Water Mass Transformation in the Southern Ocean of a Global Isopycnal Coordinate GCM

Robert Marsh James Rennell Division for Ocean Circulation and Climate, Southampton Oceanography Centre, Southampton, United Kingdom

Search for other papers by Robert Marsh in
Current site
Google Scholar
PubMed
Close
,
A. J. George Nurser James Rennell Division for Ocean Circulation and Climate, Southampton Oceanography Centre, Southampton, United Kingdom

Search for other papers by A. J. George Nurser in
Current site
Google Scholar
PubMed
Close
,
Alex P. Megann James Rennell Division for Ocean Circulation and Climate, Southampton Oceanography Centre, Southampton, United Kingdom

Search for other papers by Alex P. Megann in
Current site
Google Scholar
PubMed
Close
, and
Adrian L. New James Rennell Division for Ocean Circulation and Climate, Southampton Oceanography Centre, Southampton, United Kingdom

Search for other papers by Adrian L. New in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

A global isopycnal coordinate GCM is used to investigate the processes that drive the meridional circulation, transformation, and interocean exchange of water masses in the Southern Ocean. The noneddy-resolving model (mesh size 1.25°) includes an active mixed layer, parameterized bolus transport, and seasonally varying surface fluxes. The model gives a plausible picture of the formation and circulation of subantarctic mode water (SAMW) and Antarctic Intermediate Water (AAIW). Progressively denser versions of SAMW and AAIW form in the Indian and Pacific Oceans as the Antarctic Circumpolar Current drifts south and loses buoyancy.

SAMW forms predominantly in the Indian Ocean, at a rate of 20 Sv (Sv ≡ 106 m3 s−1), while AAIW forms mainly in the Pacific sector, at a rate of 8.5 Sv. Throughout the circumpolar zone 25°–42.5°S, there is a net formation of 11 Sv of SAMW, largely by surface cooling. This SAMW is exported northward across 25°S into the subtropical gyres. The properties, distribution, and recirculation of SAMW and AAIW compare well with observations. The authors differentiate the effects of surface fluxes and mixing in transforming water masses in two distinct circumpolar zones. South of 42.5°S, surface buoyancy gain (due to a slight dominance of freshening over cooling) and diapycnal mixing are shown to play a roughly equal role in lightening water (at a peak diapycnal flux of 9 Sv across σ = 27.3), and in forming AAIW.

The meridional overturning is computed as a function of density and decomposed. The parameterized bolus transport opposes the northward surface Ekman drift and southward deep geostrophic flow. Denser waters are not in steady state and the meridional overturning streamfunction gives a misleading impression of dense water transformation in the Southern Ocean. A “transformation streamfunction” is introduced that gives the correct (model) transformation rates; this is believed to be a powerful tool in diagnosing models that drift.

The implications for model North Atlantic Deep Water (NADW) are considerable. In the model, NADW is transported southward across 25°S in the Atlantic sector at a rate of 15.7 Sv. South of 25°S, NADW and Circumpolar Deep Water (CDW) are consumed by interior diapycnal mixing at a rate of 5.7 Sv. NADW and CDW are exported northward across 25°S in the Indo-Pacific sector at a rate of 19.5 Sv. The 9.5 Sv imbalance amounts to a steady loss of NADW and CDW from the Southern Ocean, highlighting the unsteadiness of dense water masses in the model.

Corresponding author address: Dr. Robert Marsh, James Rennell Division for Ocean Circulation, Southhampton Oceanography Centre, Empress Dock, Southhampton SO14 3ZH, United Kingdom.

Abstract

A global isopycnal coordinate GCM is used to investigate the processes that drive the meridional circulation, transformation, and interocean exchange of water masses in the Southern Ocean. The noneddy-resolving model (mesh size 1.25°) includes an active mixed layer, parameterized bolus transport, and seasonally varying surface fluxes. The model gives a plausible picture of the formation and circulation of subantarctic mode water (SAMW) and Antarctic Intermediate Water (AAIW). Progressively denser versions of SAMW and AAIW form in the Indian and Pacific Oceans as the Antarctic Circumpolar Current drifts south and loses buoyancy.

SAMW forms predominantly in the Indian Ocean, at a rate of 20 Sv (Sv ≡ 106 m3 s−1), while AAIW forms mainly in the Pacific sector, at a rate of 8.5 Sv. Throughout the circumpolar zone 25°–42.5°S, there is a net formation of 11 Sv of SAMW, largely by surface cooling. This SAMW is exported northward across 25°S into the subtropical gyres. The properties, distribution, and recirculation of SAMW and AAIW compare well with observations. The authors differentiate the effects of surface fluxes and mixing in transforming water masses in two distinct circumpolar zones. South of 42.5°S, surface buoyancy gain (due to a slight dominance of freshening over cooling) and diapycnal mixing are shown to play a roughly equal role in lightening water (at a peak diapycnal flux of 9 Sv across σ = 27.3), and in forming AAIW.

The meridional overturning is computed as a function of density and decomposed. The parameterized bolus transport opposes the northward surface Ekman drift and southward deep geostrophic flow. Denser waters are not in steady state and the meridional overturning streamfunction gives a misleading impression of dense water transformation in the Southern Ocean. A “transformation streamfunction” is introduced that gives the correct (model) transformation rates; this is believed to be a powerful tool in diagnosing models that drift.

The implications for model North Atlantic Deep Water (NADW) are considerable. In the model, NADW is transported southward across 25°S in the Atlantic sector at a rate of 15.7 Sv. South of 25°S, NADW and Circumpolar Deep Water (CDW) are consumed by interior diapycnal mixing at a rate of 5.7 Sv. NADW and CDW are exported northward across 25°S in the Indo-Pacific sector at a rate of 19.5 Sv. The 9.5 Sv imbalance amounts to a steady loss of NADW and CDW from the Southern Ocean, highlighting the unsteadiness of dense water masses in the model.

Corresponding author address: Dr. Robert Marsh, James Rennell Division for Ocean Circulation, Southhampton Oceanography Centre, Empress Dock, Southhampton SO14 3ZH, United Kingdom.

1. Introduction

Antarctic Intermediate Water (AAIW) ventilates vast areas of the lower thermocline in the World Ocean (Schmitz 1996a), particularly the South Atlantic and South Pacific. AAIW is a significant component of intermediate waters even in the North Atlantic (Schmitz 1996b). Lighter subantarctic mode waters (SAMW) form much of the upper thermocline in the Southern Ocean (McCartney 1977, 1982; Piola and Georgi 1982).

Owing to their vast extent, SAMW and AAIW may play an important role in global climate change under increasing concentrations of atmospheric CO2. For instance, Stocker and Broecker (1994) predict that 30% of excess anthropogenic CO2 sequestration occurs in the Southern Ocean.

These water masses are thought to be formed at sites along the Antarctic Circumpolar Current (ACC), as a result of winter cooling. AAIW forms during winter convection in the subantarctic zone of the southeast Pacific and Scotia Sea (McCartney 1977), while SAMW forms likewise at sites in the southeast Indian and midsoutheast Pacific Oceans (McCartney 1982). It is unclear how important is the northward drifting Antarctic surface water originating from south of the ACC (Sverdrup et al. 1942); England et al. (1993) argue that it is unimportant, even as a source of AAIW, while Ribbe and Tomczak (1997) examined SAMW formation in the Fine Resolution Antarctic Model (FRAM, an “eddy permitting” model: FRAM Group 1991) and concluded that both circumpolar cross-frontal mixing and local deep mixing play a role, an idea further supported by the modeling study of Ribbe (1997) and recent observational evidence (Rintoul and Bullister 1999).

Although most of the AAIW simply recirculates in the subtropical gyres of the Southern Oceans, a fraction warms and lightens within those gyres. Some AAIW eventually enters the North Atlantic and is, indeed, a component of the waters that cool to form North Atlantic Deep Water (NADW) in the northern North Atlantic.

Hence the formation and recirculation of SAMW and AAIW represents part of the “upper branch” of the global thermohaline circulation (Schmitz 1996a), by which deep waters are warmed and lightened as they return to the region of cooling in the northern North Atlantic. The lower branch of the thermohaline circulation involves deeper, denser water masses. There is a net flow of some 14–15 Sv (Sv ≡ 106 m3 s−1) of deep waters through the South Atlantic into the Southern Ocean (Saunders and King 1995), made up of ∼22 Sv of southward flowing NADW less ∼7 Sv of northward flowing Antarctic Bottom Water (AABW). The NADW then mixes with waters of Antarctic origin to form Circumpolar Deep Water (CDW) (Merz and Wüst 1922), which is the dominant water mass in the deep Southern Ocean.

It is unclear exactly how these deep waters lighten and enter the upper (return) branch of the thermohaline circulation. Classically, this is achieved by diapycnic mixing diffusing down heat in the Indian and Pacific Oceans (Stommel and Arons 1960). However, there may also be considerable lightening of the deep waters within the Southern Ocean, where NADW upwells close to the surface and is exposed to air–sea fluxes and enhanced mixing within, and immediately below, the mixed layer. Such lightening of NADW has indeed been found in an eddy-resolving global model (Döös and Coward 1997).

The relative importance of these two processes remains unclear. However a plausible circulation scheme is that proposed by Schmitz (1996a). He suggests that the CDW first flows into the Pacific Ocean where it is lightened slightly to form Upper Pacific Deep Water (UPDW), which then returns to the ACC as Upper Circumpolar Deep Water (UCDW). Of this, he suggests that 5 Sv upwells south of the ACC and lightens across the σ = 27.6 isopycnal to form AAIW, while a remaining 9 Sv lighten across the σ = 27.6 isopycnal in the Indian Ocean as a result of diffusive warming. Some of the resulting intermediate waters are then warmed further in the equatorial Pacific and eventually return as surface waters to the Atlantic [the “warm water path” through the Agulhas (Gordon 1986)], while other intermediate waters return directly to the Atlantic through Drake Passage [the “cold water path” (Rintoul 1991)].

Hence water mass transformation in the Southern Ocean is of great importance, both in terms of the formation of intermediate (AAIW) and thermocline (SAMW) waters, and as a possible contribution toward the lightening component of the global thermohaline circulation.

Unfortunately, the surface fluxes in this region are poorly known (Josey et al. 1999), as is the mixing, so this transformation is difficult to model. However, such modeling, even using inaccurate surface fluxes and models with imperfect physics (particularly mixing), is necessary in order to understand the processes that regulate this transformation.

Surface fluxes and mixing are sensitive to the circulation, particularly the meridional circulation: generally, poleward-moving fluid will cool, equatorward-moving fluid will warm. The meridional circulation in the Southern Ocean is not well known. However, wind stress climatologies (e.g., Hellerman and Rosenstein 1983) predict a northward Ekman drift over the ACC, which reaches ∼30 Sv at around 45°S. The ACC itself exhibits substantial meridional shifts (Gordon et al. 1982; Olbers et al. 1992), drifting southward across the Atlantic, Indian, and Pacific sectors and returning northward as a cold western boundary current (WBC) along the eastern coast of South America (the Falkland Current). North of the ACC, southward-flowing WBCs (the Brazil, Agulhas, and East Australian Currents) transport warm water from the subtropical gyres to midlatitudes. This warm water flows along the north of the ACC in each sector, ultimately recirculating northward into the gyre interiors.

If the time-mean velocity fields of OGCMs are zonally averaged on depth levels, a striking meridional circulation is obtained in the Southern Ocean [at coarse resolution (e.g., Manabe et al. 1990; Danabasoglu et al. 1994) or fine resolution (e.g., Döös and Webb 1994)]. A strong sinking of surface waters to depths of 1000–3000 m in the zone 30°–45°S, deep southward flow, and near-vertical upwelling south of ∼60°S form a closed circulation known as the Deacon cell (Bryan 1991), seemingly driven entirely by the midlatitude northward Ekman drift. Superposition of this zonal-mean Deacon cell on zonally averaged isopycnals associates strong buoyancy gain and loss with upwelling and downwelling limbs of the Deacon cell. These buoyancy forcings are, in fact, spurious. The actual time-mean, zonally averaged meridional transport on surfaces of constant density, which is linked to the buoyancy forcing, includes also (i) a term due to correlations in time of isopycnal layer thicknesses and velocities due to transient eddies (Gent and McWilliams 1990)—the “bolus transport”—and (ii) correlated zonal variations in isopycnal layer thicknesses and velocities (Döös and Webb 1994; McIntosh and McDougall 1996). Thus the actual buoyancy forcing is decoupled from the Eulerian-mean flow—the Ekman drift.

In the present study of SAMW and AAIW, we use the Miami Isopycnal Coordinate Ocean Model (MICOM: Bleck et al. 1989; Bleck and Smith 1990; Smith et al. 1990; Bleck et al. 1992), a primitive equation layered OGCM in which σ0 is the vertical (layer) coordinate [σ0 = (ρ/ρ* − 1) × 103; where ρ is the potential density at atmospheric pressure and ρ* = 1000 kg m−3]. Such a model, explicitly incorporating advection along isopycnal surfaces, does not give the spurious mixing, due to numerical and explicit diffusion, often seen in level models.

We hence define water masses by σ0. Table 1 summarizes the σ0 ranges for selected water masses of the Southern Ocean (after Schmitz 1996a). In so far as SAMW and AAIW can be thus defined, their respective σ0 ranges probably overlap, as many varieties (with widely varying σ0) have been identified (Georgi 1979;McCartney 1982; Piola and Georgi 1982). Due to compressibility effects, the deep water masses (CDW, NADW, AABW) are not well defined by σ0. In particular, the present model will fail to correctly represent AABW, which lies beneath NADW in the real world, despite the fact that AABW has a lower value of σ0 than NADW. However, in the present study we focus on the transformation of thermocline water masses, for which σ0 is an adequate vertical coordinate, and in so far as we consider the upwelling and mixing of NADW, the thermohaline circulation is reasonably well represented with σ0. We diagnose the circulation and transformation of model water masses in the Southern Ocean sector of a quasi-global implementation of MICOM (the model grid extends to 74°N). This noneddy-resolving model (with constant mesh size 1.25°) includes a parameterized “bolus velocity” (representing water mass transport by eddies), an explicit seasonal cycle of forcing, and a Kraus–Turner surface mixed layer. We consider the model state after 30 years of spinup from rest, at which time the surface layers are in steady state but the deep waters are still evolving. We focus on the following issues:

  1. the formation and recirculation of SAMW and AAIW

  2. the importance of eddy-related bolus transports in the Southern Ocean

  3. links between the meridional overturning streamfunction, water mass transformation, and surface fluxes

  4. the relative importance, for the fate of NADW, of “consumption” by mixing in the Southern Ocean and northward export into the Indian and Pacific Oceans.

The rest of this paper is laid out as follows. In section 2 we describe implementation and spinup of the model. In section 3 we describe the model Southern Ocean after 30 years of spinup, with emphasis on surface buoyancy forcing, recirculation of SAMW and AAIW, and interbasin water mass exchange. In section 4 we relate zonal-mean transformation rates to the meridional overturning streamfunction, and apportion the former between surface forcing and mixing. We explain how transformation-related and total overturning streamfunctions differ when the density structure is evolving, and how this difference allows us to quantify water mass unsteadiness. We further demonstrate the relative roles of surface forcing and mixing in the Southern Ocean. In section 5 we discuss the formation of SAMW and AAIW, the role of bolus transports in balancing the meridional circulation, and some shortcomings of the model. Section 6 summarizes our main findings.

2. The model

The model is based on MICOM version 2.4 (Bleck et al. 1992) in which potential density referenced to the surface (σ0, henceforth σ) is the vertical coordinate for isopycnic layers, which represent the ocean interior, with a Kraus–Turner type mixed layer of continuously variable density incorporated as the uppermost layer (Bleck et al. 1989). The model is implemented on a quasi-global 1.25° × 1.25° grid extending from Antarctica to 74°N (thus avoiding numerical problems associated with convergence of meridians toward the North Pole). The northern boundary at 74°N is represented by a solid vertical wall, with no thermodynamic restoration employed (i.e., there are no lateral fluxes of mass, heat, or freshwater across 74°N). The land mask has resolution 2.5° lat by 3.75° long.

Beneath the mixed layer (index 1) 19 isopycnic layers (indices 2–20) represent the ocean interior, with values of σ chosen to represent the range of global water masses. Table 2 lists the densities of layers 2–20, indicating those layers which represent the Southern Ocean water masses of interest here (note that our water-mass σ ranges vary somewhat from those of Schmitz 1996a, listed in Table 1). Fields of initial layer thickness, temperature, and salinity are obtained by vertical interpolation of the Levitus (1982) September data (with σ instabilities removed), as described in New et al. (1995).

In the rest of the paper i, j, and k denote zonal, meridional, and layer indices in the model coordinate system, and we use ρ to represent generic density, while retaining σ elsewhere when referring to model layer potential density. Sections 2a–c outline model parameterizations of particular significance in the present study, surface heat and freshwater forcing, and the 30-yr spinup.

a. Parameterizations

Model parameterizations of subgrid-scale physics are more or less as for a preceding Atlantic-only model (New et al. 1995). However, as eddy-induced transports and diapycnal mixing are believed to play important roles in the Southern Ocean, we outline further details of these parameterized model processes.

Tracers and density are advected by the resolved or slowly time-varying mean flow and also by the eddy-induced horizontal “bolus velocity” [Gent and McWilliams 1990 (GM); McDougall and McIntosh 1996],
i1520-0485-30-5-1013-e1
where h is isopycnal layer thickness and dashes and overbars represent “eddy” and “mean” quantities, respectively. This velocity is parameterized in the model by lateral diffusion of layer interface depth z(ρ). Hence Ψ*, the total horizontal bolus transport below a given isopycnal ρ, satisfies
ΨκHzρ
with Ψ* = 0 on the ocean floor. In the model, diffusivity κ does not vary in the vertical, so the bolus transport within each isopycnal layer (apart from the abyssal layer) is given by
huκHh.
This bolus transport also acts within the surface mixed layer where it varies in thickness.1 Within the densest layer, since Ψ* = 0 on the ocean floor, we have instead
huκHη,
where η is the height of the top of the densest layer. The value of κ is parameterized as κ = udL, where ud = 1 cm s−1 and L is a horizontal length scale equivalent to the 1° grid resolution (hence κ ≈ 1000 m2 s−1). Further details of MICOM layer interface diffusion can be found in Langlois et al. (1997).

Diapycnal mixing in the model arises in three different ways (Nurser et al. 1999):

  1. entrainment at the base of the model mixed layer (see Bleck et al. 1989) due to either surface forcing (surface buoyancy loss and turbulent kinetic energy input from the wind) or the convection that results from baroclinic shear (when dense water is driven over lighter water)

  2. lateral mixing of T and S in the mixed layer, which, along with entrainment, we collectively term “surface mixing”;

  3. turbulent diapycnal mixing at layer interfaces in the model interior, referred to here as “interior mixing” and represented (Hu 1996) by ek+1/2, a (positive downward) vertical velocity across the interface between the layers, k and k + 1:
    i1520-0485-30-5-1013-e5
    where κd is the coefficient of diapycnal mixing, parameterized here as κd = c/N, with c = 0.575 × 10−7 m2 s−2 and the buoyancy frequency N = −(g/ρ)(∂ρ/∂z). This parameterization yields values for κd ranging from O(0.1 cm2 s−1) in the upper thermocline to O(1.0 cm2 s−1) in the deep ocean. Applying (5) to the model thickness fields and area integrating, we can obtain a globally averaged diapycnal velocity at each layer interface. This average velocity increases with depth (reflecting the stratification-dependent parameterization for κd), ranging from 0.25–0.5 × 10−7 m s−1 in the main thermocline to ∼10−7 m s−1 in the deep ocean—the canonical value of Munk (1966).

b. Surface forcing

In a 30-yr integration, the model is forced with monthly mean climatological surface fluxes [Hellerman and Rosenstein (1983) wind stress, Esbensen and Kushnir (1981) heat fluxes and evaporation rate, and Jaeger (1976) precipitation rate]. To maintain realism in surface distributions of heat and salt it is, however, also necessary to relax the sea surface temperature (SST) and salinity to monthly mean climatological values (Levitus 1982). The method used here is similar to that of Haney (1971), although we use a fixed relaxation coefficient (λ) of 35 W m−2 K−1. The net downward heat flux is thus Qnet = Qclimλ(TTclim), the climatological flux (Qclim) plus a relaxation flux, which is proportional to the difference between model mixed layer temperature (T) and climatological SST (Tclim). The freshwater flux is formulated likewise.

As in a similar Atlantic-only model (New et al. 1995), modified surface boundary conditions are used to represent the first-order effects of sea ice: the surface fluxes of heat and freshwater and the wind mixing power (but not the wind stress) are set to zero at “ice-covered” model gridpoints, defined as such if either climatological SST or model mixed layer temperature falls below −1.8°C (the approximate freezing point for seawater). The effects of sea ice, largely insulating and shielding the mixed layer from air–sea interaction, are thus simulated for a negligible increase in model integration time.

c. Spinup

To assess the relative steadiness of model water masses during spinup, we compute time series of area-integrated layer thickness anomalies (referenced to the start of the integration at time t = 0), Δhk(t):
i1520-0485-30-5-1013-e6
where hi,j,k and ΔAi,j are layer thickness and gridbox area at each gridpoint.

Figures 1a and 1b show Δhk(t) for layers 1–20, over the 30-yr spinup, in the Southern Ocean, for zones south of 42.5°S and 25°–42.5°S, respectively (the choice of these zones, based on two distinct water mass transformation regimes, is made clear in Section 4b). Seasonal cycles of mixed layer–thermocline interaction are clear in Fig. 1. South of 42.5°S layers 7–10 display strong seasonal signals, while in the zone 25°–42.5°S layers 3–7 appear to interact with the mixed layer—indicating the ventilation of lighter layers in the equatorward zone.

The outstanding features of Figs. 1a,b, however, are the large interannual trends in the thickness of the denser unventilated layers. These trends of mean layer thickness, away from initial Levitus (1982) climatological values, are due to imbalances between formation rates, transports, and mixing rates for the water mass represented by each layer. These imbalances are further investigated in the detailed diagnosis of year 30, which follows. In the region south of 42.5°S over years 0–30, the mean thickness of layer 13 increases at an almost steady rate of 2.5 m yr−1 (area-integrated, this corresponds to a gain in volume at the rate of 6 Sv—see Fig. 16b below), while layer 15 (σ = 27.88) thins at a similar rate (Fig. 1a). The zone 25°–42.5°S is dominated over years 0–20 of the spinup by an ∼3 m yr−1 increase in the mean thickness of layer 15, while lighter layers decrease in thickness to varying degrees (Fig. 1b). However, the area-integrated thicknesses of layers that represent SAMW and AAIW (layers 8 and 9 south of 42.5°S; layers 6 and 7 in the zone 25°–42.5°S) show little change over years 20–30. This suggests that, for these model water masses, annually averaged import/export across zonal boundaries and formation/consumption rates within zones are reaching an equilibrium by year 30.

To assess the drift of layer temperature (and hence salinity), we compute monthly layer-mean temperature anomalies, ΔTk(t), in the same two zones (thickness-weighting layer temperature at each gridpoint). Figure 2 shows the 30-yr time series, over the two Southern Ocean zones, of ΔTk(t) for layers 6–16 (lighter layers are relatively thin and exhibit large seasonal temperature variability; denser layers have zero thickness). Temperature drifts are typically ±0.5°C over the 30-yr spinup. In general, intermediate waters (layers 6–10) cool while deep waters (layers 11–16) warm. As the model is isopycnic, these warming/cooling trends are exactly “balanced” by increases/decreases in salinity (temperature changes of ±0.5°C are accompanied by salinity changes of roughly ±0.1 psu). T–S curves, area-integrated in the same two zones, in years 1 and 30 (not shown) confirm that the T–S relationship for intermediate waters (σ < 27.52) shifts toward lower temperature and lower salinity over the 30-yr spinup. Although water mass characteristics do change during model spinup, these changes are reasonably small, that is, model water mass properties remain close to Levitus (1982) climatological values. Certainly, after 30 years, the properties of the intermediate waters have almost reached a steady state.

3. The model Southern Ocean in year 30: A descriptive picture

Having established the interannual trends in layer properties over a 30-yr spinup, we present model year 30 Southern Hemisphere fields of surface density flux (section 3a) and layer thickness (section 3b), to show regional details of surface forcing and recirculation of mode waters. We also consider density-partitioned transports and regional formation rates (section 3c), to illustrate the qualitative relation between import/export, and formation/consumption, of water masses in the model Southern Ocean.

a. Surface density fluxes

We consider annual-mean surface density fluxes over the Southern Hemisphere oceans. Figure 3a shows the annual-mean climatological density flux, 〈Dclim〉 (we denote annual means by angle brackets), calculated from the flux fields of Esbensen and Kushnir (1981) and Jaeger (1976), as detailed in appendix A. The flux is defined as positive downward, so that positive values imply surface density gain, or an ocean-to-atmosphere flux of buoyancy. Figure 3b shows 〈D〉, the annual-mean total density flux forcing of the model ocean in year 30, which includes the relaxation heat and freshwater fluxes. Figures 3c and 3d show, respectively, the thermal and haline components of 〈D〉.

Broadly speaking, over the Southern Hemisphere of the model, surface waters gain density (lose buoyancy) over the subtropical gyres and lose density (gain buoyancy) at mid and high latitudes. The density gain in southwest sectors of the three subtropical gyres (South Atlantic, Pacific, and Indian) is associated with the separation (in the case of the Brazil and East Australia Currents) or retroflection (in the case of the Agulhas Current) of WBCs. Relatively warm, salty waters (see Figs. 4a,b) thus move south (Fig. 4c), are cooled (again largely via relaxation fluxes), and increase in density.

In contrast, there is strong density loss (again predominantly due to relaxation) at midlatitudes in the southwest Atlantic due to the heating of cold water flowing north in the Falkland Current (Figs. 4a,c) and also east of New Zealand. South of about 45°S freshening (Fig. 3d) dominates cooling (Fig. 3c), leading to weak density loss except south of New Zealand and in the southeast Pacific. With model relaxation fluxes taken into account, thermal and haline surface forcing are found to be in opposition (i.e., freshening tends to coincide with cooling) over broad regions of the Southern Ocean (see Figs. 3c,d). This may partly result from the fact that water is cold and fresh to the south and warm and salty to the north (Figs. 4a,b). Hence water pushed south of a mean position in the Levitus climatology is freshened and cooled whereas that pushed north is warmed and salinified.

Figure 5 shows the zonally-averaged annual-mean surface density flux over the Southern Ocean due to climatological (Fig. 5a) and total year 30 (Fig. 5b) heat and freshwater fluxes. Note that the total surface density gain north of 40°–45°S is two to three times stronger than that driven solely by the climatological fluxes. Note also that, in the zone south of 40°–45°S, the haline and thermal components of 〈Dclim〉 and 〈D〉 act in opposition. The general enhancement (relative to climatology) of model density loss due to haline forcing at high latitudes may reflect the effects of spring/summer sea ice melting around Antarctica (absent from the climatological EP freshwater forcing, but implied by relaxation of surface salinity toward low values).

It is unclear to what extent the net model fluxes are more “correct” than the climatological fluxes. The climatological fluxes used here overestimate the globally averaged heat input. Moreover the surface flux observations are very sparse in the Southern Ocean (Josey et al. 1999), so the climatologies are unreliable here. Certainly the climatological flux field in Fig. 3a seems unrealistically smooth. The relaxation correction gives plausible structure in the fields, qualitatively similar to that produced in (for example) the coupled National Center for Atmospheric Research model (Gent et al. 1998), but may nevertheless show bias due to model weaknesses.

b. Ventilation sites, recirculation pathways, and mode water properties

We now consider the sources of thermocline water masses in the model, in particular the formation regions for AAIW and SAMW. McCartney (1977, 1982) identifies AAIW and SAMW formation sites in the subantarctic zone, a circumpolar band of width ∼5°, bounded to the north by the Subtropical Front (STF) and to the south by the Subantarctic Front (SAF). According to McCartney, AAIW forms during winter convection in the subantarctic zone of the southeast Pacific and Scotia Sea (McCartney 1977), while SAMW forms likewise at sites in the southeast Indian and mid/southeast Pacific Oceans (McCartney 1982).

The SAF is the northernmost of three major Southern Ocean fronts, the other two being the Polar Front (PF) and the continental water boundary (Whitworth 1988). Park et al. (1993) summarize surface temperatures and salinities that conventionally define these fronts. In the model, local gradients in surface temperature and/or salinity (Figs. 4a,b) provide some indication of the STF in the South Atlantic (where salinity varies from 34.2 to 35.8 psu in the Brazil–Falkland confluence zone), the SAF in the south Indian Ocean and east of New Zealand (defined by the 8°C isotherm), and the PF in Drake Passage (the 2°C isotherm). The model ACC (Fig. 4c) is coincident with these fronts.

Figure 6 shows Southern Ocean fields of late winter/early spring (September–November; see Fig. 7) maxima in mixed layer depth (Fig. 6a) and density (Fig. 6b). The notable feature in Fig. 6a is a band of deep mixing in the midlatitudes, where mixed layers deeper than 500 m are evident in the southeast Indian Ocean, the south Tasman Sea, and the southeast Pacific, a feature also seen in the Levitus (1982) climatology. Strong wintertime density input in the Aghulas and East Australian WBCs (Fig. 3c) weakens upper-ocean stratification. Weakly stratified waters are then strongly advected along the path of the ACC through the Indian and Pacific sectors (Fig. 4c), and further density gain leads to deep convective mixing. A similar model mixed layer evolution is reported by England et al. (1994). The density of these winter mixed layers thus increases systematically from west to east in both oceans, except to the east of New Zealand where the ACC swings northward and surface buoyancy is gained.

The field of year-round maximum mixed layer density identifies the end-of-winter water, which is subducted into the permanent thermocline of the Southern Ocean. In Fig. 6b the contours correspond to model layer densities, and thus delimit the zones in which each layer outcrops in late winter. The highest maximum surface density is between 27.74 and 27.82, in the southern Weddell Sea, indicating that the densest layer that can interact with the mixed layer (by entrainment at the mixed layer base) has σ = 27.82. In our model, this is the density of CDW and lighter NADW. Layers with σ > 27.82 (denser NADW/CDW or residual AABW) can therefore only be consumed (in the Southern Ocean) by interior diapycnal mixing. Moreover, the mixed layer is too light at high southern latitudes to permit the model to form AABW at the surface. On the other hand, layers with densities in the range 26.52 < σ < 27.22 (representing SAMW and light AAIW) do outcrop over large areas of the midlatitude Southern Ocean (see gray-shaded band in Fig. 6b), which exhibit deep winter mixing (compare Figs. 6a,b), suggesting that these layers are modified by both entrainment mixing and surface fluxes.

Figure 7 shows seasonal cycles of mixed layer depth at three selected latitudes (indicated by dotted lines in Fig. 6a), which intersect areas of year-round deepest mixing in Indian and Pacific sectors of the midlatitude Southern Ocean. In general, model mixed layer depth undergoes a strongly seasonal cycle, with deepest mixing (mixed layer depth >300 m, gray-shaded in Fig. 7) in the spring (August–December) and shallower mixing (<100 m) in late summer–autumn (January–June). A similar cycle was found in the model of England et al. (1994). At 55°S, 160°E–180°, south of New Zealand, deep mixed layers persist almost all year (with the exception of brief restratification in February–March), in agreement with observations (Heath 1981) that the water column remains nearly isothermal in the upper few hundred meters. Here SAMW may be formed over most of the year, rather than only during late winter.

Figures 8a–d show the austral spring (September) thickness of layers 6–9 (σ = 26.52, 26.80, 27.03, and 27.22), which represent light and dense varieties of SAMW and AAIW. Note the trend for thickness maxima to progressively move eastward with increasing density, again in both Indian and Pacific Oceans. These thickness maxima represent the “core” ventilation sites, and lie immediately northward of corresponding layer outcrops into the deepest winter mixed layers.

To give a qualitative picture of the recirculation of SAMW and AAIW, trajectories representing the ventilation of selected water masses are started from the positions at which appropriate model layers outcrop into the wintertime mixed layer, terminating when the layer is reentrained back into the mixed layer. We integrate year 30 monthly total (advective plus bolus) velocity fields to determine trajectories of fluid columns in individual layers. Using linear interpolation to determine velocity at arbitrary points in space and time, we time integrate these velocities with a simple Runge–Kutta scheme, updating the position of a fluid column 25 times per day. Typical trajectories of SAMW and AAIW are shown in Fig. 8. Dots indicate the position along a trajectory every five years.

Layers 6 and 7 (Figs. 8a,b) (σ = 26.52, 26.80, representative of light and dense varieties of SAMW), are ventilated in the mid/east South Pacific and south Indian Oceans. Trajectories started at the cores of several ventilation sites reveal a variety of fates: convergence on the Agulhas Current in the southwest Indian Ocean after 15 years; westward recirculation, ending with reentrainment into the mixed layer at midlatitudes of the Southern Ocean after 15–20 years; slow O(50 yr) northward spreading in the Tasman Sea; “escape” from the Southern Ocean to reach the western equatorial Pacific after 35 years (thereafter swept rapidly eastward to upwell into the mixed layer just south of the equator in the mid-Pacific). Layers 8 and 9 (Figs. 8c,d) (σ = 27.03, 27.22, representing light and dense AAIW) are ventilated in the southeast Pacific and the southwest Atlantic, respectively. Trajectories reveal that σ = 27.03 parcels take O(50 yr) to reach the western equatorial Pacific while σ = 27.22 parcels are typically swept rapidly eastward along circumpolar trajectories and reentrained into the mixed layer southwest of Australia after O(15 yr). The thickness fields of layers 8 and 9 suggest only weak northward spreading of AAIW in the South Atlantic.

How do the properties of model SAMW and AAIW compare with observations? Layers 6 and 7 (σ = 26.52 and 26.80) represent the light varieties of SAMW, which form in the Indian and Pacific Oceans. Figure 9 shows their (model) temperature and salinities in the Indian sector. Different temperature and salinity intervals are grayshaded in Fig. 9 to identify correspondence with observed modal properties. Sections of temperature and salinity at ∼32°S in the Indian Ocean (Toole and Warren 1993) indicate a light SAMW layer, between 60° and 70°E at 200–400 m with temperature and salinity in the ranges 13.0°–15.0°C and 35.30–35.50 psu. In the model, these temperature and salinity ranges are found on the σ = 26.52 isopycnal where it is most strongly ventilated in the mid–Indian Ocean (Figs. 9a,b). Likewise, SAMW with σ = 26.80, forming in the southeast Indian Ocean, can be traced by temperature and salinity in the ranges 10.5°–11.5°C and 34.85–35.15 psu. This is warmer and saltier than the observations of SAMW in the southeast Indian Ocean by Schodlok et al. (1997), who found temperature and salinity in the ranges 8.50°–9.50°C and 34.60–34.70 psu at depths of up to 600–700 m, although the T–S properties on σ = 26.80 do fall in this range where the layer outcrops south of Australia at around 45°S (Figs. 9c,d).

Model AAIW is best represented by the σ = 27.03 layer, for which the circumpolar temperature and salinity fields are shown in Figs. 10a,b. Georgi (1979) found that, in the southeast Pacific and South Atlantic, the modal properties of AAIW fall in the ranges 2.0°–7.0°C and 34.10–34.40 psu. The temperature and salinity on σ = 27.03 (Figs. 10a,b) and σ = 27.22 (not shown) do indeed fall in these ranges, over much of the Southern Ocean. In a global model based on the Geophysical Fluid Dynamics Laboratory code, England et al. (1993) show how the renewal of AAIW is concentrated in the southeast Pacific Ocean off southern Chile, from where a low salinity (34.10–34.40 psu) tongue extends north and west across the Pacific at around 1000 m. Figures 8c and 10b show that the corresponding AAIW in our model can be identified by a similar salinity minimum (34.25–34.40 psu).

Figure 10c is a meridional cross section of salinity through the upper 2000 m of the model, in the central Pacific at 160°W. This may be compared with the observations along that line presented by Reid (1965, his Fig. 3); also reproduced as Fig. II-13 of Schmitz (1996a). The AAIW salinity minimum lies at 800–1100 m, extending northward to 20°S with salinities of <34.4 psu, as in Reid’s section. However, salinities of <34.2 psu in the ACC region (around 50°S) only penetrate to ∼600 m rather than reaching 1000 m as in the observations; this may result from insufficiently deep winter mixing.

c. Interbasin transports and basin-by-basin water mass formation

Consider now the circulation across selected transects in and around the Southern Ocean. Figure 11 shows total (advective plus bolus) transports for each density-partitioned “layer.” The mixed-layer transport is apportioned (in the range 24.70 < σ < 28.13) between layers according to mixed layer density (the remaining “ML” transport is at σ < 24.70). The selected zonal transects (Figs. 11a,c,e) lie along 25°S across the South Atlantic, south Indian, and South Pacific, with meridional transects southward from the Cape of Good Hope to Antarctica along 20°E, from Tasmania to Antarctica along 146.25°E, and across Drake Passage at 65°W. The zonal South Atlantic section (a) illustrates the southward-flowing supply of NADW to the Southern Ocean and the northward return flow of water at lighter densities (the lower and upper branches of the thermohaline circulation), while the sections across the South Indian (c) and South Pacific (e) reveal the northward spreading of remnant NADW and CDW. The three meridional sections (Figs. 11b,d,f) show large eastward exchanges between the various sectors of the Southern Ocean as expected, reflecting transport by the ACC.

Across 25°S in the South Atlantic (Fig. 11a), 15.7 Sv of NADW (layers 14–16; σ = 27.82, 27.88, 27.93) flows south, balanced by a northward transport of various light and intermediate waters (layers 4–13). This is in broad agreement with the Saunders and King (1995) estimate of ∼15 Sv southward flow of deep waters through the South Atlantic into the Southern Ocean, made up of 21.7 Sv of southward-flowing NADW less 6.6 Sv of northward flowing AABW, balanced by northward fluxes of surface water (11.0 Sv) and intermediate water (4.1 Sv).

The NADW that reaches the Southern Ocean flows eastward and becomes indistinguishable from CDW. Across the section south of Africa (Fig. 11b), the combined transport of NADW and CDW (layers 13–15; σ = 27.74, 27.82, 27.88) amounts to 41.4 Sv, and the total ACC transport is 96.2 Sv, which, bearing in mind the model resolution, agrees reasonably well with the observed Drake Passage transport of ∼130 Sv (e.g., Schmitz 1996a).

The northward flow of 5.5 Sv of NADW/CDW across 25°S into the Indian Ocean, section (c), is less than recent observations suggest [e.g., Robbins and Toole (1997) estimate a deep northward flow of 11.9 Sv across ∼32°S]. Flow through the Indonesian archipelago results in a net southward transport of 25.4 Sv across (c), from the Indian to the Southern Ocean in the model, considerably larger than the hydrographic annual-mean estimate of ∼10 Sv (Fieux et al. 1996; Schmitz 1996a). The flow across 25°S in the Indian Ocean includes 11.6 Sv of very light (“ML,” σ < 24.70) “remnant” Indonesian throughflow, which has gained density across the Indian Ocean and in the Aghulas Current (Figs. 3b, 4c). After further density gain south of 25°S, these light flows pass into the Pacific sector across section (d), south of Tasmania, predominantly as part of a 28.5 Sv flow of SAMW (σ = 26.52, 26.80). The 121.7 Sv total transport across this section into the Pacific is augmented by the 25.4 Sv throughflow, which then, of course, returns northward across 25°S in the South Pacific (e). The remaining 96.3 Sv flows through the Drake Passage (f) and back into the Atlantic sector.

A 14.0 Sv northward flow of NADW/CDW in the Pacific sector is partly compensated by a 3.7 Sv southward flow of upper CDW (UCDW) (layers 12 and 13;σ = 27.64, 27.74). Applying inverse methods to transpacific sections at 28°S and 43°S, Wunsch et al. (1983) obtain a meridional circulation in the midlatitude South Pacific, dominated by a deep cell amounting to 12 Sv:NADW/CDW flows northward at depths of 3000–4000 m and mixes to a slightly lower density north of 43°S, returning southward at depths of 2000–3000 m as UCDW and North Pacific Deep Water (NPDW). Of the NADW/CDW flowing northward across 25°S in our model, 10.3 Sv does not return south. An underestimated return flow of UCDW/NPDW suggests that this deep cell is not correctly balanced by mixing in the Indo-Pacific.

We also show in Fig. 12 the water mass formation rates for each layer within the various basins, obtained as the balance of imports/exports and volumetric imbalances (further discussed in the next section). NADW/CDW (σ = 27.82, 27.88, 27.93) is consumed at rates of 4.0 Sv and 2.9 Sv in the Atlantic (Fig. 12a) and Indian (Fig. 12b) sectors, respectively, while forming at 1.2 Sv (through mixing of denser bottom water and overlying NADW/CDW layers) in the Pacific sector (Fig. 12c). South of 25°S, the net 5.7 Sv loss of NADW/CDW again compares well to an independently obtained consumption rate of 5.1 Sv due to interior diapycnal mixing (see section 4b). However, these consumption rates are much smaller than the combined 19.5 Sv northward export across 25°S in the Indian and Pacific Oceans. Note, however, that the 15.7 Sv of NADW imported into the Southern Ocean (across 25°S in the Atlantic sector) is significantly less than the combined 25.2 Sv consumed in, and exported from, the Southern Ocean, again emphasizing the importance of unsteadiness in this model at high densities. Taking this unsteadiness into account, the fate of model NADW is not unrealistic—the global circulation scheme of Schmitz (1996a) requires, for NADW, a northward export of ∼9 Sv into the Indian Ocean and a consumption of ∼5 Sv in the Southern Ocean.

At lower densities the model is, however, in balance. Figure 12b shows the formation of 19.8 Sv of SAMW (σ = 26.52, 26.80) in the Indian sector. If SAMW is instead defined to include waters of density up to 27.3, as in Speer et al. (1997), the formation rate is 24.6 Sv, in good agreement with the Speer et al. value of 25 Sv, estimated from the Da Silva (1994) flux climatology. SAMW is consumed at a rate of 15.9 Sv (Fig. 12c) in the Pacific sector, where AAIW (σ = 27.03, 27.22) forms at a rate of 8.5 Sv. The balance of AAIW in the South Atlantic is less straightforward: while 29.7 Sv is transported through Drake Passage (Fig. 11f) and 18.3 Sv continues eastward south of Africa (Fig. 11b), only 3.2 Sv is transported northward across 25°S in the Atlantic (Fig. 11a) with the remaining 7.2 Sv consumed in the Atlantic sector (Fig. 12a). The northward flow across 25°S spans a wide range of light and intermediate density classes (Fig. 11a). Note that Agulhas ring shedding is precluded by coarse model resolution, and this largely eliminates the possibility of a warm water path (Gordon 1986).

4. The meridional overturning streamfunction and water mass transformation in the Southern Ocean: Theory and diagnosis

The meridional overturning streamfunction can be related to zonally averaged water-mass transformation rates. However, in the present model, water masses are unsteady (most notably the case for dense layers, even after a 30-yr spinup—see section 2c), and care must be taken when inferring transformation rates from the overturning. In the following, we explain how the difference between total and transformation-related overturning allows us to quantify water mass unsteadiness. We also develop the theory of Walin (1982) to illustrate how water mass transformation, and hence the overturning, is related to the surface density fluxes and mixing. The model is then diagnosed to show the relative significance of resolved and bolus transports, unsteadiness, surface fluxes, and mixing for water mass transformation in the Southern Ocean.

a. Theory

1) Problems in defining a meridional streamfunction zonally integrated along density surfaces

In the Boussinesq approximation the flow of water is nondivergent,
zuwz
writing z as the horizontal gradient operator at constant z and u as the horizontal velocity. Hence it is always possible to define a meridional overturning streamfunction integrated zonally on level surfaces
i1520-0485-30-5-1013-e8a
where zb(x) is a zonally dependent bottom depth, and (8a) satisfies
i1520-0485-30-5-1013-e8b
where C(z, θ) is a closed loop with latitude θ and depth z constant.
However, problems arise in the definition of an overturning streamfunction zonally integrated on surfaces of constant density, resulting from the form of the continuity equation in density space:
i1520-0485-30-5-1013-e9
writing ρ as the horizontal gradient operator at constant ρ and defining h = −∂z/∂ρ (so hδρ is the thickness of the layer sandwiched between the ρ + δρ and ρ isopycnals) and g as a diapycnal flux (from light to dense) per unit area. The crucial point is that there is a possibility of layer inflation, ∂h/∂t ≠ 0. Hence, in density space, the flow is not nondivergent. After integration of (9) around a closed zonal loop C(ρ, θ), holding latitude θ and density ρ constant, we have
i1520-0485-30-5-1013-e10
Here term (A) represents isopycnal flow, term (B) represents diapycnal flow, and term (C) water mass unsteadiness. The unsteady term (C) on the right-hand side need not vanish. Hence we cannot in general define a streamfunction ψ to which we can relate the meridional transport on the isopycnal, and the diapycnal transport across the isopycnal, according to
i1520-0485-30-5-1013-e11a
for differentiating (11a) by y and (11b) by ρ, and summing, gives zero on the right-hand side, but the left-hand side is the unsteady term (C) in (10), which is in general not zero. As we see below, we can introduce a ψ to satisfy either (11a) or (11b), but not both simultaneously.

Note that the sign convention here is such that ψ increases with decreasing density (i.e., up through the water column) for northward flow, in agreement with the convention chosen for ψlev above. There are various approaches we could take to separate the time-dependent inflation term (C); here (below) we take a simple, physically meaningful decomposition.

2) Partitioning the meridional streamfunction into “transformation” and “unsteady” components

We now relate the meridional streamfunction zonally integrated on density surfaces to the diapycnal flux across density surfaces, making plain the intimate link between the two. Define the meridional streamfunction
i1520-0485-30-5-1013-e12
to be the total northward transport of waters denser than ρ across the line of latitude θ = ϒ (by both resolved ν and bolus ν* velocities), where ρmax is the maximum (bottom) density at that latitude. We further write
i1520-0485-30-5-1013-e13
to be the total volume of fluid with density greater than ρ lying south of θ = ϒ. Note that ψ(ϒ, ρ), V(ϒ, ρ), and the other fields defined below depend on time, despite not being written with explicit time dependence.
If we then assume incompressibility (as in MICOM), it follows that the total diapycnal flow “down across” the ρ isopycnal (from light to dense) south of θ = ϒ,
i1520-0485-30-5-1013-eq1
must appear partly as volume inflation and partly as export (see Fig. 13a) according to
GρψρVρt.
In the steady state, all of the flow across ρ must be exported, with G(ϒ, ρ) = ψ(ϒ, ρ), as noted by Speer (1997). However, the point is that, when the inflation rate ∂V(ϒ, ρ)/∂t ≠ 0, part of the diapycnal flow may be used to inflate the volume beneath ρ, and the streamfunction does not correspond exactly to the diapycnal flow, G(ϒ, ρ) ≠ ψ(ϒ, ρ). In particular,
i1520-0485-30-5-1013-e15
Hence (see Fig. 13b), a divergence of fluid denser than ρ (∂ψ/∂ϒ > 0) need not imply a light-to-dense transformation at ϒ (∂G/∂ϒ > 0). Instead, it could imply a decrease in the volume of water with density greater than ρ at ϒ (∂2V/∂ϒ∂t < 0). On the other hand, the streamfunction ψ(ϒ, ρ) as defined above does have the useful property that it satisfies (11a); namely,
ψρU,
where
i1520-0485-30-5-1013-e17
is the northward transport of waters with density between ρ and ρ + δρ. Moreover, ψ = 0 at boundaries (seafloor, coasts) and is only nonzero at the ocean surface if there is depth-integrated flow—i.e., it looks like a streamfunction!
The “transformation streamfunction,” G(ϒ, ρ), of course gives the correct diapycnal flux, ∂G/∂ϒ, of waters at latitude ϒ. However, we now have (Fig. 13c)
i1520-0485-30-5-1013-e18
[writing V = −∂V/∂ρ, so that V(ϒ, ρ)δρ is the volume of water with density between ρ and ρ + δρ south of ϒ]. Thus −∂G/∂ρ now includes the rate of inflation of the layer with density between ρ and ρ + δρ south of ϒ, together with the northward export across the latitude ϒ. Walin (1982) termed this sum the water mass formation rate M. Furthermore, of course, G(ϒ, ρ) will not in general be zero at the northern boundary of a closed basin, if the volume below the ρ isopycnal is inflating or deflating—that is, G(ϒ, ρ) does not look like a streamfunction.

So there are two candidate streamfunctions, ψ(ϒ, ρ) and G(ϒ, ρ), in terms of which we might wish to consider zonally integrated flow along surfaces of constant density. Both have deficiencies when the density structure is evolving: ψ(ϒ, ρ) gives the incorrect diapycnal flow (but correct isopycnal flow) and G(ϒ, ρ) the wrong isopycnal flow (but correct diapycnal flow). However, the important thing is that G(ϒ, ρ) has a clear physical interpretation, being the total flux of waters down across ρ, south of ϒ, and the deviation of G(ϒ, ρ) from ψ(ϒ, ρ) gives a simple and direct measure of the importance of water mass unsteadiness.

3) Diapycnal fluxes and density forcing

There must be some source of density to allow diapycnal flow across ρ, that is, G(ϒ, ρ) ≠ 0. This source is associated with either the surface outcrop of ρ or a convergence of the interior diapycnal density flux. Hence (Walin 1982; Speer and Tziperman 1992; Nurser et al. 1999):
i1520-0485-30-5-1013-e19
Here
i1520-0485-30-5-1013-e20
is the surface density gain through surface outcrops of ρ [ρ < ρm(x), where ρm(x) is the mixed layer density at x] south of latitude ϒ, where Din is the surface density gain per unit area (see appendix A), and
i1520-0485-30-5-1013-e21
is the total nonadvective density flux (positive from light to heavy) across ρ, south of ϒ, where D is the field of density flux across ρ and n = ρ/|ρ| a unit vector normal to the ρ isopycnal surface. The contribution C(ϒ, ρ) results from cabbeling: density gain through mixing on ρ (see McDougall 1987).
A striking consequence of the above is that, for the particular case in which (i) the surface fluxes dominate mixing and (ii) the ocean is in steady state, it follows that
ψρFρ
and we can infer the meridional overturning streamfunction in density coordinates from surface density forcing alone.

b. Diagnosis

1) The meridional streamfunction: Resolved and bolus transports

We calculate ψ(ϒ, σ), the meridional overturning streamfunction, as a function of σ. We choose to define ψ < 0 for an “anticlockwise” sense of overturning, as this is consistent with our direct relation between the meridional streamfunction and water mass transformation. Details of our evaluation of ψ(ϒ, σ) can be found in appendix B. An annual mean, 〈ψ(ϒ, σ)〉, is obtained by averaging 12 monthly instantaneous solutions for ψ(ϒ, σ). We also separately determine 〈ψ(ϒ, σ)〉 and 〈ψ*(ϒ, σ)〉, annual-mean overturning streamfunctions implied by mean (resolved) and bolus (parameterized, eddy-driven) transports, respectively, and discuss how these combine as a total streamfunction.

Figures 14a, 14c, and 14e show 〈ψ(ϒ, σ)〉, 〈ψ(ϒ, σ)〉, and 〈ψ*(ϒ, σ)〉 in the Southern Ocean south of 25°S. The dashed lines indicate σmaxml(ϒ), the zonal year-round maximum density of the mixed layer—a zonal maximum of the “end-of-winter” density field plotted in Fig. 6b. At latitude ϒ, water with σ < σmaxml(ϒ) outcrops into the mixed layer during part (or all) of the year, while water with σ > σmaxml(ϒ) can only change density by diapycnal mixing [hence the near-horizontal streamlines for σ > σmaxml(ϒ)]. The black mask indicates densities that are never reached at any time in the year, anywhere at a given latitude. We also present in Figs. 14b, 14d, and 14f, for the various model layers, corresponding annual-average volume fluxes northward across 42.5°S (the mean latitude of the core of the ACC), indicated by the dotted lines in Figs. 14a,c,e.

Three circulation cells are apparent in Fig. 14a. The low density cell implies density gain by lighter waters, representing circulation of the subtropical gyres. About 24 Sv of light (σ < 25.3) waters are imported from the north, gaining density in WBCs and along the midlatitude zonal extensions of those current systems (30°S to ∼42.5°S: see Figs. 3b, 4c). These waters then return northward with σ ≈ 25.8–26.8, including ∼8 Sv of SAMW (σ ≈ 26.5–26.8). This cell, since it is largely north of 42.5°S, is not evident in the layer-by-layer fluxes across 42.5°S (Fig. 14b).

A second, weaker, cell represents the lightening of NADW/CDW to AAIW, the Southern Ocean component of the thermohaline circulation: 6 Sv of water with 27.6 < σ < 27.8 flows south across 25°S, lightens to 27.3 < σ < 27.4 south of 55°S, and then returns northward. This water continues to gain buoyancy in the zone 42.5°–55°S, with density decreasing to 26.9 < σ < 27.2 (values typical of dense SAMW and light AAIW). In this range of density the flow ultimately exits the Southern Ocean across 25°S. This cell is also evident in the layer-by-layer flow across 42.5°S (Fig. 14b) that shows southward flowing layers with 27.38 ⩽ σ ⩽ 27.74 balanced by northward return flows at σ ⩽ 27.22.

Although there is no AABW formation in the model, Fig. 14a suggests a third circulation cell, extending almost to 60°S, involving a ∼2 Sv northward export of very dense waters (σ > 27.8) balanced by input of slightly lighter waters. The export of these very dense waters across 42.5°S reaches 3 Sv (Fig. 14b) and includes (unrealistic) northward transport in the density range of NADW (layers 14, 15; σ = 27.82, 27.88). We find that this spurious cell is due to the effect of unsteadiness and disappears when this is taken into account (see below).

In Figs. 14c and 14e, we decompose 〈ψ(ϒ, σ)〉 into 〈ψ(ϒ, σ)〉 and 〈ψ*(ϒ, σ)〉, annual-mean overturning streamfunctions implied by mean and bolus transports, respectively. Note that the diapycnal flows implied by 〈ψ(ϒ, σ)〉 and 〈ψ*(ϒ, σ)〉 are entirely spurious: the purpose of this decomposition is simply to show the strength of isopycnal bolus transport relative to other key quantities such as the zonally integrated Ekman drift. A structure very similar to the Deacon cell appears in Fig. 14c as an overturning of ∼24 Sv centered on σ = 27.5 at 45°S. Note, however, that this is not the Deacon cell as defined by Döös and Webb (1994), as averaging has been performed on isopycnals, not levels. The strength of this cell (18 Sv at 42.5°S), however, is in fact comparable to the zonally integrated Ekman drift (∼30 Sv at 42.5°S), suggesting a physical explanation similar to that for the Deacon cell.

To emphasize this further, we include in Fig. 14d the layer-by-layer advective fluxes (gray bars) prior to the apportionment of mixed layer fluxes into density classes, as well as the apportioned fluxes (black bars) used to generate the streamfunction in Fig. 14c. We see that indeed there is a very strong northward flux in the mixed layer—the ∼30 Sv Ekman flux. It is this flux, after apportionment into the lighter layers (layers 2–7; 24.70 < σ < 26.80), that gives the northward component of the cell (Fig. 14c; the black bars in Fig. 14d). Since these light layers carry southward transport where they are in the thermocline (gray bars for layers 5–7 in Fig. 14d), the total northward transport of our “quasi-Deacon cell” only reaches 18 Sv (rather then 30 Sv). The return southward transport is mostly carried in layer 15 (σ = 27.88), which is generally the bottom layer. So, apportioning the mixed layer flow into density classes has not eliminated this cell in the resolved flow, as it did in FRAM (Döös and Webb 1994), but merely weakened it. Since this cell does not appear in the “total” flow (Fig. 14a), it must be counterbalanced by the bolus transport.

Figure 15 shows an annual-mean, zonally averaged meridional section of layer interfaces across the Southern Ocean, illustrating how model layers tend to shoal toward Antarctica. Hence the bolus transport streamfunction
i1520-0485-30-5-1013-e23
(where j is the unit vector in the direction of increasing latitude) is positive throughout the Southern Ocean (Fig. 14e): deep layers move northward and shallow layers southward. The eddy-driven overturning cell thus opposes the Deacon cell–like structure in the resolved flow. Peak bolus transports coincide with the steepest sloping isopycnals. The rapid equatorward decrease in the thickness of layer 15 (labeled in Fig. 15), σ = 27.88, gives a particularly strong northward bolus transport of that layer in the latitude range 40°–60°S, reaching 16 Sv at 42.5°S (Fig. 14f). The compensating bolus transports in the upper layers 5–13 are southward (Figs. 14e,f). The bolus transport in the mixed layer is negligible at 42.5°S, so we do not show the gray bars in Fig. 14f. At 42.5°S, layer 6 (σ = 26.52) is the layer that (on zonal average) lies immediately below the mixed layer, and hence is thinning toward the south most rapidly. It therefore has the strongest southward bolus transport. Farther north, the corresponding layer immediately below the mixed layer is lighter (see Fig. 15), so the strongest bolus transport occurs at lower densities (Fig. 14e).

Compared to Figs. 14c,d and 14e,f, there is relatively weak circulation at σ > 27.8 in Figs. 14a,b; that is, the resolved and bolus transports tend to cancel at high density. Strong southward deep flows (Figs. 14c,d) are supported by topography feeling a westward pressure drag (from the topography) that balances the surface eastward wind stress. These southward flows are balanced by northward bolus transport (Figs. 14e,f) in the bottom layers. Since the depth-integrated bolus transport disappears, lighter waters experience southward bolus transports (Figs. 14e,f), which weaken but do not entirely cancel (Figs. 14a,b) the northward Ekman drift (Figs. 14c,d).

2) The meridional streamfunction: Unsteadiness and water mass transformation

We now relate ψ(ϒ, σ), the total meridional overturning streamfunction, to the water-mass transformation streamfunction 〈G(ϒ, σ)〉 discussed in section 4a, and examine the impacts of unsteadiness on these streamfunctions.

Figure 16a shows 〈∂V(ϒ, σ)/∂t〉, the rate of change of fluid denser than ρ south of latitude ϒ [see (14)], and Fig. 16b the layer-by-layer rate of change of volume south of 42.5°S. A cell reaching −6 Sv at 25°S, σ ≈ 27.8, is evident. Water with density greater than ∼27.8 is being lost from the Southern Ocean and replaced by waters in the density range 27.6 < σ < 27.8. However, lighter waters (σ < 27.6) are close to equilibrium. The vestigial AABW cell of strength 5 Sv apparent in Fig. 14a, also appears in Fig. 16a, indicating that it is driven by volumetric unsteadiness at densities slightly greater/less than σ = 27.8.

Figure 16c shows 〈G(ϒ, σ)〉, the transformation streamfunction [see again (14)], while Fig. 16d shows layer-by-layer formation rates south of 42.5°S, the sum of northward layer flows (Fig. 14b) and rates of change in layer volume (Fig. 16b). Note that the AABW cell in Figs. 14a,b is not evident in Figs. 16c,d since there is no transformation at such densities in the model. The AABW cell is simply a consequence of volumetric unsteadiness in waters with σ around 27.8, as shown in Figs. 16a,b. It is now evident that NADW waters (σ = 27.82, 27.88) are indeed exported to, and lightened in, the Southern Ocean. The northward flow of these waters (Figs. 14a,b) results largely from unsteady processes associated with model nonequilibrium.

Agreement between the transformation streamfunction 〈G(ϒ, σ)〉 and total streamfunction 〈ψ(ϒ, σ)〉 is, however, very good for σ < 27.6, showing that the light and intermediate waters are reasonably near to a steady state. Note that dense SAMW and light AAIW (layers 7 and 8, with σ = 26.80 and 27.03) is being formed south of 42.5°S (Fig. 16d) at a combined rate of 5.3 Sv.

In Fig. 16e we plot 〈F(ϒ, σ)〉, the annual-average surface-forced volume flux across σ south of ϒ [see (20)]. The dominant feature in Fig. 16e is a very strong meridional gradient in 〈F(ϒ, σ)〉 for light waters (density up to σ ≈ 26.0). At σ = 26.0, 〈F(ϒ, σ)〉 increases from −6 Sv to +28 Sv moving north from 42.5°S to 25°S. This results from the strong surface density input in this latitude range (Fig. 5b), which is taken up predominantly by waters with σ ≈ 26.0, a typical mixed layer density at these latitudes. An equatorward return branch in the density range 25.8–26.9 at 25°S completes the southern half of a major surface-forced overturning cell. Weaker cells are found at higher latitudes and densities: a positive cell of 6 Sv, centered on ∼45°S, σ ≈ 26.4; a negative cell of −10 Sv at 50°–60°S, spanning the density range 26.8 < σ < 27.3.

Surface density forcing thus drives a ∼30 Sv production of waters with 25.8 < σ < 26.8, including ∼8 Sv of SAMW with 26.5 < σ < 26.8. This is the result of surface density gain over the WBCs acting on light waters that are imported from the north. Associated with the weaker postive cell is a ∼2 Sv equatorward return flow of AAIW (σ ≈ 27.1). At low densities, the surface-forced transformation (Fig. 16e) is generally very similar to the total transformation (Fig. 16c), indicating that surface fluxes tend to dominate mixing in that density range. However, the high-latitude lightening cells in Figs. 16c and 16e look somewhat different, suggesting (not surprisingly) that mixing does play a substantial role at higher densities.

3) Water mass transformation: Surface fluxes versus mixing

We now differentiate more precisely between the relative roles, in water mass transformation, of surface fluxes and interior mixing. We consider two zones. The first zone is south of 42.5°S, the latitude at which the Ekman drift reaches its maximum strength. Here there is a generally negative surface density flux (see Figs. 3 and 5). NADW upwells and mixes in this zone to form lighter CDW, some of which reaches the surface, further gaining buoyancy to form AAIW. The second zone is 25°–42.5°S, bounded equatorward by the centers of the subtropical gyres. This is a region of surface density gain and SAMW formation (see Figs. 3, 5, and 8a,b). Consider first the zone south of 42.5°S.

The (total) annual-average water-mass transformation rate, 〈G(σ)〉 [i.e., 〈G(ϒ, σ)〉|ϒ=42.5°], is indicated by the thick solid curve in Fig. 17a. It is negative at all σ, indicating that the diapycnal volume flux is toward lower density: all water masses south of 42.5°S in the model are being lightened. The diapycnal flux reaches a peak of 9.3 Sv across σ = 27.3.

The dotted curve in Fig. 17a indicates 〈F(σ)〉, the transformation rate driven by surface density forcing south of 42.5°S. As was seen for 〈F(σ)〉 (Fig. 16e), surface forcing south of 42.5°S drives generally negative (dense to light) transformation, reaching −6.2 Sv and −6.6 Sv across σ = 26.0 and 27.3, respectively. The gray curve in Fig. 17a, computed as a residual,
GmixσGσFσDσσCσ
is the implied transformation rate due to mixing south of 42.5°S.2 As one would expect, 〈Gmix(σ)〉 is positive at low density (σ < 26.5) and negative at high density (σ > 26.5), transforming light and dense water toward an intermediate density of σ = 26.5, that is, mixing.

Figure 17b shows the thermal (solid curve) and haline (dotted curve) components of 〈F(σ)〉. Separately, each of these is strong, implying diapycnal fluxes of more than 30 Sv across σ = 27.0. However, the thermal forcing (net cooling) almost cancels the haline forcing (net freshwater gain) over this region, although the overall net forcing is weakly negative, that is, haline forcing is slightly dominant (see also Fig. 5). This suggests that surface fluxes south of 42.5°S are important in cooling and freshening water on isopycnals.

The implied mixing in Fig. 17a is made up (Nurser et al. 1999) of lateral mixing within the mixed layer and entrainment at the mixed layer base (collectively termed “surface mixing”), and diapycnal mixing across isopycnic layer interfaces (“interior mixing”). Layers with density in the range σ > 27.74 (NADW and dense CDW) do not outcrop in the model Southern Ocean (see Fig. 6b), and therefore can only be transformed by interior mixing in the thermocline. Figure 17a indeed shows how lightening of waters denser than 27.5 (forming CDW) is achieved by mixing rather than surface forcing, at a rate of ∼4 Sv. Further lightening to form AAIW (27.0 < σ < 27.4) is achieved by a combination of mixing and surface buoyancy forcing in the mixed layer.

To further investigate dense water transformation, we use (5) to compute annual-mean diapycnal velocity fields at the interfaces between isopycnic (interior) layers. Over large areas of the Southern Ocean, diapycnal velocities at the interfaces between dense layers are in the range −0.6 to −0.8 (×10−7 m s−1). Integrating the differences between diapycnal velocity at the upper and lower interfaces of a layer over year 30, we obtain annual changes in layer thickness due to interior diapycnal mixing. Consider the case of layer 15 (model NADW). Fields of layer 15 thickness change due to mixing (not shown) indicate an annual 2.0–2.5 m thickness decrease over wide expanses of the Southern Ocean. Integrating layer thickness change regionally, we further obtain layer formation rates due to this mixing. Figure 18a shows the formation rates due to interior mixing south of 42.5°S. While 2.0–2.5 m thickness changes are small increments of “total” layer 15 thickness (typically 500–1500 m: see Fig. 15), the area-integrated annual-mean layer 15 consumption rate is 2.7 Sv, in reasonable agreement with the rate of 3.8 Sv implied by volume fluxes across, and unsteadiness south of, 42.5°S (Fig. 16d). In fact, for layers 14–16, which do not outcrop in the Southern Ocean, there is good correspondence between the formation rates in Figs. 18a and 16d. Formation rates due to the “total” mixing, from the residual mixing transformation rates (Fig. 17a), are shown in Fig. 18b. Differences between the formation rates due to total and interior mixing (Fig. 18c) are attributed to “surface mixing.” Despite the uncertainities inherent in our estimations, it is clear that surface mixing dominates interior mixing except at high densities.

We again emphasize here how important it is to differentiate between water mass formation and transports at high density. For these water masses, formation rates due to diffusive diapycnal mixing are of the same order as the unsteady volumetric changes. For example, although mixing accounts for ∼3 Sv consumption of layer 15 (Fig. 16d), the volume of this layer south of 42.5°S declines at the greater rate of ∼6 Sv (Fig. 16b).

Figure 19a shows the transformation rate within the equatorward zone, 25°–42.5°S, 〈G(ϒ, σ)〉|ϒ=25° − 〈G(ϒ, σ)〉|ϒ=42.5°. In contrast to the region south of 42.5°S, the transformation rates in Fig. 19a are positive at almost all σ (i.e., most water masses in the zone 25°–42.5°S are being transformed to higher density), peaking at 28.8 Sv across σ = 25.0. Figure 19a also shows that this net light-to-dense conversion is predominantly surface forced (the dotted curve largely follows the solid curve), although the effect of implied mixing (gray curve) is significant. Interior diapycnal mixing of deep layers still acts, as it did south of 42.5°S, to transform dense water (σ > 27.2) toward lower density. Note (Fig. 19b) that, in contrast to the region south of 42.5°S, thermal and haline surface forcing over the region 25°–42.5°S are generally of the same sign (i.e., net cooling is accompanied by evaporation) and that thermal forcing dominates haline forcing. This surface density gain is also clear in Fig. 5.

Our transformation rates are subject to sampling errors, arising from the use of instantaneous monthly model fields to represent the seasonal cycle (rather than calculating fully time-integrated diagnostics “online” as the model runs, which is logistically prohibitive). We feel, however, that the results are sufficently accurate for the purposes of this paper. To test the robustness of the diagnosis we repeated our calculations on years 26–29 of the model spinup and found no notable difference in regional volume fluxes, etc.

5. Discussion

In the following, we consider the extent to which an isopycnic model (based on MICOM) is successful in forming and maintaining SAMW and AAIW. We also discuss how the meridional circulation of the model Southern Ocean is explained in terms of a Deacon cell and bolus transports. Given these successes, we then consider shortcomings of the model.

a. The formation of SAMW and AAIW

The model provides a plausible picture of SAMW and AAIW formation and recirculation. In agreement with the observations reviewed by McCartney (1982), progressively denser varieties of model SAMW and AAIW form eastward of the Greenwich meridian, with modal potential density in the thermocline ranging from 26.52 in the mid–south Indian Ocean to 27.22 in the southwest Atlantic. Modal temperatures and salinities of SAMW and AAIW are close to observed values in the mid–Indian Ocean (Toole and Warren 1993) and in the southeast Pacific (Georgi 1979). However, mode waters south of Australia are warmer and saltier than recently observed SAMW (Schodlok et al. 1997). An eastward increase in winter surface density in both Indian and Pacific sectors of the Southern Ocean is associated with systematic buoyancy loss along the track of the ACC as it drifts southward across these sectors. This is also consistent with the eastward deepening of winter mixed layers in both south Indian and South Pacific Oceans, as shown in the Levitus (1982) climatology.

Our model supports the ideas of wintertime formation of AAIW and SAMW, with the strongest sources toward the east of south Indian and South Pacific sectors where mixed layers are deepest (McCartney 1977, 1982). Superposed onto the large-scale recirculating flow of SAMW/AAIW, there is a smaller component of AAIW originating from upwelling of NADW and upper CDW (Sverdrup et al. 1942). Exactly how these two components of AAIW combine together is unclear, but it seems that mixing occurs extensively along the ACC in winter mixed layers.

Previous GCM studies have highlighted problems that lead to incorrect representation of SAMW and AAIW. In an early version of the GFDL model without seasonal cycling, Toggweiler et al. (1989) found SAMW formation to be concentrated in two zones of intense convection adjacent to Australia, while AAIW failed to form. The Southern Ocean in that model was therefore dominated by light SAMW of southeast Indian Ocean origin, and T–S relationships in the southeast Pacific and South Atlantic showed thermocline water masses to be too light, warm, and salty.

In more recent GFDL model studies, progress has been made toward a better representation of the Southern Ocean (England 1993; England et al. 1993; England et al. 1994). However, problems inherent to such level-coordinate models remain. The prevalence of deep convection (due to density instabilities) at high latitudes (south of 60°S) is a major error in Bryan–Cox type models (e.g., England et al. 1994). The introduction of the GM scheme does, however, alleviate this problem (Danabasoglu and McWilliams 1995).

Without GM, the representation of AAIW in such models has proved problematic. England (1993) finds that AAIW is also sensitive to the level of horizontal diffusion (strong diffusion weakens the front at which AAIW subducts), the chosen geometry of Drake Passage (a wide Drake Passage strengthens the Deacon cell at the expense of equatorward-spreading AAIW), and the vertical buoyancy flux due to NADW upwelling in the Southern Ocean (enhanced NADW export to the Southern Ocean strengthens the halocline, allowing a more realistic low-salinity tongue of AAIW to form). Furthermore, England et al. (1993) find the formation of AAIW in the same model to be strongly dependent on the choice of a lateral diffusion scheme. Enhanced Cartesian mixing in an along-isopycnal sense brings more freshwater from the surface to intermediate model depth levels, increasing AAIW formation rates. By contrast, isopycnal advection and mixing are naturally represented in MICOM, which includes no spurious horizontal mixing terms. Layer properties are therefore largely conserved over long trajectories and timescales. This amounts to an important advantage over the GFDL model, for which mixing must be carefully tuned in order to achieve correct representation of AAIW.

Selected trajectories reveal that model SAMW and AAIW forms in the subtropics and midlatitudes of the Southern Ocean and then either follows a circumpolar path before reentrainment into the winter mixed layer north of the ACC or recirculates around one of the anticyclonic subtropical gyres in the Southern Hemisphere, on 10–50-yr timescales. This agrees with the schematic SAMW and AAIW ventilation patterns of McCartney (1982). Our trajectories do not, however, account for the diapycnal mixing of adjacent layers in the main thermocline. Associated with this mixing are upward velocities of around 0.3 × 10−7 m s−1, implying 50-yr vertical displacements of typically 50 m. So, while the isopycnal trajectories calculated here do accurately represent the O(10–50 yr) ventilated pathways followed by subducted water, it should be appreciated that diapycnal mixing substantially determines the fate of a water parcel on longer timescales.

Ribbe and Tomczak (1997) recently used an “off-line” diffusion and advection model to advect an ideal tracer with the mean velocity field of FRAM, with the purpose of demonstrating the formation mechanisms and sites, and likely pathways, for SAMW. In that study, the use of mean fields (for logistical reasons) of convection depth and velocity implicitly omits the seasonal cycle and bolus transports (by transient eddies), effects which we have shown to be integral to the formation and recirculation of SAMW.

In using a seasonally forced isopycnic model that parameterizes bolus transports, we feel that we are able to more successfully represent the formation and recirculation of SAMW and AAIW. Table 3 summarizes the water-mass formation rates for SAMW and AAIW, in circumpolar zones and by ocean sectors. SAMW forms mainly in the zone 25°–42.5°S, at a rate of 11.3 Sv, while AAIW forms more strongly south of 42.5°S, at the somewhat weaker rate of 2.7 Sv. Regionally, SAMW (σ = 26.52, 26.80) forms predominantly in the south Indian Ocean, at a rate of 19.8 Sv, while AAIW (σ = 27.03, 27.22) forms mainly in the Pacific, at a rate of 8.5 Sv. Our Indian sector SAMW formation rate is in close agreement with an estimate of 25 Sv obtained by Speer et al. (1997) from the air–sea fluxes of Da Silva et al. (1994), but considerably more than Schmitz’s (1996a) asserted 4 Sv of USAMW (upper SAMW defined by the σ-range 26.5–26.8) forming and recirculating in the Indian Ocean. Our South Pacific AAIW formation rate also compares favorably with Speer et al. (1997), who obtain a formation rate of 7 Sv in the σ-range 26.95–27.15, while Schmitz (1996a) similarly suggests the formation of 11 Sv of UPIW and LOIW (upper and lower intermediate waters with σ in the range 26.8–27.5) in the southeast Pacific.

b. The Deacon cell and bolus transports

Our density-averaged streamfunction shows, in agreement with the circulation inferred by hydrography and consistent with plausible buoyancy forcing,

  1. the import of dense NADW, lightening south of 42.5°S, and the northward export of AAIW

  2. the import of light waters in the WBCs, cooled north of the ACC (25°–42.5°S) to form SAMW, which is exported northward.

We computed the meridional overturning streamfunction from the “resolved” velocities and layer thicknesses, zonally averaged along isopycnal surfaces. This showed a Deacon cell–like structure with near-surface (light) northward Ekman drift and southward deep (dense) flows resting against topography, supporting an overturning of up to 24 Sv in the zone 30°–70°S. However, the bolus velocity associated with transient (parameterized) eddies transports light water southward and dense water northward so as to give an opposing overturning of ∼16 Sv. This gives a net overturning of only ∼8 Sv for the deeper waters (in the sense of southward moving dense waters, northward moving lighter waters), balanced by buoyancy forcing. This weakening of the Deacon cell–like structure is also seen in coarser resolution models—with both level (Danabasoglu et al. 1994) and isopycnal (Hu 1997) vertical coordinates—which employ a parameterized bolus transport.

Our results are, in these respects, broadly similar to those obtained with the idealized conceptual models of Johnson and Bryden (1989) and Marshall (1997). However, because the present model includes substantial diapycnal forcing, fluid can change density, and the cancellation between the Deacon cell and bolus transports is not complete. Note also that, with the parameterization of bolus velocity employed here, bolus transport within the mixed layer is actually northward because zonal-mean mixed layer depth increases to the south (Fig. 15), and therefore cannot balance the northward Ekman drift, as in the conceptual model of Johnson and Bryden (1989). Instead, the compensatory flow is in the layers immediately below the mixed layer.

This result differs from that found in the eddy-permitting FRAM model in which zonal correlations between isopycnal thickness and velocity balanced the zonal-mean circulation (Stevens and Ivchenko 1997), so that the Deacon cell was not evident in the overturning streamfunction plotted on density surfaces (Döös and Webb 1994). Why these results should differ so much is unclear. Compared to satellite altimeter observations (e.g., Chelton et al. 1990), the transient eddy activity is underestimated in FRAM (Stevens and Killworth 1992) for which the horizontal resolution is sufficient only to “permit,” rather than fully resolve, the eddies. On the other hand, parameterized bolus transports in the present model may be too strong in quiet regions of the ocean away from the ACC. The value of the interface diffusion coefficient used here (1000 m2 s−1) may be correct for eddy-rich regions (e.g., Lee et al. 1997), but may be unrealistically high away from the regions of strong eddy activity.

c. Shortcomings of the model

In the model, surface heat and freshwater fluxes comprise a seasonally varying climatological component, together with relaxation to (seasonally varying) surface temperature and salinity. Unfortunately observations are sparse in the Southern Ocean (e.g., Josey et al. 1999), so climatologies are unreliable. Also, the modeled Southern Ocean circulation may be unrealistic—for example, the ACC may be too broad (given the coarse horizontal resolution)—resulting in further sea surface temperature and salinity errors.

In the model Southern Ocean, the zonally averaged total heat flux is dominated by the relaxation term (see Fig. 5), with cooling in the zone 25°–42.5°S at rates of 10–20 W m−2. Spanning such a broad range of latitude, this suggests that cooling may be underestimated here in the climatology. In contrast, the climatological freshwater flux seems surprisingly accurate. South of 42.5°S, freshening (due to precipitation and/or ice melting) slightly dominates cooling, giving weak buoyancy gain, while cooling dominates in the zone 25°–42.5°S (see again Fig. 5), driving strong buoyancy loss. There are no systematic shifts in the model buoyancy forcing (see Figs. 3a,b), which might be expected if the WBCs or ACC were mispositioned. It seems, moreover, that the field of climatological density forcing (Fig. 3a) is too smooth and that realistic details in the model forcing field (Fig. 3b) are, in a sense, “restored” by the relaxation scheme (e.g., due to climatological low salinity water off southern Chile, eastward extension of model WBCs at midlatitudes, etc.).

The intermediate and deep circulation of the model should be treated skeptically. In common with many other models, we use only a simple ice model and (poorly known) monthly mean climatological surface forcing. Furthermore, cabbeling is prohibited in this version of MICOM, thereby excluding another possible buoyancy sink. Hence the surface waters nowhere get dense enough for AABW to form. Furthermore, the choice of σ0 as a vertical coordinate would lead to an ambiguity in the representation and dynamics of model AABW, even if it was correctly formed. AABW lies beneath NADW in the real world, despite that AABW has a lower value of σ0 than NADW. The inclusion in the model of an arbitrary reference depth for σ, and thermobaric effects, is presently under development. Recent results suggest that σ2 (σ referenced to 2000 m) might be a more suitable isopycnal coordinate (Sun 1997; Sun et al. 1999), with implications for dynamical balances in the Southern Ocean. Sun (1997) shows that, in global models based on σ0 and σ2 versions of MICOM, layer interfaces in the circumpolar zone 45°–60°S slope much more steeply to the surface in the σ2 model. The σ2 model thus features stronger circumpolar transports (e.g., a more realistic Drake Passage transport of 130 Sv) and, by inference, stronger bolus transports.

Finally, the parameterization of interior diapycnal mixing by the Hu (1996) scheme may be inadequate. That scheme has been recently superseded by the physically consistent scheme of McDougall and Dewar (1998), but the latter was not available when the present model was implemented. Poorly parameterized mixing may be a particular problem in the Pacific, owing to the large area and weak thermohaline forcing (i.e., slower meridional circulation) of that ocean. The consequent failure to form upper Circumpolar Deep Water (UCDW) (the end product of NADW/CDW mixing in the Pacific Ocean) and/or export this water mass to the Southern Ocean is a serious model flaw. Most of the ∼12 Sv of deep water thought to flow across 25°S into the South Pacific does not return as surface water but as UCDW (Callahan 1972; Wunsch et al. 1983). It may be that UCDW outcrops at high southern latitudes and is driven equatorward in the Ekman layer, which is freshened and lightened by surface fluxes to form SAMW and AAIW in the southeast Pacific (Schmitz 1996a); that is, lightening of deep waters to form AAIW may, in the real ocean, be a two-stage process: the first stage in the deep Pacific and the second near the surface in the Southern Ocean. Poor representation of UCDW in the present model may thus have implications for the realism of the thermohaline circulation. The model shows strong dense-to-light transformation of intermediate and dense waters within the Southern Ocean (up to 9.3 Sv south of 42.5°S; Fig. 17a). In the model most of the lightening of these intermediate waters is accomplished by mixing rather than surface forcing. Surface forcing does, however, have a great impact in freshening and cooling the intermediate waters in the model (Fig. 17b).

Taking these flaws into consideration, it is no surprise that the fate of NADW (closing the thermohaline circulation) is poorly represented in the present model. While 15.7 Sv of NADW flows into the Southern Ocean from the South Atlantic, 5.7 Sv of CDW/NADW is lightened by mixing processes in the Southern Ocean (south of 25°S), and 19.5 Sv is apparently exported across 25°S into the Indian and Pacific Oceans. Clearly then, the dense water masses are significantly unsteady, and we cannot therefore use the present model integration to determine whether, in the real world, NADW upwells globally (Drijfhout et al. 1996) or principally in the Southern Ocean (Döös and Coward 1997). All we can say is that, in the model, both processes appear to operate. Note however that the lightening of 5.7 Sv of CDW/NADW in the Southern Ocean agrees well (though perhaps coincidentally!) with Schmitz’s (1996a) estimate of 5 Sv.

6. Summary

We have used an intermediate-resolution (1.25°) isopycnal coordinate OGCM of a quasi-global ocean to investigate the processes involved in both the circulation and transformation of water masses in the Southern Ocean. The model includes an active mixed layer, parameterization of the bolus transport driven by transient eddies, and seasonally varying surface fluxes. A large-scale circulation scenario emerges. NADW is transported southward across 25°S in the Atlantic sector at a rate of 15.7 Sv, and is partly mixed with waters of Antarctic origin to form CDW in the Southern Ocean. South of 25°S, both NADW and CDW are, on average, consumed by interior diapycnal mixing at a rate of 5.7 Sv, eventually to be transformed into intermediate waters such as AAIW. NADW and CDW are also exported northward across 25°S in the Indo-Pacific sector at a rate of 19.5 Sv. The 9.5 Sv imbalance amounts to a steady loss of NADW and CDW from the Southern Ocean, highlighting an unsteadiness of these dense water masses in the model. Farther to the north, light surface waters drift southward and lose buoyancy, subducting as the SAMW which recirculates around the subtropical gyres of the south Indian and South Pacific Oceans.

The meridional overturning is computed as a function of density and decomposed. The bolus transport opposes a structure similar to (but in fact different from) the Deacon cell of Döös and Webb (1994). A problem, though, is that while realistically initialized, the model lacks the correct physics and surface forcing to maintain this initial state, and so drifts. Denser water masses in particular are volumetrically unsteady. As a consequence, the meridional overturning streamfunction integrated zonally on density surfaces gives a misleading impression of dense water mass transformation in the Southern Ocean. We have shown here, however, that it is relatively simple to diagnose a “transformation streamfunction” that exposes the extent of water mass unsteadiness. We believe that this streamfunction is a useful diagnostic for models that are studied while not in steady state.

The model simulates plausible formation and recirculation of SAMW and AAIW over a short (30-yr) spinup. Progressively denser versions of SAMW and AAIW form in the Indian and Pacific as the Antarctic Circumpolar Current drifts south and loses buoyancy. SAMW forms predominantly in the Indian sector, at a rate of 19.8 Sv, while AAIW forms mainly in the Pacific sector, at a rate of 8.5 Sv. The properties, distribution, and recirculation of SAMW and AAIW compare well with observations. We differentiate the effects of surface fluxes and mixing in transforming water masses in two distinct circumpolar zones. In the zone 25°–42.5°S, 11.3 Sv of SAMW is formed, largely by surface cooling, and exported northward across 25°S into the subtropical gyres. South of 42.5°S, surface buoyancy gain (due to a slight dominance of freshening over cooling) and diapycnal mixing are shown to play a roughly equal role in lightening water, forming dense SAMW and AAIW at a combined rate of 4.5 Sv.

Overall, we have established the respective roles played by bolus transports, water mass unsteadiness, and diapycnal mixing in the balances of water masses in a model Southern Ocean. Taking the unsteadiness into account, we conclude that both surface forcing and mixing play important roles in transforming water masses in the Southern Ocean.

Acknowledgments

This work has been carried out partially with the support of the Hadley Centre for Climate Prediction and Research, part of the U.K. Meteorological Office. We also gratefully acknowledge generous access to the Department of the Environment supercomputer facilities at the Hadley Centre. We would like to thank David Webb, Peter Killworth, David Marshall, and Trevor McDougall for helpful discussions and suggestions. Thanks are also due to Ian Udall, for originally developing the method used to determine trajectories from model velocity fields, to Luke West, for supporting the GIM project, and to Lisa Redbourn, for assistance with some of the figures. We are especially grateful to two anonymous reviewers for their thorough reviews and constructive comments, which enabled us to improve the paper considerably.

REFERENCES

  • Bleck, R., and L. T. Smith, 1990: A wind-driven isopycnic coordinate model of the North Atlantic and equatorial Atlantic Ocean. 1. Model development and supporting experiments. J. Geophys. Res.,95, 3273–3286.

  • ——, H. P. Hanson, D. Hu, and E. B. Krauss, 1989: Mixed-layer thermocline interaction in a three-dimensional isopycnic coordinate model. J. Phys. Oceanogr.,19, 1417–1439.

  • ——, C. Rooth, D. Hu, and L. T. Smith, 1992: Salinity-driven thermocline transients in a wind- and thermohaline-forced isopycnic coordinate model of the North Atlantic. J. Phys. Oceanogr.,22, 1486–1505.

  • Bryan, K., 1991: Ocean circulation models. Strategies for Future Climate Research, M. Latif, Ed., Max-Planck Institut für Meteorologie, 265–286.

  • Callahan, J. E., 1972: The structure and circulation of deep water in the Antarctic. Deep-Sea Res.,19, 563–575.

  • Chelton, D. B., M. G. Schlax, D. L. Witter, and J. G. Richman, 1990:Geosat altimeter observations of the surface circulation of the Southern Ocean. J. Geophys. Res.,95, 17 877–17 903.

  • Danabasoglu, G., and J. C. McWilliams, 1995: Sensitivity of the global ocean circulation to parameterizations of mesoscale tracer transports. J. Climate,8, 2967–2987.

  • ——, ——, and P. R. Gent, 1994: The role of mesoscale tracer transports in the global ocean circulation. Science,264, 1123–1126.

  • Da Silva, A. M., C. C. Young, and S. Levitus, 1994: Atlas of Surface Marine Data. Vol. 1: Algorithms and Procedures. NOAA Atlas Series, 74 pp.

  • 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.

  • ——, and A. Coward, 1997: The Southern Ocean as the major upwelling zone of North Atlantic Deep Water. Int. WOCE Newslett., No. 27, 3–4, 17.

  • Drijfhout, S. S., E. Maier-Reimer, and U. Mikolajewicz, 1996: Tracing the conveyor belt in the Hamburg large-scale geostrophic ocean general circulation model. J. Geophys. Res.,101, 22 563–22 575.

  • England, M. H., 1993: Representing the global-scale water masses in ocean general circulation models. J. Phys. Oceanogr.,23, 1523–1552.

  • ——, J. S. Godfrey, A. C. Hirst, and M. Tomczak, 1993: The mechanism for Antarctic Intermediate Water renewal in a World Ocean model. J. Phys. Oceanogr.,23, 1553–1560.

  • ——, V. Garcon, and J.-F. Minster, 1994: Chlorofluorocarbon uptake in a World Ocean model. Part 1: Sensitivity to the surface gas forcing. J. Geophys. Res.,99, 25 215–25 233.

  • Esbensen, S. K., and Y. Kushnir, 1981: The heat budget of the global ocean: An atlas based on estimates from surface marine observations. Rep. 29, Climatic Research Institute, Oregon State University, Corvallis, OR, 27 pp.

  • Fieux, M., R. Molcard, and A. G. Ilahude, 1996: Geostrophic transport of the Pacific–Indian Oceans throughflow. J. Geophys. Res.,101, 12 421–12 432.

  • The FRAM Group (D. J. Webb and others), 1991: An eddy-resolving model of the Southern Ocean. Eos, Trans. Amer. Geophys. Union,72, 169, 174–175.

  • Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr.,20, 150–155.

  • ——, F. O. Bryan, G. Danabasoglu, S. C. Doney, W. R. Holland, W. G. Large, and J. C. McWilliams, 1998: The NCAR climate system model global ocean component. J. Climate,11, 1287–1306.

  • Georgi, D. T., 1979: Modal properties of Antarctic Intermediate Water in the southeast Pacific and the South Atlantic. J. Phys. Oceanogr.,9, 456–468.

  • Gordon, A. L., 1986: Interocean exchange of thermocline water. J. Geophys. Res.,91, 5037–5046.

  • ——, E. Molinelli, and T. Baker, 1982: Southern Ocean Atlas. Columbia University Press, 34 pp. and 248 plates.

  • Haney, R. L., 1971: Surface thermal boundary conditions for ocean circulation models. J. Phys. Oceanogr.,1, 241–248.

  • Heath, R. A., 1981: Oceanic fronts around southern New Zealand. Deep-Sea Res.,28, 547–560.

  • Hellerman, S., and M. Rosenstein, 1983: Normal monthly wind stress over the world ocean with error estimates. J. Phys. Oceanogr.,13, 1093–1104.

  • Hu, D., 1996: The computation of diapycnic diffusivity and advective scalar fluxes in multilayer isopycnic-coordinate ocean models. Mon. Wea. Rev.,124, 1834–1851.

  • ——, 1997: Global-scale water masses, meridional circulation, and heat transport simulated with a global isopycnal ocean model. J. Phys. Oceanogr.,27, 96–120.

  • Jaeger, L., 1976: Monatskarten des Niedersclags für die ganze Erde. Ber. Dtsch. Wetterdienstes 18, No. 139, 38 pp.

  • Johnson, G. C., and H. L. Bryden, 1989: On the size of the Antarctic Circumpolar Current. Deep-Sea Res.,36, 39–53.

  • Josey, S. A., E. C. Kent, and P. K. Taylor, 1999: New insights into the ocean heat budget closure problem from analysis of the SOC air–sea flux climatology. J. Climate,12, 2856–2880.

  • Langlois, G., 1997: Miami Isopycnic Coordinate Ocean Model (MICOM). User’s manual. Revised and translated by D. Brydon, R. Bleck, and S. Dean. [Available online from http://www.acl.lanl.gov/CHAMMP/micom.html.].

  • Lee, M.-M., D. P. Marshall, and R. G. Williams, 1997: On the eddy transfer of tracers: Advective or diffusive. J. Mar. Res.,55, 483–505.

  • Levitus, S., 1982: Climatological Atlas of the World Ocean. NOAA Prof. Paper No. 13, U.S. Govt. Printing Office, 173 pp.

  • Manabe, S., K. Bryan, and M. L. Spelman, 1990: Transient response of a global ocean-atmosphere model to a doubling of atmospheric carbon dioxide. J. Phys. Oceanogr.,20, 722–749.

  • Marshall, D. P., 1997: Subduction of water masses in an eddying ocean. J. Mar. Res.,55, 201–222.

  • McCartney, M. S., 1977: Subantarctic mode water. A voyage of discovery, supplement to Deep-Sea Res.,24 (Suppl. to George Deacon 70th Anniversary Volume), 103–119.

  • ——, 1982: The subtropical recirculation of Mode Water. J. Mar. Res.,40 (Suppl.), 427–464.

  • McDougall, T. J., 1987: Thermobaricity, cabbeling and water-mass conversion. J. Geophys. Res.,92, 5448–5464.

  • ——, and P. C. McIntosh, 1996: The temporal-residual-mean velocity. Part I: Derivation and the scalar conservation equations. J. Phys. Oceanogr.,26, 2653–2665.

  • ——, and W. K. Dewar, 1998: Vertical mixing and cabbeling in layered models. J. Phys. Oceanogr.,28, 1458–1480.

  • McIntosh, P. C., and T. J. McDougall, 1996: Isopycnal averaging and the residual mean circulation. J. Phys. Oceanogr.,26, 1655–1660.

  • Merz, A., and G. Wüst, 1922: Die Atlantische Vertikalzirkulation. Zeitschrift der Gesellschaft für Erdkunde zu Berlin, Vol. 1922, 35 pp.

  • Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res.,13, 707–730.

  • New, A. L., R. Bleck, Y. Jia, R. Marsh, M. Huddleston, and S. Barnard, 1995: An isopycnic model study of the North Atlantic. Part I: Model experiment. J. Phys. Oceanogr.,25, 2667–2699.

  • Nurser, A. J. G., R. Marsh, and R. G. Williams, 1999: Diagnosing water mass formation from air–sea fluxes and surface mixing. J. Phys. Oceanogr.,29, 1468–1487.

  • Olbers, D., V. Gouretski, G. Seiss, and J. Schröter, 1992: Hydrographic Atlas of the Southern Ocean. Alfred Wegener Institute, 82 pp.

  • Park, Y.-H., L. Gamberoni, and E. Charriaud, 1993: Frontal structure, water masses, and circulation in the Crozet Basin. J. Geophys. Res.,98, 12 361–12 385.

  • Piola, A. R., and D. T. Georgi, 1982: Circumpolar properties of Antarctic Intermediate Water and Subantarctic Mode Water. Deep-Sea Res.,29, 687–711.

  • Reid, J. L., 1965: Intermediate waters of the Pacific Ocean. The Johns Hopkins Oceanographic Studies, No. 2, The Johns Hopkins Press, 85 pp. and 32 Figs.

  • Ribbe, J., 1997: What drives mid-latitude convection and the formation of Subantarctic Mode Water? Int. WOCE Newslett., No. 29, 20–26.

  • ——, and M. Tomczak, 1997: On convection and the formation of Subantarctic Mode Water in the Fine Resolution Antarctic Model (FRAM). J. Mar. Syst.,13, 137–154.

  • Rintoul, S. R., 1991: South Atlantic interbasin exchange. J. Geophys. Res.,96, 2675–2692.

  • ——, and J. L. Bullister, 1999: A late winter meridional section from Tasmania to Antarctica. Deep-Sea Res.,46, 1417–1454.

  • Robbins, P. E., and J. M. Toole, 1997: The dissolved silica budget as a constraint on the meridional overturning circulation of the Indian Ocean. Deep-Sea Res.,44, 879–906.

  • Saunders, P. M., and B. A. King, 1995: Oceanic fluxes on the WOCE A11 section. J. Phys. Oceanogr.,25, 1942–1958.

  • Schmitz, W. J., 1996a: On the World Ocean circulation. Vol. II: The Pacific and Indian Oceans—A global update. Woods Hole Oceanogr. Inst. Tech. Rep. WHOI-96-08, 241 pp. [Available from Woods Hole Oceanographic Institute, Woods Hole, MA 02543.].

  • ——, 1996b: On the World Ocean circulation. Vol. I: Some global features—North Atlantic circulation. Woods Hole Oceanog. Inst. Tech. Rept. WHOI-96-03, 150 pp. [Available from Woods Hole Oceanographic Institute, Woods Hole, MA 02543.].

  • Schodlok, M. P., M. Tomczak, and N. White, 1997: Deep sections through the South Australian Basin and across the Australian-Antarctic Discordance. Geophys. Res. Lett.,24, 2785–2788.

  • Smith, L. T., R. Bleck, and D. B. Boudra, 1990: A wind-driven isopycnic coordinate model of the North Atlantic and equatorial Atlantic Ocean. Part II: The Atlantic basin experiments. J. Geophys. Res.,95, 13 105–13 128.

  • Speer, K. G., 1997: A note on average cross-isopycnal mixing in the North Atlantic Ocean. Deep-Sea Res.,44, 1981–1990.

  • ——, and E. Tziperman, 1992: Rates of water mass formation in the North Atlantic Ocean. J. Phys. Oceanogr.,22, 93–104.

  • ——, S. Rintoul, and B. Sloyan, 1997: Subantarctic mode water formation by air–sea fluxes. Int. WOCE Newslett., No. 29, 29–31.

  • Stevens, D. P., and P. D. Killworth, 1992: The distribution of kinetic energy in the Southern Ocean: A comparison between observations and an eddy-resolving general circulation model. Philos. Trans. Roy. Soc. London,338B, 251–257.

  • ——, and V. Ivchenko, 1997: The zonal momentum balance in an eddy-resolving general-circulation model of the Southern Ocean. Quart. J. Roy. Meteor. Soc.,123, 929–951.

  • Stocker, T. F., and W. S. Broecker, 1994: Carbon uptake experiments with a zonally-averaged global ocean circulation model. Tellus,46B, 103–122.

  • Stommel, H., and A. Arons, 1960: On the abyssal circulation of the World Ocean. Part I: Stationary planetary flow patterns on a sphere. Deep-Sea Res.,8, 140–154.

  • Sun, S., 1997: Compressibility effects in the Miami Isopycnic Coordinate Ocean Model. Ph.D. dissertation, University of Miami, Miami, FL, 138 pp.

  • ——, R. Bleck, C. G. Rooth, J. Dukowicz, E. P. Chassignet, and P. D. Killworth, 1999: Inclusion of thermobaricity in isopycnic-coordinate ocean models. J. Phys. Oceanogr.29, 2719–2729.

  • Sverdrup, H. U., M. W. Johnson, and R. H. Fleming, 1942: The Oceans: Their Physics, Chemistry and General Biology. Prentice-Hall, 1087 pp.

  • Toggweiler, J. R., K. Dixon, and K. Bryan, 1989: Simulations of radiocarbon in a coarse-resolution World Ocean model. Part II:Distributions of bomb-produced Carbon-14. J. Geophys. Res.,94, 8243–8264.

  • Toole, J. M., and B. A. Warren, 1993: A hydrographic section across the subtropical South Indian Ocean. Deep-Sea Res.,40, 1973–2019.

  • Treguier, A. M., I. M. Held, and V. D. Larichev, 1997: On the parameterization of quasigeostrophic eddies in primitive equation ocean models. J. Phys. Oceanogr.,27, 567–580.

  • Walin, G., 1982: On the relation between sea-surface heat flow and thermal circulation in the ocean. Tellus,34, 187–195.

  • Whitworth, T., III, 1988: The Antarctic Circumpolar Current. Oceanus,31, 53–58.

  • Wunsch, C., D. Hu, and B. Grant, 1983: Mass, heat, salt and nutrient fluxes in the South Pacific Ocean. J. Phys. Oceanogr.,13, 725–753.

APPENDIX A

Evaluating the Surface Density Flux

The annual-mean surface-forced diapycnal density flux across an isopycnal σ:
i1520-0485-30-5-1013-ea1
is the integral over the surface outcrop—wherever σ is less than local mixed layer density σi,j,mml [sampled by Γ(σ,σi,j,mml), which takes the value Γ(σ,σi,j,mml) = 1 if σ < σi,j,mml]—of the surface density flux, Di,j,m, expressed in terms of heat and freshwater fluxes:
i1520-0485-30-5-1013-ea2
where α = −ρ−1(∂ρ/∂T) and β = ρ−1(∂ρ/∂S) are the thermal expansion and haline contraction coefficients; Cp and ρo are the specific heat capacity and the mean density of seawater; and Q, E, P, T, S are the net heat flux, evaporation rate, precipitation rate, surface temperature, and fractional salinity.

APPENDIX B

Evaluating the Meridional Overturning Streamfunction as a Function of Density

To compute ψj, σN), the meridional overturning streamfunction as a function of latitude ϒj, and arbitrary potential density σN, we choose 29 values of σN, including the 19 model layer densities (so that layers 2 < k < 20 take a density index 11 < n < 29), plus 10 “lighter” values (1 < n < 10) at intervals of 0.5 in the range 20.0 ⩽ σN ⩽ 24.0 (water of this density can only exist in the mixed layer of the model). The streamfunction is obtained by integrating density-partitioned layer volume transport zonally, remembering that, with a horizontal resolution of 1.25°, the model has 288 gridpoints in each zonal band:
i1520-0485-30-5-1013-eb1
where υi,j,k is the meridional component of current and hυi,j,k is the layer thickness at υ points. Here Π1(σn, σi,jml) and Π2(σn−1, σi,jml) linearly partition the contribution of mixed layer volume flux between adjacent density bins:
i1520-0485-30-5-1013-eq2
and σi,jml is the “upstream” mixed layer density, taken as either σi,j−1ml if υi,j,1 < 0 (southward flow), or σi,jml if υi,j,1 > 0 (northward flow). We can substitute the meridional component of the bolus velocity, υ*i,j,k and υi,j,k + υ*i,j,k for υi,j,k in (B1) to determine a meridional overturning implied by bolus and net (advective plus bolus) transports, respectively.

APPENDIX C

Evaluating the Meridional Density Flux Due to Lateral Mixing of Mixed Layer Density

Temperature and salinity are laterally diffused in the model mixed layer, implying a corresponding lateral flux of density. Across a latitude ϒj, ΔDlatj, σ), an instantaneous meridional flux of density across the isopycnal σ is computed as
i1520-0485-30-5-1013-ec1
where uT,Sd is a constant temperature/salinity diffusion velocity, taken as 1 cm s−1; Δxυi,j and Δyυi,j are the zonal and meridional side lengths of a gridbox at ϒj; hυi,j,1 is the mixed layer depth at ϒj; and the sampling function Γ(σ,σi,jml, σi,j−1ml) = 1 if σi,jmlσ < σi,j−1ml or σi,jmlσ > σi,j−1ml (σi,j−1ml and σi,jml are mixed layer density at adjacent gridpoints north and south of ϒj). A horizontal diffusivity of mixed layer density across ϒj, given by κmlH = uT,SdΔxυi,j, is typically 1000 m2 s−1. We are concerned here with southward fluxes (into regions of interest), hence the negative sign that precedes (C1).
As a consequence of lateral mixing of mixed layer density across ϒj, there is an annual mean total flux of density across σ, into the region south of ϒj, 〈Dlatj, σ)〉, obtained by integrating fluxes across isopycnals greater that σ and averaging monthly (index m) data over a year:
i1520-0485-30-5-1013-ec2
A corresponding transformation rate 〈Glatj, σ)〉, follows as the divergence of (C2) with respect to σ.

Fig. 1.
Fig. 1.

Year 0–30 time series of area-integrated monthly layer thickness anomalies [Δhk(t), in m] (a) south of 42.5°S and (b) in the zone 25°–42.5°S.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 2.
Fig. 2.

Year 0–30 time series of area-integrated monthly layer 6–16 temperature anomalies [ΔTk(t), in °C] (a) south of 42.5°S and (b) in the zone 25°–42.5°S.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 3.
Fig. 3.

Southern Hemisphere fields of the annual-mean surface density flux (10−6 kg m−2 s−1, roughly equivalent to a cooling rate of 15 W m−2, gray-shaded where flux is negative): (a) climatological and (b) year 30 net (climatological plus relaxation) flux and (c) thermal and (d) haline contribution to the net flux.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 4.
Fig. 4.

Year 30 austral spring (1 Sep) fields of mixed layer (a) temperature T (shaded where T < 0°C), (b) salinity S (shaded where S < 34.0 psu), and (c) currents (at every second gridpoint).

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 5.
Fig. 5.

Zonally averaged surface density flux (10−6 kg m−2 s−1) due to thermal (dotted curve), haline (dashed curve), and net forcing (solid curve): (a) climatology and (b) model year 30.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 6.
Fig. 6.

Year 30 fields of (a) maximum mixed-layer depth (H, in m, gray-shaded where H > 300 m), (b) maximum mixed-layer potential density (gray-shaded in the SAMW and AAIW density range).

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 7.
Fig. 7.

Year 29/30 (Apr–Mar) seasonal cycle of circumpolar mixed layer depth at (a) 35°S, (b) 45°S, and (c) 55°S.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 8.
Fig. 8.

Austral spring (1 Sep) thickness (m) fields for (a)–(d) layers 6, 7, 8, and 9 (σθ = 26.52, 26.80, 27.03, and 27.22, respectively) superimposed with selected trajectories (dots indicate every fifth year along each trajectory).

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 9.
Fig. 9.

Temperature and salinity fields for 1 Sep for SAMW layers 6 (a,b) and 7 (c,d) in the Indian sector. [Gray-shaded in the following temperature and salinity ranges: >13.0°C and >35.15 psu (layer 6);10.5°–13.0°C and 34.85–35.15 psu (layer 7).]

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 10.
Fig. 10.

Fields of (a) temperature and (b) salinity for 1 Sep for AAIW layer 8 (gray-shaded in the temperature and salinity ranges, 2.0°–7.0°C and 34.10–34.40 psu, respectively), (c) salinity in the upper 2000 m along 160°W from Antarctica to the equator [see dashed line in (b)], interpolated from layer salinity values and gray-shaded where salinity is less than 34.40 psu.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 11.
Fig. 11.

Density-partitioned (ML: water lighter than 24.70) transport (Sv, positive eastward/northward) (a) across 25°S in the South Atlantic, (b) south of the Cape of Good Hope, (c) across 25°S in the Indian Ocean, (d) south of Tasmania, (e) across 25°S in the South Pacific, and (f) through Drake Passage [sections are shown on map at top of figure].

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 12.
Fig. 12.

Density-partitioned annual-mean formation rates in the three Southern Ocean sectors: (a) Atlantic, (b) Indian, and (c) Pacific.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 13.
Fig. 13.

Schematic meridional sections across the Southern Ocean showing (a) G(ϒ, ρ), the diapycnal flux “down” across the ρ isopycnal (south of ϒ), V(ϒ, ρ), the volume below the ρ isopycnal and south of ϒ, and ψ(ϒ, ρ), the northward flow across ϒ of waters denser than ρ; (b) the relationship between meridional convergences of ψ and G; and (c) how the convergence of the diapycnal flux is related to water mass formation, made up of export (U δρ) and changes in the volume V(ϒ, ρ) δρ of the isopycnic layer between the isopycnals ρ and ρ + δρ.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 14.
Fig. 14.

Year 30 mean meridional overturning streamfunctions as a function of σ (south of 25°S, contoured at 2-Sv intervals, positive values gray-shaded, σ-scale doubled for σ > 27.0) and density-partitioned (as Fig. 11) annual-mean volume fluxes [U(σ) in Sv per density class, positive = northward] across 42.5°S [vertical dotted lines in (a), (c), (e)]: (a) and (b) total (mean plus bolus) transport, 〈ψ(ϒ, σ)〉 and 〈U(σ)〉; (c) and (d) mean (advective) transport, 〈ψ(ϒ, σ)〉 and 〈U(σ)〉; (e) and (f) bolus transport, 〈ψ*(ϒ, σ)〉 and 〈U*(σ)〉. In (a), (c), (e) the dashed lines indicate σmaxml(ϒ), the year-round maximum density of the mixed layer. In (d), the gray bars show the layer-by-layer transport before partition of the mixed layer flow and the dashed line indicates the split between density-partitioned “layers” that do/do not include mixed layer contributions.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 15.
Fig. 15.

Year 30 mean zonally averaged section south of 25°S showing layer interfaces and indicating layers 6 and 15. Vertical dotted line lies at 42.5°S.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 16.
Fig. 16.

As Fig. 14 but showing (a), (b) the rate-of-change in volume of water denser than σ, 〈∂V(ϒ, σ)/∂t〉, and density-partitioned rates of change (south of 42.5°S), ∂V(σ)/∂t · δσ; (c), (d) the transformation streamfunction 〈G(ϒ, σ)〉 and formation rates M and (e) the streamfunction implied by net surface density fluxes, 〈F(ϒ, σ)〉.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 17.
Fig. 17.

Transformation rates (Sv) south of 42.5°S: (a) 〈G(σ)〉 (thick solid curve), 〈F(σ)〉 (due to surface forcing, dotted curve), and 〈Gdiff(σ)〉 (due to implied mixing, gray curve); (b) thermal (solid curve) and haline (dashed curve) components of 〈F(σ)〉. (The gray dashed bar at σ = 24.70 denotes the density of the lightest thermocline waters resolved by the model.)

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 18.
Fig. 18.

Layer-by-layer formation rates south of 42.5°S, due to (a) interior diapycnal mixing, (b) total mixing, and (c) implied “surface mixing.”

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Fig. 19.
Fig. 19.

As in Fig. 17 but for the zone 25°–42.5°S.

Citation: Journal of Physical Oceanography 30, 5; 10.1175/1520-0485(2000)030<1013:WMTITS>2.0.CO;2

Table 1.

Selected water masses of the Southern Ocean (after Schmitz 1996a).

Table 1.
Table 2.

Model-layer densities.

Table 2.
Table 3.

Model SAMW and AAIW formation rates (Sv) in and around the Southern Ocean.

Table 3.

1

Treguier et al. (1997) suggest a treatment of the GM parameterization in a level model in which the mixed-layer bolus transport is set by the slope of the isopycnals immediately below the mixed layer. This only gives the same mixed layer bolus transport as MICOM where there is a density “jump” at the mixed layer base (so that the isopycnal slopes immediately below are similar to the slope of the base).

2

We include in 〈Gmix(σ)〉 a contribution Glat from diffusive fluxes southward across 42.5°S; this term is small, with a minimum (peak) value of −1.3 Sv at σ = 26.5. See appendix C.

Save
  • Bleck, R., and L. T. Smith, 1990: A wind-driven isopycnic coordinate model of the North Atlantic and equatorial Atlantic Ocean. 1. Model development and supporting experiments. J. Geophys. Res.,95, 3273–3286.

  • ——, H. P. Hanson, D. Hu, and E. B. Krauss, 1989: Mixed-layer thermocline interaction in a three-dimensional isopycnic coordinate model. J. Phys. Oceanogr.,19, 1417–1439.

  • ——, C. Rooth, D. Hu, and L. T. Smith, 1992: Salinity-driven thermocline transients in a wind- and thermohaline-forced isopycnic coordinate model of the North Atlantic. J. Phys. Oceanogr.,22, 1486–1505.

  • Bryan, K., 1991: Ocean circulation models. Strategies for Future Climate Research, M. Latif, Ed., Max-Planck Institut für Meteorologie, 265–286.

  • Callahan, J. E., 1972: The structure and circulation of deep water in the Antarctic. Deep-Sea Res.,19, 563–575.

  • Chelton, D. B., M. G. Schlax, D. L. Witter, and J. G. Richman, 1990:Geosat altimeter observations of the surface circulation of the Southern Ocean. J. Geophys. Res.,95, 17 877–17 903.

  • Danabasoglu, G., and J. C. McWilliams, 1995: Sensitivity of the global ocean circulation to parameterizations of mesoscale tracer transports. J. Climate,8, 2967–2987.

  • ——, ——, and P. R. Gent, 1994: The role of mesoscale tracer transports in the global ocean circulation. Science,264, 1123–1126.

  • Da Silva, A. M., C. C. Young, and S. Levitus, 1994: Atlas of Surface Marine Data. Vol. 1: Algorithms and Procedures. NOAA Atlas Series, 74 pp.

  • 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.

  • ——, and A. Coward, 1997: The Southern Ocean as the major upwelling zone of North Atlantic Deep Water. Int. WOCE Newslett., No. 27, 3–4, 17.

  • Drijfhout, S. S., E. Maier-Reimer, and U. Mikolajewicz, 1996: Tracing the conveyor belt in the Hamburg large-scale geostrophic ocean general circulation model. J. Geophys. Res.,101, 22 563–22 575.

  • England, M. H., 1993: Representing the global-scale water masses in ocean general circulation models. J. Phys. Oceanogr.,23, 1523–1552.

  • ——, J. S. Godfrey, A. C. Hirst, and M. Tomczak, 1993: The mechanism for Antarctic Intermediate Water renewal in a World Ocean model. J. Phys. Oceanogr.,23, 1553–1560.

  • ——, V. Garcon, and J.-F. Minster, 1994: Chlorofluorocarbon uptake in a World Ocean model. Part 1: Sensitivity to the surface gas forcing. J. Geophys. Res.,99, 25 215–25 233.

  • Esbensen, S. K., and Y. Kushnir, 1981: The heat budget of the global ocean: An atlas based on estimates from surface marine observations. Rep. 29, Climatic Research Institute, Oregon State University, Corvallis, OR, 27 pp.

  • Fieux, M., R. Molcard, and A. G. Ilahude, 1996: Geostrophic transport of the Pacific–Indian Oceans throughflow. J. Geophys. Res.,101, 12 421–12 432.

  • The FRAM Group (D. J. Webb and others), 1991: An eddy-resolving model of the Southern Ocean. Eos, Trans. Amer. Geophys. Union,72, 169, 174–175.

  • Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr.,20, 150–155.

  • ——, F. O. Bryan, G. Danabasoglu, S. C. Doney, W. R. Holland, W. G. Large, and J. C. McWilliams, 1998: The NCAR climate system model global ocean component. J. Climate,11, 1287–1306.

  • Georgi, D. T., 1979: Modal properties of Antarctic Intermediate Water in the southeast Pacific and the South Atlantic. J. Phys. Oceanogr.,9, 456–468.

  • Gordon, A. L., 1986: Interocean exchange of thermocline water. J. Geophys. Res.,91, 5037–5046.

  • ——, E. Molinelli, and T. Baker, 1982: Southern Ocean Atlas. Columbia University Press, 34 pp. and 248 plates.

  • Haney, R. L., 1971: Surface thermal boundary conditions for ocean circulation models. J. Phys. Oceanogr.,1, 241–248.

  • Heath, R. A., 1981: Oceanic fronts around southern New Zealand. Deep-Sea Res.,28, 547–560.

  • Hellerman, S., and M. Rosenstein, 1983: Normal monthly wind stress over the world ocean with error estimates. J. Phys. Oceanogr.,13, 1093–1104.

  • Hu, D., 1996: The computation of diapycnic diffusivity and advective scalar fluxes in multilayer isopycnic-coordinate ocean models. Mon. Wea. Rev.,124, 1834–1851.

  • ——, 1997: Global-scale water masses, meridional circulation, and heat transport simulated with a global isopycnal ocean model. J. Phys. Oceanogr.,27, 96–120.

  • Jaeger, L., 1976: Monatskarten des Niedersclags für die ganze Erde. Ber. Dtsch. Wetterdienstes 18, No. 139, 38 pp.

  • Johnson, G. C., and H. L. Bryden, 1989: On the size of the Antarctic Circumpolar Current. Deep-Sea Res.,36, 39–53.

  • Josey, S. A., E. C. Kent, and P. K. Taylor, 1999: New insights into the ocean heat budget closure problem from analysis of the SOC air–sea flux climatology. J. Climate,12, 2856–2880.

  • Langlois, G., 1997: Miami Isopycnic Coordinate Ocean Model (MICOM). User’s manual. Revised and translated by D. Brydon, R. Bleck, and S. Dean. [Available online from http://www.acl.lanl.gov/CHAMMP/micom.html.].

  • Lee, M.-M., D. P. Marshall, and R. G. Williams, 1997: On the eddy transfer of tracers: Advective or diffusive. J. Mar. Res.,55, 483–505.

  • Levitus, S., 1982: Climatological Atlas of the World Ocean. NOAA Prof. Paper No. 13, U.S. Govt. Printing Office, 173 pp.

  • Manabe, S., K. Bryan, and M. L. Spelman, 1990: Transient response of a global ocean-atmosphere model to a doubling of atmospheric carbon dioxide. J. Phys. Oceanogr.,20, 722–749.

  • Marshall, D. P., 1997: Subduction of water masses in an eddying ocean. J. Mar. Res.,55, 201–222.

  • McCartney, M. S., 1977: Subantarctic mode water. A voyage of discovery, supplement to Deep-Sea Res.,24 (Suppl. to George Deacon 70th Anniversary Volume), 103–119.

  • ——, 1982: The subtropical recirculation of Mode Water. J. Mar. Res.,40 (Suppl.), 427–464.

  • McDougall, T. J., 1987: Thermobaricity, cabbeling and water-mass conversion. J. Geophys. Res.,92, 5448–5464.

  • ——, and P. C. McIntosh, 1996: The temporal-residual-mean velocity. Part I: Derivation and the scalar conservation equations. J. Phys. Oceanogr.,26, 2653–2665.

  • ——, and W. K. Dewar, 1998: Vertical mixing and cabbeling in layered models. J. Phys. Oceanogr.,28, 1458–1480.

  • McIntosh, P. C., and T. J. McDougall, 1996: Isopycnal averaging and the residual mean circulation. J. Phys. Oceanogr.,26, 1655–1660.

  • Merz, A., and G. Wüst, 1922: Die Atlantische Vertikalzirkulation. Zeitschrift der Gesellschaft für Erdkunde zu Berlin, Vol. 1922, 35 pp.

  • Munk, W. H., 1966: Abyssal recipes. Deep-Sea Res.,13, 707–730.

  • New, A. L., R. Bleck, Y. Jia, R. Marsh, M. Huddleston, and S. Barnard, 1995: An isopycnic model study of the North Atlantic. Part I: Model experiment. J. Phys. Oceanogr.,25, 2667–2699.

  • Nurser, A. J. G., R. Marsh, and R. G. Williams, 1999: Diagnosing water mass formation from air–sea fluxes and surface mixing. J. Phys. Oceanogr.,29, 1468–1487.

  • Olbers, D., V. Gouretski, G. Seiss, and J. Schröter, 1992: Hydrographic Atlas of the Southern Ocean. Alfred Wegener Institute, 82 pp.

  • Park, Y.-H., L. Gamberoni, and E. Charriaud, 1993: Frontal structure, water masses, and circulation in the Crozet Basin. J. Geophys. Res.,98, 12 361–12 385.

  • Piola, A. R., and D. T. Georgi, 1982: Circumpolar properties of Antarctic Intermediate Water and Subantarctic Mode Water. Deep-Sea Res.,29, 687–711.

  • Reid, J. L., 1965: Intermediate waters of the Pacific Ocean. The Johns Hopkins Oceanographic Studies, No. 2, The Johns Hopkins Press, 85 pp. and 32 Figs.

  • Ribbe, J., 1997: What drives mid-latitude convection and the formation of Subantarctic Mode Water? Int. WOCE Newslett., No. 29, 20–26.

  • ——, and M. Tomczak, 1997: On convection and the formation of Subantarctic Mode Water in the Fine Resolution Antarctic Model (FRAM). J. Mar. Syst.,13, 137–154.

  • Rintoul, S. R., 1991: South Atlantic interbasin exchange. J. Geophys. Res.,96, 2675–2692.

  • ——, and J. L. Bullister, 1999: A late winter meridional section from Tasmania to Antarctica. Deep-Sea Res.,46, 1417–1454.

  • Robbins, P. E., and J. M. Toole, 1997: The dissolved silica budget as a constraint on the meridional overturning circulation of the Indian Ocean. Deep-Sea Res.,44, 879–906.

  • Saunders, P. M., and B. A. King, 1995: Oceanic fluxes on the WOCE A11 section. J. Phys. Oceanogr.,25, 1942–1958.

  • Schmitz, W. J., 1996a: On the World Ocean circulation. Vol. II: The Pacific and Indian Oceans—A global update. Woods Hole Oceanogr. Inst. Tech. Rep. WHOI-96-08, 241 pp. [Available from Woods Hole Oceanographic Institute, Woods Hole, MA 02543.].

  • ——, 1996b: On the World Ocean circulation. Vol. I: Some global features—North Atlantic circulation. Woods Hole Oceanog. Inst. Tech. Rept. WHOI-96-03, 150 pp. [Available from Woods Hole Oceanographic Institute, Woods Hole, MA 02543.].

  • Schodlok, M. P., M. Tomczak, and N. White, 1997: Deep sections through the South Australian Basin and across the Australian-Antarctic Discordance. Geophys. Res. Lett.,24, 2785–2788.

  • Smith, L. T., R. Bleck, and D. B. Boudra, 1990: A wind-driven isopycnic coordinate model of the North Atlantic and equatorial Atlantic Ocean. Part II: The Atlantic basin experiments. J. Geophys. Res.,95, 13 105–13 128.

  • Speer, K. G., 1997: A note on average cross-isopycnal mixing in the North Atlantic Ocean. Deep-Sea Res.,44, 1981–1990.

  • ——, and E. Tziperman, 1992: Rates of water mass formation in the North Atlantic Ocean. J. Phys. Oceanogr.,22, 93–104.

  • ——, S. Rintoul, and B. Sloyan, 1997: Subantarctic mode water formation by air–sea fluxes. Int. WOCE Newslett., No. 29, 29–31.

  • Stevens, D. P., and P. D. Killworth, 1992: The distribution of kinetic energy in the Southern Ocean: A comparison between observations and an eddy-resolving general circulation model. Philos. Trans. Roy. Soc. London,338B, 251–257.

  • ——, and V. Ivchenko, 1997: The zonal momentum balance in an eddy-resolving general-circulation model of the Southern Ocean. Quart. J. Roy. Meteor. Soc.,123, 929–951.

  • Stocker, T. F., and W. S. Broecker, 1994: Carbon uptake experiments with a zonally-averaged global ocean circulation model. Tellus,46B, 103–122.

  • Stommel, H., and A. Arons, 1960: On the abyssal circulation of the World Ocean. Part I: Stationary planetary flow patterns on a sphere. Deep-Sea Res.,8, 140–154.

  • Sun, S., 1997: Compressibility effects in the Miami Isopycnic Coordinate Ocean Model. Ph.D. dissertation, University of Miami, Miami, FL, 138 pp.

  • ——, R. Bleck, C. G. Rooth, J. Dukowicz, E. P. Chassignet, and P. D. Killworth, 1999: Inclusion of thermobaricity in isopycnic-coordinate ocean models. J. Phys. Oceanogr.29, 2719–2729.

  • Sverdrup, H. U., M. W. Johnson, and R. H. Fleming, 1942: The Oceans: Their Physics, Chemistry and General Biology. Prentice-Hall, 1087 pp.

  • Toggweiler, J. R., K. Dixon, and K. Bryan, 1989: Simulations of radiocarbon in a coarse-resolution World Ocean model. Part II:Distributions of bomb-produced Carbon-14. J. Geophys. Res.,94, 8243–8264.

  • Toole, J. M., and B. A. Warren, 1993: A hydrographic section across the subtropical South Indian Ocean. Deep-Sea Res.,40, 1973–2019.

  • Treguier, A. M., I. M. Held, and V. D. Larichev, 1997: On the parameterization of quasigeostrophic eddies in primitive equation ocean models. J. Phys. Oceanogr.,27, 567–580.

  • Walin, G., 1982: On the relation between sea-surface heat flow and thermal circulation in the ocean. Tellus,34, 187–195.

  • Whitworth, T., III, 1988: The Antarctic Circumpolar Current. Oceanus,31, 53–58.

  • Wunsch, C., D. Hu, and B. Grant, 1983: Mass, heat, salt and nutrient fluxes in the South Pacific Ocean. J. Phys. Oceanogr.,13, 725–753.

  • Fig. 1.

    Year 0–30 time series of area-integrated monthly layer thickness anomalies [Δhk(t), in m] (a) south of 42.5°S and (b) in the zone 25°–42.5°S.

  • Fig. 2.

    Year 0–30 time series of area-integrated monthly layer 6–16 temperature anomalies [ΔTk(t), in °C] (a) south of 42.5°S and (b) in the zone 25°–42.5°S.

  • Fig. 3.

    Southern Hemisphere fields of the annual-mean surface density flux (10−6 kg m−2 s−1, roughly equivalent to a cooling rate of 15 W m−2, gray-shaded where flux is negative): (a) climatological and (b) year 30 net (climatological plus relaxation) flux and (c) thermal and (d) haline contribution to the net flux.

  • Fig. 4.

    Year 30 austral spring (1 Sep) fields of mixed layer (a) temperature T (shaded where T < 0°C), (b) salinity S (shaded where S < 34.0 psu), and (c) currents (at every second gridpoint).

  • Fig. 5.

    Zonally averaged surface density flux (10−6 kg m−2 s−1) due to thermal (dotted curve), haline (dashed curve), and net forcing (solid curve): (a) climatology and (b) model year 30.

  • Fig. 6.

    Year 30 fields of (a) maximum mixed-layer depth (H, in m, gray-shaded where H > 300 m), (b) maximum mixed-layer potential density (gray-shaded in the SAMW and AAIW density range).

  • Fig. 7.

    Year 29/30 (Apr–Mar) seasonal cycle of circumpolar mixed layer depth at (a) 35°S, (b) 45°S, and (c) 55°S.

  • Fig. 8.

    Austral spring (1 Sep) thickness (m) fields for (a)–(d) layers 6, 7, 8, and 9 (σθ = 26.52, 26.80, 27.03, and 27.22, respectively) superimposed with selected trajectories (dots indicate every fifth year along each trajectory).

  • Fig. 9.

    Temperature and salinity fields for 1 Sep for SAMW layers 6 (a,b) and 7 (c,d) in the Indian sector. [Gray-shaded in the following temperature and salinity ranges: >13.0°C and >35.15 psu (layer 6);10.5°–13.0°C and 34.85–35.15 psu (layer 7).]

  • Fig. 10.

    Fields of (a) temperature and (b) salinity for 1 Sep for AAIW layer 8 (gray-shaded in the temperature and salinity ranges, 2.0°–7.0°C and 34.10–34.40 psu, respectively), (c) salinity in the upper 2000 m along 160°W from Antarctica to the equator [see dashed line in (b)], interpolated from layer salinity values and gray-shaded where salinity is less than 34.40 psu.

  • Fig. 11.

    Density-partitioned (ML: water lighter than 24.70) transport (Sv, positive eastward/northward) (a) across 25°S in the South Atlantic, (b) south of the Cape of Good Hope, (c) across 25°S in the Indian Ocean, (d) south of Tasmania, (e) across 25°S in the South Pacific, and (f) through Drake Passage [sections are shown on map at top of figure].

  • Fig. 12.

    Density-partitioned annual-mean formation rates in the three Southern Ocean sectors: (a) Atlantic, (b) Indian, and (c) Pacific.

  • Fig. 13.

    Schematic meridional sections across the Southern Ocean showing (a) G(ϒ, ρ), the diapycnal flux “down” across the ρ isopycnal (south of ϒ), V(ϒ, ρ), the volume below the ρ isopycnal and south of ϒ, and ψ(ϒ, ρ), the northward flow across ϒ of waters denser than ρ; (b) the relationship between meridional convergences of ψ and G; and (c) how the convergence of the diapycnal flux is related to water mass formation, made up of export (U δρ) and changes in the volume V(ϒ, ρ) δρ of the isopycnic layer between the isopycnals ρ and ρ + δρ.

  • Fig. 14.

    Year 30 mean meridional overturning streamfunctions as a function of σ (south of 25°S, contoured at 2-Sv intervals, positive values gray-shaded, σ-scale doubled for σ > 27.0) and density-partitioned (as Fig. 11) annual-mean volume fluxes [U(σ) in Sv per density class, positive = northward] across 42.5°S [vertical dotted lines in (a), (c), (e)]: (a) and (b) total (mean plus bolus) transport, 〈ψ(ϒ, σ)〉 and 〈U(σ)〉; (c) and (d) mean (advective) transport, 〈ψ(ϒ, σ)〉 and 〈U(σ)〉; (e) and (f) bolus transport, 〈ψ*(ϒ, σ)〉 and 〈U*(σ)〉. In (a), (c), (e) the dashed lines indicate σmaxml(ϒ), the year-round maximum density of the mixed layer. In (d), the gray bars show the layer-by-layer transport before partition of the mixed layer flow and the dashed line indicates the split between density-partitioned “layers” that do/do not include mixed layer contributions.

  • Fig. 15.

    Year 30 mean zonally averaged section south of 25°S showing layer interfaces and indicating layers 6 and 15. Vertical dotted line lies at 42.5°S.

  • Fig. 16.

    As Fig. 14 but showing (a), (b) the rate-of-change in volume of water denser than σ, 〈∂V(ϒ, σ)/∂t〉, and density-partitioned rates of change (south of 42.5°S), ∂V(σ)/∂t · δσ; (c), (d) the transformation streamfunction 〈G(ϒ, σ)〉 and formation rates M and (e) the streamfunction implied by net surface density fluxes, 〈F(ϒ, σ)〉.

  • Fig. 17.

    Transformation rates (Sv) south of 42.5°S: (a) 〈G(σ)〉 (thick solid curve), 〈F(σ)〉 (due to surface forcing, dotted curve), and 〈Gdiff(σ)〉 (due to implied mixing, gray curve); (b) thermal (solid curve) and haline (dashed curve) components of 〈F(σ)〉. (The gray dashed bar at σ = 24.70 denotes the density of the lightest thermocline waters resolved by the model.)

  • Fig. 18.

    Layer-by-layer formation rates south of 42.5°S, due to (a) interior diapycnal mixing, (b) total mixing, and (c) implied “surface mixing.”

  • Fig. 19.

    As in Fig. 17 but for the zone 25°–42.5°S.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 798 154 7
PDF Downloads 504 101 3