583–638.
Efron, B., and G. Gong, 1983: A leisurely look at the bootstrap, the jackknife, and cross-validation. Amer. Stat, 37 , 36–48.
Egbert, G. D., A. F. Bennett, and M. G. G. Foreman, 1994: TOPEX/POSEIDON tides estimated using a global inverse model. J. Geophys. Res, 99 , 24821–24852.
Fukumori, I., 1991: Circulation about the Mediterranean Tongue: An analysis of an EOF-based model ocean. Progress in Oceanography, Vol. 27, Pergamon, 197–224.
Garrett, C., 2001: An isopycnal view of near-boundary mixing and associated flows. J. Phys. Oceanogr, 31 , 138–142.
Gill, A. E., 1982: Atmosphere–Ocean Dynamics. Academic Press, 662 pp.
Hautala, S. L., and S. C. Riser, 1993: A nonconservative β-spiral determination of the deep circulation in the Eastern South Pacific. J. Phys. Oceanogr, 23 , 1975–2000.
Hogg, N. G., and W. B. Owens, 1999: Direct measurements of the deep circulation within the Brazil Basin. Deep-Sea Res. II, 46 , 335–353.
Hogg, N. G., P. Biscaye, E. Gardner, and W. J. Schmitz, 1982: On the transport of Antarctic Bottom Water in the Verma Channel. J. Mar. Res, 40 , (Suppl),. 231–263.
Huq, P., and R. E. Britter, 1995: Turbulence evolution and mixing in a two layer stably stratified fluid. J. Fluid Mech, 285 , 41–67.
Ivey, G. N., and J. Imberger, 1991: On the nature of turbulence in a stratified fluid. Part I: The energetics of mixing. J. Phys. Oceanogr, 21 , 650–658.
Jackett, D. R., and T. J. McDougall, 1997: A neutral density variable for the world's oceans. J. Phys. Oceanogr, 27 , 237–263.
Kunze, E., and T. B. Sanford, 1996: Abyssal mixing: Where it is not. J. Phys. Oceanogr, 26 , 2286–2296.
Ledwell, J. R., E. T. Montgomery, K. L. Polzin, L. C. St. Laurent, R. W. Schmitt, and J. M. Toole, 2000: Mixing over rough topography in the Brazil Basin. Nature, 403 , 179–182.
Menke, W., 1989: Geophysical Data Analysis: Discrete Inverse Theory. Academic Press, 285 pp.
McDougall, T. J., 1987: Thermobaricity, cabbeling, and water-mass conversion. J. Geophys. Res, 92 , 5448–5464.
McDougall, T. J., 1988: Neutral-surface potential vorticity. Progress in Oceanography, Vol. 20, Pergamon, 185–221.
McDougall, T. J., 1995: The influence of ocean mixing on the absolute velocity vector. J. Phys. Oceanogr, 25 , 705–725.
Morris, M., M. M. Hall, L. C. St. Laurent, and N. Hogg, 2001: Abyssal mixing in the Brazil Basin. J. Phys. Oceanogr.. 31 , 3331–3348.
Moum, J. N., 1996: Efficiency of mixing in the main thermocline. J. Geophys. Res, 101 , 12057–12069.
O'Dwyer, J., and R. G. Williams, 1997: The climatological distribution of potential vorticity over the abyssal ocean. J. Phys. Oceanogr, 27 , 2488–2506.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr, 10 , 83–89.
Pedlosky, J., 1996: Ocean Circulation Theory. Springer-Verlag, 453 pp.
Polzin, K. L., and E. T. Montgomery, 1996: Microstructure profiling with the high resolution profiler. Proc. ONR Workshop on Microstructure Sensors,. Mt. Hood, OR, Office of Naval Research, 109–115.
Polzin, K. L., J. M. Toole, J. R. Ledwell, and R. W. Schmitt, 1997: Spatial variability of turbulent mixing in the abyssal ocean. Science, 276 , 93–96.
Reid, J. L., 1989: On the total geostrophic circulation of the South Atlantic Ocean: Flow patterns, tracers, and transports. Progress in Oceanography, Vol. 23, Pergamon, 149–244.
Roemmich, D., S. Hautala, and D. Rudnick, 1996: Northward abyssal transport through the Samoan Passage and adjacent regions. J. Geophys Res, 101 , 14039–14055.
Saunders, P. M., 1987: Flow though Discovery Gap. J. Phys. Oceanogr, 17 , 631–643.
Schmitt, R. W., J. M. Toole, R. L. Koehler, E. C. Mellinger, and K. W. Doherty, 1988: The development of a fine- and microstructure profiler. J. Atmos. Oceanic Technol, 5 , 484–500.
Smith, W. H. F., and D. T. Sandwell, 1997: Global sea floor topography from satellite altimetry and ship depth soundings. Science, 277 , 1956–1962.
Spall, M. A., 2001: Large-scale circulations forced by localized mixing over a sloping bottom. J. Phys. Oceanogr.. 31 , 2369–2384.
Speer, K. G., and W. Zenk, 1993: The flow of Antarctic Bottom Water into the Brazil Basin. J. Phys. Oceanogr, 23 , 2667–2682.
Speer, K. G., G. Siedler, and L. Talley, 1995: The Namib Col Current. Deep-Sea Res, 42 , 1933–1950.
Stigebrandt, A., and J. Aure, 1989: Vertical mixing in basin waters of fjords. J. Phys. Oceanogr, 19 , 917–926.
St. Laurent, L., 1999: Diapycnal advection by double diffusion and turbulence in the ocean. Ph. D. dissertation, MIT–WHOI Joint Program in Oceanography, 139 pp.
St. Laurent, L., and R. W. Schmitt, 1999: The contribution of salt fingers to vertical mixing in the North Atlantic Tracer Release Experiment. J. Phys. Oceanogr, 29 , 1404–1424.
Stommel, H., 1957: The abyssal circulation of the ocean. Nature, 180 , 733–734.
Stommel, H., and A. B. Arons, 1960: On the abyssal circulation of the World Ocean—I: Stationary planetary flow patterns on a sphere. Deep-Sea Res, 6 , 140–154.
Stommel, H., and F. Schott, 1977: The beta spiral and the determination of the absolute velocity field from hydrographic station data. Deep-Sea Res, 24 , 325–329.
Stommel, H., A. B. Arons, and A. J. Faller, 1958: Some examples of stationary planetary flow patterns in bounded basins. Tellus, 10 , 1–11.
Toole, J. M., K. L. Polzin, and R. W. Schmitt, 1994: Estimates of diapycnal mixing in the abyssal ocean. Science, 264 , 1120–1123.
Toole, J. M., R. W. Schmitt, K. L. Polzin, and E. Kunze, 1997a: Near-boundary mixing above the flanks of a midlatitude seamount. J. Geophys. Res, 102 , 947–959.
Toole, J. M., J. R. Ledwell, K. L. Polzin, R. W. Schmitt, E. T. Montgomery, L. St. Laurent, and W. B. Owens, 1997b: The Brazil basin Tracer Release Experiment. Int. WOCE Newslett, 28 , 25–28.
Whitehead, J. A., and L. V. Worthington, 1982: The flux and mixing rates of Antarctic Bottom Water within the North Atlantic. J. Geophys. Res, 87 , 7903–7924.
Wunsch, C., 1996: The Ocean Circulation Inverse Problem. Cambridge University Press, 442 pp.
Yamazaki, H., and T. R. Osborn, 1990: Dissipation estimates for stratified turbulence. J. Geophys. Res, 95 , 9739–9744.
Zhang, H-M., and N. G. Hogg, 1992: Circulation and water mass balance in the Brazil Basin. J. Mar. Res, 50 , 385–420.