Skip to main content

Advertisement

Log in

Formation of subantarctic mode water in the southeastern Indian Ocean

  • Original paper
  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

Subantarctic Mode Water (SAMW) is the name given to the relatively deep surface mixed layers found directly north of the Subantarctic Front in the Southern Ocean, and their extension into the thermocline as weakly stratified or low potential vorticity water masses. The objective of this study is to begin an investigation into the mechanisms controlling SAMW formation, through a heat budget calculation. ARGO profiling floats provide estimates of temperature and salinity typically in the upper 2,000 m and the horizontal velocity at various parking depths. These data are used to estimate terms in the mode water heat budget; in addition, mode water circulation is determined with ARGO data and earlier ALACE float data, and climatological hydrography. We find a rapid transition to thicker layers in the central South Indian Ocean, at about 70°S, associated with a reversal of the horizontal eddy heat diffusion in the surface layer and the meridional expansion of the ACC as it rounds the Kerguelen Plateau. These effects are ultimately related to the bathymetry of the region, leading to the seat of formation in the region southwest of Australia. Upstream of this region, the dominant terms in the heat budget are the air–sea flux, eddy diffusion, and Ekman heat transport, all having approximately equal importance. Within the formation area, the Ekman contribution dominates and leads to a downstream evolution of mode water properties.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Notes

  1. http://www.ifremer.fr/coriolis/cdc/argo.htm and http://www.argo.ucsd.edu

  2. http://www.cdc.noaa.gov/.

  3. http://www.ifremer.fr/cersat/.

  4. http://www.ssmi.com/.

  5. http://www.aoml.noaa.gov/.

  6. The eddy diffusion coefficient has been computed using the single particle diffusivity tensor relation described by Davis (1991): \(k_{{jk}} \left( {{\rm x} = {\left\langle {v_{j} \prime {\left( {\left. {t_{0} } \right|x,\,t_{0} } \right)}d_{k} \prime {\left( {\left. {t_{0} - t} \right|{\rm x},\,t_{0} } \right)}} \right\rangle }} \right.\), where v′ and d′ are the departures from the Lagrangian mean velocity and displacement, respectively; t 0 is the initial time and the brackets indicate averaging over the ensemble of the particles; the notation a (t|x 0, t 0) represents the value of a at time t of a particle passing through x 0 at time t 0.

References

  • Brainerd KE, Gregg MC (1995) Surface mixed and mixing layer depth. Deep Sea Res, Part II 42(A):1521–1543

    Google Scholar 

  • Caldeira K, Duffy PB (2000) The role of the Southern Ocean in uptake and storage of anthropogenic carbon dioxide. Science 287:620–622

    Article  Google Scholar 

  • Chaigneau A, Morrow R (2002) Surface temperature and salinity variations between Tasmania and Antarctica, 1993–1999. J Geophys Res 107(C12):SRF22 1–9

    Article  Google Scholar 

  • Davis RE (1991) Observing the general circulation with floats. Deep Sea Res, Part II 38:S531–S571

    Article  Google Scholar 

  • Davis RE (2005) Intermediate-depth circulation of the Indian and South Pacific Oceans measured by autonomous floats. J Phys Oceanogr 35:583–707

    Article  Google Scholar 

  • Fine RA (1993) Circulation of Antarctic intermediate water in the South Indian Ocean. Deep Sea Res, Part I 40:2021–2042

    Article  Google Scholar 

  • Gill AE (1982) Atmosphere–ocean dynamics. Academic, New York

    Google Scholar 

  • Gille ST (2003a) Float observations of the Southern Ocean: part 1. Estimating mean fields, bottom velocities, and topographic steering. J Phys Oceanogr 33:1167–1181

    Article  Google Scholar 

  • Gille ST (2003b) Float observations of the Southern Ocean: part 2. Eddy fluxes. J Phys Oceanogr 33:1182–1196

    Article  Google Scholar 

  • Gille ST, Kelly KA (1996) Scales of spatial and temporal variability in the Southern Ocean. J Geophys Res 101:8759–8773

    Article  Google Scholar 

  • Gordon AL, Molinelli EJ (1982) Southern ocean atlas. Columbia Univ. Press, New York, p 200

    Google Scholar 

  • Hanawa K, Talley L (2001) In: Siedler G, Church J (eds) Ocean circulation and climate, chapter mode waters. International Geophysics Series, Academic, New York, pp 373–386

    Google Scholar 

  • Ichikawa Y, Takatsuki Y, Mizuno K, Shikama N, Takeuchi K (2002) Estimation of drifting velocity and error at parking depth for the ARGO float. Technical report, Argo Technical Report, FY2001

  • Karstensen J, Quadfasel D (2002) Formation of Southern hemisphere thermocline waters: water mass conversion and subduction. J Phys Oceanogr 32:3020–3038

    Article  Google Scholar 

  • Keffer T (1985) The ventilation of the world’s oceans: maps of the potential vorticity field. J Phys Oceanogr 15:509–523

    Article  Google Scholar 

  • Le Traon PY, Morrow R (2001) Satellite altimetry and earth sciences, chapter Ocean currents and eddies. Academic, New York, pp 171–215

    Google Scholar 

  • MacCready P, Quay P (2001) Biological export flux in the Southern Ocean estimated from a climatological nitrate budget. Deep Sea Res, Part II 49:4299–4322

    Article  Google Scholar 

  • McCarthy MC, Talley LD (1999) Three-dimensional potential vorticity structure in the Indian Ocean. J Geophys Res 104:13251–13267

    Article  Google Scholar 

  • McCartney MS (1977) Subantarctic mode water. In: Angel M (ed) A voyage of discovery, Pergamon, New York, pp 103–119

    Google Scholar 

  • McCartney MS (1982) The subtropical recirculation of mode waters. J Mar Res 40(suppl):427–464

    Google Scholar 

  • Metzl N, Tilbrook B, Poisson A (1999) The annual fCO2 cycle and the air–sea CO2 flux in the sub-Antarctic Ocean. Tellus 51B:849–861

    Google Scholar 

  • Nagata Y, Michida Y, Umimura Y (1988) Variation of position and structures of the oceanic fronts in the Indian Ocean sector of the Southern Ocean in the period from 1965 to 1987. In: D. Sahrhage (ed) Antarctic ocean and resources variability, Springer, Berlin Heidelberg New York, pp 92–98

    Google Scholar 

  • Oh IS, Zhurbas V, Park W (2000) Estimating horizontal diffusivity in the East Sea (Sea of Japan) and the Northwest Pacific from satellite-tracked drifter data. J Geophys Res 105:6483–6492

    Article  Google Scholar 

  • Orsi AH, Whitworth T III, Nowlin WD Jr (1995) On the meridional extent and fronts of the Antarctic Circumpolar Current. Deep Sea Res, Part II 42:641–673

    Article  Google Scholar 

  • Qui B (2000) Interannual variability of the Kuroshio Extension system and its impact on the wintertime SST field. J Phys Oceanogr 30:1486–1502

    Article  Google Scholar 

  • Ribbe J (1999) On wind-driven mid-latitude convection in ocean general circulation models. Tellus 51A:505–516

    Google Scholar 

  • Rintoul SR, England MH (2002) Ekman transport dominates air–sea fluxes in driving variability of Subantarctic Mode Water. J Phys Oceanogr 32:1308–1321

    Article  Google Scholar 

  • Rintoul SR, Trull TW (2001) Seasonal evolution of the mixed layer in the Subantarctic Zone south of Australia. J Geophys Res 106:31447–31462

    Article  Google Scholar 

  • Sabine C, Feely RA, Gruber N, Key RM, Lee K, Bulister JL, Wannikhof R, Wong CS, Wallace DWR, Tilbrook B, MIllero FJ, Peng T-H, Kozry A, Ono T, Rios AF (2004) The oceanic sink for anthropogenic co2. Science 305:367–371

    Article  Google Scholar 

  • Sandwell DT, Zhang B (1989) Global mesoscale variability from the Geosat Exact Repeat Mission: correlation with ocean depth. J Geophys Res 94:17971–17984

    Article  Google Scholar 

  • Sarmiento JL, Gruber N, Brezinski MA, Dunne JP (2004) High-latitude controls of thermocline nutrients and lowlatitude biological productivity. Nature 427:56–60

    Article  Google Scholar 

  • Sarmiento JL, Orr JC (1991) Three-dimensional ocean model simulations of the impact of Southern Ocean nutrient depletion on atmospheric CO2 and ocean chemistry. Limnol Oceangr 36:1928–1950

    Article  Google Scholar 

  • Schodlock MP, Tomczak M, White N (1997) Deep sections through the South Australian Basin and across the Australian–Antarctic discordance. Geophys Res Lett 24:2785–2788

    Article  Google Scholar 

  • Speer KG, Isemer HJ, Biastoch A (1995) Water mass formation from revised COADS data. J Phys Oceanogr 25:2444–2457

    Article  Google Scholar 

  • Stevenson JW, Niiler PP (1983) Upper ocean heat budget during the Hawaii to Tahiti Shuttle Experiment. J Phys Oceanogr 13:1894–1907

    Article  Google Scholar 

  • Stramma L (1992) The south Indian Ocean current. J Phys Oceanogr 22:421–430

    Article  Google Scholar 

  • Talley LD (1999) Antarctic intermediate water in the South formation of Subantarctic Mode Water in the southeastern Indian Ocean. In: Wefer G, Berger WH, Siedler G, Webb D (eds) The South Atlantic: present and past circulation, Springer, Berlin Heidelberg New York, pp 219–238

    Google Scholar 

  • Thomson RE, Fine IV (2003) Estimating mixed layer depth from oceanic profile data. J Atmos Ocean Technol 20(2):319–329

    Article  Google Scholar 

  • Wong APS, Bindoff NL, Church J (1999) Large-scale freshening of intermediate waters in the Pacific and Indian oceans. Nature 400:440–443

    Article  Google Scholar 

  • Wong APS, Johnson G, Owens W (2003) Delayed-mode calibration of autonomous CTD profiling float salinity data by theta–S climatology. J Atmos Ocean Technol 20(2):308–318

    Article  Google Scholar 

Download references

Acknowledgements

The ARGO data were collected and made freely available by the International ARGO Project and the national programs that contribute to it (http://www.argo.ucsd.edu, http://argo.jcommops.org). ARGO is a pilot program of the Global Ocean Observing System. This study was supported by the French PATOM and Coriolis programs and by NSF grants OCE 0336697 and OCE 0117618 to K. Speer.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean-Baptiste Sallée.

Additional information

Responsible Editor: Bernard Barnier

Appendix

Appendix

Appendix I: objective analysis of the float displacements

We obtained 8,849 velocity components from the ARGO database and 11,667 from the ALACE/PALACE database. We deduced from them the velocities at 400 m by adding the climatological shear between their drifting depth and 400 m. The velocities were averaged into 2° longitude by 1° latitude bins. Values formed by at least five data points are retained.

We mapped the velocity field and a stream function from the ARGO data in the southern Indian Ocean using an objective analysis following Gille (2003a).We seek an estimate \(\widehat{\psi }\) of the true streamfunction ψ using the key relation:

$$\widehat{\psi } = P{\left( {A + {{\in }} I} \right)}^{{ - 1}} \phi $$
(3)

with:

$$P = {\left\langle {\phi ,\psi } \right\rangle }$$
(4)
$$A = {\left\langle {\phi ,\phi ^{T} } \right\rangle }$$
(5)

where 〈.〉 is a scalar product, and Φ is a column vector containing all of the u and v measurements: Φ=[u 1, u 2,...u n, v 1, v 2,...v n ]. A is the covariance matrix of the measurement. We add an error ε to its diagonal, which represents the effective increase in autocovariance due to measurement noise. This simplification assumes that the measurement noise is independent between two different positions and between u and v at the same position.

As shown by Gille and Kelly (1996), in the Southern Ocean we can apply a Gaussian correlation function in space. Gille (2003a) shows that an isotropic decorrelation scale of 495 km gives the best results for an objective analysis in the Southern Ocean. In this study, we chose an isotropic decorrelation scale of 400 km.

The mean 10-day current velocities estimated from ARGO data at the parking depth contain errors, mainly due to the float drift at the surface. Furthermore, the float drift during the descent and ascent phase is also unknown. A study on the error in the drifting velocity at the float parking depth by Ichikawa et al. (2002) estimates this error to be between 10 to 25%. To be sure to overestimate this error, we chose 50% error level for the float velocities.

Appendix II: definition of the mixed layer depth

Accurate estimation of the mixed layer depth is a crucial part of the heat budget calculation. According to Thomson and Fine (2003), the threshold method with a finite difference criterion better approximates the “true” mixed layer depth, compared to the integral and regression methods. Experimental studies (see Brainerd and Gregg 1995) also find that the mixed layer depth based on a difference criterion is more stable because the gradient criterion method requires high-resolution profiles, which resolve the sharp vertical gradients. High-resolution in situ profiles are not always available, and ARGO profiles with a 10–20 m vertical resolution in the mixed layer cannot resolve sharp vertical gradients.

We chose to use a difference criterion method. The usual parameters used with this method are temperature and density; temperature is normally used when there is no equivalent salinity profiles. Most ARGO floats offer temperature and salinity profiles. We examined three different methods to find the mixed layer depth:

  • a temperature difference criterion with a threshold 0.1°C; ΔT≤0.1°C

  • a density difference criterion with a threshold 0.01 kg m−3; Δσ≤0.01 kg m−3

  • a density difference criterion with a threshold 0.03 kg m−3; Δσ≤0.03 kg m−3

The first measurement closest to the surface was chosen as the reference for each profile if this measurement is in the range from 0 to 10 m. If not the profile is rejected.

Consider first the difference between the two density criteria: Δσ≤0.01 and Δσ≤0.03. The results obtained by these two criteria are similar over most of the southern Indian Ocean (not shown). Nevertheless, some 3% of the profiles reveal mixed layer depths, which vary by more than 50%. When these exceptional profiles are examined, the 0.03 threshold is consistently associated with the mixed layer that we seek (i.e., the mixed layer depth for monthly averaged applications).

In the southern Indian Ocean, there are big differences in the mixed layer depth depending on whether we use the temperature or the density criteria. To visually highlight these differences, we calculated the average of the mixed depth layer in bins (2° latitude by 4° longitude) for each criterion and then mapped the differences. Figure 16 shows the difference between the method ΔT≤0.1 and Δσ≤0.03.

Fig. 16
figure 16

Left pannel Normalized depth differences of the mixed layer depth identified by the Δσ≤0.03 and the ΔT≤0.1 criteria. A negative difference occurs when the ΔT≤0.1 criterion is shallower than the Δσ≤0.03 criteria. Right panel Selected density profiles in the region of mode water formation (upper) and south of the SAF (lower), which show large mixed layer depth differences using the density or temperature criteria. The red points correspond to the Δσ≤0.03 criterion and the black points correspond to the ΔT≤0.1 criterion

South of the SAF, the ΔT≤0.1 criterion shows a mixed layer depth far deeper than the density criteria; here there are small vertical temperature gradients but a strong halocline near the surface separating the fresh Antarctic Surface Waters from the saltier deep waters. In the region of mode water formation east of the Kerguelen Plateau between 80 and 120°E, the temperature criterion is shallower than the density criterion where the surface T–S compensates. Figure 16 also shows selected profiles in these two areas where there is a maximum depth difference between the two methods (ΔDepth>100 m). Clearly in both areas the density difference method is much better in identifying the mixed layer depth. For the following work, we will apply the 0.03-density criterion to find the mixed layer depth.

Appendix III: detailed heat budget calculations

Ekman heat transport

For the advective term of the heat budget Eq. 1, v the horizontal mean velocity can be decomposed into the Ekman and geostrophic components as v=v E +v g .The Ekman heat transport correspond to:

$${\int {{\bf v}_{{\it{E}}} \cdot \rho C_{p} \nabla } }Tdz$$

integrated over the mixed layer. Here

$${\bf v}_{{\it{E}}} \cdot \nabla T$$

can be written:

$${\bf v}_{{\it{E}}} \cdot \nabla T = u_{{\it{E}}} \frac{{\partial T}} {{\partial x}} + v_{{\it{E}}} \frac{{\partial T}} {{\partial y}} + w_{{\it{E}}} \frac{{\partial T}} {{\partial z}}$$

Assuming that the Ekman layer is included in the mixed layer, ρC p has no dependency with the vertical and within the mixed layer \(\frac{{\partial T}}{{\partial z}} = 0\). Thus:

$${\int_{{\it{EL}}} {{\bf v}_{{\it{E}}} \cdot \rho C_{{\text{p}}} \nabla Tdz} } = \rho C_{{\text{p}}} {\left( {U_{{\it{E}}} \frac{{\partial T}} {{\partial x}} + V_{{\it{E}}} \frac{{\partial T}} {{\partial y}}} \right)}$$

In the southern Indian Ocean, the Ekman transport is mainly northward (positive) with regard to the strong westerly winds, and ∇T is dominated by the strong (positive) meridional temperature gradient. So we expect \(V_{{\rm{E}}} \frac{{\partial T}} {{\partial y}}\) to be much greater than \(U_{E} \frac{{\partial T}}{{\partial x}}\) Consequently, the Ekman induced heat flux can be approximated by \(\rho C_{{\rm{p}}} V_{{\rm{E}}} \frac{{\partial T}} {{\partial y}}\). We note from this equation that a positive Ekman heat transport (as in our case) will induce a negative temperature tendency, or cooling. In other words, a positive Ekman heat transport can induce a negative effective Ekman heat flux at the base of the Ekman layer.

Heat content variations from ARGO floats

The heat content variation is calculated for each ARGO float within the formation region, using pairs of float profiles separated by 30 days (three float cycles) that remain in our zone. The temperature is integrated down to the deepest mixed layer depth, which occurs for the two profiles.

Because the floats do not identically sample the water column, every profile has been vertically interpolated onto a regular 10-db grid. The shallowest value of each profile has also been extended to the surface.

Several techniques were applied to evaluate the heat content variation. A first possibility was to calculate a mean temperature profile for every month, which was used to calculate a mean mixed layer depth and thus a mean heat content. This method shows a poor ability to represent the cycle of the wintertime enhanced convection and was not retained. Instead, we adopted a second method where all of the 30-day heat content variations available for the different floats were binned to form the monthly average.

For each 30-day heat content calculation, the forcing terms (Ekman and air sea fluxes) are interpolated onto the 10-day float positions and then averaged over the 30 days. In other words, considering four profiles from the same float, Ekman advection and air sea fluxes are interpolated onto the first, the second, and the third profile and then averaged, and the heat content variation is computed between the first and the fourth profile. The average of the forcing terms and the variation of the heat content are stored with the date of the fourth profile. The monthly averages shown in Figs. 8 and 10 are the average of all float pairs within the formation region whose fourth profile occurs during the given month.

One problem encountered with this calculation concerns the Lagrangian behavior of our floats. The heat budget calculation assumes a local change of the mixed layer in the same location, but for our calculation, the floats move between two samples. The movement of the float from one water mass to another introduces a change of heat content due to the float’s advection. To assure that the floats remain in the same local area and water masses and to reduce this “relative advection,” we developed a series of tests.

  1. 1.

    First, we elimated pairs of profiles when their difference in SST (considering time and position) was 1°C greater than the local change of SST (at the last profile position) during the same period. This reduces the effects of floats close to the SAF crossing large SST gradients due to meanders or eddies.

  2. 2.

    We also added a deep density criteria. If the difference between the density at 1,000 m is greater than 0.1 kg m−3, we consider that the float doesn’t sample the same water mass.

  3. 3.

    We also calculated the monthly rms of the forcing terms and of the ARGO heat content variation. We then removed the individual ARGO profiles whose value plus one rms lay outside the range of the forcing terms plus one rms. We visually checked that this criterion removed most profiles crossing the SAF.

The net effect of these tests was to eliminate 36% of data pairs.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Sallée, JB., Wienders, N., Speer, K. et al. Formation of subantarctic mode water in the southeastern Indian Ocean. Ocean Dynamics 56, 525–542 (2006). https://doi.org/10.1007/s10236-005-0054-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-005-0054-x

Keywords

Navigation