noaa's atlantic oceanographic and meteorological laboratory - … · 2009-10-14 · the...

19
1902 VOLUME 33 JOURNAL OF PHYSICAL OCEANOGRAPHY q 2003 American Meteorological Society Large-Scale Vertical and Horizontal Circulation in the North Atlantic Ocean RICK LUMPKIN Cooperative Institute for Marine and Atmospheric Studies, University of Miami, Miami, Florida KEVIN SPEER Department of Oceanography, The Florida State University, Tallahassee, Florida (Manuscript received 4 March 2002, in final form 25 February 2003) ABSTRACT Observations of large-scale hydrography, air–sea forcing, and regional circulation from numerous studies are combined by inverse methods to determine the basin-scale circulation, average diapycnal mixing, and adjustments to air–sea forcing of the North Atlantic Ocean. Dense overflows through the Denmark Strait and Faroe Bank channels are explicitly included and are associated with strong vertical and lateral circulation and mixing. These processes in the far northern Atlantic play a fundamental role in the meridional overturning circulation for the entire ocean, accompanied by an upper cell of mode-water and intermediate-water circulation. The two cells converge roughly at the mean depth of the midocean ridge crest. The Labrador Sea Water layer lies within this convergence. South of the overflow region, model-derived mean diapycnal diffusivities are O(10 25 m 2 s 21 ) or smaller at the base of the thermocline, and diapycnal advection is driven primarily by air–sea transformation on outcropping layers. 1. Introduction The meridional circulation in the North Atlantic Ocean results from the interaction of several primary elements: western boundary currents (WBCs), the mod- ification of northward-flowing Antarctic Intermediate and Bottom Waters (AAIW and AABW, respectively), and air–sea formation of subtropical and subpolar mode waters (STMW and SPMW, respectively) and Labrador Sea Water (LSW). Another primary element of merid- ional flow is the overflow of dense water from the Greenland–Iceland–Norwegian (GIN) Seas that ulti- mately forms the dense core of lower North Atlantic Deep Water (NADW). The overflow component has a long history of observation, and recent work (cf. Girton et al. 2001) goes much further toward a quantification of mean and variable outflow. But this component has not been incorporated into large-scale estimates of cir- culation except indirectly via tracer budgets (Smethie et al. 2000; Orsi et al. 2002) or as bulk transfers in coarse box models (Worthington 1976; McCartney and Talley 1982; McCartney 1992; Schmitz and McCartney 1993; Dickson and Brown 1994). The replenishment of subtropical and subpolar mode waters is intimately tied to the strength of overturning, but also to diapycnal mixing, so concurrent estimates Corresponding author address: Rick Lumpkin, NOAA/AOML/ PhOD, 4301 Rickenbacker Cswy., Miami, FL 33149-1097. E-mail: [email protected] of both flow and mixing are needed. How much mid- depth deep water is supplied from below via upwelling of dense Antarctic Bottom Water and overflow, as com- pared with the supply from above, via entrainment and convection in the mixed layer? Diapycnal mixing is nec- essary to form and support the NADW outflow, but how is it distributed? We answer these questions with circulation estimates based on the box inverse model framework, which com- bines observations from hydrographic sections, current meters, and air–sea flux estimates. Unknown features of the circulation are inferred from prior property flux con- vergences. Absolute meridional transports are thus sen- sitive to boundary conditions at either end of the model. While many models have used net constraints set by the Pacific-to-Arctic Bering Strait exchange, no basin- or global-scale inverse model has taken advantage of the more detailed observations which can constrain mean exchanges between the North Atlantic and GIN Seas through the Greenland–Iceland–Scotland (GIS) choke point. Early models (Roemmich and Wunsch 1985; Rin- toul and Wunsch 1991) focused on the tropical cell, subtropical gyre, and the development of the model framework, while more recent global models (Macdon- ald 1998, hereinafter M98; Ganachaud 1999, hereinafter G99) extend only to 488N in the Atlantic—overturning across this latitude then becomes more sensitive to the choice of reference level since no explicit information closes the streamfunction farther north except bulk mass conservation. Because much of this transformation is

Upload: others

Post on 15-Aug-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1902 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

q 2003 American Meteorological Society

Large-Scale Vertical and Horizontal Circulation in the North Atlantic Ocean

RICK LUMPKIN

Cooperative Institute for Marine and Atmospheric Studies, University of Miami, Miami, Florida

KEVIN SPEER

Department of Oceanography, The Florida State University, Tallahassee, Florida

(Manuscript received 4 March 2002, in final form 25 February 2003)

ABSTRACT

Observations of large-scale hydrography, air–sea forcing, and regional circulation from numerous studies arecombined by inverse methods to determine the basin-scale circulation, average diapycnal mixing, and adjustmentsto air–sea forcing of the North Atlantic Ocean. Dense overflows through the Denmark Strait and Faroe Bankchannels are explicitly included and are associated with strong vertical and lateral circulation and mixing. Theseprocesses in the far northern Atlantic play a fundamental role in the meridional overturning circulation for theentire ocean, accompanied by an upper cell of mode-water and intermediate-water circulation. The two cellsconverge roughly at the mean depth of the midocean ridge crest. The Labrador Sea Water layer lies within thisconvergence. South of the overflow region, model-derived mean diapycnal diffusivities are O(1025 m2 s21) orsmaller at the base of the thermocline, and diapycnal advection is driven primarily by air–sea transformationon outcropping layers.

1. Introduction

The meridional circulation in the North AtlanticOcean results from the interaction of several primaryelements: western boundary currents (WBCs), the mod-ification of northward-flowing Antarctic Intermediateand Bottom Waters (AAIW and AABW, respectively),and air–sea formation of subtropical and subpolar modewaters (STMW and SPMW, respectively) and LabradorSea Water (LSW). Another primary element of merid-ional flow is the overflow of dense water from theGreenland–Iceland–Norwegian (GIN) Seas that ulti-mately forms the dense core of lower North AtlanticDeep Water (NADW). The overflow component has along history of observation, and recent work (cf. Girtonet al. 2001) goes much further toward a quantificationof mean and variable outflow. But this component hasnot been incorporated into large-scale estimates of cir-culation except indirectly via tracer budgets (Smethieet al. 2000; Orsi et al. 2002) or as bulk transfers incoarse box models (Worthington 1976; McCartney andTalley 1982; McCartney 1992; Schmitz and McCartney1993; Dickson and Brown 1994).

The replenishment of subtropical and subpolar modewaters is intimately tied to the strength of overturning,but also to diapycnal mixing, so concurrent estimates

Corresponding author address: Rick Lumpkin, NOAA/AOML/PhOD, 4301 Rickenbacker Cswy., Miami, FL 33149-1097.E-mail: [email protected]

of both flow and mixing are needed. How much mid-depth deep water is supplied from below via upwellingof dense Antarctic Bottom Water and overflow, as com-pared with the supply from above, via entrainment andconvection in the mixed layer? Diapycnal mixing is nec-essary to form and support the NADW outflow, but howis it distributed?

We answer these questions with circulation estimatesbased on the box inverse model framework, which com-bines observations from hydrographic sections, currentmeters, and air–sea flux estimates. Unknown features ofthe circulation are inferred from prior property flux con-vergences. Absolute meridional transports are thus sen-sitive to boundary conditions at either end of the model.While many models have used net constraints set by thePacific-to-Arctic Bering Strait exchange, no basin- orglobal-scale inverse model has taken advantage of themore detailed observations which can constrain meanexchanges between the North Atlantic and GIN Seasthrough the Greenland–Iceland–Scotland (GIS) chokepoint. Early models (Roemmich and Wunsch 1985; Rin-toul and Wunsch 1991) focused on the tropical cell,subtropical gyre, and the development of the modelframework, while more recent global models (Macdon-ald 1998, hereinafter M98; Ganachaud 1999, hereinafterG99) extend only to 488N in the Atlantic—overturningacross this latitude then becomes more sensitive to thechoice of reference level since no explicit informationcloses the streamfunction farther north except bulk massconservation. Because much of this transformation is

Page 2: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1903L U M P K I N A N D S P E E R

driven by heat loss to the atmosphere, box models mustalso include air–sea fluxes if they are to attempt to dis-tinguish this from other parts of the diapycnal transport.

Property transports in the ocean are naturally dividedinto components along and across constant density sur-faces since the ocean is stratified and advection usuallydominates mixing. While there has been a trend to con-vergent values of isopycnal transports (heat, for in-stance) across the principal transatlantic hydrographiclines, the diapycnal fluxes inferred from large-scalestudies have been largely arbitrarily constrained or in-significant. More recent box inverse box calculationsexplicitly seek to resolve the diapycnal components aswell as the isopycnal components by including air–seafluxes of heat and fresh water and by building a degreeof freedom into diapycnal fluxes that decouples themfrom ad hoc parameterizations (Sloyan and Rintoul2001). Our approach in this study is similar, and we goto some length to represent in the prior circulation con-straints derived by previous investigators. Various bal-ances are then imposed to different degrees, that is, massbalance in layers, Ekman fluxes near the surface, andso on, with the goal of producing an estimate of themean circulation consistent with the suite of observa-tions. The reasonable existence of such a (record) meanis postulated and justified a posteriori by stable modelsolutions. Variability studies will be presented sepa-rately, but a sequence of inversions conducted with al-ternative repeats of two hydrographic lines demonstratesthe robustness of the mean circulation derived for theWorld Ocean Circulation Experiment (WOCE) period.

In the following sections, we review diapycnal trans-formation by air–sea forcing and mixing (section two),describe construction of the box inverse model used inthis study (section 3), discuss the results of the inversion(section 4), and make concluding remarks (section 5).In appendixes we present model-specific details: air–seaand river flux products used (appendix A), choices ofprior model errors and unknown variances (appendixB), and the treatment of individual hydrographic sec-tions (appendix C).

2. Transport and transformation in an isopycnallayer

In this study, isopycnal and diapycnal transports areinferred from property budgets within isopycnal oceaniclayers. This formalism was developed from the iso-thermal construction of Walin (1982) and has been re-viewed recently by Large and Nurser (2001).

Consider a layer of the ocean between isopycnal in-terfaces r 5 ro and r 5 ro 1 Dr within a domainbounded by bathymetry and a vertical control sectionrunning from one coastline to another. The layer volumeDV changes according to

] DV 5 C(r ) 2 C(r 1 Dr) 1 A(r )t o o o

2 A(r 1 Dr) 1 F , (1)o fw

where C is the overturning streamfunction across thecontrol section, A(ro) is the net diapycnal volume transportacross the interface r 5 ro (defined positive toward in-creasing density), and Ffw is the freshwater flux (P 2 E)integrated over the layer’s outcrop area. The layer bud-get of a property with concentration C is

] (^C&DV ) 5 2Dr[^C] C& 1 ] (^C&A) 1 ] D ]t r r r c

1 F 1 S , (2)c c

where ^ & is the layer or interface mean, Dc(ro) is theinterior flux of C across the interface r 5 ro due todiffusion and to eddies (e.g., covariations of C and dia-pycnal advection across the interface), Fc is the air–seainput of the property at the layer outcrop, Sc is interiorsources/sinks, and ]r is shorthand for the finite differ-ential between the two interfaces. In the special case C5 r, (2) becomes

r] DV 5 Dr[2r] (C 1 A) 2 A(r) 2 ] D 1 F ]t r r r r

1 r F , (3)fw fw

where 5 ro 1 Dr/2, interior sources of densificationr(e.g., cabbeling) have been neglected, and Fr is thetransformation of density (buoyancy) due to air–seafluxes of heat f h and freshwater f fw:

gaF 5 2 f (x, y, t)r E h[ rCpoutcrop

gr bSfw1 f (x, y, t) dx dy (4)fw ]r(1 2 S)

(cf. Speer and Tziperman 1990). By taking (3) 2 3 (1)rand dropping the negligible terms associated with directinput by Ffw, one obtains the transformation equation

A 5 F 2 ] D .r r r (5)

By substituting (5) into (1), one derives the formationequation

] DV 5 2D ] (C 1 F 2 ] D ).t r r r r r (6)

When the control section is a line of hydrographic casts,the streamfunction C can be divided into four com-ponents,

C 5 C 1 C 1 C 1 C9,Ek tw ref (7)

representing the directly wind-driven transport, the ther-mal wind transport calculated with respect to an as-sumed reference level, the unknown reference compo-nent arising from nonzero transport on the referencelevel, and a bolus component that we shall neglect.

The terms Fr, CEk, and Ctw can be estimated fromobservations, leaving several unknowns in (6): the massstorage term (left-hand side), produced by interannualvariations in diapycnal transformation and finite exporttime; the reference transport Cref; and the mixing ]rDr,which may be unequal to dispersion-based estimates ata point. Additional assumptions are needed to extract

Page 3: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1904 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

TABLE 1. Hydrographic sections used in the control inversion. Dates are for the casts shown in Fig. 1, not for the entire cruise. Modelparameters |y ref|ap (cm s21) and Tap (Sv) are defined in appendix B: (a) Tap 5 0.25 Sv for entire Iceland–Scotland ridge composite section,(b) |y ref|ap 5 30 cm s21 in the Straits of Florida. IFR: Iceland–Faroe Rise, FBC: Faroe Bank Channel, WTR: Wyville–Thompson Ridge, PB:Porcupine Bank.

Section Chief scientist, research vessel Dates |y ref|ap Tap

AR18IFR (west)IFR (east)FBC, WTRFaroe Bank

S-A. Malmberg, SœmundssonO. Astthorsson, SœmundssonHeinasonHeinasonChallenger

7–8 Sep 199513–14 Jun 19989–12 Jul 1987

11–13 Jun 198610–14 Jun 1988

1010101010

0.250.25a

0.25a

0.25a

0.25a

AR7E/A1EAR19/A2AR16/AR6A5A8

A. Sy, ValdivaA. Sy, GaussE. Hagen, HumboldtG. Parrilla, HesperidesT. Muller, Meteor

14–27 Sep 19928–25 Jul 1993

20–22 Sep 199219 Jul–15 Aug 19921 Apr–7 May 1994

52

101b

1

0.510.2522

an absolute C across a given latitude. Two recent ap-proaches are representative. Marsh (2000) calculatedyearly streamfunctions from time-varying fields of Fr

north of 158N in the Atlantic, equal to C at 158N in thelimit of zero storage and mixing. Speer (1997) also as-sumed zero storage and calculated a similar, but time-mean, Fr north of 118S, along with CEk at this latitude.He used the 118S Oceanus hydrographic section (Speeret al. 1996) to estimate Ctw and adjusted C 1 Fr withinan estimated error range to determine the minimum nec-essary mixing ]rDr. Speer (1997) found that interiormixing was necessary to account for the export of denseNADW, while air–sea formation alone could accountfor the export at subtropical thermocline densities (su

5 26.6–27.6).

3. A box inverse model

In this study, we extend the analysis of Speer (1997)by incorporating several hydrographic sections toachieve gyre-scale resolution of diapycnal and isopycnaltransport. Explicit air–sea forcing is calculated from theComprehensive Ocean–Atmosphere Data Set (COADS)and National Centers for Environmental Prediction–Na-tional Center for Atmospheric Research (NCEP–NCAR)reanalysis products (appendix A). We seek adjustmentsto the prior (thermal wind plus Ekman) flow field con-sistent with property conservation in isopycnal layers.The statistical framework for solving this problem inthe steady-state limit is the box inverse model (cf.Wunsch 1996).

Earlier studies have applied this framework to NorthAtlantic hydrographic sections (Wunsch 1984; Roem-mich and Wunsch 1985; Rintoul and Wunsch 1991;M98), including several of the WOCE sections used inthis study (G99). However, none of these studies hasapplied a priori heat and freshwater fluxes at the air–sea interface of the North Atlantic. Instead, heat con-servation is not enforced in outcropping layers, and themass loss associated with E 2 P is treated as an un-known. Although these models’ net heat transport di-vergences are consistent with integrated heat fluxes

(Ganachaud and Wunsch 2000), it is not clear if indi-vidual layer budgets are consistent. The freshwater flux-es inferred by these models are not consistent with theintegrated E 2 P loss of O(0.5 Sv) (118S–248N; G99),but this too can be enforced in the layer budgets. Inaddition to missing available information on the directinput of heat and freshwater, these inversions must solvefor the diapycnal advective transport A as an unknownto lowest order. They do not use (5) to divide A intotwo components, the unknown interior mixing and theestimable air–sea density transformation Fr. In theNorth Atlantic, Fr reaches ;30 Sv (Sv [ 106 m3 s21),converges water into subtropical and subpolar mode wa-ter densities and plays a dominant role in closing theupper limb of the overturning circulation (Schmitt et al.1989; Speer and Tziperman 1990; Speer 1997; Marsh2000). Although air–sea flux products contain errors dueto bulk parameterization and sampling biases, the gyre-scale structure of Fr is robust to lowest order in theNorth Atlantic. Out model calculates explicit layer bylayer adjustments to account for errors, thus producinga set of adjusted heat and freshwater fluxes consistentwith continuity and hydrography.

Neglecting air–sea transformation is problematicalwhere many layers outcrop, for example, in the subpolarAtlantic where cooling produces subpolar mode waterand, at the extreme end of SPMW, Labrador Sea Water(McCartney and Talley 1982, 1984), which is exportedsouthward at upper NADW densities. The box modelsof M98 and G99 circumvent this issue in the NorthAtlantic by choosing a nominal 488N section as theirnorthernmost boundary. However, in the absence of in-formation north of 488N, the overturning (and associatedproperties such as heat transport) becomes a strongerfunction of the prior thermal wind reference level andtransport error bars are magnified (see section 3b). Thebest estimates of absolute transports must be based uponan integration of divergences from a well-constrainedboundary, for example, the bathymetric choke point ofthe GIS ridges where extensive current observations(Ross 1984; Saunders 1990; Østerhus et al. 1999; Han-sen et al. 1999; Girton et al. 2001; Price et al. 2001)

Page 4: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1905L U M P K I N A N D S P E E R

FIG. 1. Geometry of the box inverse model, superimposed on bathymetry (m).

constrain exchanges with the GIN Seas and ArcticOcean (cf. Mauritzen 1996).

a. Model construction

Hydrographic sections (Table 1) are used to dividethe North Atlantic into four boxes (Fig. 1). Each boxis divided into 45 isopycnal layers separated by inter-faces of constant neutral density g n (Jackett and Mc-Dougall 1997). These interfaces were defined such thatthe 35 densest layers contain equal volumes of the globalocean according to the SAC/Hamburg climatology(Gouretski and Jancke 1998). Increased density reso-lution at g n , 26.5 helps to resolve transformation of

outcropping water masses in the Tropics and subtropics.To facilitate presenting and discussing model results,we refer to blocks of layers (Table 2) illustrated in Fig.2. Upper deep layers span the density range of LabradorSea Water in the subpolar gyre, and include the high-salinity core of upper NADW passing southward across118S. Lower deep layers span the density range of lowerNADW and intermediate Nordic Seas Overflow Water(NSOW) and are unventilated in the North Atlantic.Bottom layers include Antarctic Bottom Water and thedensest NSOW.

The model solves conservation equations based on(2) and (5). In layer i bounded by hydrographic stationpairs j 5 1, . . . , N, these take the form

N

Dx C (y 1 b ) dz 1 (1 1 e*)C C 1 C A 2 C A 1 F 1 F* 1 (] D ) 2 (] D ) ø 0,O j E j ref, j j Ek,i j j,10 m i21 i21 i i c,i c,i r c i21 r c i[ ]j51 i(8)

where Dxj is the distance between station pair j, thevertical integral spans the pair-dependent thickness oflayer i, Cj(z) is the pair-averaged profile of property C,y ref is a reference level adjustment to the prior thermalwind profile bj(z), e* is a fractional adjustment to eachsection’s Ekman transport, i is the interface-averagedC

value of C within the box, and F* is an adjustment tothe air–sea input of C. For mass conservation, C 5 r,]rD 5 0 (no diffusion of mass), and Fc 5 r fwFfw. Forheat anomaly (thermal energy with respect to 08C) con-servation, C 5 ruCp and Fc 5 Fh. For salt anomalyconservation, C 5 r(S 2 35)/1000 and Fc 5 0. For all

Page 5: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1906 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

TABLE 2. Blocks of isopycnal layers referred to in the text: ‘‘layers’’ indicates layer numbers in the inverse model, ‘‘g n’’ is the neutraldensity at the upper interface, ‘‘potential density’’ shows a corresponding value of s at 488N, and ‘‘water masses’’ within the blocks areindicated (see Fig. 2).

Name Layers g n Potential density Water masses

Surface (SFC)Thermocline (TC)Intermediate (INT)Upper deep (UD)Lower deep (LD)Bottom (BOT)

1–78–11

12–2021–2425–3637–45

Surface25.927.1527.8827.9828.12

su 5 26.0su 5 27.10s1.5 5 34.59s2.5 5 39.16s4 5 45.89

STMWAAIW, MOW, SPMW, AWLSW, NADW, NSOWNADW, NSOWAABW, NSOW

FIG. 2. Neutral density layers in the inverse model. Layers (numbered at right) are separated by interfacesof constant g n (left). Shading: zonally averaged salinity from SAC/Hamburg climatology. Black dashed line:densest water east of the Mid-Atlantic and Reykjanes Ridges. In addition to those defined in the text, labeledwater masses are Mediterranean Outflow Water (MOW) and Atlantic Water (AW). Bars at left denote blocksof layers (see Table 2).

property budgets, the diapycnal advection A is given byAi 5 (Fr 1 2 ]rDr)i, where (Fr 1 )i is calculatedF* F*r r

from adjusted heat and freshwater fluxes at the layer’soutcrops. Gauss–Markov estimation (Wunsch 1996) isused to solve for the unknowns: y ref for each station pair,e* for each hydrographic section, and for out-F* F*fw h

cropping layers in each box, and ]rD for density, heat,and salt across each interface within each box. Detailsregarding prior unknown variances and model errors aregiven in appendix B. Within each box, the net (all layers)mass, heat, and salt were also conserved via explicitconstraints. Additional section-dependent constraints,such as layer exchanges with the GIN Seas across theGIS ridges, were imposed upon the solution as docu-mented in appendix C.

Both this model and that of Sloyan and Rintoul (2001,hereinafter SR01) have explicit air–sea forcing. How-

ever, the property budgets (8) are constructed differ-ently. We assume that a property-independent diapycnaltransport A advects all properties from one layer to an-other, as in G99; SR01 assumed an independent dia-pycnal transfer for each property. This distinction issomewhat arbitrary, however, because we include anadditional property-dependent diffusive or eddy flux,unlike the property independent diffusivity of G99, andthus our equations contain the same degrees of freedomas those of SR01. A more fundamental distinction isthat SR01 added the explicit interior (within box) Ek-man diapycnal advection AEk to their budgets, yieldingthe prior diapycnal transport A 5 Fr 1 AEk. Their in-version simultaneously adjusted Fr and mixing, param-eterized by an effective diapycnal velocity w*. However,in the derivation of the volume budget (6), the net dia-pycnal advection A, which implicitly includes the in-

Page 6: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1907L U M P K I N A N D S P E E R

terior Ekman component, is replaced by Fr 2 ]rDr. Theadditional explicit AEk term in SR01’s volume budgetmust be combined with their w* term to infer mixing]rDr. Thus, SR01 have found a solution from the priorchoice 2]rDr 5 AEk, w* 5 0, A 5 Fr 1 AEk, while wechoose the minimal mixing prior 2]rDr 5 0, A 5 Fr.In oceanic layers where the primary balance is A ; AEk

5 F; that is, where net transformation is dominated bya surface buoyancy flux consistent with the mean Ekmantransport across isopycnals, their choice would inap-propriately double the prior transformation and drivethe solution toward spuriously large mixing. In oceaniclayers where strong mixing is associated with AEk andFr is relatively small, their prior choice would be moreappropriate while ours would tend to produce minimalestimates of both A and the interior mixing. In summary,our construction reflects a prior choice that will tend toproduce weaker overturning and less mixing than thatof SR01 wherever it is applied, and can be considereda minimum mixing scheme.

b. Experimental inversions

In addition to the ‘‘control’’ inversion constructed asin Fig. 1, a number of experimental runs were con-ducted. Two of these were particularly revealing:

1) EFFECT OF INFORMATION NORTH OF 488N

Previous large-scale box inversions (M98; G99) haveused the 488N section as their northern boundary. Byincluding explicit hydrographic information farthernorth and requiring that the model be consistent withestimates of GIN Seas overflows and exchanges, ourmodel is better able to constrain property transports andis less sensitive to the initial reference level choice at488N. To demonstrate this, we ran suite of two-box in-versions, consisting of the equatorial and subtropicalboxes (118S–488N). Bering Strait throughflow-derivedconstraints for net volume and salt (Wijffels et al. 1992)were applied at the 488N northern boundary, and all elsewas identical to the control (four box) inversion in-cluding explicit air–sea forcing in the two boxes (notpresent in previous inversions). Because the two-boxinversions lack information regarding subpolar air–seaforcing, exchanges with the GIN Seas and conservationin subpolar layers, their transports are less well con-strained and more strongly dependent upon the initialreference level choice than those of the control inversion(Fig. 3). For very dense initial reference levels, the four-box inversion was not able to satisfy the constraintswithin the prior choices of the inversion; these resultsare not shown in Fig. 3.

The M98 and G99 inversions included hydrographicsections and continuity south of 118S, not included inour control inversion. However, because there is nosouthern choke point comparable to the GIS ridges, norSouth Atlantic outcropping and air–sea-driven trans-

formation comparable to the subpolar North Atlantic,this information does not affect transports across 118Sas dramatically as in Fig. 3. In an experimental inversionwith an additional South Atlantic box (bounded to thesouth by the 308–458S line A11), overturning across118S had a maximum strength of 15.8 6 2.1 Sv (16.46 2.1 Sv in the control run) and the net transport across118S was 0.68 6 0.10 PW (0.69 6 0.10 PW control),not significantly different within the nearly identical er-ror bars.

2) EFFECT OF HYDROGRAPHIC VARIABILITY

Recent analyses of hydrographic repeats at 488N(Koltermann et al. 1999), 608N (Bersch et al. 1999) andin the deep boundary current off Cape Farvel (Bacon1998) have deduced significant variations in North At-lantic property transports. In the steady-state inversionof this study, these variations must be absorbed by mod-el error terms or cancelled by adjustments to the initiallevel of no motion [e.g., 1000 m in Bacon (1998)]. Toassess the impact of hydrographic variations upon thesolution, we reran the model with the 16 possible com-binations of four 488N A2/AR19 repeats (1993, 1996,1998, 1999) and four 608N A1E/AR7E repeats (1991,1992, 1995, 1996), everything else identical to the con-trol run. The 16-ensemble-mean diapycnal transfor-mation differs by less than one standard deviation fromthe control inversion in all layers (Fig. 4), demonstratingthat, when given identical constraints for North Atlantic/GIN Seas exchanges, the model is able to produce arobust time-mean circulation via acceptable adjustmentsto the unknowns and absorption by model error terms.The inverse solution is altered more dramatically whenone makes large changes to the boundary condition con-straints, for example, by doubling or halfing the over-flow through the Denmark Strait channel. Observedchanges in the overflow are smaller (Girton et al. 2001)but experimental runs such as these demonstrate theimportance of choke-point observations for monitoringoverturning variability.

4. North Atlantic circulation and transformation

The inversion was able to derive a set of unknownswithin the range of prior variance (appendix B) whichsatisfied property conservation and additional, section-dependent constraints (appendix C) within specified pri-or model error. Neighboring solutions, in the sense ofsmall adjustments to initial conditions, constraints, layerdefinitions, and model errors, were very similar and inthis sense the result is considered stable. Figure 5 showsthe adjusted air–sea density, heat, and freshwater inputsin each box derived by the model. In most layers theadjusted air–sea inputs are not significantly differentthan given by the University of Wisconsin—Milwaukee(UWM)/COADS or NCEP–NCAR reanalysis products,although they tend to match more closely the NCEP–

Page 7: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1908 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

FIG. 3. (top) Net heat transport, (middle) overturning strength, and (bottom) southward transportof lower NADW across 488N, as a function of thermal wind reference level (horizontal axis; blackline marks the value used in the control inversion). Results from the four-box inversion are whitebullets, with black standard error bars. Results from a two-box inversion (the equatorial andsubtropical boxes) are asterisks, with gray standard error bars.

NCAR values. The interior mixing ]rD is shown in Fig.6, as a function of model layer (see Table 2). Figure 7shows ]rDr converted to an effective diffusivity ^k&using layer mean values from the SAC/Hamburg cli-matology (Gouretski and Jancke 1998); error bars donot include uncertainties in these climatological values.Some very dense layers present in the hydrography donot exist in the climatology; for these, the volume wasextrapolated and the error bars in Fig. 7 were magnified.

Net transports of volume, salt, and heat after inversionare given in Table 3. By construction, divergences inheat and volume transport are consistent with the ad-

justed air–sea heat and freshwater inputs. The equatorialbox gains 0.41 6 0.07 PW from the adjusted heat fluxes,while the subtropical, subpolar, and overflow boxes lose0.37 6 0.04, 0.17 6 0.03 and 0.14 6 0.01 PW, re-spectively.

Figure 8 shows the distribution of volume transportin the model layers, integrated bottom to top (i.e., theoverturning streamfunction C). A schematic of the two-dimensional meridional overturning circulation, Fig. 9,is produced by linearly interpolating C between the sec-tions and extrapolation to zero north of the nominal668N overflow ridge sections. Because this schematic

Page 8: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1909L U M P K I N A N D S P E E R

FIG. 4. (left) Interior and (right) air–sea diapycnal transformation in the (top) overflow and (bottom) subpolar boxes.Bullets indicate the ensemble mean of 16 inversions using combinations of repeat hydrographic lines at 488 and 608N;horizontal lines show their standard deviations. The control inversion is a heavy black line, with the standard errorshaded.

is presented in model layer (density) coordinates, dia-pycnal transformation is indicated by nonhorizontalstreamlines. Figure 10 presents a top view of the cir-culation, showing separately the western boundary andinterior components of the shallow (white) and deep(black) pathways.

In the remainder of this section, we compare the re-sults of our inversion with those of M98 and G99, wherethe domains of these inversions overlap (118S–488N).We then discuss the pathways of isopycnal circulationand diapycnal transformation that define the three-di-mensional circulation of the North Atlantic in our so-lution.

a. Comparison with previous inversions

Over the domain 118S–488N, our model circulationloses a net 0.3 6 0.1 Sv to integrated E 2 P 2 R,consistent (by construction) with estimated air–sea flux-es and riverine input. In contrast, G99 gains 0.2 6 0.2Sv (M98 gains 0.05 Sv) over this domain, presumablyaffected by his prior choice of 0 Sv.

In the subtopical box, the gross features of the me-

ridional overturning circulation derived by our inversemodel are similar to those of G99. The northward trans-port of 15.6 6 1.3 Sv surface, thermocline, and inter-mediate water at 248N and 488N (16 6 2 Sv in G99)is balanced by an equal southward transport of NADWat 488N, increased slightly at 248N by upwelling of 0.96 2.1 Sv AABW (G99: 1 6 1 Sv) within the subtropicalbox. There is no significant net exchange between thenorthward and southward overturning limbs within thesubtropical box, as in G99 but unlike M98 in which adivergence of 6 Sv in the northward layers was fed byupwelling. The southward flow of NADW across 488Nis substantially less than M98’s 25 6 1 Sv, primarilyin the layers g n . 28.12 (u , 2.0) where our solutionhas 0.5 6 1.0 Sv crossing 488N, while M98 found 106 1 Sv. As only 3.7 6 0.3 overflows the GIS ridges atthese extreme densities, and upwelling in the overflowbox reduces this to 2.2 6 0.6 Sv flowing past CapeFarvel (608N), it is unclear what process could feed 10Sv at these densities. The discrepancy with M98 maybe due to temporal variability (pre-WOCE section inM98, WOCE here) or a bottom triangle problem in M98(G99, pg. 118). In our solution, the deep western bound-

Page 9: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1910 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

FIG. 5. Air–sea fluxes of (top) buoyancy, (middle) freshwater and (bottom) heat in each of the four boxes of the inverse model; shadingis standard error. The buoyancy flux Fr is calculated from the freshwater and heat fluxes. Dashed lines: prior fluxes from UWM/COADS.Dotted lines: prior fluxes from NCEP–NCAR reanalysis. Positive Fr corresponds to a diapycnal advection toward greater density.

ary current (DWBC) is twice the strength of the north-ward deep interior flow. At 488N our solution shows a14 Sv deep cyclonic recirculation in the NewfoundlandBasin (Fig. 10), not present in the McCartney (1992) orSchmitz and McCartney (1993) schemes but qualita-tively similar to M98’s 5 Sv recirculation. At 248N, M98found a southward transport of 19 6 2 Sv NADW, sim-ilar to the 17.6 6 2.7 Sv in our model and in earlierestimates (e.g., ;18 Sv, Wunsch and Grant 1982; 20 65 Sv, Roemmich and Wunsch 1985). The primary dif-ference between our results and those of G99 in thesubtropical box is that transports in our solution are ingeneral more tightly constrained because of the prop-agation of information from the subpolar gyre (see sec-tion 3b). Other major differences are that we resolvetransport and transformation at densities lighter than g n

5 26.44, where air–sea-driven advection must be bal-anced by significant heat and salt diffusion, and we donot find significantly nonzero downwelling in the upper

deep density range 27.92 , g n , 28.0, present in G99.Last, our solution requires a different relative combi-nation of salt and heat mixing in the thermocline layers(discussed below), which would not be accommodatedby G99’s property-independent diffusivity.

Within the equatorial box, our solution and M98, G99find that 2–4 Sv of AABW is converted to lower deepdensities, associated with a diffusivity of O(1023 m2 s21)(Fig. 7). This conversion increases the DWBC transportof NADW at 118S from its value at 248N. However,while both M98 and G99 show conversion of northward-flowing AAIW to southward-flowing NADW (5 Sv inM98, 2 Sv in G99) between 118S and 248N, we findnegligible downwelling across the AAIW/NADW in-terface. As a consequence, we find a smaller transportof NADW across 118S: 19.6 6 2.4 Sv, as comparedwith 28 6 1 in M98 and 23 6 4 in G99. In the coreNADW density range 27.5 # g n # 28.11, Speer et al.(1996) inferred a southward transport of 14 Sv across

Page 10: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1911L U M P K I N A N D S P E E R

FIG. 6. Interior diapycnal transports of (top) volume and diffusive or eddy fluxes of (middle) salt anomaly and (bottom) heat anomaly inthe four boxes; shading is standard error. Positive values are toward decreasing neutral density (upward).

118S; our adjusted field carries 18.2 6 2.5 Sv, com-parable to G99’s 20 6 4 Sv in these layers. The upper,northward-flowing branch of the M98 and G99 solutionswere stronger at surface densities, while ours is strongerat intermediate densities, yielding a smaller net heattransport across 118S in our solution (0.69 6 0.1 PW,vs 0.9 6 0.4 PW in G99).

b. The three-dimensional North Atlantic circulation

The North Atlantic overturning circulation is fed by16 Sv of surface to upper intermediate water (g n , 27.6)passing north across 118S. This is carried primarily bythe North Brazil Current, as the net interior (non-WBC)transport in these layers is 0.1 6 4.1 Sv. Surface waterabsorbs 0.41 6 0.07 PW from the atmosphere withinthe equatorial box, producing a strong buoyancy gainin the warm pool layers, which must be balanced bymixing (Niiler and Stevenson 1982). In the model so-lution, a 16 Sv mixing-driven downwelling opposes theair-sea-driven transformation across g n 5 22.6. Down-

ward heat diffusion reaches a maximum of 0.4 PW inthese layers (Fig. 6). At densities g n 5 23–24, the ad-justed air–sea heat input is significantly smaller than theintegrated UWM/COADS fluxes but remains consistentwith NCEP.

Water in the upper limb of the overturning cell entersthe subtropical box through the Straits of Florida (30Sv; Schott et al. 1988), the Antilles Current east of theBahamas (19 6 9 Sv) and via Ekman transport across248N (4.3 Sv). Much of this recirculates southward inthe subtropical gyre (Fig. 10), leaving a net northwardtransport of 16.6 6 1.5 Sv across 248N (g n , 27.74).Within the subtropical box, outcropping water at g n .24.7 loses a net 0.40 6 0.04 PW, concentrated in theGulf Stream and its recirculation. The associated air–sea transformation reaches a peak of 19.2 6 3.3 Sv atg n 5 26.4, producing an air–sea buoyancy convergence(formation) of 19 Sv in STMW densities (Speer andTziperman 1990) and downward-plunging streamlineswithin the subtropical box (Fig. 9). Some of this waterreenters the equatorial box in the subtropical gyre where

Page 11: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1912 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

FIG. 7. Effective diapycnal diffusivities (horizontal axes) as a function of neutral density (vertical axes) in the four boxes,calculated from the interior mixing component of the diapycnal volume transport A; shading is standard error.

TABLE 3. Vertically integrated transports of volume, heat, and salt across the hydrographic sections in the inverse model solution. ‘‘NetGIS’’ is the combined flux across the Denmark Strait and Iceland–Scotland sections. Positive values are northward (westward for AR16).

Section Volume (Sv) Heat (PW) Salt (106 kg s21)

Net GIS 0.1 6 0.3 0.30 6 0.03 7 6 13AR18 (Denmark Strait)Iceland–Scotland

24.0 6 0.74.1 6 0.7

0.05 6 0.020.25 6 0.02

2142 6 27149 6 25

AR7E (608N)AR19 (488N)AR16 (Gulf of Cadiz)

0.1 6 0.620.5 6 1.0

0.0 6 0.2

0.44 6 0.030.60 6 0.04

20.04 6 0.02

6 6 26210 6 36

0 6 9

A5 (248N; total) 21.0 6 1.4 1.17 6 0.08 231 6 50A5 (Florida Strait)A5 (Bahamas–Africa)

29.9 6 1.6230.9 6 2.1

2.42 6 0.1121.24 6 0.10

1148 6 3421162 6 48

A8 (118S) 20.7 6 1.2 0.69 6 0.10 218 6 59

it is heated, reconverted to surface layers, and returnedto the subtropical box, creating a diapycnal signature ofthis predominantly isopycnal recirculation. The remain-der of the STMW continues northward in the GulfStream extension and leaves the subtropical box across488N. Within the subtropical box, the heat balance ofthe upper thermocline and surface layers g n , 26.72,which include STMW, is primarily a balance between

advective input through the Straits of Florida (0.58 60.9 PW relative to the mean temperature of the interfaceg n 5 26.72) and loss by advection across 248N Bahamasto Africa (0.14 6 0.11 PW), 488N (0.04 6 0.04 PW)and loss to the atmosphere (0.36 6 0.03 PW absoluteheat loss). A diffusive or eddy flux of 0.04 6 0.22 PWacross g n 5 26.72 (which outcrops in the North AtlanticCurrent) is needed to close this budget but is not for-

Page 12: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1913L U M P K I N A N D S P E E R

FIG. 8. Overturning streamfunction across hydrographic sections A8 (118S), A5 (248N), AR19 (488N), AR7E (608N), andthe combined GIS ridge sections (668N). Positive is northward. Standard error bars are indicated by shading.

mally significant. In order to generate the observed char-acteristics of STMW, a different relative combination ofsalt and heat diffusion is required: while heat diffusionis negligible in the thermocline layers, salt diffusiontransfers 5 3 106 kg s21 salt out of mode water layersto lower density (Fig. 6), that is, acts to reduce thesalinity increase caused by net E 2 P.

Across the Gulf of Cadiz section, 2.2 6 0.6 Sv over-turning is associated with the conversion of g n , 27.4water to the range 27.4 , g n , 27.96 within the Med-iterranean Sea and Strait of Gibraltar (Baringer and Price1997). Figure 9 reflects this transformation as an ap-parent downwelling at dense thermocline to interme-diate layers between 248 and 488N. This transformationis not explicitly resolved in the model domain—in fact,there is no significant mixing in these subsurface layerswithin the subtropical box (Fig. 6).

Northward flow across 488N is concentrated in thewestern Atlantic, on both sides of a strong anticyclonicrecirculation in the western boundary (the Mann Eddy;Mann 1967). The net northward transport at 488N is16.1 6 1.6 Sv for g n , 27.84. This water enters thesubpolar gyre (the combined subpolar and overflow box-

es), where it loses 0.31 6 0.03 PW to the atmosphere.The associated air–sea transformation Fr peaks at 12.66 1.1 Sv (g n 5 27.41) and drops to zero by g n 5 28.04,yielding a formation of 12.6 Sv of SPMW. Of this, 3.16 1.6 Sv is at extreme SPMW (e.g., Labrador Sea Wa-ter) densities of g n . 27.88 (su ø 27.74 at 488N). Anadditional 5.8 6 3.1 Sv of water is advected to layersdenser than g n . 27.88 by interior mixing, representingmixed layer entrainment, eddy stirring across outcrop-ping isopycnals, and warm water entrainment at theoverflows; 2.5 6 2.3 Sv continues to densities greaterthan g n 5 27.98 (su ø 27.81). The resulting mixing-driven advective convergence is the same strength asair–sea formation in the LSW density range (27.91 ,g n , 27.98), and the combined formation (air–sea plusmixing) accounts for nearly all of the 5.9 6 2.0 Sv,which is exported southward across 488N in this densityrange. Subtracting the 0.6 Sv that enters the overflowbox over the GIS ridges and interpolating to extract thetotal formation rate in the range 27.74 , su , 27.80yields a value of 4.7 6 2.0 Sv, within the estimates ofmean LSW production from CFC inventories in thesame su range (4.4–5.6 Sv; Rhein et al. 2002).

Page 13: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1914 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

FIG. 9. Side view of the North Atlantic meridional overturning, contoured in 2-Sv intervals, superimposed on zonally averaged(top) salinity and (bottom) oxygen (mL L21) calculated from climatology (Gouretski and Jancke 1998). Light gray curve: densestoutcropping layer, estimated from COADS climatology. Dark gray curve: crest of the Mid-Atlantic Ridge, including the AzoresPlateau and Iceland.

A total of 6.5 6 0.7 Sv Atlantic water at g n , 27.79leaves the domain across the GIS ridges, via the Irmin-ger and Norwegian Atlantic Currents. Within the GINSeas, this water is converted to NSOW, which returnsto the North Atlantic through channels in the GIS ridges(cf. Mauritzen 1996). The overflow spans a broad rangeof densities; the densest, 1.6 6 0.2 Sv at g n . 28.12through the Faroe Bank Channel and 1.5 6 0.3 Sv atg n . 28.23 through the Denmark Strait channel, is notpresent in AR7E (608N) and therefore must be mixedto lighter layers entirely within the overflow box. Atdensities 27.6 , g n , 28.08 (27.5 , su , 27.89 at608N) within the overflow box, interior mixing trans-forms water to denser layers. The resulting mixing-driv-en convergence centered on g n ø 28.1 is a robust resultof the inversion (e.g., Fig. 4), consistent with detrain-ment of the dense overflow and entrainment of less-dense water, in the overflow plumes descending fromthe mouths of the Denmark Strait and Faroe Bank chan-nels (cf. Speer and Tziperman 1990).

At densities g n . 27.96 (su . 27.8), 13.1 6 1.7 Svcrosses 608N in the southward deep boundary currentagainst Cape Farvel, Greenland, consistent with Dick-son and Brown (1994). The net outflow across 608N in

these layers is 9.8 6 1.9 Sv, fed by 5.6 6 0.5 Sv over-flow across the GIS ridges plus entrainment of 3.0 62.3 Sv from lighter layers. Thus, the model results implythat entrainment is the same size as lateral isopycnaltransport in these layers and that entrainment and de-trainment within these layers significantly affect thedensity structure of water exported in the deep boundarycurrent.

5. Conclusions

Using a statistical best-fit method we have estimatedlarge-scale circulation, transports, the distribution ofmixing, and air–sea fluxes, which are simultaneouslyconsistent with the North Atlantic hydrography and awide range of constraints from regional studies. Themethod handles explicitly the effect of air–sea trans-formation where layers outcrop. This allows us, for in-stance, to extend the model beyond 488N without aban-doning resolution of outcropping layers. The result ofthis added information is to produce better-constrainedestimates of property transports.

In the inverse solution, average mixing (Fig. 7) hasa well-determined structure related to air–sea transfor-

Page 14: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1915L U M P K I N A N D S P E E R

FIG. 10. Circulation of thermocline (white arrows; g n , 27.6) and deep (black arrows; g n . 27.6) waterin the inverse solution. Numbers indicate transport (Sv) where currents cross the hydrographic sections;values in parentheses are not well-determined in the solution. The strength of the net overturning exchangebetween these layers is given in panels; north of 118S, larger values are found for denser interfaces.

mation, overflow, and deep upwelling. The solutionshows two major overturning cells: a thermocline plusintermediate water cell and a dense overflow cell. Theoverflow component is ;40% of the total. A third,smaller cell is associated with the inflow and upwellingof AABW. All cells ultimately depend on air–sea forcingat the surface, but the overflow presents a strong anddistinct control; in the overflow box, vertical and hor-izontal transports are the same strength, and intense re-circulation around the boundary and within the adjacentIrminger and Iceland basins absorbs and transmits over-flow water. Hints of a shallow tropical cell are present,and significant diapycnal transports occur in the equa-torial box, but we are unable to distinguish mixing atthe equator from that occurring elsewhere within thebox, and so this aspect of the solution has not beenpursued here.

The net isopycnal volume transports across each sec-

tion and diapycnal transports across layer interfaces maybe combined to yield bulk volume fluxes into/out ofeach block of layers in each box. The layers’ volumedivided by this flux gives a mean residence time, up-dating Worthington’s (1976) values (Table 4; theseshould not be equated with age, i.e., time since out-cropping). The smallest residence time is the ;10 daysrequired to flush bottom layers (g n . 28.127) in theoverflow box—this densest overflow is vigorously de-trained after flowing through the Faroe Bank Channeland rapidly advected in the deep boundary current fromthe Denmark Strait channel to Cape Farvel. The largestwell-defined residence times are in lower deep layers,which require ;100 years to renew their large volumeswithin the equatorial and subtropical boxes, and in in-termediate layers of the equatorial box, which require;150 years because they are near the turning densityof the overturning circulation. Mean residence times in

Page 15: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1916 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

TABLE 4. Bulk volume fluxes ( ), volumes (V ), and mean residence times (Tres, range of values from the standard error of ) for layer˙ ˙V Vblocks (see Table 2).

Box TC INT UD LD BOT

Overflow V (Sv):V (106 km3):Tres (yr):

3.5 6 0.50.04

0.3–0.4

12.7 6 2.21.61

3.4–4.9

6.5 6 2.20.96

3.5–7.1

6.0 6 2.00.68

2.7–5.4

3.7 6 0.60.003

0.002–0.031Subpolar V (Sv):

V (106 km3):Tres (yr):

8.4 6 1.80.28

0.87–1.34

7.4 6 1.92.55

8.7–15

6.6 6 2.23.36

12–24

7.5 6 2.24.53

15–27

2.2 6 0.60.28

3.2–5.5Subtropical V (Sv):

V (106 km3):Tres (yr):

21.2 6 1.36.68

9.4–11

6.0 6 0.710.9652–66

6.3 6 2.48.63

31–70

8.7 6 3.226.2870–150

1.7 6 1.32.6828–210

Equatorial V (Sv):V (106 km3):Tres (yr):

22.3 6 1.28.97

13–14

4.8 6 1.523.03

120–220

6.0 6 2.010.2240–81

13.2 6 2.345.4493–130

4.6 6 2.75.4224–90

upper deep layers are 20–50 years throughout most ofthe North Atlantic.

Acknowledgments. We thank A. Tandon and A.Thurnherr for their input and assistance throughout thisproject. Author RL thanks the hosts and participants ofthe 2001 WOCE Young Investigators’ Workshop. Theinverse model was developed from DOBOX code (Mor-gan 1995) and generous assistance from B. Sloyanhelped us to initiate the study. The study also benefitedfrom communications with C. Wunsch, P. Saunders, C.Mauritzen, and S. Josey. NCEP reanalysis data wereprovided by the NOAA–CIRES Climate DiagnosticsCenter (http://www.cdc.noaa.gov/). RL and KS receivedsupport from NSF grant OCE-9906657.

APPENDIX A

Air–Sea and Land–Sea Fluxes

a. Heat and freshwater

The monthly mean climatologies were the UWM/COADS (da Silva et al. 1994) and the 22 April 2002version NCEP–NCAR reanalysis products. For eachmonth, sea surface g n was calculated from SST and SSSfields (the latter from Levitus and Boyer 1994), and thenthe net heat and freshwater (P 2 E) fluxes were summedfor each outcropping model layer. Accumulation overthe 12 months produced an estimate of the annual meanfluxes into each layer. Air–sea transformation Fr wascalculated with (4) for each month and accumulatedsimilarly.

b. Rivers

The 60 largest rivers flowing into the Atlantic (total:0.33 Sv) were extracted from the Global Runoff DataCenter database (Grabs et al. 1996). This freshwaterinput was included in the prior volume and buoyancybudgets by adding each river’s transport to the UWM/COADS or NCEP–NCAR grid point closest to the rivermouth.

c. Wind stress

Monthly mean wind stresses were used from theNCEP–NCAR reanalysis and from the SouthamptonOceanography Centre reanalysis of COADS, which usesa larger drag coefficient than UWM/COADS at highwind speeds to match more closely direct observations(Josey et al. 2002). For each layer the Ekman transportacross a section was calculated for the 12 climatologicalmonths and assumed to advect monthly temperature andsalinity at 10-m depth (Levitus and Boyer 1994).

APPENDIX B

Prior Unknown Variances and Model Errors

Prior variances for the unknowns and model errorswere contained in matrices Rxx, Rnn (cf. Wunsch 1996).In the absence of knowing the full covariance structureof these prior matrices, we took both as diagonal.

Adjustments to reference velocities were required tobe in a range 6 | y ref | ap. Interior (nonboundary) valuesof | y ref | ap were chosen to be section-dependent con-stants of 1–10 cm s21 (see Table 1), with larger valuesfor sections dominated by shallow flow over sharp ba-thymetry. In the western boundaries of the 118S, 248N,and 488N sections, | y ref | ap was increased to 40 cm s21

to allow strong WBC adjustment. In addition, at eachstation pair a second value of | y ref | ap was calculated toaccommodate a 6250-dbar uncertainty in the referencelevel, an error range inferred from analysis of a GCM(Ganachaud 1999). The larger of the two values wasused for each pair.

Prior values for the air–sea inputs of heat, freshwater,and momentum were taken as the mean of the COADS-and NCEP-derived calculations. The squared differencebetween the integrated flux estimates from the two prod-ucts was used as the permitted variance of the F*, e*adjustment. This choice accommodates both COADSand NCEP fields in the prior error.

The interior (subsurface) mixing of density, salt, andheat were given prior values of zero. The allowed var-iance of the density mixing terms ]rDr were calculated

Page 16: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1917L U M P K I N A N D S P E E R

from the prior transformation equation for each layer.After some experimentation we added a uniform 2 Svto the allowed adjustment within each layer in order toachieve a solution with lower overall mixing, potentiallyat the expense of higher mixing in some individual lay-ers. Variances for the unknown diffusion and eddy trans-ports of heat and salt were calculated similarly, that is,the value needed to satisfy the prior budgets, consistentwith our goal of seeking a minimum-mixing solution.

By applying a box inverse model to the output of aforward model, Ganachaud (1999) has shown that netvolume conservation should be less strictly imposedthan in earlier efforts, primarily due to temporal vari-ations in the upper-ocean baroclinic structure of low-latitude sections. In this study, each section is assigneda net error Tap (see Table 1) due to asynopticity. Theprior variance from exact net volume conservation with-in a box is given by the sum of the for the sections2T ap

bounding that box. To make an observation-based es-timate of individual layer errors due to asynopticity, wecalculated the combination of bottom-referenced ther-mal wind and reference level velocity offset, which yield31 Sv for the August 1992 and February 1998 WOCErepeats of the Straits of Florida section. Within indi-vidual isopycnal layers, the difference in volume trans-port ranged from 1.2 Sv (lightest layer) to ,0.2 Sv forlayers denser than g n 5 25.3. We used these values asthe prior errors on volume conservation within layers,with 0.2 Sv assigned to all layers at g n . 25.3 (layers7–45, representing 98.9% of the ocean’s volume in thestudy domain). To estimate the model error in heat con-servation, we multiplied the layers’ volume error by thesum of the box- and layer-dependent mean potentialtemperature ^u& and twice the standard deviation of uwithin the layer, in the bounding hydrographic sections.The second term represents possible correlations be-tween potential temperature anomalies and eddy massfluxes (Ganachaud 1999). A similar procedure was usedfor salt anomaly conservation, using the salinity anom-aly (S 2 35)/1000.

APPENDIX C

Hydrographic Sections

For each hydrographic section (Table 1), a thermalwind reference level (isobaric or isopycnal) or set oflevels were chosen based on existing literature and netcontinuity arguments. For some station pairs, nonzeroinitial reference velocities were included so that the pri-or velocity field was consistent with additional, section-dependent constraints in particular layers. Often in theliterature, net transports are given for a range of su oru. We convert su to g n (and thus to model layer number)using the scatter plot of su versus g n for the section.Temperature-defined layer transports were identified bythe layer-mean u–S relation of the section.

a. Canadian Archipelago throughflow

A transport of 0.9 6 0.2 Sv, (31 6 6) 3 106 kg s21

salt was added to the mass and salt budgets of the sub-polar box. This volume transport matches the BeringStrait throughflow (0.8 Sv; Coachman and Aagaard1988) combined with 0.1 Sv net P 1 R 2 E and icemelt in the Arctic Ocean. Applying this transport towater in the Canadian Archipelago (mean salinity 33.6psu in the SAC/Hamburg climatology) gives the cor-responding salt transport. The inversion was also re-quired to match the Bering Strait salt transport, (26.76 19) 3 106 kg s21, via the combined southward saltinput through the Denmark Strait and Iceland–Scotlandsections and the Canadian Archipelago input (Wijffelset al. 1992).

b. Denmark Strait

Thermal wind for WOCE section AR18 was calcu-lated relative to a bottom level of no motion. Withoutadditional information, the model fails to produce anacceptable overflow. The inverse model is ‘‘informed’’of the bathymetrically controlled flow via additionalconstraints:

• No net mass transport into GIN Seas: 0 6 0.35 Svacross the Greenland–Scotland ridges (AR18 and theIceland–Scotland section), all layers.

• Denmark Strait overflow: 2.0 6 0.2 Sv southward inlayers 38–45 (g n . 28.14) (Girton et al. 2001).

• Denmark Strait overflow: 0.9 6 0.1 Sv southward,layers 27–37 (28.01 , g n , 28.14) (Girton et al.2001).

• Polar water in the East Greenland Current: 1.6 6 0.5Sv southward, westernmost 15 station pairs, layers 1–15 (g n , 27.6) (Malmberg et al. 1972; Mauritzen1996).

• Irminger Current branch: 0.9 6 0.4 Sv northward ineasternmost five station pairs, layers 1–15 (g n , 27.6)(Kristmannsson et al. 1989; Mauritzen 1996).

c. Iceland–Scotland Ridge

Because of the complex bathymetry between Icelandand Scotland, an ideal hydrographic section separatingthe North Atlantic Ocean and the Norwegian Sea wouldrun southeast from Iceland along the Iceland–Faroe Riseto the Faroe Shelf, southwest across the Faroe BankChannel (FBC) to the Faroe Bank, then southeast alongthe Wyville–Thompson Ridge to the Hebridian Shelfoff the Scottish coast. Unfortunately, no such sectionexists in the WOCE database. We assembled a com-posite section from the International Council for theExploration of the Sea (ICES) Service Hydrographiquedata base. Three segments, all from the North Atlanticand Norwegian Sea Exchange (NANSEN) experiment,span most of the ideal path but terminate on the Iceland–

Page 17: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1918 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

Faroe Rise 230 km from the Iceland coast. This gapwas filled by a segment from the Variability of Ex-changes in the Northern Seas (VEINS) experiment, col-lected 10 yr after the 1986–88 NANSEN sections. Ther-mal wind was calculated relative to a level-of-no-motionat g n 5 27.88, separating northward-flowing Atlanticwater and southward-flowing overflow. Despite the tem-poral separation between the VEINS and NANSEN seg-ments, they had nearly identical vertical profiles of tem-perature and salinity where they were joined, producinga maximum thermal wind of 0.5 cm s21 (the near-con-current pairs on either side have maxima of 1.9 and 5.7cm s21). Section-dependent constraints imposed on theinverse solution were:

• Faroe Bank Channel Overflow: 1.4 6 0.2 Sv south-ward, layers 38–45 (g n . 28.14) (Saunders 1990;Mauritzen 1996; Østerhus et al. 1999; Price et al.2001).

• Overflow across Iceland–Faroe Rise (including NorthIcelandic Winter Water and Arctic Intermediate Wa-ter): 0.85 6 0.3 Sv southward, westernmost 14 stationpairs, layers 21–45 (g n . 27.88) (Saunders 1990;Mauritzen 1996).

• Atlantic water across Iceland–Faroe Rise: 4.0 6 0.4Sv northward, westernmost 14 station pairs, layers 1–20 (g n , 27.88) (Hansen et al. 1999).

• Norwegian Atlantic Current between Faroe Islandsand Scotland: 3.0 6 0.4 Sv northward, easternmost18 station pairs, layers 1–20 (g n , 27.88) (Gould etal. 1985; Mauritzen 1996; Hansen et al. 1999).

d. Greenland to Ireland, 608N

The second repeat of WOCE line AR7E/A1E (1992)ran from Cape Farvel, Greenland, to the PorcupineBank, west of Ireland, passing over the Reykjanes Ridgeapproximately halfway between Iceland and the CharlieGibbs Fracture Zone and also passing close to the south-ern edge of the Rockall Plateau. Ice conditions againstthe Greenland coast were favorable during this cruise,allowing water in the East Greenland Current (EGC) tobe sampled to 38 km offshore. The eastern end of AR7Elies on the Porcupine Bank (an extension of the Hebri-dian Shelf ) at a depth of 320 m, but still 240 km fromthe Irish coast. Thus, while northward-flowing Atlanticwater is sampled in this section, an area of potentiallysignificant transport over the shelf is missed. This gapwas filled with a segment of the previous year’s AR7Erepeat, which began 60 km west of the Irish coast buthad encountered ice 45 km farther offshore of Green-land. Water properties on the Porcupine Bank had notchanged significantly between 1991 and 1992: at theirclosest casts, salinity and temperature have smaller dif-ferences than similarly spaced 1992 casts farther weston the bank. Thermal wind was calculated relative tog n 5 27.88. Section-dependent constraints imposed onthe inverse solution were:

• EGC, Greenland to 2000-m isobath: 15 6 2 Sv south-ward, westernmost five station pairs, all layers (Bacon1997).

• Total EGC strength: 32.2 6 3.2 Sv southward, west-ernmost seven station pairs, all layers (Bersch 1995).

e. North America to Europe, 488N

Thermal wind was calculated relative to the no-mo-tion level g n 5 27.88. No section-dependent constraintswere imposed.

f. Gulf of Cadiz

Thermal wind was calculated relative to a bottomreference level. Section-dependent constraints were:

• No net transport across section: 0 6 0.25 Sv, alllayers.

• Mediterranean Water outflow: 2.0 6 0.5 Sv westward,layers 15–20 (27.52 , g n , 27.88) (Baringer andPrice 1997).

• No net flow at densities greater than North AtlanticCentral Water: 0 6 0.125 Sv, layers 23–45 (g n .27.94).

g. North America to Europe, 248N

For the Straits of Florida segment of this section,thermal wind was calculated relative to a northward 28.5cm s21 at the bottom, producing a prior transport of 31Sv. For the rest of the section (Bahamas to Africa),thermal wind was calculated relative to an isobaric levelof no motion at 3000 dbar (Rintoul and Wunsch 1991).The inverse model was constrained to retain a net trans-port of 31 6 2 Sv northward in the Straits of Florida(Schott et al. 1988).

h. South America to Africa, 118S

Thermal wind was calculated relative to the isobaricreference levels of Speer et al. (1996): 1100 m in thewestern boundary (west of 35.58W), 3800 m in the Bra-zil Basin (17.678–35.58W), 2400 m over the Mid-At-lantic Ridge (10.178–17.678W), and 4000 m in the An-gola Basin (east of 10.178W). Section-dependent con-straints were:

• Bering Strait volume transport minus integrated NorthAtlantic E 2 P 2 R across 118S: 0.58 6 2.0 Svsouthward, all layers.

• Bering Strait salt flux across 118S: (26.7 6 72.9) 3106 kg s21 salt southward, all layers.

• Antarctic Bottom Water, supplying 2.14 6 1 Sv intothe Guiana Basin (Hall et al. 1997) and 1.22 6 0.25Sv through the Romanche and Chain Fracture Zones(Mercier and Speer 1998): 3.4 6 1 Sv northward,layers 37–45 (g n . 28.12).

Page 18: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

SEPTEMBER 2003 1919L U M P K I N A N D S P E E R

• No net flow of lower deep water east of the Mid-Atlantic Ridge: 0 6 2 Sv, pairs 57–109, layers 35–45 (g n . 28.10).

REFERENCES

Bacon, S., 1997: Circulation and fluxes in the North Atlantic betweenGreenland and Ireland. J. Phys. Oceanogr., 27, 1420–1435.

——, 1998: Decadal variability in the outflow from the Nordic seasto the deep Atlantic Ocean. Nature, 394, 871–874.

Baringer, M. O., and J. F. Price, 1997: Mixing and spreading of theMediterranean outflow. J. Phys. Oceanogr., 27, 1654–1692.

Bersch, M., 1995: On the circulation of the northeastern North At-lantic. Deep-Sea Res., 42A, 1583–1607.

——, J. Meincke, and A. Sy, 1999: Interannual thermohaline changesin the northern North Atlantic 1991-1996. Deep-Sea Res., 46B,55–75.

Coachman, L., and K. Aagaard, 1988: Transports through BeringStrait: Annual and interannual variability. J. Geophys. Res., 93,15 535–15 539.

da Silva, A. M., C. C. Young, and S. Levitus, 1994: Algorithms andProcedures. Atlas of Surface Marine Data 1994. Vol. 1. NOAAAtlas NESDIS 1, 51 pp.

Dickson, R. R., and J. Brown, 1994: The production of North AtlanticDeep Water: Sources, rates, and pathways. J. Geophys. Res., 99,12 319–12 341.

Ganachaud, A. S., 1999: Large scale oceanic circulation and fluxesof freshwater, heat, nutrients and oxygen. Ph.D. thesis, MIT–WHOI Joint Program, 249 pp.

——, and C. Wunsch, 2000: Improved estimates of global oceancirculation, heat transport and mixing from hydrographic data.Nature, 408, 453–457.

Girton, J. B., T. B. Sanford, and R. H. Kase, 2001: Synoptic sectionsof the Denmark Strait Overflow. Geophys. Res. Lett., 28, 1619–1622.

Gould, W. J., L. Loynes, and J. Backhaus, 1985: Seasonality in slopecurrent transports N. W. of Shetland. ICES CM, C:7, 1–13.

Gouretski, V., and K. Jancke, 1998: A new World Ocean climatology:Optimal interpolation of historical and WOCE hydrographic dataon neutral surfaces. WOCE Rep. 162/98. [Available from WOCESpecial Analysis Center, Max Planck Institute, Hamburg, Ger-many.]

Grabs, W., T. de Couet, and J. Pauler, 1996: Freshwater fluxes fromcontinents into the world oceans based on data of the GlobalRunoff Data Base. Global Runoff Data Centre Tech Rep. 10,Koblenz, Germany, 49 pp.

Hall, M. M., M. McCartney, and J. A. Whitehead, 1997: AntarcticBottom Water flux in the equatorial western Atlantic. J. Phys.Oceanogr., 27, 1903–1926.

Hansen, B., K. M. H. Larsen, S. Osterhus, B. Turrell, and S. Jonsson,1999: The Atlantic Water Inflow to the Nordic Seas. Int. WOCENewsl., 35, 33–35.

Jackett, D. R., and T. J. McDougall, 1997: A neutral density variablefor the world’s oceans. J. Phys. Oceanogr., 27, 237–263.

Josey, S. A., E. C. Kent, and P. K. Taylor, 2002: On the wind stressforcing of the ocean in the SOC climatology: Comparisons withthe NCEP–NCAR, ECMWF, UWM/COADS and Hellerman andRosenstein datasets. J. Phys. Oceanogr., 32, 1993–2019.

Koltermann, P., A. Sokov, V. Tereschenkov, S. Dobroliubov, K. Lor-bacher, and A. Sy, 1999: Decadal changes in the thermohalinecirculation of the North Atlantic. Deep-Sea Res., 46B, 109–138.

Kristmannsson, S. S., S.-A. Malmberg, and J. Briem, 1989: Inflowof warm Atlantic water to the sub-Arctic Iceland Sea. Rapp. PvReun. Cons. Int. Explor. Mer, 188, 74.

Large, W. G., and A. J. Nurser, 2001: Ocean surface water masstransformation. Ocean Circulation and Climate, G. Siedler etal., Eds., International Geophysics Series, Vol. 77, AcademicPress, 317–336.

Levitus, S., and T. Boyer, 1994: Temperature. Vol. 4, World OceanAtlas 1994, NOAA Atlas NESDIS 4, 117 pp.

Macdonald, A. M., 1998: The global ocean circulation: A hydro-graphic estimate and regional analysis. Progress in Oceanog-raphy, Vol. 41, Pergamon, 281–382.

Malmberg, S. A., H. G. Gade, and H. E. Sweers, 1972: Currentvelocities and volume transports in the East Greenland Currentoff Cape Nordenskjold in August–September 1965. Proc. Int.Sea Ice Conf., Reykjavik, Iceland, 130–139.

Mann, C. R., 1967: The termination of the Gulf Stream and thebeginning of the North Atlantic Current. Deep-Sea Res., 14, 337–359.

Marsh, R., 2000: Recent variability of the North Atlantic thermo-haline circulation inferred from surface heat and freshwater flux-es. J. Climate, 13, 3239–3260.

Mauritzen, C., 1996: Production of dense overflow waters feedingthe North Atlantic across the Greenland–Scotland Ridge. Deep-Sea Res., 43A, 769–835.

McCartney, M. S., 1992: Recirculating components to the deepboundary current of the northern North Atlantic. Progress inOceanography, Vol. 29, Pergamon, 283–383.

——, and L. D. Talley, 1982: The subpolar mode water of the NorthAtlantic Ocean. J. Phys. Oceanogr., 12, 1169–1188.

Mercier, H., and K. G. Speer, 1998: Transport of bottom water in theRomanche Fracture Zone and the Chain Fracture Zone. J. Phys.Oceanogr., 28, 779–790.

Morgan, P. P., 1995: Box inverse modelling with DOBOX 4.2. CSIROMarine Laboratories, Rep. 225, 26 pp.

Niiler, P. P., and J. Stevenson, 1982: The heat budget of tropical oceanwarm-water pools. J. Mar. Res., 40 (suppl.), 465–480.

Orsi, A. H., W. J. Smethie, and J. L. Bullister, 2002: On the totalinput of Antarctic waters to the deep ocean: A preliminary es-timate from chlorofluorocarbon measurements. J.Geophys. Res.,107 (C8), 31-1–31-14.

Østerhus, S., B. Hansen, R. Kristiansen, and P. Lundberg, 1999: Theoverflow through the Faroe Bank Channel. Int. WOCE Newsl.,35, 35–37.

Price, J., T. Sanford, C. Mauritzen, and M. Prater, cited 2001: A reportof XCP and CTD data taken during June, 2001, RRS Discoverycruise 247B: A process study of the Faroe Bank Channel over-flow. [Available online at http://www.whoi.edu/science/PO/people/jprice/oflow/FBC.pdf.]

Rhein, M., and Coauthors, 2002: Labrador Sea Water: Pathways,CFC-inventory and formation rates. J. Phys. Oceanogr., 32, 648–665.

Rintoul, S. R., and C. Wunsch, 1991: Mass, heat, oxygen and nutrientfluxes and budgets in the North Atlantic Ocean. Deep-Sea Res.,38 (suppl.), S355–S377.

Roemmich, D., and C. Wunsch, 1985: Two transatlantic sections:Meridional circulation and heat flux in the subtropical NorthAtlantic Ocean. Deep-Sea Res., 32, 619–664.

Ross, C. K., 1984: Temperature–salinity characteristics of the ‘‘over-flow’’ water in Denmark Strait during ‘‘OVERFLOW ‘73’’. ICESMarine Science Symp.: Hydrobiological Variability in the NorthAtlantic and Adjacent Seas, 185, 111–119.

Saunders, P. M., 1990: Cold outflow from the Faroe Bank Channel.J. Phys. Oceanogr., 20, 29–43.

Schmitt, R. W., P. S. Bogden, and C. E. Dorman, 1989: Evaporationminus precipitation and density fluxes for the North Atlantic. J.Phys. Oceanogr., 19, 1208–1221.

Schmitz, W. J., and M. S. McCartney, 1993: On the North Atlanticcirculation. Rev. Geophys., 31, 29–49.

Schott, F. A., T. N. Lee, and R. Zantopp, 1988: Variability of structureand transport of the Florida Current in the period range of daysto seasonal. J. Phys. Oceanogr., 18, 1209–1230.

Sloyan, B. M., and S. R. Rintoul, 2001: The Southern Ocean limbof the global deep overturning circulation. J. Phys. Oceanogr.,31, 143–173.

Smethie, W. M., D. W. Chipman, J. H. Swift, and K. P. Koltermann,2000: Chlorofluoromethanes in the Arctic Mediterranean Seas:

Page 19: NOAA's Atlantic Oceanographic and Meteorological Laboratory - … · 2009-10-14 · The meridional circulation in the North Atlantic Ocean results from the interaction of several

1920 VOLUME 33J O U R N A L O F P H Y S I C A L O C E A N O G R A P H Y

Evidence for formation of bottom water in the Eurasian Basinand deep-water exchange through Fram Strait. Deep-Sea Res.,35A, 347–369.

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

——, and E. Tziperman, 1990: Convection from a source in an oceanbasin. Deep-Sea Res., 37, 431–446.

——, J. Holfort, T. Reynaud, and G. Siedler, 1996: South Atlanticheat transport at 118S. The South Atlantic: Present and PastCirculation, G. Wefer et al., Eds., Springer-Verlag, 105–120.

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

Wijffels, S. E., R. W. Schmitt, H. L. Bryden, and A. Stigebrandt,

1992: Transport of freshwater by the oceans. J. Phys. Oceanogr.,22, 155–162.

Worthington, L. V., 1976: On the North Atlantic circulation. JohnsHopkins Oceanographic Studies 6, Woods Hole OceanographicInstitution, Rep. 6, 110 pp.

Wunsch, C., 1984: An eclectic Atlantic Ocean circulation model. PartI: The meridional flux of heat. J. Phys. Oceanogr., 14, 1712–1733.

——, 1996: The Ocean Circulation Inverse Problem. Cambridge Uni-versity Press, 442 pp.

——, and B. Grant, 1982: Towards the general circulation of theNorth Atlantic Ocean. Progress in Oceanography, Vol. 11, Per-gamon, 1–59.