<<

DECEMBER 2006 S I M E E T A L . 2253

A Decomposition of the Atlantic Meridional Overturning

LOUISE C. SIME* School of Environmental Sciences, University of East Anglia, Norwich, United Kingdom

DAVID P. STEVENS School of Mathematics, University of East Anglia, Norwich, United Kingdom

KAREN J. HEYWOOD AND KEVIN I. C. OLIVER School of Environmental Sciences, University of East Anglia, Norwich, United Kingdom

(Manuscript received 5 July 2005, in final form 5 May 2006)

ABSTRACT

A decomposition of meridional overturning circulation (MOC) cells into geostrophic vertical shears, Ekman, and bottom –dependent (or external mode) circulation components is presented. The decomposition requires the following information: 1) a density profile wherever changes to construct the vertical shears component, 2) the zonal-mean zonal stress for the Ekman component, and 3) the mean depth-independent velocity information over each isobath to construct the external mode. The decomposition is applied to the third-generation Hadley Centre Coupled General Circulation Model (HadCM3) to determine the meridional variability of these individual components within the Atlantic Ocean. The external mode component is shown to be extremely important where western boundary currents impinge on topography, and also in the area of the overflows. The explains the shape of the external mode MOC component to first order, but the time variability of the external mode exhibits only a very weak dependence on the wind stress curl. Thus, the Sverdrup balance cannot be used to determine the external mode changes when examining temporal change in the MOC. The vertical shears component allows the time-mean and the time-variable upper North Atlantic MOC cell to be deduced at 25°S and 50°N. A stronger dependency on the external mode and Ekman components between 8° and 35°N and in the regions of the overflows means that hydrographic sections need to be supplemented by bottom pressure and wind stress information at these latitudes. At the decadal time scale, variability in is less important than that in geostrophic shears. In the Southern Hemisphere the vertical shears component is dominant at all time scales, suggesting that hydrographic sections alone may be suitable for deducing change in the MOC at these latitudes.

1. Introduction and aims North Atlantic and lower Antarctic MOC cells both play a significant role in the distribution of heat within The meridional overturning circulation (MOC) in the the Atlantic Ocean system. Each cell is made up of a Atlantic Ocean strongly influences the of north- combination of -driven baroclinic and baro- western Europe, and is a key part of the planetary cli- tropic geostrophic flows, wind-driven Ekman trans- mate system (e.g., Vellinga and Wood 2002). The upper ports, and other ageostrophic components. Under- standing the contribution of each MOC component and * Current affiliation: Physical Sciences Division, British Antarc- their interactions is critical for determining how the tic Survey, Cambridge, United Kingdom. MOC may respond to changes in atmospheric forcing over a range of time scales. Zonal hydrographic sections offer a direct means of Corresponding author address: Louise C. Sime, Physical Sci- ences Division, British Antarctic Survey, Cambridge CB3 0ET, calculating MOC and heat transport in the Atlantic United Kingdom. (e.g., Bryan 1962). The basin has zero net meridional E-mail: [email protected] volume transport, apart from a small Bering Strait

© 2006 American Meteorological Society

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC

JPO2974 2254 JOURNAL OF PHYSICAL VOLUME 36 throughflow and surface freshwater fluxes. This makes MOC variability on the optimum design for a monitor- such calculations easier in the Atlantic as compared ing array at 26°N, with the purpose of detecting any with other ocean basins, where throughflows compli- trend in MOC transport (Hirschi et al. 2003; Baehr et cate the situation. However, one-off or repeat hydro- al. 2004). Here, we are interested in the components graphic zonal sections are likely to be influenced by underlying variability in the MOC at seasonal-to- seasonal and lower-frequency fluctuations in the MOC decadal time scales. We present a generalized decom- components. This study aims to clarify this question of position of the MOC and apply this to latitudes be- hydrographic section representativeness in the calcula- tween 30°S and 70°N. Additionally, three frequency tion of the MOC at all latitudes in the Atlantic. This is bands, each characterized by time variability of a dif- achieved by deriving and applying a decomposition of ferent nature, are considered in turn, generating a com- MOC components to a coupled global circulation plete picture of the variability of these components model (GCM). within both hemispheres. This provides a quantitative, Hall and Bryden (1982) decomposed heat transport if model-based, estimate of the errors associated with across a hydrographic section into contributions from the time-mean assumption in analysis of hydrographic, the barotropic, Ekman, and baroclinic components. wind, and barotropic flow fields, and has the potential This method has subsequently been used by a number to assist in identifying improvements on this approach. of authors (e.g., Saunders and King 1995; Böning et al. 2001). In this paper we further develop a decomposi- 2. Model description: HadCM3 tion originally employed by Lee and Marotzke (1998) and Baehr et al. (2004). This decomposition allows us to The ocean component of HadCM3 is a 20-level Cox- separate out geostrophic vertical shears, Ekman, and (1984) type ocean model on a 1.25° latitude ϫ 1.25° bottom pressure–dependent (or external mode) circu- longitude Arakawa B grid (Gordon et al. 2000), and is lation components. Of these components, the variabil- based on the primitive equations. The vertical levels are ity of the external mode component of the MOC is not distributed to have enhanced resolution near the ocean well known and is sometimes neglected. Consequently, surface. Each high-latitude ocean grid box can have the impact of the external mode on reference-level in- partial ice cover. The atmospheric component of formation used in observational studies of the MOC is HadCM3 also has a regular latitude–longitude grid with also not well understood (e.g., Koltermann et al. 1999). a lower horizontal resolution of 2.5° ϫ 3.75°, and 19 This decomposition represents a chance to determine hybrid coordinate levels in the vertical (Pope et al. the latitudinal variability in sensitivity to reference- 2000). The Bering Strait is closed in the model (Par- level velocity information. We present the decomposi- daens et al. 2003) ensuring no net Atlantic meridional tion in a generalized form, and in this study apply it to transport (except for insignificant surface fluxes). the zonal integral of the circulation components in a Roberts and Wood (1997) showed that Bryan–Cox- GCM. type models can be sensitive to the depth of the various GCMs are the best available tool to investigate how channels along the Greenland–Iceland–Scotland Ridge. the components of Atlantic MOC interact on seasonal Many of these channels are subgrid scale and so three and longer time scales. We use the control integration routes (one grid point wide on the velocity grid) of the third-generation Hadley Centre Coupled Ocean– through the ridge represent the total of the subgrid- Atmosphere GCM (HadCM3), a global non-flux- scale channels in HadCM3. These lead to a long-term adjusted coupled atmosphere–ocean GCM. HadCM3 mean outflow of approximately 8.5 Sv (1 Sv ϵ 106 has been shown to represent a stable realistic oceanic m3 sϪ1) of deep water from the Greenland–Iceland– and atmospheric climatology (Gordon et al. 2000). Ad- Norwegian Sea in the coupled simulation. This com- ditionally, Böning et al. (2001) noted that this class of pares to the observed outflow of around 5–6Svby Cox- (1984) type models does not differ substantially Dickson and Brown (1994), or 4–8 Sv by Bacon (1998). from other classes of models in their seasonal anoma- Convective adjustment in the region of the Denmark lies. Willebrand et al. (2001), however, showed that the Strait and Iceland–Scotland Ridge is modeled so as to annual-mean patterns of MOC transport and the struc- represent the downslope mixing of the overflow water, ture of the western boundary currents do differ. Using allowing it to find its level of neutral buoyancy instead 100 yr of the HadCM3 control run allows us to examine of convectively mixing it through the . seasonal-to-decadal variability in MOC components, Thorpe et al. (2004) find that although the density without contending with any significant model drift in structure in the Labrador Sea does depend upon the properties or behavior. specifics of how the mixing of these overflows are mod- Previous authors have focused on the impact of eled, the (global) and climate

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2255

responses are not sensitive to these details. While this ponent represents deviations from the vertical mean. If model setup is still not fully realistic, it has enabled the water column is in hydrostatic equilibrium, such HadCM3 to capture the observed separation of deep- deviations can only be caused by horizontal density gra- water sources (Gordon et al. 2000; Thorpe et al. 2001). dients. The models are coupled once per day. The atmo- Here, the cross-sectional velocity ␷ is decomposed as spheric model is run with fixed sea surface tempera- follows: tures through the day and the various forcing fluxes are ␷͑x, y, z, t͒ ϭ ␷͑y, t͒ ϩ ␷␳͑x, y, z, t͒ ϩ ␷␶͑x, y, z, t͒ accumulated each time step. At the end of the day these fluxes are passed to the ocean ϩ ␷ ͑x, y, t͒ ϩ ␷ ͑x, y, z, t͒, ͑1͒ * E model, which is then integrated forward in time (Gor- don et al. 2000). The fields used in this study are all where x, y, and z are the along-sectional, cross- monthly averages, so the highest temporal resolution sectional, and vertical (positive upward) coordinates, model behavior is not examined. respectively, and t is time; ␷ is the section-mean veloc- The model climatology is similar to that observed, ity, ␷␳ is the geostrophic vertical shears component, ␷␶ with an excellent match between model and observa- is the Ekman component, ␷ is the depth-independent * ␷ tion estimates of poleward oceanic heat transport (Coo- external mode component, and E contains small ageo- per and Gordon 2002). Additionally, the run used in strophic terms and errors resulting from imperfect sam- this study exhibits realistic low-frequency variability pling. Our terminology follows that of Lee and Ma- from El Niño–Southern Oscillation and North Atlantic rotzke (1998). Oscillation events (Collins et al. 2001). However, some The various components are most useful if their sec- salinity and drifts occur throughout the tion integrals are each equal to zero, because this fa- 1400-yr control run (Gordon et al. 2000), with some cilitates calculations of their contribution to overturn- depth dependence in the drift of these properties (Par- ing and gyre transports, as well as heat and freshwater daens et al. 2003). For this reason we restrict analysis to fluxes. For this reason, we define our components 100 yr of the run to avoid the need to detrend results. In slightly differently in comparison with those of Hall and the North Atlantic, slight cooling and freshening still Bryden (1982). The section-integrated Ekman compo- occurs over the period of the run we use, but this is nent is unlikely to be zero, but, if the section encloses a insignificant when compared with 100-yr interannual part of the ocean, volume conservation requires a bal- model variability. Other noted model deficiencies are ancing return flow. Following Baehr et al. (2004), we in the simulated front over the subtract a barotropic velocity of equal magnitude from region, which is weaker than observed, this term to obtain the Ekman component ␷␶; the im- and the zonal and meridonal wind stresses are weaker plications of this are discussed below. If the vertical than observed in midlatitudes (Dong and Sutton shears term is taken to be the zero vertical-mean geo- 2002b). The associated underestimation of the wind strophic velocity, then its column integral (and there- stress magnitude is also seen in the atmosphere-only fore its section integral) is zero. Defined in this way, ␷␳ HadAM3 model. See Pope et al. (2000) and Gordon et can be inferred from the temperature and salinity field, al. (2000) for more details about internal dynamics and without any additional information. The depth- basic climatology. independent flow is separated into a section mean ␷, and the residual once the mean is removed is ␷ . Sec- * 3. Method: A decomposition for cross-sectional tion-mean velocity ␷ is the only component that may velocities and fluxes integrate over the section to a nonzero volume trans- It is useful to decompose the velocity across any sec- port, although in the HadCM3 Atlantic Ocean, the clo- tion into geostrophic and Ekman components (Bryan sure of the Bering Strait leads to ␷ ϭ 0, simplifying our 1962), because it is reasonable to assume that tidal (av- decomposition. Using this decomposition we are able eraged over a tidal cycle), frictional, and other ageo- to quantify the role in overturning of nonuniform strophic contributions are small. The geostrophic com- depth-independent external mode, which can be an elu- ponent can then be partitioned into barotropic and sive quantity. ␷ baroclinic components (Hall and Bryden 1982). The Nonzero E may arise because of imperfect sampling barotropic component represents the vertically aver- or the presence of ageostrophic terms (other than the aged velocity, and is therefore a function of horizontal, Ekman component). In this study, however, the use of but not vertical, location, although Wunsch (1996) ar- depth-independent velocity, rather than pressure, fields gued that barotropic flow should strictly be defined as means that depth-independent ageostrophic flow is not ␷ ␷ uniform in all spatial dimensions. The baroclinic com- included in E.If is obtained using the same data and

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2256 JOURNAL OF VOLUME 36

assumptions that are used in obtaining each compo- ten in terms of a to emphasize that nent, as is typically the case in hydrographic section this response is geostrophic. We anticipate that the Ek- ␷ ϭ ␷ inversions, E 0 would be obtained. In this study, is man-compensating return flow is poorly represented as instead obtained directly from HadCM3 (which does being uniform in the along-sectional dimension. Devia- not assume geostrophic balance), and the output den- tions from this are contained in ␷ ; thus, we would * sity and velocity fields have each been monthly aver- expect correlation between ␷␶ and ␷ . Nevertheless, * ␷ aged. Thus, we anticipate that E 0, but is still small. any improvement on this approach would need to be Herein, x is taken to be eastward and y is taken to be made a posteriori. northward. The northward velocity is given by Substituting (1) into (2) and subtracting ␷␳, ␷␶, and ␷ ץ ץ ϩ ␶ Ϫ [H(z h␶) x/h␶] ( p␶ / x) leaves us with 1 Ѩp͑x, y, z, t͒ H͓z ϩ h␶͑x, y͔͒␶ ͑x, y͒ ␷ ϭ ͭͫ ͬ Ϫ x ͮ ␳ f Ѩx h␶͑x, y͒ 1 Ѩp Ѩp␶ 0 ␷ ϭ ͩ m Ϫ ͪ ϩ ␷ Ϫ ␷ ͑ ͒ ␳ Ѩ Ѩ a . 7 ϩ ␷ ͑ ͒ ϩ ␷ ͑ ͒ ͑ ͒ * 0 f x x a x, y, t E x, y, z, t , 2 where p is pressure, ␶ is the zonal wind stress, ␳ is a Because HadCM3 outputs velocities rather than x 0 ␷ reference density, f is the parameter, h␶ is the pressure gradients, the vertical-mean velocity m is a ␷ more useful variable than the vertical-mean horizontal Ekman-layer depth, and a contains any depth-inde- pendent ageostrophic flow; H is the Heaviside function pressure gradient, so we write ϩ and so H(z h␶) gives 1 in the (top 10-m Ѩ ␶ 1 pm x layer of the model) and 0 elsewhere. The geostrophic ␷ ϭ ͩ Ϫ ͪ ϩ ␷ . ͑8͒ m ␳ f Ѩx h a term can be decomposed as 0 b Substituting into (7), Ѩp Ѩp ͑x, y, t͒ z Ѩ␳͑x, y, Z, t͒ ϭ m Ϫ Ѩ Ѩ gͫ͵ Ѩ dZͬ x x Ϫ ͑ ͒ x 1 ␶ L ␶ hb x,y ␷ ϭ ͑␷ Ϫ ␷͒ ϩ ͩ x Ϫ 0 xͪ ͑ ͒ m ␳ . 9 * 0 f hb A Ѩp␳ ϩ ͩ ͪ͑x, y, t͒, ͑3͒ Ѩx The first part contains the anomaly in vertically aver- aged velocity, relative to the section mean. Subtracting where hb is the local bathymetric depth, g is the accel- the second part removes the anomaly in vertically av- ץ ץ eration resulting from gravity, pm/ x is the vertical- eraged Ekman velocity, caused by horizontal variations mean horizontal pressure gradient in the water column, in zonal wind stress or bathymetric depth. This leaves and the anomaly in depth-independent flow, the section in- 0 z ␷ ␷ Ѩp␳ g Ѩ␳͑x, y, Z, t͒ tegral of which is zero. We note that while ␶ and ␳ are ͩ ͪ ϭϪ ͫ͵ ͵ dZ dzͬ. ͑4͒ defined as the Ekman and vertical shears components, Ѩx h Ϫ Ϫ Ѩx b hb hb wind- and density gradient–driven circulation plays a strong role in determining ␷ , but their contributions If the vertical shears circulation is defined as circulation * that is directly driven by temperature and salinity gra- cannot be extracted from instantaneous velocity, den- dients, then the last two terms in (3) form this thermo- sity, and wind stress fields. haline component—the only terms containing ␳. Thus, The HadCM3 output overturning streamfunction is for the vertical shears velocity (␷␳), we can write z xe͑y,Z͒ z ␺͑y, z, t͒ ϭϪ͵ ͵ ␷͑x, y, Z, t͒ dx dZ, ͑10͒ 1 Ѩ␳͑x, y, Z, t͒ Ѩp␳ Ϫ ␷ ϭϪ ͵ Ϫ ͑ ͒ h xw͑y,Z͒ ␳ ͭͫg dZͬ ͮ. 5 ␳ f Ѩx Ѩx 0 Ϫhb where xe(y, Z) and xw(y, Z) are the locations of eastern Integrating the Ekman transport over the section and and western boundaries, respectively, at depth Z, and h dividing by area provides us with a uniform (in both the is the maximum bathymetric depth of the section. The vertical and along-sectional dimensions) velocity to be maximum in ␺ represents the Atlantic cell and is de- removed from the Ekman velocity to yield the Ekman ␺ fined as N ; the location at which this occurs is zN. The component minimum (maximum negative) value below zN repre- sents the Antarctic cell and is defined as ␺ , with loca- 1 H͑z ϩ h␶͒␶ Ѩp␶ Ѩp␶ L ␶ S ␷ ϭϪ ͫ x Ϫ ͬ ϭ 0 x ͑ ͒ ␺ ␶ ␳ Ѩ , Ѩ , 6 tion zS. Figure 1a shows the 100-yr mean . 0 f h␶ x x A Similarly, we integrate ␷␳, ␷␶, and ␷ along the sec- * ␶ where x is the zonal-mean zonal wind stress and L0 is tion and vertically to obtain the overturning stream- the basin width at the surface. The component is writ- functions of the three components,

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2257

FIG. 1. The 100-yr mean MOC diagnostics for the Atlantic Ocean in the HadCM3 control run: (a) ␺, the total overturning, (b) ␺␳, the thermohaline or vertical shears component, (c) ␺␶, ␺ , the external mode component. Gray shading and positive the Ekman component, and (d) * values indicate clockwise (looking west) overturning; white indicates anticlockwise overturn- ing. Contour intervals are 2 Sv. The overturning is not decomposed within 5° of the equator.

Ն Ϫ 1 z Z the region where hb(x) z. These components are ␺ ͑ ͒ ϭ ␳͑ Ј ͒ Ј ␳ y, z, t ␳ ͭ͵ ͫg͵ ˆ y, Z , t dZ ͬ depicted in Figs. 1b–1d. 0 f Ϫh Ϫh The above definitions of L, pˆ ␳, and ␳ˆ are based on the assumption that the section crosses a single basin, with Ϫ pˆ ␳͑y, Z, t͒ dZͮ, ͑11͒ depth monotonically increasing from 0 to h and then monotonically decreasing to 0. In practice, sections where tend to be divided into subbasins by topographic fea- tures, particularly in the deep ocean. The analysis is x ͑y,z͒ e Ѩp␳ ␳ ͑ ͒ ϭ ͵ ␳ ϭ ␳͑ ͒ Ϫ ␳͑ ͒ equally applicable in such cases. The terms pˆ ␳ and ˆ are pˆ ␳ y, z, t dx, ˆ x , z x , z , Ѩx e w xw͑y,z͒ the sum of the zonal pressure and density differences, respectively, in each subbasin, and L is the sum of sub- and h is the maximum bathymetric depth of the section, basin widths.

z The information required to construct each compo- 1 L H͑Z ϩ h␶͒␶ L͑y, Z͒L ␶ ␺ ͑ ͒ ϭ 0 x Ϫ 0 x ␶ y, z, t ␳ ͵ dZ, nent is as follows. A density profile is required wher- 0 f Ϫh h␶ A ever bathymetry changes to reconstruct ␺␳. The zonal- ␺ ͑12͒ mean zonal wind stress is required for ␶. The mean velocity [from bottom pressure, or from suitably aver- z aged lowered acoustic Doppler current profiler ␺ ͑y, z, t͒ ϭϪ͵ L͑y, Z͒ * Ϫh (LADCP) data] as well as the Ekman contribution to this, over each isobath is needed to reconstruct ␺ .A * 1 ␶ L ␶ ␺ ϫ ͭ͑␷ Ϫ ␷ ͒ ϩ ͫͩ x ͪ Ϫ 0 xͬͮ nonzero value of occurs if there is a correlation be- ˜ m m ␳ dZ, * 0 f hb A tween depth-independent flow and bathymetric depth. ͑13͒ For example, northward depth-independent flow over shallow isobaths and southward depth-independent where L(y, z) ϭ x (y, z) Ϫ x (y, z), and a tilde ϳ flow over deep isobaths would yield positive ␺ be- e w * indicates the along-sectional average of a property over tween these levels. (If isopycnal coordinates were used,

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2258 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

␺ ␺ FIG. 2. Components of the (a) North Atlantic N and (b) Antarctic MOC S with inter- and intra-annual variability. The shading shows the range of interannual variability at each lati- tude. The dotted lines show the range of seasonal variation. The depth of the maximum (c)

North Atlantic overturning zN, and (d) Antarctic overturning zS with inter- and intra-annual variability. Here (a) and (b) show observations of MK2005: McDonagh and King (2005), G2003: Ganachaud (2003b), LS2003: Lumpkin and Speer (2003), and T2003: Talley (2003).

␺ would also depend on the correlation between rents. The Antarctic cell occupies the lower portion of * depth-independent flow and isopycnal layer thickness.) the basin. It is fed by waters transformed in the South- ␺ ϭ ␺ ␺ ϭ ␺ We can now define ␳N ␳(zN), ␶N ␶(zN), ern Ocean. The northern limit of the cell varies zonally ␺ ϭ ␺ (z ), ␺␳ ϭϪ␺␳(z ), ␺␶ ϭϪ␺␶ (z ), and *N * N S S S S and temporally, but on average is at 60°N (Fig. 2b). The ␺ ϭϪ␺ (z ), where z and z are the depth of the *S * S N S horizontal boundary between the two cells is not con- maximum in North Atlantic and Antarctic overturning, stant with latitude, but on average is about 3000 m in

respectively. We also define terms representing the ef- depth (Fig. 1a and Fig. 2c). By defining zN and zS as the fect of nonzero ␷ , ␺ ϭ ␺ Ϫ (␺␳ ϩ ␺␶ ϩ ␺ ), and E EN N N N *N temporally variable vertical location of the maximum in Ϫ␺ ϭ ␺ Ϫ (␺␳ ϩ ␺␶ ϩ ␺ ). It worth noting that ES S S S *S the Atlantic (Fig. 2c) and Antarctic (Fig. 2d) cells, respec- the maximum and minimum of each overturning com- tively, we both aid investigation of the meridional and

ponent is liable to occur at z zN and z zS. Changes temporal MOC strength and structure of the cells, and in the location of the maximum overturning stream- enable comparison with observational MOC estimates. function will affect the magnitudes of the various com- Observational Atlantic MOC estimates are derived ponents as defined here. For example, if zS increases in from zonal hydrographic sections. Recent studies in- ␺ depth, the magnitude of ␶S will decrease, independent clude those from inverse-box-model studies (McDon- ␶ ␺ of any change in x, because ␶ decreases monotonically agh and King 2005; Ganachaud 2003b; Lumpkin and from 10-m depth to the bottom. These definitions are Speer 2003) and a manual adjustment of geostrophic useful in allowing a discussion of the MOC and its com- (plus Ekman) velocities for mass conservation (Talley ponents in relation to published observational estimates. 2003). Estimates of North Atlantic MOC from these studies are shown on Fig. 2a. For figure clarity, the 4. The structure of the MOC cells uncertainties associated with the estimates are not shown. When these uncertainties [of around 3–4Svfor HadCM3 depicts an Atlantic MOC with two distinct Ganachaud (2003b)] are taken into consideration, the cells, consistent with our current understanding of the estimates generally match the model overturning. Esti- Atlantic circulation. The upper Atlantic cell is fed with mates from Lumpkin and Speer (2003), calculated spe- North Atlantic Deep Water (NADW) that originates cifically for the North Atlantic, are indistinguishable from dense water masses formed in the Nordic and from the model overturning. HadCM3 therefore seems the Labrador Sea. The Nordic seas overflows entrain to represent a good approximation of our current un- overlying water while flowing southward to attain a derstanding of the time-mean North Atlantic MOC cell maximum strength of 18–22 Sv between 25° and 45°N (Gordon et al. 2000). For the Antarctic MOC (Fig. 2b) (Fig. 2a). The warmer waters of the North Atlantic Cur- the estimates of McDonagh and King (2005) and Talley rent flow north replacing these deep southward cur- (2003) agree with the model circulation. However,

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2259

those from Ganachaud (2003b) are generally weaker division of the MOC into these components against lati- than those from the HadCM3 Antarctic MOC. Salinity tude for both the North Atlantic and Antarctic cells. ␺ and temperature drifts in the Atlantic during the model The largest component of N over most latitudes is ␺ run mostly relate to a buildup of Antarctic Bottom Wa- ␳N, or the vertical shears component (Fig. 3a). We ␺ ter in the Atlantic, suggesting that the model has a calculate this to be between 97% and 145% of N for all slightly too strong Antarctic MOC (Pardaens et al. Southern Hemisphere latitudes, and between 62% and ␺ 2003). Evidence therefore suggests that the mean 101% between 35° and 62°N. Between 5° and 19°N ␶N HadCM3 Antarctic cell is probably stronger than that (the Ekman component) provides the largest contribu- in reality. tion (Fig. 3c). Between 19° and 35°N, where the Gulf ␺ The time-mean N hides considerable interannual Stream overlies shallow topography and the Deep and seasonal variability (Fig. 2a). At 26°N, the mean of Western overlies deep topography, ␺ is 16 Sv, the seasonal range is 20 Sv, and the high- the external mode (␺ ) component is dominant (Fig. N *N ␴ ␺ pass standard deviation ( ␺N) is 2 Sv. The range in 3e). It ranges between 45% and a peak of 131% of N annual-mean ␺ is 4 Sv, and the low-pass standard de- at 28°N. The shape of ␺ corresponds closely to the N *N viation is 1 Sv. These values seem smaller than the gyre (not shown), and the 28°N latitude corresponds to ␺ observed decrease in N of 30% at 26°N over 45 yr the maximum in North Atlantic gyre transport (with a found by Bryden et al. (2005), though it is worth noting 100-yr mean of 50-Sv gyre transport). that aliasing seasonal into decadal trends in HadCM3 In the region of maximum Antarctic MOC strength, ␺ ␺ could give a similar change in magnitude, so the range ␳S is the largest component of S (Fig. 3b). Where the ␺ in N in HadCM3 could be similar to that found in the Antarctic cell is weaker south of 20°S and north of 40°N Bryden et al. (2005) study. The cell center z (Fig. 2c) (Fig. 2b), the external mode ␺ is the largest propor- N *S ϳ ␺ ␺ has a 200 m range of interannual variability, and its tion of S. The Ekman ␶S component is small every- average depth is meridionally quite constant, lying be- where relative to the vertical shears and external mode tween 500 and 750 m. It becomes a little deeper at 30°N contributions.

and there is more variability associated with zN north of This division into contributions from each compo- ␺ 60°N, where the N is smaller. The Antarctic overturn- nent tells us about the information required to assess ␺ ing cell weakens north of 20°N. At 26°N S has a mean the mean strength of the MOC in the Atlantic. By se- of 7 Sv (Fig. 2b), while the seasonal range can be up to lecting latitudes that minimize Ekman or external mode 12 Sv and the annual variation up to 5 Sv. The cell contributions, we can minimize the systematic errors

depth zS (Fig. 2d) changes from 3200 to 4000 m be- that can arise either from using uncertain bottom pres- tween 8° and 16°N. sure (or velocity) data for the calculation of reference- It is unclear whether the time-varying HadCM3 cir- level velocities or from poorly known wind fields. Thus, culation matches real ocean variability, because the ob- Fig. 3 indicates that hydrographic sections, unsupported servational evidence is limited to hydrographic sections by bottom pressure (or other reference velocity in- repeated only a few times in the last 50 yr (e.g., Kol- formation) and wind stress information, around 25°S termann et al. 1999; Bryden et al. 2005). Most studies or 50°N will tend to succeed in time-mean North At- presume that hydrographic sections depict the mean lantic MOC reconstructions because the mean North ocean state, though some evidence (e.g., Thurnherr and Atlantic MOC has external mode and Ekman contri- Speer 2004) suggests that synoptic hydrographic sec- butions of less than 20% of the total overturning. In tions may not be representative of any mean ocean contrast, North Atlantic MOC reconstructions between state. But, because there are few repeat data of suffi- 8° and 35°N require the time-mean Ekman and external cient quantity, insight is limited on real ocean time vari- mode contributions to be ascertained. For the Antarctic ability, and it is difficult to assess whether HadCM3 cell, latitude of 20°N confines the non–vertical shears oceanic variability is a good representation of real contribution to less than 20%. Thus, for reconstructions ocean variability. based on long-term-averaged , careful selection of particular zonal sections appears to 5. Components of the MOCs allow 80% or greater of the mean MOC of both the North Atlantic and Antarctic cells to be deduced solely a. Time-mean results by the geostrophic vertical shears component [Eq. The decomposition presented in section 3 allows us (11)]. ␺ ␺ to ascribe the North Atlantic and Antarctic MOC over The error term EN is shown in Fig. 3g and ES is in any time scale to vertical shears, and Ekman and ex- Fig. 3h. The mean error associated with the North At- ternal mode components. Over 100 yr, Fig. 3 shows the lantic cell decomposition is 6% and the maximum is

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2260 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

FIG. 3. Components of the Atlantic and Antarctic MOC with inter- and intra-annual vari- ability. The shading shows the interannual variability at each latitude. The dotted lines show ␺ the seasonal variation. (a) North Atlantic vertical shears, ␳N; (b) Antarctic vertical shears, ␺ ␺ ␺ ␳S; (c) North Atlantic Ekman, ␶N; (d) Antarctic Ekman, ␶S; (e) North Atlantic external mode, ␺ ; (f) Antarctic external mode, ␺ ; (g) North Atlantic error term, ␺ ; and (h) *N *S EN ␺ Antarctic error term, ES.

␺ 16% at 32°N. For the Antarctic cell, the mean of ES is Meinen 2001), and the NAC separates from the coast 13% and the maximum is 35% at 57°N, where the Ant- farther north than in the real ocean. These differences arctic cell is very weak. This confirms our expectation between HadCM3 and the ocean imply that there is not ␺ ␺ that E 0, but is still small when compared with N necessarily direct equivalence between model and indi- ␺ and S. The error terms represent a combination of both ageostrophic vertical shears and errors in the cal- ABLE ␴ culation related to using monthly average model out- T 1. Mean and std dev (e.g., ␺␳N is the std dev of the North ␺ put. As such they are not analyzed here to determine Atlantic vertical shears ␳N component) of MOC components (Sv) at latitudes commonly used for hydrographic sections. the ageostrophic component, because with the current dataset we cannot separate out these two components Low pass High pass of the error term. (pentadal) (seasonal)

Table 1 (first four columns) shows the breakdown of Lat ␺␳ ␺␶ ␺ ␴␺ ␴␺ ␴␺ ␴␺ ␴␺ ␴␺ N N *N ␳N ␶N N ␳N ␶N N time-mean MOC component at latitudes where hydro- * * 30°S 18.8 Ϫ0.7 Ϫ3.7 0.5 0.2 0.2 0.7 0.9 0.9 graphic section has been repeatedly occupied. It con- 11°S 21.3 Ϫ8.3 1.7 0.6 0.3 0.2 2.2 1.0 0.6 firms that ␺␶ and ␺ are small at around 30°S and Ϫ N *N 24°N 3.5 4.3 16.5 0.7 0.3 0.7 0.9 1.5 1.3 48°N. However, we note that the 1.25° resolution of the 36°N 13.9 Ϫ1.9 7.8 1.1 0.4 0.9 0.8 1.7 0.9 model cannot produce a North Atlantic Current (NAC) 48°N 17.5 Ϫ1.1 1.2 1.9 0.3 2.0 0.3 1.0 0.5 that is as narrow as that observed (Gordon et al. 2000; 62°N 8.2 0.0 5.1 1.0 0.1 0.9 0.6 0.3 0.9

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2261

FIG. 4. Mean (a) June–August MOC streamfunction anomaly, and (b) MOC streamfunction ␺ anomaly for the 15 yr with highest basin-averaged N. Components of (a) are shown in (c) ␺ ␺ ␺ vertical shear ␳, (e) Ekman ␶, and (g) external mode *. The components of (b) are shown ␺ ␺ ␺ in (d) ␳, (f) ␶, and (h) *. Contour interval is 0.5 Sv. Gray shading and positive values indicate clockwise (looking west) overturning; white indicates anticlockwise overturning. The components are not calculated for the equator region from 5°Sto5°N.

vidual hydrographic sections, making the applicability b. Time-varying components of the MOCs of the results to inverse studies (e.g., Ganachaud 2003b; Lumpkin and Speer 2003) less straightforward. Addi- Section 5a indicates that a long-term mean of the tionally, we note that these results do not necessarily MOC can be deduced using differing sets of observa- imply that latitudes with small ␺ or ␺␶ are best for tions at different latitudes. We now address the impor- *N N MOC monitoring. The MOC is generally monitored for tant question (e.g., Bryden et al. 2005) of temporal vari- variations through time and, as Table 1 shows, a smaller ability. Figure 4a shows the mean June–August (JJA) time-mean component does not necessarily mean a seasonal anomaly of the MOC and Fig. 4b shows the smaller time-variable component. mean anomaly in MOC from the 15 yr with highest

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2262 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

␺ ␺ ␺ ␺ ␺ ␺ ␺ FIG. 5. Correlation coefficients for (a) N and ␳N, (b) S and ␳S, (c) N and ␶N, (d) S and ␺␶ , (e) ␺ and ␺ , and (f) ␺ and ␺ . The components are calculated using monthly data S N *N S *S fields and are filtered into seasonal, annual-to-pentadal, and longer-than-pentadal frequen- cies. Each set of correlations is calculated using 100 yr of control run. Any correlations that have a significance levels lower than 95% have been removed.

␺ ␺ basinwide N. The seasonal anomalous MOC largely 4e and 5c show that ␶N explains the largest part of the consists of two counter rotating full-depth cells. In con- time variability of the MOC in the Northern Hemi- trast, the annual MOC anomaly is formed of one basin- sphere at seasonal (and generally higher than pentadal) length full-depth cell. The anomalous MOC structures frequencies. For the common North Atlantic monitor- ␴ shown are supported by empirical orthogonal function ing latitudes shown in Table 1, high-pass-filtered ␺␶N is ␴ ␴ (EOF) analysis (not shown). Less than 10% of the larger than high-pass ␺ N or ␺ N, indicating that the ␳ * anomalous MOC variability is explained by EOFs that largest time variability is in the Ekman, directly wind ␺ exhibit more than one cell in the vertical dimension, driven, component. These ␶N variations tend to ex- indicating that only a small proportion of MOC vari- plain more than 80% of the seasonal MOC variation. ability is related to positively correlated North Atlantic This supports Dong and Sutton (2001) who found that and Antarctic anomalous cell overturning. The vast ma- average zonal wind forcing is the most important jority of the MOC anomaly relates to full-depth (nega- boundary condition on overturning at seasonal-to-5-yr ␺ ␺ tively correlated anomalous N and S) cells. time scales south of the overflows. On longer time ␺ Figure 4 shows the decomposition of the anomalous scales, Figs. 4d and 5a and Table 1 show that ␳N ex- MOC. Figure 5 shows bandpass-filtered correlations plains the largest proportion of MOC variation. How- ␺ between the whole North Atlantic and Antarctic cells ever, in the Southern Hemisphere ␳N is generally the and their components. For the North Atlantic cell, Figs. most important component at all time scales; thus, high-

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2263

␺ ␺ frequency hydrographic observations are necessary to yr. In the interim 1–5-yr band, ␶S and ␳S show very detect variation in the North Atlantic MOC. The ␺ similar levels of correlation with ␺ . The Antarctic ex- *N S component tends to have a larger standard deviation ternal mode contribution ␺ is significant at all fre- *S (Table 1) and a greater correlation (Figs. 4g and 5e) quencies in the subtropics of the Northern Hemisphere. ␺ with N at shorter time scales, except in the region of At subannual time scales it is also significant in the the overflows between 62° and 64°N where ␺ ex- Southern Hemisphere. *N plains most MOC variance at all time scales (Figs. 5a, 5c, and 5e). 6. The relationship between, and causes of, the Monitoring of low-frequency MOC variations (e.g., MOC components Koltermann et al. 1999; Bryden et al. 2005) requires different observations at different latitudes, this in spite We have examined the division of the North Atlantic of the fact that total low-pass MOC is actually almost and Antarctic cell strengths into vertical shears, and ␴ Ekman and external mode contributions above. It is constant at all latitudes in Table 1 (low-pass ␺N for all latitudes in Table 1 is 0.6 Sv Ϯ 0.1 Sv, not shown). clear that our vertical shears component is driven solely by baroclinic geostrophic motions, is controlled by local Low-pass ␺␳ and ␺ tend to be negatively correlated N *N in the Northern Hemisphere. This compensation be- zonal gradients in density (although these gradients tween the components appears to contribute to higher may themselves be generated by wind stress), and, in our definition, has no expression in bottom pressure. low-pass external mode standard deviation ␴␺ in the *N Northern Hemisphere relative to the Southern Hemi- Similarly, the Ekman component is solely controlled by sphere where the components are uncorrelated. Lower the zonally averaged zonal wind. The relationships be- ␴ tween these components-, and the drivers of the exter- low-pass ␺ N implies that there should be potentially * nal mode component ␺ are poorly known. Thus, it is smaller errors in MOC calculation in the Southern * Hemisphere resulting from unknown reference-level worth briefly exploring some interrelationships be- ␺ tween the components. velocities at these latitudes, because ␳N carries most of the low-pass variability. a. Ekman compensation and MOC components The ability to detect low-frequency MOC variability partly depends on the ability to filter high-frequency The question of whether Ekman-driven cells tend to (seasonal) variability in the components. Generally, it is be closed by near-surface baroclinic or barotropic flows considered easier to average or high-pass filter the Ek- has been generally examined indirectly through heat man component, because repeat wind observations are flux considerations by previous authors (e.g., Böning et easier to obtain than repeat ocean density observations. al. 2001). Over longer time scales these heat flux calcu- Strong seasonality in the hydrographic section observa- lations indicate that Ekman cells must be closed near tions therefore has the potential to bias MOC esti- the surface (e.g., Baehr et al. 2004), indicating that ␺␶ ␺ ␺ mates. Seasonal MOC variation in ␳N is lowest at 48°N and ␳ will be negatively correlated at some lower fre- (Table 1), suggesting that to minimize seasonality of quencies. Three models presented by Böning et al. ␺ ␳N this latitude is useful. However, the large contribu- (2001) (geopotential-, isopycnal-, and sigma-type mod- tion of low-frequency ␺ (Table 1) at this latitude is els) support the Gill and Niiler (1973) and Bryan (1982) *N liable to be a disadvantage. Provided that hydrographic hypothesis that the density structure does not evolve sections can be adequately high-pass filtered, or per- fast enough to compensate for the seasonal changes in haps observed in a representative season, the results Ekman transport, but can do so at longer time scales. above imply that a southern latitude, perhaps around At shorter time scales Bryan (1982) suggests that the 25°S, is potentially a good latitude for monitoring low- seasonal Ekman transport is compensated by nearly frequency MOC variations, because there is less of a depth-independent return flows. This leaves open the need to obtain bottom pressure or other reference-level possibility that Ekman compensatory flow contribu- velocity information (at least in HadCM3), and sea- tions may arise within the external mode ␺ at these * ␺ sonal variation in ␳N is smaller than at most other shorter time scales. This question of frequency and lati- latitudes (Table 1). tudinally dependent compensation is also very relevant For the Antarctic MOC the seasonal anomalies (Figs. to the question of the seasonality of vertical shears flow, ␺ 4a, 4e, and 5d) are most closely correlated with ␶S at and is therefore critical to the correct interpretation of ␺ most latitudes, with a stronger correlation with ␳S (Fig. transport estimates from one-time (or repeat) hydro- ␺ 5b) only between the equator and 20°N. The ␳S com- graphic sections at different latitudes. Thus, the time ponent becomes the most significantly correlated com- scale over which Ekman compensation occurs in either ponent at virtually all latitudes at time scales beyond 5 ␺␳ or ␺ is worth examining in HadCM3. *

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2264 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

FIG. 6. (a) Contoured correlation coefficient between pentadal (low-pass filtered) ␺␶ and ␺␳. Shaded contours with labeled white contours show regions of negative correlation; the shading interval is 0.1. Black-line (unfilled) contours show regions of positive correlation. For clarity only Ϫ0.7, Ϫ0.3, 0.3, and 0.7 are picked out. (b) Contoured correlation between seasonal ␺␶ and ␺ . Shaded contours with labeled white contours show regions of (high-pass filtered) * positive correlation; the shading interval is 0.1. Black-line (unfilled) contours show regions of negative correlation. For clarity Ϫ0.3, 0.3, and 0.5 are picked out.

Hydrographic sections are not available at the sea- sated by nearly depth-independent return flows, at least sonal time scale to allow for direct observational infer- outside the Tropics. Additional evidence comes from ences on basin-wide scales, and subbasin studies such as Figs. 4a, 4e, and 4g, where the shape of the ␺ closely * those by Meinen (2001) or Thurnherr and Speer (2004) mirrors the seasonal MOC anomaly in ␺␶, suggesting are of limited applicability to schemes that require full that the anomalous cells tend to be closed by the ex- zonal sections for closure. Figure 6a shows the correla- ternal mode. At low frequencies there are almost no tion coefficients between low-pass-filtered (pentadal regions of significant correlations between ␺␶ and ␺ * and lower frequency) ␺␶ and ␺␳. There is a clear strong (not shown); however, interestingly, at frequencies be- negative correlation in the tropical regions, particularly tween 1 and 5 yr the correlation coefficients between ␺␶ in the surface and middepth waters, indicating shallow and ␺ rise, generally maintaining the same shape as * closure of Ekman-driven surface cells at low frequen- that shown in Fig. 6b, but with R generally 0.1–0.2 cies. This correlation drops to insignificant levels for higher than the high-pass correlation. In HadCM3 the high-pass-filtered time series everywhere, except in the barotropic external mode closure of the Ekman cell upper (0–2000-m depth) waters within Ϸ8°–10° of the appears to be strongest at frequencies between 1 and equator, where R ϷϪ0.7 at z Ϸ 500 m (not shown). 5 yr. This result supports the reduction of the importance of Wind stress can generate zonal, as well as meridional, baroclinic geostrophic constraints on inversions within Ekman transport, and is a potential driver of density the upper parts of tropical hydrographic sections (as variability within the North Atlantic. However, in suggested by Ganachaud 2003a,b), because single hy- HadCM3 we find that the zonal Ekman flow generated drographic sections will be “contaminated” by these by the meridional wind stress is not significantly corre- rapid baroclinic adjustments. Mid-bandpass-filtered (1– lated at any frequency with the vertical shears or the 5-yr frequencies, not shown) correlation coefficients re- external mode MOC components at any latitude out- main high in this tropical band, and a strong band of side of the Tropics (not shown). The small regions of shallow (z Ϸ 100–500 m) negative correlation (R ϭ significant correlation in the Tropics reflect the shallow Ϫ0.5) extends north to 40°N, suggesting strong baro- wind-driven tropical cell that is present in this region. clinic Ekman closure at midlatitudes for interannual The ␺␶ component is positively correlated with the me- Ekman anomalies. Thus, mean annual meridional Ek- ridional wind stress, but this is simply a reflection of the man transports remain important over a large range of tendency for meridional and zonal components of wind latitudes. Without adequate Ekman information, the stress to increase or decrease simultaneously. Thus, it uncertainty in baroclinic transports in inverse models appears that zonal Ekman flow is not a significant con- would significantly increase. Figure 6b shows the cor- tributory component of anomalous Atlantic MOC. relation coefficients between high-pass-filtered (sea- b. The external mode sonal) ␺␶ and ␺ . Strong correlations (R Ͼ 0.7) support * the Gill and Niiler (1973) and Bryan (1982) hypothesis The external mode component of meridional over- that the high-frequency Ekman variations are compen- turning is dependent on nonuniform topography in the

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2265

boundary flow between 0 and 1000 m, we obtain the ␺ strong time-dependent Sverdrup overturning s␷N shown in Fig. 7b. Although this calculation contradicts the steady-state flat-bottomed Sverdrup assumptions, it clarifies the role played by the wind stress curl in the overturning, and demonstrates that Sverdrup transport, with non-curl-driven closure, explains the meridional shape of ␺ to first order. We note that our closure *N assumptions may not be well justified at some latitudes, ␺ which may partially explain why s␷N is much weaker than ␺ . At many latitudes a deeper southward- *N flowing western boundary current, which is not in re- ␺ sponse to a Sverdrup interior, is omitted from s␷N by the simplifying closure assumption that only permits non-Sverdrup balance transport in the upper waters. Bryden and Imawaki (2001) also find that the Sverdrup transport is much weaker than the observed transport. Thus, Fig. 7b only approximates Fig. 3e using the top- 1000-m closure assumption, although it resembles the shape (though again not magnitude) of the HadCM3 FIG. 7. Mean, interannual limits, and seasonal limits of (a) the gyre transport much more closely (not shown). Alter- zonal mean wind stress curl and (b) the overturning resulting from ␺ native calculations confining non-Sverdrup balance the component s␷N. The gray shading indicates the interannual variability at each latitude. The dashed lines show the seasonal transport to the westernmost oceanic HadCM3 grid cell variation. or grid cells (which produce a meridionally varying clo- sure depth) produce results that bear no resemblance to the external mode, and experiments using shallower or zonal direction and the depth-averaged flow that passes deeper meridionally constant closures give results with over the topography. Strong flows over variable topog- less semblance to the external mode than those in Fig. raphy, such as midlatitude gyre circulation, are liable to 7b (not shown). These results suggest that no simple produce large ␺ . For example, on the western bound- alternative closure gives a significantly better represen- * ary region between 20° and 40°N there is a particularly tation of the non-Sverdrup western boundary flows. strong ␺ associated with strong upper water current The correlations between time-varying ␺ and ␺ ␷ *N *N s N speeds over shallow topography (not shown). explain less than 10% of the total variance over most of Various authors (Lee and Marotzke 1998; Kolter- the regions where wind stress curl effects are large (Fig. mann et al. 1999) have suggested that the external 8a). The Sverdrup balance explains a larger proportion mode is, in part, a generalized version of the Sverdrup of the time variability in the regions of the tropical and relation with time dependence and important frictional subtropical gyres, with the correlation diminishing to- ␺ effects near the western boundary. Figure 7a shows that ward the poles. In the Northern Hemisphere, s␷N ex- the region of the strongest ␺ (Fig. 3e) is associated plains up to a maximum of 80% of the variance in the *N with the strongest zonally averaged wind stress curl. 1–5-yr-frequency band at 10°N. Over most of the rest of ␺ Theory suggests that response to wind stress curl at the North Atlantic s␷N does not explain more than midlatitudes should be primarily barotropic, and should Ϸ30% of the variability, and for the whole Atlantic the be rapidly transmitted to the western boundary by mean explained variability in all frequency bands is barotropic Rossby waves (Böning et al. 2001), ensuring 11%. Because the Sverdrup balance is tending to leave that the Sverdrup balance is valid over time periods of 70%–90% of the external mode variability unex- a few months. This implies that seasonal variation in the plained, it is probably not adequate to use the Sverdrup external mode should be observable in the Sverdrup balance to determine the decadal change in the external balance if it represents ␺ , though weaker stratifica- mode (the approach taken by Koltermann et al. 1999). *N tion and shallower topography at high latitudes will c. The relationship between the external mode and tend to erode the relationship north of the subtropical the vertical shears components gyre (Webb and Suginohara 2001). When we use the Sverdrup balance to calculate an Because bottom pressure gradients are the result of overturning by closing the transport with western column-integrated density, we expect a relationship be-

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2266 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

FIG. 8. Correlation coefficients for (a) ␺ and ␺ ␷ , (b) ␺ and ␺␳ , (c) ␺ and ␺␶ , and *N s N *N N *N N ␺ ␺ (d) ␳N and ␶N. The components are calculated using monthly data fields. They are filtered into seasonal, annual-to-pentadal, and longer-than-pentadal frequencies. Each set of correla- tions is calculated using 100 yr of control run. Any correlations that have a significance level lower than 95% have been removed.

tween ␺ and ␺␳ . Figure 8b shows that in the North cant negative correlations in the regions north of 20°N, *N N Atlantic the majority of ␺ variance is explained by where the NAC is confined to the western boundary of *N ␺ ␳N, indicating a significant relationship between the the Atlantic. Midbandpass correlations show a very components because they are both dependent on zonal similar pattern, though the correlations are a little gradients in density. This can be seen by considering stronger than the low-pass correlations in the tropical the approximate form of the depth-averaged regions between 15°S and 15°N. High-pass-filtered time equation,

Ѩ g 0 0 Ѩ␳ ␤␷ Ӎ ͩ ͵ ͵ dZ dzͪ * Ѩy ␳ h Ѩx 0 b Ϫhb z Ѩ 0 0 Ѩ␳ Ѩ ␶ g x Ϫ ͩ ͵ ͵ dZ dzͪ Ϫ ͩ ͪ Ѩx ␳ h Ѩy Ѩy ␳ h 0 b Ϫhb z 0 b Ѩ ␶ ϩ y ͑ ͒ Ѩ ͩ␳ ͪ. 14 x 0hb

The depth-independent horizontal flow largely depends upon the joint effect of the baroclinity, relief term, and wind stress curl. This provides a link between ␺ , ␺␳ , *N N ␺ and ␶N. Figures 8a and 8b imply that while ␺ in a time- *N mean sense may primarily be forced by wind stress curl, it is gradients in density, imposed on top of bottom topography, that are important in determining the ex- FIG. 9. The correlation between pentadal and lower-frequency (low-pass filtered) ␺ and ␺␳. Shaded contours with labeled white ternal mode component ␺ . This is supported by Fig. 9, * * contours show regions of negative correlation; the shading inter- which demonstrates a very strong negative correlation val is 0.1. Black-line (unfilled) contours show regions of positive between low-pass-filtered ␺ and ␺␳. There are signifi- correlation. For clarity only Ϫ0.9, Ϫ0.5, 0.5, and 0.9 are picked out. *

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2267

series show much weaker, mostly insignificant, correla- phy, bottom pressure (or suitably averaged LADCP), tion, except again in the tropical regions. This result and wind stress, respectively. We suggest that for two disagrees with Meinen (2001) who found a weak posi- reasons the decomposition we present is slightly pref- tive correlation of R ϭ 0.29 between baroclinic and erable to one using a barotropic ␷␳ closure. First, the barotropic northward flow (totaling Ϸ95 Sv) in the re- correlations between ␺ and ␺␳ are slightly higher, and gion at about 42°N. However, given that Meinen’s re- because the density field tends to be the best-known sults are from an open-ended short section and do not part of any hydrographic section, it is preferable to relate directly to MOC, it is difficult to interpret the know in detail the component of ␺ that explains the possible discrepancy between his results and the cur- largest part of the time variability (without being de- rent study. pendent on knowledge of bottom pressure or other ref- ⌻he gradients in density that support ␺␳ will also be erence velocities); second, this velocity decomposition dependent on the zonal expression of meridional dif- is also directly comparable with previous decomposi- ferences in surface buoyancy forcing. Detailed investi- tions (Hall and Bryden 1982; Lee and Marotzke 1998). gation of the nature of the relationship between ␺␳, ␺ , * and meridional differences in surface buoyancy forcing 7. Summary and conclusions is outside the scope of this paper [see Haines and Old (2005) for a study of thermal surface forcing and ther- We have provided the explicit observational require- mal North Atlantic oceanic variability in HadCM3]. ments of the decomposition outlined by Lee and Ma- This correlation between ␺ and ␺␳ does highlight rotzke (1998) and Baehr et al. (2004) and an examina- *N N a difficulty in whether we should use a barotropic or tion of the interdependent nature of each of the com- nonuniform flow to close the ␷␳ velocity term, and this ponents. The application of the decomposition to the decision changes the balance of velocities between the Atlantic Ocean of HadCM3 quantifies the components ␷␳ and ␷ terms. Using zero bottom velocities, that is, a at all latitudes and frequencies, and therefore the accu- * uniform barotropic assumption to close ␷␳, where ␷␳ is racy of the detection of the time-mean (e.g., Lumpkin calculated using cross-basin gradients, generates a dif- and Speer 2003) and time-variable (e.g., Bryden et al. ferent balance to the MOC decomposition. This alter- 2005; Koltermann et al. 1999) MOC cells. native method means we would obtain a difference in The division into contributions from each component the component velocity fields between the equator and indicates the information that is required to assess the ⌵ ␺ 28° . When examined in terms of N, a large propor- mean strength of the MOC cells in the Atlantic. By tion (up to 15 Sv) of the MOC passes from ␺ to ␺␳ selecting latitudes that minimize Ekman or external *N N (not shown). The Antarctic cell distribution of MOC mode contributions, we minimize the systematic errors also changes significantly. Almost all Northern Hemi- that will arise from using uncertain bottom pressure (or sphere ␺ (up to 9 Sv) switches from ␺␳ to ␺ while in reference-level velocity) data, or errors from poorly S S *S ␺ the Southern Hemisphere, most of the S MOC known wind fields. HadCM3 indicates that hydro- switches from ␺ to ␺␳ . Thus, the choice of a baro- graphic sections, unsupported by bottom pressure in- *S S tropic or a nonuniform closure for ␺ and ␺␳ makes a formation and wind stress information, at Ϸ25°S are * significant difference to the shape of the MOC decom- good for deducing both time-mean and time-varying position. Interestingly, in favor of our presented (non- MOC. This is because first, the mean North Atlantic ␺ uniform) decomposition, the correlations between N MOC has external mode and Ekman contributions of ␺ and ␳N are generally (slightly) higher and the correla- less than 20% of the total overturning; and second, the ␺ ␺ tions between S and ␳S are much higher than in the proportion of the MOC variability resulting from the case of the uniform barotropic closure with zero bottom external mode (and direct wind forcing) is also lower in velocities. Moreover, the meridionally averaged corre- the Southern Hemisphere. This implies that there lation between ␺␳ and ␺ for both cells remains ap- should be potentially smaller errors in MOC calculation * proximately equal for either method (not shown). This because of unknown reference-level velocities at these highlights that it is not possible, using either barotropic latitudes. This is despite the fact that total non-Ekman or nonuniform flow to close the ␷␳ velocity term, to MOC high-frequency variation is actually higher in the entirely divide geostrophic flow into that obtained from Southern Hemisphere. Low-frequency variation of the density gradients and that from bottom pressure gradi- external mode is relatively small in the Southern Hemi- ␺ ents, because both are dependent on density gradients sphere. Provided that low-frequency (nonseasonal) ␳N [see Eq. (14)]. This does not detract from our assertion can either be adequately represented by hydrographic that the decomposition (as outlined in section 3) quan- observations, which itself is not a trivial issue (Thurn- tifies the proportions of MOC dependent on hydrogra- herr and Speer 2004), or observed in a representative

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2268 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

season, a southern latitude around 25°S is potentially a frequencies. This has implications for the representa- good latitude for monitoring low-frequency MOC tiveness of hydrographic sections in tropical regions as variations, because it minimizes the need for bottom suggested by Ganachaud (2003a) and arguably ob- pressure or other reference-level velocity information. served by Thurnherr and Speer (2004). Midlatitude We note that this study does not include an examina- baroclinic closure becomes larger at time scales be- tion of the effect of including information from instru- tween 1 and 5 yr. At higher (seasonal) frequencies, out- ment arrays such as the Florida Strait cable (e.g., Bar- side the Tropics, the depth-independent external mode inger and Larsen 2001), which will alter the balance of tends to close the Ekman cell. Zonal Ekman flows do available information at some latitudes (Baehr et al. not correlate (at zero lag) with the vertical shears or 2004). external mode component at any latitude or frequency. For the Antarctic cell, a latitude of 20°N confines the (Except possibly for the shallow tropical cell, which we nonvertical shears contribution to less than 20%. Thus, cannot investigate here because the decomposition for reconstructions based on long-term-averaged hy- breaks down where beta effects begin to dominate drography, careful selection of particular zonal sections within Ϸ5° of the equator.) appears to allow 80% or greater of the mean MOC of Sverdrup balance explains the shape of the external both the North Atlantic and Antarctic cells to be de- mode MOC component to first order, however it is duced solely from hydrographic sections, with no addi- much weaker than the model external mode, or gyre, tional information requirements. backing up Bryden and Imawaki (2001). “Second or- The external mode component (␺ ) has been shown der” deviations in the shape of ␺ and ␺ ␷ are de- * *N s N ␺ to be extremely important where Western Boundary pendent on fluctuations in ␳N. Correlations between Currents impinge on topography, and also in the area time-varying external mode and the Sverdrup balance of the overflows. This term allows us to obtain a quan- explain less than 10% of the variance over regions titative assessment of the effect of bottom pressure in- where wind stress curl effects are large. The Sverdrup formation at these locations on different time scales. At balance does tend to explain a higher proportion of the worst case, omission can result in time-mean MOC er- time variability at lower frequencies, but even here it rors in excess of 100% for the North Atlantic in the does not explain more than 30% of the external mode 25°–32°⌵ belt and at 62°⌵. ⌻hese estimates quantify variability. Thus, it is probably not adequate to use the the importance of bottom pressure and other arrays at Sverdrup balance to determine the external mode these latitudes (e.g., Lee et al. 1996; Bacon 1998; Bar- changes for detecting decadal change in the MOC (as inger and Larsen 2001). The Antarctic cell external tried by Koltermann et al. 1999). So, while it is impor- mode is particularly important around 22°S, where it tant to include the effects of changes in the external contributes 70% of the time-mean Antarctic MOC, and mode for detecting interannual or decadal change in 75% at ϳ48°N. We note that where reference levels are the MOC, this cannot be done with the Sverdrup bal- used (e.g., Bacon 1998; Koltermann et al. 1999) to mini- ance alone. mize the impact of the external mode, error estimates For most of the Northern Hemisphere there is a very associated with the omission of ␺ will not be directly strong negative correlation between the vertical shears * comparable with the values of ␺ obtained in this study. and the external mode contribution. This is related to * We have shown that the mean seasonal and mean the confinement of currents to the western boundary interannual anomalous MOC tends consist of full-depth shelf region. When boundary currents move off the cells, with two cells for the seasonal MOC and just one shelf they change from contributing to the external cell spanning the whole North Atlantic for the mean mode to the vertical shears component. We note that if interannual anomaly. These anomalous cells explain an alternative closure is used, where the external mode 90% of the time variability of HadCM3 North Atlantic is calculated from bottom velocities (giving a uniform MOC. As Dong and Sutton (2001, 2002a) suggest, the barotropic closure to the vertical shears component), Ekman transport explains the largest part of the time then there is no change in this strong interdependence variability of the MOC in the Northern Hemisphere on of ␺ and ␺␳. Therefore, perhaps the components * seasonal-to-5-yr time scales. Over longer time scales should be seen as providing more of an observational, ␺ the vertical shears component ␳N explains most vari- rather than a dynamic, decomposition. ␺ ability. However, ␳N is the most important component Previous authors (Hirschi et al. 2003; Baehr et al. at all time scales in the Southern Hemisphere, where 2004; Vellinga and Wood 2004) have looked at the spe- variation in the Ekman transport is less important. cific instrumental setup, using predeployed (e.g., Bar- We find that wind-generated Ekman transport cells inger and Larsen 2001) and new (Hirschi et al. 2003; are closed near the surface in tropical regions at all Baehr et al. 2004) instrumentation, that could be used

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC DECEMBER 2006 S I M E E T A L . 2269 to detect change in the Atlantic MOC. We have atmosphere system to a sudden change in the Thermohaline demonstrated here, in more general terms, the obser- Circulation. Geophys. Res. Lett., 29, 1728, doi:10.1029/ vational requirements for the deduction of MOC 2002GL015229. ——, and ——, 2002b: Variability in North Atlantic heat content strengths from one-time or repeat hydrographics sec- and heat transport in a coupled ocean–atmosphere GCM. tions, without addressing these same instrumental is- Climate Dyn., 19, 485–497. sues. Future work will use inversions to investigate how Ganachaud, A. S., 2003a: Error budget of inverse box models: including additional assumptions and information af- The North Atlantic. J. Atmos. Oceanic Technol., 20, 1641– fects the deduction of decadal changes in MOC, heat, 1655. ——, 2003b: Large-scale mass transports, water mass formation, and freshwater transport. and diffusivities estimated from World Ocean Circulation Experiment (WOCE) hydrographic data. J. Geophys. Res., Acknowledgments. We thank the Natural Environ- 108, 3213, doi:10.1029/2002JC001565. mental Research Council for funding this work through Gill, A. E., and P. P. Niiler, 1973: The theory of the seasonal the Coupled Ocean-Atmosphere Processes and Euro- variability in the ocean. Deep-Sea Res., 20, 141–178. pean Climate (COAPEC) Programme, Grant NER/T/ Gordon, C., C. Cooper, C. A. Senior, H. Banks, J. M. Gregory, T. C. Johns, J. F. B. Mitchell, and R. A. Wood, 2000: The S/2000/00303. simulation of SST, sea ice extents and ocean heat transports in a version of the Hadley Centre coupled model without flux REFERENCES adjustments. Climate Dyn., 16 (2–3), 147–168. Haines, K., and C. Old, 2005: Diagnosing natural variability of Bacon, S., 1998: Decadal variability in the outflow from the Nor- North Atlantic water masses in HadCM3. J. Climate, 18, dic seas to the Deep Atlantic Ocean. Nature, 394, 871–874. 1925–1941. Baehr, J., J. Hirschi, J. O. Beismann, and J. Marotzke, 2004: Moni- Hall, M. M., and H. L. Bryden, 1982: Direct estimates and mecha- toring the meridional overturning circulation in the North nisms of ocean heat transport. Deep-Sea Res., 29, 339–359. Atlantic: A model-based array design study. J. Mar. Res., 62, Hirschi, J., J. Baehr, J. Marotzke, J. Stark, S. Cunningham, and 283–312. J. O. Beismann, 2003: A monitoring design for the Atlantic Baringer, M. O., and J. C. Larsen, 2001: Sixteen years of Florida meridional overturning circulation. Geophys. Res. Lett., 30, Current transport at 27°N. Geophys. Res. Lett., 28, 3179– 1413, doi:10.1029/2002GL016776. 3182. Koltermann, K. P., A. V. Sokov, V. Tereschenkov, S. A. Dobro- Böning, C. W., C. Dieterich, B. Barnier, and Y. L. Jia, 2001: Sea- liubov, K. Loracher, and A. Sy, 1999: Decadal changes in the sonal cycle of meridional heat transport in the subtropical thermohaline circulation of the North Atlantic. Deep-Sea North Atlantic: A model intercomparison in relation to ob- Res. II, 46 (1–2), 109–138. servations near 25°N. Progress in Oceanography, Vol. 48, Lee, T., and J. Marotzke, 1998: Seasonal cycles of meridional Pergamon, 231–253. overturning and heat transport of the Indian Ocean. J. Phys. Bryan, K., 1962: Measurements of meridional heat transport by Oceanogr., 28, 923–943. ocean currents. J. Geophys. Res., 67, 3403–3414. Lee, T. N., W. E. Jones, R. J. Zantopp, and E. R. Fillenbaum, ——, 1982: Seasonal variation in meridional overturning and pole- 1996: Moored observations of western boundary current vari- ward heat transport in the Atlantic and Pacific : A ability and thermohaline circulation at 26.5°N in the subtropi- model study. J. Mar. Res., 40, 39–53. cal North Atlantic. J. Phys. Oceanogr., 26, 962–983. Bryden, H. L., and S. Imawaki, 2001: Ocean heat transport. Ocean Lumpkin, R., and K. Speer, 2003: Large-scale vertical and hori- Circulation and Climate: Observing and Modelling the Global zontal circulation in the North Atlantic Ocean. J. Phys. Ocean, G. Siedler, J. Church, and J. Gould, Eds., Academic Oceanogr., 33, 1902–1920. Press, 455–474. McDonagh, E. L., and B. A. King, 2005: Oceanic fluxes in the ——, H. R. Longworth, and S. A. Cunningham, 2005: Slowing of South Atlantic. J. Phys. Oceanogr., 35, 109–122. the Atlantic meridional overturning circulation at 25 degrees Meinen, C. S., 2001: Structure of the North Atlantic current in N. Nature, 438, 655–657. stream-coordinates and the circulation in the Newfoundland Collins, M., S. F. B. Tett, and C. Cooper, 2001: The internal cli- basin. Deep-Sea Res., 48, 1553–1580. mate variability of HadCM3, a version of the Hadley Centre Pardaens, A. K., H. T. Banks, J. M. Gregory, and P. R. Rowntree, coupled model without flux adjustments. Climate Dyn., 17, 2003: Freshwater transports in HadCM3. Climate Dyn., 21, 61–81. 177–195. Cooper, C., and C. Gordon, 2002: North Atlantic oceanic decadal Pope, V. D., M. L. Gallani, P. R. Rowntree, and R. A. Stratton, variability in the Hadley Centre coupled model. J. Climate, 2000: The impact of new physical parametrizations in the 15, 45–72. Hadley Centre climate model: HadAM3. Climate Dyn., 16 Cox, M. D., 1984: A primitive equation 3-dimensional model of (2–3), 123–146. the ocean. GFDL Ocean Group Tech. Rep. 1, 143 pp. Roberts, M. J., and R. A. Wood, 1997: Topographic sensitivity Dickson, R. R., and J. Brown, 1994: The production of North studies with a Bryan–Cox-type ocean model. J. Phys. Ocean- Atlantic deep water: Sources, rates, and pathways. J. Geo- ogr., 27, 823–836. phys. Res., 99 (C6), 12 319–12 341. Saunders, P. M., and B. A. King, 1995: Oceanic fluxes on the Dong, B. W., and R. T. Sutton, 2001: The dominant mechanisms WOCE A11 section. J. Phys. Oceanogr., 25, 1942–1958. of variability in Atlantic Ocean heat transport in a coupled Talley, L. D., 2003: Shallow, intermediate, and deep overturning ocean-atmosphere GCM. Geophys. Res. Lett., 28, 2445–2448. component of the global heat budget. J. Phys. Oceanogr., 33, ——, and ——, 2002a: Adjustment of the coupled ocean– 530–560.

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC 2270 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 36

Thorpe, R. B., J. M. Gregory, T. C. Johns, R. A. Wood, and J. F. collapse of Atlantic thermohaline circulation. Climate B. Mitchell, 2001: Mechanisms determining Atlantic thermo- Change, 54 (3), 251–267. haline circulation response to greenhouse gas forcing in a ——, and ——, 2004: Timely detection of anthropogenic change non-flux-adjusted coupled climate model. J. Climate, 14, in the Atlantic meridional overturning circulation. Geophys. 3102–3116. Res. Lett., 31, L14203, doi:10.1029/2004GL020306. ——, R. A. Wood, and J. F. B. Mitchell, 2004: Sensitivity of the Webb, D. J., and N. Suginohara, 2001: The interior circulation of the ocean. Ocean Circulation and Climate: Observing and modelled thermohaline circulation to the parameterisation of Modelling the Global Ocean, mixing across the Greenland–Scotland ridge. Ocean Modell., G. Siedler, J. Church, and J. Gould, Eds., Academic Press, 205–214. 7 (3–4), 259–268. Willebrand, J., and Coauthors, 2001: Circulation characteristics in Thurnherr, A. M., and K. G. Speer, 2004: Representativeness of three -permitting models of the North Atlantic. Progress meridional hydrographic sections in the western South At- in Oceanography, Vol. 48, Pergamon, 123–161. lantic. J. Mar. Res., 62, 37–65. Wunsch, C., 1996: The Ocean Circulation Inverse Problem. Cam- Vellinga, M., and R. A. Wood, 2002: Global climate impacts of a bridge University Press, 442 pp.

Unauthenticated | Downloaded 09/24/21 07:44 PM UTC