direct observation of subtropical mode water circulation in the western north atlantic ocean

75
Author’s Accepted Manuscript Direct Observation of Subtropical Mode Water Circulation in the Western North Atlantic Ocean David M. Fratantoni, Young-Oh Kwon, Benjamin A. Hodges PII: S0967-0645(13)00084-2 DOI: http://dx.doi.org/10.1016/j.dsr2.2013.02.027 Reference: DSRII3313 To appear in: Deep-Sea Research II Cite this article as: David M. Fratantoni, Young-Oh Kwon and Benjamin A. Hodges, Direct Observation of Subtropical Mode Water Circulation in the Western North Atlantic Ocean, Deep-Sea Research II, http://dx.doi.org/10.1016/j.dsr2.2013.02.027 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting galley proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. www.elsevier.com/locate/dsr2

Upload: benjamin-a

Post on 04-Dec-2016

221 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

Author’s Accepted Manuscript

Direct Observation of Subtropical Mode WaterCirculation in the Western North Atlantic Ocean

David M. Fratantoni, Young-Oh Kwon, Benjamin A.Hodges

PII: S0967-0645(13)00084-2DOI: http://dx.doi.org/10.1016/j.dsr2.2013.02.027Reference: DSRII3313

To appear in: Deep-Sea Research II

Cite this article as: David M. Fratantoni, Young-Oh Kwon and Benjamin A. Hodges,Direct Observation of Subtropical Mode Water Circulation in the Western North AtlanticOcean, Deep-Sea Research II, http://dx.doi.org/10.1016/j.dsr2.2013.02.027

This is a PDF file of an unedited manuscript that has been accepted for publication. As aservice to our customers we are providing this early version of the manuscript. Themanuscript will undergo copyediting, typesetting, and review of the resulting galley proofbefore it is published in its final citable form. Please note that during the production processerrors may be discovered which could affect the content, and all legal disclaimers that applyto the journal pertain.

www.elsevier.com/locate/dsr2

Page 2: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

Direct Observation of Subtropical Mode Water 1Circulation in the Western North Atlantic Ocean 2

3456

David M. Fratantoni 7Young-Oh Kwon 8

Benjamin A. Hodges 9101112

Physical Oceanography Department 13Woods Hole Oceanographic Institution 14

Woods Hole, MA 02543 1516171819

Revised 22 February 2013 20212223

Submitted to Deep-Sea Research II Special Issue on 24Subtropical Mode Water in the North Atlantic Ocean 25

26272829

Corresponding Author:3031

Dr. David M. Fratantoni 32Autonomous Systems Laboratory 33Physical Oceanography Department 34Woods Hole Oceanographic Institution 35Woods Hole, MA 02543 36

37Telephone: (508) 289-2908 38Fax: (508) 457-2181 39Email: [email protected] 40

41

Page 3: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

2

Abstract42

43

Eighteen Degree Water (EDW) is the dominant subtropical mode water of the North 44

Atlantic subtropical gyre and is hypothesized as an interannual reservoir of anomalous 45

heat, nutrients and CO2. Although isolated beneath the stratified upper-ocean at the end 46

of each winter, EDW may reemerge in subsequent years to influence mixed layer 47

properties and consequently air-sea interaction and primary productivity. Here we report 48

on recent quasi-Lagrangian measurements of EDW circulation and stratification in the 49

western subtropical gyre using an array of acoustically-tracked, isotherm-following, 50

bobbing profiling floats programmed to track and intensively sample the vertically-51

homogenized EDW layer and directly measure velocity on the 18.5˚ C isothermal 52

surface. 53

54

The majority of the CLIVAR Mode Water Dynamics Experiment (CLIMODE) bobbers 55

drifted within the subtropical gyre for 2.5–3.5 years, many exhibiting complex looping 56

patterns indicative of an energetic eddy field. Bobber-derived Lagrangian integral time 57

and length scales (3 days, 68 km) associated with motion on 18.5˚ C were consistent with 58

previous measurements in the Gulf Stream extension region and fall between previous 59

estimates at the ocean surface and thermocline depth. Several bobbers provided evidence 60

of long-lived submesoscale coherent vortices associated with substantial EDW thickness. 61

While the relative importance of such vortices remains to be determined, our 62

observations indicate that these features can have a profound effect on EDW distribution.63

EDW thickness (defined using a vertical temperature gradient criterion) exhibits seasonal 64

Page 4: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

3

changes in opposition to a layer bounded by the 17˚ C and 19˚ C isotherms. In particular, 65

EDW thickness is generally greatest in winter (as a result of buoyancy-forced 66

convection), while the 17˚–19˚ C layer is thickest in summer consistent with seasonal 67

Ekman pumping. Contrary to previous hypotheses, the bobber data suggests that a 68

substantial fraction of subducted EDW is isolated from the atmosphere for periods of less 69

than 24 months. Seasonal-to-biennial reemergence (principally within the recirculation 70

region south of the Gulf Stream) appears to be a common scenario which should be 71

considered when assessing the climatic and biogeochemical consequences of EDW. 72

1. Introduction 73

74

Eighteen Degree Water (EDW) is the archetype for the anomalously thick and vertically 75

homogenous mode waters typical of all subtropical western boundary current systems 76

(e.g. Hanawa and Talley, 2001). EDW, the subtropical mode water of the northwestern 77

Atlantic Ocean (e.g. Worthington, 1959), is notable for both its vertical homogeneity (a 78

minimally-stratified layer 200–500 m thick; Fig. 1) and horizontal extent (~40° in 79

longitude and ~20° in latitude). EDW is in direct contact with the atmosphere only 80

during wintertime and only in limited areas within the Gulf Stream and an adjacent 81

recirculation to the south (Talley and Raymer, 1982; Alfultis and Cornillon, 2001; Kwon, 82

2003; Joyce et al., 2009; Joyce et al., 2012, this issue). Elsewhere, and for the remainder 83

of the year, EDW is isolated from direct contact with the atmosphere by the stratified 84

upper ocean. 85

86

Page 5: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

4

Observational analyses (e.g. Kwon and Riser, 2004) and numerical modeling studies (e.g. 87

Old and Haines, 2006; Forget et al., 2011) have shown that the large volume of EDW 88

formed each winter and isolated below the seasonal thermocline has the potential to store 89

and integrate (i.e. “remember”) the consequences of air-sea exchange for several years.90

As a result, EDW impacts climate and marine ecosystems by serving as an interannual 91

reservoir of anomalous heat, nutrients and CO2 (Bates et al., 2002; Kwon and Riser, 92

2004; Dong and Kelly, 2004; Palter et al., 2005; Dong et al., 2007). A notable 93

characteristic of EDW is that its volume is anti-correlated with the total upper ocean heat 94

content in the subtropical gyre, so that EDW is actually a deficit heat reservoir (Kwon, 95

2003; Dong et al., 2007). That is, winters with greater air-sea cooling result in a thicker 96

and slightly cooler variant of EDW, and vice versa (Talley and Raymer, 1982; Kwon and 97

Riser, 2004).98

99

EDW, formed near the surface during a period of high primary productivity (Siegel et al., 100

2002), is nutrient-poor. A recent analysis of historical hydrographic data (Palter et al., 101

2005) has suggested that EDW formation and advection can influence biogeochemical 102

processes in the subtropical gyre by introducing variability in the subsurface nutrient 103

reservoir. In particular, Palter et al., (2005) suggest that both spatial and interannual 104

changes in EDW formation may have consequences for subsequent productivity 105

hypothesized to be limited by nutrient delivery from below the euphotic zone. 106

107

The mechanisms and consequences of EDW formation have been studied and debated for 108

decades (e.g. Worthington, 1959, 1972; Warren, 1972; Speer and Tziperman, 1992; 109

Page 6: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

5

Thomas, 2005; Thomas and Joyce, 2010; Joyce, 2009; see Marshall et al., 2009, for a 110

review), yet substantial questions remain regarding the rates and mechanisms of EDW 111

formation, dispersal, and decay. In particular, little is known of the character and 112

intensity of the circulation at the depth of the EDW layer and the role played by meso- 113

and submesoscale phenomena unresolved by previous measurement programs. In 114

addition, while several estimates of the EDW formation rate have been proposed in recent 115

years (typically an annualized 5–15 Sv; e.g. Speer and Tziperman, 1992; Kwon and 116

Riser, 2004; Forget et al., 2011; Joyce, 2012; Joyce et al., 2012, this issue), less attention 117

has been paid to the mechanisms responsible for the required comparable rate of EDW 118

destruction (i.e. restratification). Hypothesizing that meso- and submesoscale circulation 119

features contribute to both the wide dispersal and eventual destruction of EDW, we 120

designed an experiment to directly measure the character and intensity of circulation at 121

the depth of the EDW thermostad with appropriate temporal and spatial resolution. 122

123

The 2005–2007 CLIVAR Mode Water Dynamics Experiment (CLIMODE) field 124

campaign (Marshall et al., 2009; see http://climode.org) resulted in a unique and 125

comprehensive array of EDW-related measurements including air-sea fluxes, detailed 126

hydrographic and velocity observations within the Gulf Stream, and gyre-scale 127

measurements of upper-ocean circulation and stratification. The program incorporated 5 128

research cruises conducted between November 2005 and November 2007, including 2 129

extensive mid-winter cruises in the Gulf Stream. As one component of this effort, 40 130

acoustically-tracked, bobbing, profiling floats (“bobbers”) were used to explore, in a 131

Lagrangian sense, the processes responsible for the dispersal and eventual destruction of 132

Page 7: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

6

EDW in the western North Atlantic subtropical gyre. In this contribution we describe the 133

bobber measurement program and present unique measurements of the northwestern 134

Atlantic circulation at the depth of the EDW thermostad during the CLIMODE 135

observational period. 136

137

The remainder of this article is structured as follows: In Section 2 we describe our 138

instrumentation and sampling methods. Examples of the fundamental trajectory and 139

stratification measurements collected by the CLIMODE bobber array are presented in 140

Section 3. In Section 4 we synthesize these measurements into basin-scale composites to 141

infer general patterns of EDW circulation in the northwest Atlantic. Finally, in Section 5 142

we provide a brief summary of our findings and suggest avenues for further study. 143

144

2. Methods 145

146

During the 1991–1993 Subduction Experiment in the eastern North Atlantic, 147

acoustically-tracked SOFAR floats (Rossby et al., 1975) were deployed with active 148

buoyancy control to track the motion and evolution of a subducting slab of water (James 149

Price, pers. comm.; see Joyce et al., 1998). The CLIMODE bobbers were initially 150

inspired by this earlier work of Price. Rossby and Prater (2005) subsequently used 151

isopycnal RAFOS floats equipped with modest buoyancy control to explore changes in 152

layer thicknesses along Lagrangian trajectories in the North Atlantic Current. 153

154

Page 8: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

7

The CLIMODE bobbers are modified APEX (Autonomous Profiling Explorer; Teledyne 155

Webb Research Corp., E. Falmouth, MA) profiling floats comparable in form and 156

function to those comprising the current global Argo array (e.g. Davis et al., 1992; Davis 157

et al., 2001; Roemmich et al., 2009). Briefly, a profiling float is a 1.5 m, 25 kg, satellite-158

tracked quasi-Lagrangian drifter which periodically descends and ascends vertically 159

through the ocean by changing its effective density (ratio of mass to volume, i.e. 160

buoyancy) relative to that of the surrounding water. In practice, this is done by altering 161

the volume of the float: Fluid (generally oil) is moved via pump or piston between a 162

reservoir within an aluminum pressure hull and a flexible external bladder. The 163

(constant) total mass of a bobber is set in the laboratory based on anticipated oceanic 164

temperature and density conditions. For a float correctly ballasted for its region of 165

operation a volume change of 200 cm3 results in an ascent/descent rate of approximately 166

10 cm/s.167

168

The differences between an “Argo float” and a “CLIMODE bobber” are few but 169

significant. Several of these differences are visually apparent (Fig. 2) while others are 170

software-based and evident only in the behavior of the deployed instrument. EDW 171

generally occupies only the upper few hundred meters of the water column. Thus floats 172

intended to remain within the EDW layer must be capable of achieving and maintaining 173

neutral buoyancy at depths much shallower than the 1000 m parking depth typical of 174

Argo floats, including within the minimally-stratified and turbulent wintertime mixed 175

layer. To this end the CLIMODE bobber’s aluminum pressure hull was designed to be 176

Page 9: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

8

shorter, thinner, and lighter than typical and substantially greater onboard energy (in the 177

form of primary lithium batteries) was allocated to active buoyancy control. 178

179

Unlike the standard conductivity-temperature-depth (CTD) instrument carried by Argo 180

floats (e.g. Roemmich et al., 2009) the bobbers were equipped with temperature and 181

pressure sensors only. In the western North Atlantic the relationship between 182

temperature and salinity near the EDW thermostad is fairly robust (Fig. 3) allowing 183

density to be reasonably approximated from temperature alone. While others have noted 184

interesting and relevant regional variation in salinity along density surfaces in this area 185

(e.g. Joyce et al., 2012, this issue), a conscious decision was made to neglect direct 186

measurement of salinity in favor of constructing and deploying a larger total number of 187

bobbers.188

189

In addition to temperature and pressure sensors, the bobbers also carried an external low-190

frequency hydrophone and an internal RAFOS navigation system (Rossby et al., 1986; 191

Seascan, Inc., Falmouth, MA) which detected daily acoustic transmissions from an array 192

of moored sound sources. Arrival times of these transmissions were used post-193

experiment to determine each bobber’s daily geographic position. Position records were 194

subsequently differentiated in time to compute daily average horizontal (isothermal) 195

bobber velocities. 196

197

The goal of this program was to directly measure the dispersal of EDW as deep 198

wintertime mixed layers are subducted beneath the subtropical gyre stratification and/or 199

Page 10: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

9

capped by the seasonal thermocline. The bobber mission was thus designed to maximize 200

the number of temperature and acoustically-derived position/velocity measurements 201

within the EDW layer, with only occasional measurements above and below this layer. 202

While Argo floats typically drift isobarically at 1000 dbar for 10 days between 203

surfacings, the 40 CLIMODE bobbers were programmed to actively track the 18.5˚ C 204

temperature surface. Analysis of previous measurements in the region suggested that this 205

isotherm was, with some interannual variation, grossly representative of the vertical 206

center of the minimally-stratified EDW layer in the western Atlantic (Fig. 3). The 207

bobber’s onboard controller was programmed with a feedback loop that, informed by 208

hourly-measured in-situ temperature, adjusted the float’s buoyancy by controlling the 209

amount of oil pumped into the external bladder. The bobbers were programmed to 210

closely (+/- 0.1˚ C) follow the 18.5˚ C isotherm continually for 3 days, then “bob” 211

quickly between the 17˚ C and 20˚C isotherms before returning again to 18.5˚ C. Bobs 212

were completed as rapidly as possible (using the full range of the available volume 213

change) to maximize the time spent on the 18.5˚ C isotherm and minimize contamination 214

of the horizontal (isothermal) velocity record. 215

216

Specific behavioral rules were devised in case a bobber should fail to locate one of its 217

target isotherms. For example, each bobber was instructed to maintain an isobaric 218

position of 50 dbar should the 18.5˚C isotherm be unattainable (e.g. during wintertime 219

surface exposure of the EDW layer). The alternative – drifting at the surface – would 220

have contaminated the position/velocity record due to the direct action of wind and 221

surface waves on the bobber. The drift/bob cycle was repeated 10 times over 30 days, 222

Page 11: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

10

punctuated with a final descent to 1000 m before surfacing to transmit temperature, 223

pressure, RAFOS, and engineering data via satellite. Bobbers remained on the surface 224

for approximately 12 hours each month during CLS/Argos satellite data transmission. 225

226

The bobbers were acoustically tracked using an array of 4 moored sound sources (Fig. 4) 227

built at the Graduate School of Oceanography, University of Rhode Island (T. Rossby, 228

pers. comm.). Each source emitted a daily, 80 second, 260 Hz swept signal detectable for 229

hundreds to thousands of kilometers. The source moorings were deployed in November 230

2005 and recovered in November 2007. Additional sound sources deployed in the North 231

Atlantic by investigators from the U.S., France, and Germany were occasionally heard by 232

the bobbers and were used during post-experiment tracking. Time-series measurements 233

of temperature, salinity, and velocity were recorded at several depths on each source 234

mooring. Details of the CLIMODE subsurface moored measurements are described 235

elsewhere (e.g. Davis et al., 2012, this issue). 236

237

In summary, each bobber yielded daily-resolved position, velocity, and pressure; 3-day-238

resolved EDW stratification and thickness; and monthly-resolved 0–1000 m temperature 239

profiles. Vertical resolution of temperature ranged from 5 m near the surface to 20 m 240

below 500 m depth. 241

242

243

244

245

Page 12: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

11

246

3. Results 247

3.1 Bobber Performance and Data Return 248

249

From an engineering perspective overall bobber performance was excellent: 28 of 40 250

bobbers had a 100% success rate (actual/expected satellite transmissions of valid data) 251

and 38 of 40 had a success rate greater than 90%. Bobber lifetimes were substantially 252

greater than anticipated: The half-life of the bobber array exceeded 4 years, almost twice 253

as long than the overall duration of the CLIMODE field campaign. A timeline 254

illustrating the lifetime of the bobber array is shown in Fig. 5. 255

256

During the CLIMODE experiment acoustic detection ranges were reduced versus more 257

typical deep RAFOS float applications due to the shallow depth of the bobbers (well 258

above the SOFAR channel and occasionally within the mixed layer) and the strong lateral 259

density gradients associated with the Gulf Stream and its rings. Nevertheless, source 260

transmissions were detected at ranges as large as 2000 km (Fig. 6). Most valid positions 261

were determined using sources at ranges less than 1000 km. Post-experiment acoustic 262

tracking (Wooding et al., 2005) enabled position and velocity estimation with a temporal 263

resolution of 1 day and geographic uncertainty (relative to CLS/Argos satellite-derived 264

surface position) of approximately 5 km (see Fratantoni et al., 2010). Errors in absolute 265

positioning are due primarily to poorly-resolved non-constant clock drift on both sources 266

and receivers. Thirty-seven of 40 bobbers were acoustically tracked over all or part of 267

Page 13: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

12

their lifetime. Additional details regarding bobber performance and the acoustic tracking 268

procedure can be found in Fratantoni et al., (2010). 269

270

Bobbers were primarily deployed in winter (January 2006 and February 2007) along the 271

southern edge of the Gulf Stream (Fig. 7). Additional bobbers were deployed during 272

mooring deployment and turn-around cruises in November 2005 and November 2006. 273

Care was taken to deploy bobbers south of the Gulf Stream axis to minimize the loss of 274

instruments from the subtropical gyre to the slope water north of the Gulf Stream. 275

276

The bobbers returned two independent data types: Temperature profiles and acoustically-277

derived position/velocity measurements. The profile data described herein includes that 278

obtained between July 2005 and May 2009. Due to the limited deployment duration of 279

the CLIMODE sound source moorings, acoustic tracking was only available between 280

November 2005 and November 2007. A census of available data (Fig. 8) demonstrates 281

that the total number of monthly temperature profiles (bobs and deep profiles combined) 282

increased with each deployment cruise until the final bobbers were launched in March 283

2007, and then slowly declined. The number of valid RAFOS velocity measurements 284

likewise increased with time until the source array was recovered in late November 2007. 285

The distribution of temperature profiles by month is fairly consistent. Due to preferential 286

wintertime deployments and a shorter 2-year measurement window, RAFOS velocity 287

measurements are skewed slightly towards the spring and summer seasons. 288

289

Page 14: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

13

Every third day, each bobber rapidly descended from its parking depth on the 18.5˚ C 290

isotherm to the depth of the 17˚ C isotherm, ascended to the 20˚ C isotherm, and then 291

returned to the 18.5˚ C isotherm. As expected, the bounding depths of these bobs varied 292

with both season and latitude. Most bobs sampled a vertical interval from approximately 293

100 m to 500 m depth (Fig. 9). The dominant modes shown in the bob top, bottom, and 294

extent histograms approximate the regional- and time-averaged distributions of the 17˚295

and 19˚ C isotherms and their vertical separation, respectively. Where the 20˚ C isotherm 296

did not exist the bobber was allowed to profile to the surface – values corresponding to a 297

bob-top depth of zero have been suppressed in Fig. 9. The near-surface mode of the 298

“bob-bottom” histogram results from situations in which neither bounding isotherm 299

existed – a state which can only be realized after the bobber completes a finite vertical 300

excursion to measure the stratification. 301

302

3.2 Bobber Trajectories 303

304

Fig. 10 summarizes the trajectories of the CLIMODE bobbers. In Fig. 10a, only initial 305

and final positions are shown to illustrate the overall displacement of the bobbers from 306

their deployment locations. In Fig. 10b, low-resolution Argos satellite tracking provides 307

roughly 30-day position resolution of each bobber trajectory. The great majority of 308

bobbers exhibited a net southward displacement and remained within the western 309

subtropical gyre for at least 2.5–3.5 years. Several bobbers did escape the subtropics by 310

moving northward across the Gulf Stream / North Atlantic Current east of the Grand 311

Banks between 30–40˚ W. The fraction of bobbers escaping the subtropical gyre (3 of 312

Page 15: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

14

40) is comparable to that found by Kwon and Riser (2005; 7 of 71), and also similar to 313

surface drifter studies by Brambilla and Talley (2006) and Rypina et al., (2011). 314

315

The most adventurous bobber drifted through the eastern Norwegian Sea and was last 316

heard from in the Arctic Ocean at 79˚ N near Spitsbergen. A single bobber passed 317

southward into the Caribbean Sea through the Anegada Passage east of Puerto Rico and 318

nearly reached the Yucatan Strait. Although most bobbers were contained within the 319

western subtropical gyre, 8 eventually drifted eastward across the mid-Atlantic ridge with 320

2 reaching the Canary Basin. 321

322

As described above, the bobbers collected both position/velocity measurements (daily) 323

and temperature profiles (every third day). Fig. 11 summarizes the geographic 324

distribution of these measurements in a smaller region of the western subtropical gyre 325

within which our analyses will focus. Fig. 11a illustrates the spatial distribution of 326

stratification measurements including both every third day EDW-centric bobs (blue) and 327

monthly 1000 m deep profiles (red). Profile locations are presented using the best-328

available positioning method: acoustic RAFOS tracking if available, and linearly-329

interpolated Argos satellite tracking otherwise. As per the experiment design, 330

measurement density is greatest to the south of the Gulf Stream and diminishes rapidly 331

east of 45˚ W (the mid-Atlantic ridge) and south of 30˚ N.332

333

Fig. 11b presents a composite of all valid acoustically-tracked bobber trajectory 334

segments. This figure provides an unprecedented overview of the general character of 335

Page 16: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

15

circulation on a temperature surface associated with the vertical center of the EDW layer. 336

Immediately notable in Fig. 11b are loops and swirls associated with bobbers trapped 337

within various recirculations, eddies, and (in a few cases) long-lived, translating, coherent 338

vortices. Several of these features will be examined in greater detail below. 339

340

Trajectory information near the Gulf Stream and mid-Atlantic Ridge is occasionally 341

fragmented due to difficulties encountered in post-experiment acoustic tracking. 342

Throughout this contribution only trajectory information collected from bobbers drifting 343

on the appropriate temperature surface (18.5˚ C +/- 0.1˚ C) will be used for display and 344

analysis. In order to most effectively describe the trajectory of each bobber over its 345

complete lifetime, we use composite best-available trajectories consisting of high-346

resolution RAFOS-tracked position data supplemented by Argos satellite tracking when 347

required. A sample of 16 such composite trajectories illustrating the substantial variation 348

in regional circulation patterns is shown in Fig. 12 (c.f. Kwon and Riser, 2005). 349

350

The Gulf Stream’s zonal extension defines the northern extent of the subtropical gyre and 351

is the dominant circulation feature in the northwest Atlantic. Several bobbers deployed 352

just south of the Gulf Stream were swept rapidly eastward (with speeds exceeding 70 353

cm/s) to 50–60˚ W before slowly drifting into the gyre interior (e.g. Bobbers 448, 721, 354

766 in Fig. 12). Bobbers deployed slightly further south within the moored array (e.g. 355

577, 585, 604 in Fig. 12) did not encounter the eastward Gulf Stream and drifted 356

generally southward and/or westward. A total of 13 of 40 bobbers drifted sufficiently 357

south and west to re-enter the northward western boundary current flow (Antilles 358

Page 17: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

16

Current, Florida Current, or Gulf Stream depending on latitude) and recirculate within the 359

subtropical gyre. Six examples of this behavior are shown in Fig. 12 (e.g. 474, 518, 585).360

A large fraction of the trajectories shown in Fig. 12 demonstrate substantial southward 361

movement between 50–60˚ W, consistent with altimeter-derived mean dynamic 362

topography (e.g. Jayne, 2006; see below Fig. 17). 363

364

3.3 Case Studies: Stratification and Along-Track EDW Thickness Evolution 365

366

Four interesting examples of combined trajectory and stratification measurements are 367

shown in Fig. 13a and Fig. 13b. In addition to the observed quantities of position, 368

temperature and pressure, these figures also illustrates the derived RAFOS velocities and 369

the time variation of EDW thickness (see Appendix A) computed along the bobber 370

trajectory. 371

372

Bobber 448: Long trajectory, minimal EDW thickness. This bobber (Fig. 13a) 373

experienced 3 full annual cycles during its lifetime but only tracked a significant amount 374

of EDW during the first 3 months of its mission. Deployed just south of the Gulf Stream 375

in January 2006, this bobber was swept eastward for several days before drifting south 376

into the gyre interior near 50˚ W. Initially drifting isothermally at 300 m on 18.5˚ C, the 377

bobber rose to within 50 m of the surface while in the Gulf Stream before moving rapidly 378

southward beneath the warmer surface waters of the subtropical gyre. Upon deployment 379

the local EDW thickness was approximately 150 m. This increased to 400 m in the 380

wintertime Gulf Stream but decayed to near zero as the bobber moved southward to 30˚381

Page 18: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

17

N over the next 3 months. By July 2006 the bobber was south of 30˚ N on the western 382

flank of the mid-Atlantic Ridge and the local EDW thickness had vanished completely. 383

Note that the thickness of the 17˚–19˚ C layer alone (gray curve in panel (d)) does not 384

adequately capture the evolution of EDW thickness – the temperature gradient criterion 385

(Appendix A) allows us to discriminate the minimally-stratified mode water from the 386

background stratification. This distinction is important, as potential vorticity 387

conservation (for example, f/H where f is the Coriolis parameter and H the local layer 388

thickness – an approximation valid for low-Rossby number, large-scale circulation) 389

implies that the 17˚–19˚ C layer must thin to compensate for the reduction in planetary 390

vorticity as the water parcel (i.e. the bobber) is advected equatorward. 391

392

Bobber 721: Minimal translation, annual reemergence: Though deployed at almost the 393

same time and position as the previous example, Bobber 721 (Fig. 13a) remained north of 394

32˚ N for most of its lifetime and tracked a substantial layer of EDW through 2 annual 395

cycles. EDW thickness increased from 200 m to 450 m during the first winter, then 396

remained steady at 150–200 m. The bobber was re-exposed to the mixed layer along the 397

southern edge of the Gulf Stream in March–April 2007 resulting in a second transient 398

thickness increase. EDW thickness decreased (and bobber depth increased) as it 399

migrated southwestward towards Bermuda in September 2007.400

401

Bobber 598: Gradual EDW decay, western boundary encounter: Also deployed south of 402

the Gulf Stream in January 2006, Bobber 598 (Fig. 13b) illustrates several scenarios 403

observed frequently during this experiment: (a) extended and vigorous recirculation 404

Page 19: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

18

between 35˚ N and the Gulf Stream, (b) southwestward translation including a variety of 405

looping behaviors characteristic of mesoscale eddies accompanied by gradual reduction 406

in local EDW thickness, and (c) an encounter with the western boundary and re-407

absorption into the Gulf Stream system. An initial EDW thickness of 200 m decayed to 408

less than 50 m over 8 months (November 2007–August 2008) as the bobber moved 409

southwestward from the recirculation region towards the Bahamas. Although poorly 410

resolved by monthly Argos-only tracking after November 2007, it is evident that the 411

bobber re-entered the western boundary current and eventually crossed to the north side 412

of the Gulf Stream where the surface temperature was substantially lower.413

414

Bobber 474: Long-lived anticyclonic eddy, western boundary encounter 415

As a final example, Bobber 474 (Fig. 13b) was deployed in November 2006 at 300 m 416

depth in a 200 m thick EDW layer near the southern edge of the Gulf Stream. Following 417

deployment the bobber moved southwestward while occasionally looping with no 418

preferred sense of rotation. The local EDW thickness remained near 300 m for almost a 419

year following a brief wintertime thickening to nearly 500 m. In late summer of 2007 420

this bobber began looping anticyclonically and remained trapped within a long-lived 421

translating eddy for more than 4 months. A more detailed view of the looping portion of 422

this trajectory is shown in Fig. 14. Subsequent investigation of this and similar features 423

in the context of satellite altimetry and co-located surface drifters (Park et al., 2012) 424

suggests a strong correlation between surface vortices and those revealed by the bobbers 425

at EDW depth. 426

427

Page 20: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

19

4. Basin-Scale Synthesis 428

429

To investigate low-frequency, basin-scale aspects of EDW circulation the bobber 430

measurements were averaged in spatial bins to construct quasi-Eulerian fields of velocity, 431

eddy kinetic energy (EKE), isotherm depth, and EDW thickness. Although the bobber 432

measurements cover only a fraction of the North Atlantic, they do provide a unique and 433

high-resolution snapshot of EDW structure and circulation during the CLIMODE field 434

experiment. Additional contemporary measurements including ongoing Argo profiling 435

float observations provide a broader spatial and temporal context for the bobbers, and are 436

examined in separate contributions (e.g. Park et al., 2012) and in future work. 437

438

4.1 Velocity Statistics and Mean Fields 439

440

The fields presented below are based on observations grouped into 2o x2o boxes. In any 441

such analysis there is a trade-off between spatial resolution, areal coverage, and statistical 442

reliability of the box-averaged quantities. This box size was chosen subjectively to 443

provide a reasonable depiction of the circulation while ensuring that most boxes contain 444

sufficient data to form statistically reliable mean values. To that end, 90-day non-445

overlapping segments of isothermal bobber velocity were used to estimate the Lagrangian 446

time and length scales associated with EDW circulation in the western North Atlantic 447

(see e.g. Lumpkin et al., 2002). The valid isothermal velocity observations were 448

distributed vertically with a mean depth of 185 m and standard deviation of 80 m. The 449

velocity autocorrelation function (Davis, 1991; see Lumpkin et al., 2002 for the specific 450

Page 21: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

20

definition used herein) was computed for each of 75 such segments, then ensemble 451

averaged and integrated to the first zero crossing to yield a Lagrangian integral timescale 452

estimate (TL ; see Lumpkin et al., 2002) of approximately 3.2 days. No significant 453

differences were noted between TL computations using meridional vs. zonal components 454

of velocity. The Lagrangian integral length scale (LL) is related to TL via a characteristic 455

speed, LL=u’TL. Taking u’ as the RMS velocity magnitude averaged over the 75 90-day 456

trajectory segments (approximately 22 cm/s) yields a length scale of approximately 68 457

km, or roughly twice the local first-baroclinic deformation radius (30–40 km over much 458

of the study area as reported by Chelton et al., 1998).459

460

A sensitivity study was conducted to determine the potential impact of different segment 461

lengths on these calculations. Segment lengths varying from 30–120 days resulted in 462

time and length scales which differed by less than 7%. Based on the observed ability of 463

monthly-surfacing bobbers to track translating features (e.g. Bobber 474; Fig. 14) we are 464

confident that monthly bobber surfacing (resulting in removal and reinsertion of the 465

bobber from a tracked water parcel) does not significantly affect these calculations. 466

467

The estimated time, length, and velocity scales within the EDW layer are consistent with 468

previous float and drifter measurements in the same general region as summarized by 469

Lumpkin et al., (2002) (Fig. 15). The bobber scale estimates fall between prior 470

measurements at the surface and 700 m depth. In their analysis, Lumpkin et al., (2002) 471

defined adjacent geographical regions encompassing the Gulf Stream extension and the 472

MODE region (e.g. Rossby et al., 1983) in the southwest subtropical gyre. The bobber-473

Page 22: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

21

derived integral scales are most similar to previous measurements in the Gulf Stream 474

extension region. Note that there are fewer bobber measurements in the less energetic 475

MODE region than in the vicinity of the Gulf Stream, and thus that the aggregate scale 476

estimates reported here are biased towards the latter region. Further analysis, not 477

included herein, is required to determine the degree of spatial and/or temporal variability 478

in these parameter estimates. 479

480

The integral time and length scales were used to quantify the independence (i.e. the 481

number of degrees of freedom(DOF)) of individual observations contributing to spatially-482

binned velocity and temperature fields. The DOF associated with each grid cell was 483

computed by following individual bobbers as they translated through a 2o x2o rectilinear 484

grid. Measurements obtained in a given cell were judged to be independent if (a) they 485

resulted from different bobbers, or (b) they resulted from the same bobber but that bobber 486

remained in the box for more than one Lagrangian integral timescale (taken to be 3 days). 487

The raw distribution of all RAFOS-derived velocity measurements and the resulting DOF 488

distribution for valid, isothermal velocity measurements are shown in Fig. 16. Data489

density, and thus DOF, in a given area is a function of both the initial deployment 490

strategy and the intensity of the regional circulation. For the record-length mean fields 491

shown below, binned values obtained in regions with less than 10 DOF are ignored. 492

493

The record-length mean velocity field on the 18.5˚ C isotherm is shown in Fig. 17. Note 494

that this field does not represent the circulation of EDW specifically (the abundance and 495

distribution of which will be described below) but rather identifies the broad patterns of 496

Page 23: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

22

circulation on an isothermal surface coincident with the expected vertical center of EDW. 497

For comparison, we also show in Fig. 17 the mean dynamic topography of the ocean 498

surface derived from satellite altimetry (Jayne, 2006). 499

500

At the northern extent of the domain the eastward flow of the Gulf Stream extension is 501

readily apparent with mean values exceeding 15 cm/s. Over the remainder of the 502

sampled domain the flow is generally southwest at speeds of 5–10 cm/s. Between 50˚–503

55˚ W the flow turns cyclonically from weak westward flow to strong (6–8 cm/s) 504

southward flow over a short distance. A comparison of this velocity field with the 505

climatological mean 18.5˚ C isotherm depth (Fig. 7; see also Fig. 20 a) indicates that this 506

direction change is generally coincident with the location of a strong gradient in isotherm 507

depth near 58˚ W. 508

509

The densely sampled region in the center of Fig. 17, east of Bermuda near 32˚ N, 60˚ W, 510

is interesting for its lack of significant flow. Though perhaps overemphasized in our 511

coarse 2˚x2˚ presentation, it appears that modest southwestward flow bifurcates into two 512

branches near 34˚ N, 57˚ W, leaving an apparent void over the Bermuda Rise east of 513

Bermuda. A closer examination of non-significant mean velocities in this area (i.e. those 514

grid cells with magnitude smaller than their standard deviation) indicates flow is 515

generally southwestward in this region, but at a slower mean rate and with enhanced 516

variability. The Bermuda Rise is well-known among paleoceanographers for its high 517

sedimentation rate, believed to be caused by recirculating bottom currents associated with 518

the westward Gulf Stream return flow (e.g. Laine and Hollister, 1981; Lund and Keigwin, 519

Page 24: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

23

1994). The bobber observations appear to support the notion of a regionally-distinct 520

circulation character over the Bermuda Rise, and are coincident with a relative “flat spot” 521

in the mean dynamic topography contours suggestive of reduced mean flow. 522

523

The EKE distribution associated with the 2˚x2˚ averaged velocities is shown in Fig. 18 524

and compared with previously reported observations using 15-m drogued surface drifter 525

data (Fratantoni, 2001). EKE values on the 18.5˚ C isotherm are, as expected, 526

substantially lower than surface values in the vicinity of the Gulf Stream. Both surface 527

and subsurface fields exhibit a strong meridional gradient trending towards comparable, 528

low EKE values in the gyre interior south of 33˚ N. Maximum EKE values observed at 529

the EDW core depth are near 1500 cm2/s2. Recent work by Park et al., (2012) examines 530

similarities between surface and subsurface eddy characteristics in this region. 531

532

Inhomogeneous deployment of drifting instruments in regions of strong eddy variability 533

can introduce an array bias (Davis, 1991) on an estimate of the mean velocity field. For 534

the substantially inhomogeneous bobber deployment strategy adopted during the 535

CLIMODE field campaign we estimated the array bias associated with the mean velocity 536

field shown in Fig. 17. Following Lavender et al., (2005), the array bias was computed 537

as the product of a horizontal (isothermal) diffusivity parameter and the horizontal 538

(isothermal) gradient of bobber concentration: uarray = -k grad(ln C). The spatially-539

varying isothermal diffusivity k was estimated as the product of <u’v’> (comparable to 540

the EKE field shown in Fig. 18) and the integral timescale derived above (3 days). 541

Bobber concentration C was approximated by the DOF field shown in Fig. 16b. The 542

Page 25: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

24

resulting vector velocity field describing the inherent array bias is shown in Fig. 19. Bias 543

velocity magnitudes exceed 5 cm/s in close proximity to the Gulf Stream (where velocity 544

variance and thus k is large) and along the edges of the sampled domain, especially in the 545

far southwest corner. Overall, however, the magnitude of the array bias over the vast 546

majority of the study region is small (less than 1 cm/s) and does not fundamentally 547

impact the interpretation of either Fig. 17 or those to follow. 548

549

4.2 Spatial Distribution of EDW 550

551

Temperature observations obtained by the CLIMODE bobbers were spatially and 552

temporally averaged in a manner similar to the velocity observations described above. A 553

2˚x2˚ grid was used to generate record-length mean depictions of the depth of the 18.5˚ C 554

isotherm, the vertical thickness of the 17˚–19˚ C layer, and the EDW column integral or 555

“thickness” as defined in Appendix A. These results are shown in Fig. 20.556

557

The mean topography of the 18.5˚ C isotherm obtained from the bobber measurements 558

(Fig. 20a) is quite similar to that derived from long-term climatology (Fig. 7) in both 559

structure and magnitude. In particular, the mean 18.5˚ C isotherm was found to be 560

shallowest (50–100 m) in the northeast of the domain near 35˚ N, 50˚ W, and deepest 561

(250–350 m) just south of the Gulf Stream and northeast of the Bahamas. These 2 562

relatively deep boundary regions are separated by a shallower zonal ridge extending 563

westward along 32˚ N in both the bobber measurements and the climatology. This zonal 564

ridge has similarities to the C-shaped surface dynamic height field shown by Reid (1978) 565

Page 26: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

25

inferred from hydrographic data, and to a weaker but similar feature in the satellite-566

derived absolute dynamic height field (Rio et al., 2005; Kwon and Riser, 2005). Note, 567

however, that the C-shape quickly disappears in the subsurface dynamic heights, 568

especially for the EDW layer (Kwon and Riser, 2004). The pronounced zonal gradient in 569

isotherm depth near 57˚ W is coincident with enhanced southward flow as shown in Fig. 570

17.571

572

Fig. 20b represents the record-length mean thickness of the vertical interval bounded by 573

the 17˚ C and 19˚ C isotherms. When compared to our best estimate of mean EDW 574

thickness (Fig. 20c), which incorporates the temperature gradient criterion of Kwon and 575

Riser (2004), several similarities and differences may be noted. First, the typical 576

magnitude of the layer thickness differs by approximately 25% – the gradient criterion is 577

substantially more restrictive. The general spatial patterns evident in both fields are 578

comparable, with maximal thickness in the northeast of the domain and a zonal ridge 579

along 35˚ N. Fig. 20c indicates that the region of thickest EDW is limited to a band 580

extending approximately 8–10˚ south of the Gulf Stream axis and eastward to the mid-581

Atlantic Ridge – outside of this band EDW thickness vanishes rapidly. In contrast, the 582

17˚–19˚ C layer thickness remains non-zero well to the south and east of this band and, 583

by construction, can never entirely vanish. Overall the thinning of EDW is noticeably 584

greater in the western reaches of this zonal band when compared to the 17˚–19˚ C field.585

We interpret this as illustrating the progressive destruction of EDW with distance from 586

the formation region. 587

588

Page 27: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

26

While the bobber sampling is neither uniform nor extensive, the majority of the observed 589

annual mean EDW volume is clearly isolated within a region bounded by the Gulf 590

Stream, the mid-Atlantic Ridge, and 25˚ N latitude. The narrow extension of the thick 591

EDW layer in the southwest corner of the domain is the result of a few bobbers looping 592

within coherent vortices in this area (see below, Section 4.3). A strong lateral gradient in 593

EDW thickness is evident extending southward from the Gulf Stream at 45˚ W to 30˚ N, 594

and then westward along 30˚ N. Bobber trajectories (Fig. 12) and velocities (Fig. 17) 595

indicate significant mean flow across this thickness gradient. 596

597

Away from the direct influence of the atmosphere the EDW thickness should decrease 598

with time (and distance from the formation site) as vertical and horizontal mixing 599

processes erode the layer’s homogeneity (i.e. cause restratification). If wintertime 600

atmospheric forcing along the southern edge of the Gulf Stream extension is the 601

proximate cause for locally thick layers of EDW (e.g. Marshall et al., 2009) then Fig. 20 602

suggests that 2 segments of the Gulf Stream, 65˚–75˚ W and 45˚–55˚ W, may be 603

important EDW source regions: Local EDW thickness is greatest adjacent to the Gulf 604

Stream in these areas. These 2 separate local maxima for the thickness of EDW (or the 605

winter outcropping regions) are consistent with the previous observational studies (Talley 606

and Raymer, 1982; Kwon and Riser, 2004; Joyce, 2012) as well as recent modeling 607

studies (Douglass et al., 2012; Maze et al., 2012). In particular Joyce (2012) articulated 608

that the western maximum is likely related to the non-advective EDW formation model 609

of Warren (1972) and the eastern one is more consistent with the combination of the 610

strong lateral advection and the air-sea cooling as described by Worthington (1959, 611

Page 28: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

27

1976). In addition, McGrath et al., (2010) reported that the 2 regions are separated by the 612

minimum cross-frontal flow from the north across the Gulf Stream based on the surface 613

drifter observations. 614

615

Seasonal subsets of the bobber temperature data were binned and mapped as���������Fig.616

21 provides additional evidence of relatively localized EDW thickness maxima which we 617

may interpret as either (a) regions of EDW formation, or (b) regions in which “old” EDW 618

transported within the Gulf Stream is re-injected into the gyre interior. Note particularly 619

the 2 segments of the wintertime Gulf Stream, identified in the paragraph above, adjacent 620

to which lie substantial local maxima in EDW thickness extending southward into the 621

gyre interior (Fig. 21e). Interestingly, Joyce et al., (2012, this issue) claim unusually 622

large EDW production in a briefly-sampled region between 67˚–52˚ W – a region notable 623

as a local minimum in EDW thickness in Fig. 21e. In contrast, the bobber-derived EDW 624

thickness fields agree well with the inferred spatial pattern of EDW formation rate 625

reported by Douglass et al., (2012, this issue) in a climatologically-forced numerical 626

model. In that simulation there are 2 notable, isolated regions of EDW formation: A 627

dominant region extending from 45˚–55˚ W, and a less vigorous region near 70˚ W. 628

629

Finally, it is interesting to note that the 17˚–19˚ C and EDW layers exhibit seasonal layer 630

thickness changes in the opposite sense: EDW thickness is generally greatest in winter, 631

while the 17˚–19˚ C layer is thickest in summer (see Fig. 21c–f). This is due to two 632

fundamentally different forcing mechanisms: EDW layer thickness is dominated by 633

buoyancy forcing during the winter formation period and gradually decreases during the 634

Page 29: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

28

rest of the year as EDW disperses and decays (Kwon and Riser, 2004; Forget et al., 635

2011). In contrast, the 17˚–19˚ C layer thickness is maximal in summer due to seasonal 636

wind forcing and Ekman pumping (see, e.g., Trenberth et al., 1990).637

638

4.3 EDW Circulation 639

640

The two-dimensional fields presented above separately illustrate the mean circulation at 641

the central depth of the EDW and the mean thickness of the EDW layer. In contrast, Fig. 642

22 was constructed by bin-averaging (as above) daily values of the product of EDW 643

thickness (h) and RAFOS-derived velocity (u) on the 18.5˚ C isotherm. The resulting 644

mean EDW thickness flux, <uh>, field approximates the mean horizontal (isothermal) 645

transport of EDW. The general pattern of EDW transport is similar to the velocity field 646

shown in Fig. 17, with mean flow from the northeast to the southwest over much of the 647

sampled region. This broad pattern is generally consistent with transport streamfunctions 648

constructed by Kwon and Riser (2005) for the upper 900 m of the water column using 649

Argo float hydrographic data.650

651

Fig. 23 provides an integrated view of the record-length mean EDW thickness flux as a 652

function of latitude and longitude. As can be inferred from the vector field in Fig. 22, 653

southward transport of EDW is dominant at all latitudes within the sampled domain. In 654

this annual representation, we observe nearly 20 Sv of EDW flowing southward across 655

32–34˚ N with substantial meridional divergence between 34–38˚ N. This region of 656

divergence is consistent with the relatively thick zonal band of EDW along 35˚ N noted 657

Page 30: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

29

above. Westward transport increases from east to west reaching a maximum near 10 Sv at 658

70˚ W. 659

660

The depiction of a relatively narrow southward extension of the thick EDW layer in the 661

southwest corner of the domain results almost entirely from a few bobbers (most notably 662

474; see Fig. 13b) which moved slowly through this area while looping within coherent663

vortices, and eventually experienced abrupt thinning of the EDW as the floats entered the 664

strongly sheared western boundary current. The ability of submesoscale coherent 665

vortices (e.g. McWilliams, 1985) to transport relatively undiluted ocean properties over 666

substantial distances has been noted in several instances, most notably in the case of 667

Mediterranean water eddies (Meddies; see Richardson et al., 2000, for a review).668

Meddies contribute substantially to the southwestward dispersal of saline water at mid-669

depth in the eastern Atlantic and have been suggested as comparable in importance to 670

downgradient diffusion in maintaining the overall shape and intensity of the 671

Mediterranean salt tongue (Wang and Dewar, 2003). While the relative importance of 672

submesoscale coherent vortices in the dispersal of subducted EDW remains to be 673

determined, it is clear from our limited observations that such features can have a 674

profound effect on EDW distribution. 675

676

677

678

679

Page 31: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

30

680

4.4 EDW Reemergence and Destruction 681

682

The CLIMODE bobbers spent all but approximately 12 hours of each 30-day period on 683

their target isothermal surface. As such, they reasonably approximate the trajectory of an 684

18.5˚ C parcel of water. As shown graphically in the case studies of Fig. 13a and Fig. 685

13b, the three-dimensional bobber trajectories collectively illustrate the subduction, 686

translation, and eventual reemergence of water parcels. We identified a total of 55 687

examples in which a bobber, initially in the mixed layer, was at some time later located 688

beneath the stratified upper ocean, i.e. “subducted.” In this analysis no distinction was 689

made regarding the horizontal motion of the bobber: The assumed “subduction” could 690

therefore occur via motion along a sloping isothermal surface into an area of persistent 691

stratification, -or- via local restratification above a stationary bobber.692

693

Fig. 24 presents a histogram summarizing these results. Shown is the interval between a 694

bobber leaving and rejoining the surface mixed layer, with the requirement that these 695

times are separated by a period of at least 6 months during which the bobber reported a 696

stratified environment with surface temperature greater than 20˚ C. Note that the number 697

of examples exceeds the total number of bobbers: Those bobbers that recycled within the 698

subtropical gyre (i.e. experienced more than one cycle of subduction and reemergence) 699

are included more than once in this tally. 700

701

Page 32: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

31

There exists an inherent bias in this analysis due to the finite number of bobbers 702

available, our targeted deployment strategy in the EDW formation region, and the limited 703

time span of this experiment. However, our limited observations do clearly suggest that, 704

contrary to previous hypotheses, subducted EDW is isolated from the atmosphere for 705

only a short period of time. This is an unexpected result as previous data analysis efforts 706

and modeling studies (e.g. Kwon and Riser, 2004; Forget et al., 2011; Douglass et al., 707

2012, this issue) have suggested EDW residence times of several years. One possible 708

explanation for this discrepancy may be that mode water formed in different regions 709

and/or different time periods is sequestered for varying durations. Such variance in the 710

reemergence timescale is supported by, for example, the modeling study of Douglass et 711

al., 2012, this issue, (see their Fig. 11) which reveals a highly skewed distribution of 712

mode water ages with a peak near 2 years but an extended tail extending beyond 10 713

years. Given the temporally-limited span of our in-situ observations (4 years) and their 714

geographically non-uniform deployment, it is not surprising that we are unable to resolve 715

such a distribution in full. However, the bobber observations clearly demonstrate that in 716

the region of greatest EDW volume (i.e. Fig. 20c) in the northwestern subtropical gyre, 717

short-term EDW sequestration (i.e. less than 24 months) is the more common scenario, 718

and thus is the most relevant when considering climatic and biogeochemical 719

consequences.720

721

A careful analysis of the rates of EDW formation and destruction requires a well-resolved 722

time series of basin-spanning EDW volume. Unfortunately, while spatial distribution of 723

bobber-derived EDW thickness measurements is unusually dense for a short period of 724

Page 33: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

32

time, the aggregate observational density in both space and time is inadequate for a 725

careful volumetric approach using the bobbers alone. Rather than applying extensive 726

temporal and spatial interpolation to a sparse data set, we qualitatively examined all 727

bobber-derived EDW thickness measurements to approximate the seasonality and rate of 728

EDW thickness creation and destruction. 729

730

Fig. 25 illustrates a composite annual cycle of EDW thickness computed from all bobber 731

temperature measurements during the CLIMODE field campaign. EDW thickness 732

observations were binned using a 20 m vertical interval and a 10-day time period. The 733

resulting 4-year average provides insight into the temporal variation of EDW thickness 734

across the western subtropical Atlantic. In particular, we note that prior to late February 735

there are extremely few observations of EDW layers with thickness greater than 300 m. 736

In approximately the first week of March the number of thick EDW layers measurements 737

increases dramatically. This general pattern is consistent with the volumetric analysis of 738

Forget et al., (2011) derived from assimilating numerical model results, and with 739

CLIMODE moored observations (Davis et al., 2012, this issue).740

741

Also interesting is the bimodal distribution of EDW thickness between March and May.742

During this 3-month period there appears to be a distinction between “new” EDW 743

(formed since late February) and “old” EDW. Approximately 6–9 months are required 744

for the extremely thick “new” EDW layers to decay to background thickness levels. 745

Based on the bobber-derived circulation and thickness maps we infer that much of this 746

decay (i.e. restratification) occurs as mode water is transported southwestward in the 747

Page 34: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

33

subtropical gyre. It is apparent from the bobber trajectories that the western boundary 748

(and the Gulf Stream) set an upper limit on the lifetime of a subducted EDW parcel. 749

750

5. Summary and Conclusions 751

752

In this contribution we have presented initial results from a novel Lagrangian study of 753

subtropical mode water stratification and circulation in the western North Atlantic. We 754

illustrated the development and application of an array of acoustically-tracked, isotherm-755

following, bobbing profiling floats and described their application to a 4-year study of 756

EDW dispersal and evolution. 757

758

The key results of this study may be summarized as follows: 759

760

1. The CLIMODE bobber program was an overwhelming technical success: 38 of 761

40 instruments achieved a success rate greater than 90%, and 37 of 40 were 762

acoustically tracked for at least a portion of their lifetime (10–26 months). 763

Operational endurance of the bobber array was substantially greater than 764

anticipated with a half-life exceeding 4 years. 765

766

2. The bobber-derived Lagrangian integral time and length scales (3 days, 68 km) 767

associated with motion on the 18.5˚ C isothermal surface are consistent with 768

previous measurements in the Gulf Stream extension region and fall between 769

previous Lagrangian estimates at the ocean surface and thermocline depth. The 770

Page 35: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

34

first direct observations of circulation within the EDW layer are indicative of 771

substantial eddy activity both near the Gulf Stream and throughout the subtropical 772

gyre. Evidence of long-lived submesoscale features associated with substantial 773

EDW thickness suggests that coherent vortices may play an important role in the 774

dispersal of mode water within the subtropical gyre. 775

776

3. We find that EDW (defined as a minimally-stratified layer) exhibits opposing 777

seasonal thickness changes to a layer bounded simply by the 17˚–19˚ C isotherms. 778

EDW thickness is generally greatest in winter (as a result of buoyancy-forced 779

convection), while the 17˚–19˚ C layer is thickest in summer (resulting from 780

Ekman pumping – vertical stretching – associated with basin-scale seasonal wind 781

forcing.)782

783

4. Finally, our limited observations suggest that, contrary to previous hypotheses, 784

the majority of subducted EDW is isolated from the atmosphere for periods less 785

than 24 months. Although our targeted deployment strategy imposed a spatial 786

bias, and we were unable to directly observe longer-term EDW sequestration, 787

seasonal-to-biennial reemergence appears to be a common scenario which should 788

be explored further and considered when assessing the climatic and 789

biogeochemical consequences of EDW. 790

791

We anticipate further analyses of these data in the context of longer-term Argo float 792

measurements, moored observations, and numerical modeling exercises. In particular, 793

Page 36: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

35

with our new understanding of the scope of the EDW pool in the subtropical western 794

Atlantic it may be advantageous to explore numerical simulations with and without 795

explicitly resolved submesoscale eddies. In addition to facilitating the widespread 796

dispersal of EDW, such translating eddies may also contribute to the large variance in 797

EDW age / reemergence time by transporting and isolating mode water away from the 798

dissipative western boundary. 799

Appendix A: EDW Thickness Estimation 800

801

Temporal and spatial variations over the expanse of the western Atlantic make it 802

challenging to uniquely define EDW based on measured water properties (e.g. Joyce, 803

2012). Worthington (1959) defined EDW in terms of a specific temperature and salinity 804

range (17.9˚ C +/- 0.3˚ C; 36.5 +/- 0.1). More recent work by Klein and Hogg (1996), 805

Alfultis and Cornillon (2001), and Kwon and Riser (2004) have explicitly incorporated 806

the vertically-homogenous nature of the EDW layer by including in their analyses 807

potential vorticity and/or vertical temperature gradient criteria. These latter definitions 808

are somewhat more flexible because, as noted by Talley and Raymer (1982) and Klein 809

and Hogg (1996), the specific temperature-salinity (T-S) characteristics of EDW may 810

vary interannually. 811

812

In this contribution we adopt the Kwon and Riser (2004) EDW definition based on 813

temperature and temperature gradient. Peng et al., (2006) used Argo float profiles to 814

explore the sensitivity of EDW volume estimates to variations in the choice of 815

temperature gradient threshold. They found that a slightly relaxed vertical temperature 816

Page 37: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

36

gradient criterion vs. that used by Kwon and Riser (2004) (0.01˚ C/m vs. 0.006˚ C/m) 817

resulted in a factor-of-two increase in inferred EDW volume. For ease in comparison 818

with previous observational results we apply here the original Kwon and Riser (2004) 819

criteria. Specifically, we label a particular vertical segment of a temperature profile as 820

EDW if (a) the temperature falls between 17˚–19˚ C and (b) the local vertical temperature 821

gradient is less than 0.006˚ C/m. 822

823

A complication that arises from the use of fixed criteria as above is the possibility of 824

identifying multiple, inconsequentially-different, layers separated by regions of slightly 825

different temperature or temperature gradient (see e.g. Fig. 1). While such multi-layered 826

structures may be both real and relevant, for this basin-scale overview we integrate the 827

total EDW contribution of such layers using the following procedure: Each temperature 828

profile (with nominal 5–20 m resolution) is vertically interpolated using a shape-829

preserving cubic spline to a uniform 1 m interval. The vertical temperature gradient is 830

computed from the interpolated profile. We define the local EDW thickness (or more 831

precisely, the column integral of water with EDW characteristics) as the sum of all 1 m 832

bins in the profile meeting both the temperature range and temperature gradient criteria 833

described above. This method is simple, objective, provides thickness estimates that are 834

robust to small variations in both temperature and temperature gradient about the 835

threshold criteria, and yields a conservative lower bound on the actual EDW thickness 836

and volume using the Kwon and Riser (2004) definition. 837

838

Page 38: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

37

Acknowledgements 839

840

The authors acknowledge the logistical and scientific cooperation of our CLIMODE 841

collaborators, the technical support of Mr. John Lund, and the able assistance of the 842

captains and crew of the WHOI research vessels Oceanus, Knorr, and Atlantis. Ellyn 843

Montgomery, Terry McKee, and Heather Furey contributed significantly to the 844

development of software and methods for bobber tracking and data processing. We thank 845

Fiamma Straneo (WHOI) and Bernadette Sloyan (CSIRO) for substantial contributions to 846

the planning of this field experiment. We acknowledge the careful work of three 847

anonymous reviewers whose thoughtful comments and suggestions greatly improved this 848

manuscript. This work was supported by the U.S. National Science Foundation under 849

grant nos. OCE-0424492 and OCE-0961090. 850

851

Page 39: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

38

References852

853

Alfultis, M.A., P. Cornillon, 2001. Annual and Interannual Changes in the North Atlantic 854

STMW Layer Properties. J. Phys. Oceanogr. 31, 2066–2086. 855

856

Bates, N.R., A.C. Pequignet, R.J. Johnson, N. Gruber, 2002. A short-term sink for 857

atmospheric CO2 in subtropical mode water of the North Atlantic Ocean. Nature 420, 858

489–493.859

860

Brambilla, E., L.D. Talley, 2006. Surface drifter exchange between the North Atlantic 861

subtropical and subpolar gyres. J. Geophys. Res. 111(C07026), 16 pp., 862

doi:10.1029/2005JC003146.863

864

Chelton, D.B., R.A. de Szoeke, M.G., Schlax, K. El Naggar, N. Siwertz, 1998.865

Geographical variability of the first baroclinic Rossby radius of deformation. J. Phys. 866

Oceanogr. 28, 433–460.867

868

Curry, R.G., 1996. Hydrobase–A database of hydrographic stations and tools for 869

climatological analysis. Woods Hole Oceanographic Institution Technical Report 870

WHOI-96-01, 44 pp. 871

Page 40: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

39

872

Davis, R.E., 1991. Observing the general circulation with floats. Deep-Sea Res. 38, 873

Suppl. 1, S531–S571.874

875

Davis, R.E., J.T. Sherman, J. Dufour, 2001. Profiling ALACEs and other advances in 876

autonomous subsurface floats. J. Atmos. Ocean. Tech. 18, 982–993. 877

878

Davis, R.E., D.C. Webb, L.A. Regier, J. Dufour, 1992. The Autonomous Lagrangian 879

Circulation Explorer (ALACE). J. Atm. Oceanic Tech. 9, 264–285. 880

881

Davis, X.J., F. Straneo, Y.-O. Kwon, K.A. Kelly, J.M. Toole, 2012. Evolution and 882

Formation of North Atlantic Eighteen Degree Water in the Sargasso Sea from Moored 883

Data. Deep-Sea Res. II, this issue. 884

885

Dong, S., S.T. Gille, J. Sprintall, 2007. An Assessment of the Southern Ocean Mixed 886

Layer Heat Budget. J. Climate 20, 4425–4442. 887

888

Dong, S., K.A. Kelly, 2004. Heat budget in the Gulf Stream region: the importance of 889

heat storage and advection. J. Phys. Oceanog. 34, 1214–1231. 890

Page 41: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

40

891

Douglass, E.M., Y.-O. Kwon, S.R. Jayne, 2012. A comparison of North Pacific and 892

North Atlantic Subtropical Mode Waters in a Climatologically-Forced Model. Deep-Sea 893

Res. II, this issue. 894

895

Forget, G., G. Maze, M. Buckley, J. Marshall, 2011. Estimated Seasonal Cycle of North 896

Atlantic Eighteen Degree Water Volume. J. Phys. Oceanog. 41(2), 269–286. 897

898

Fratantoni, D.M., 2001. North Atlantic surface circulation during the 1990’s observed 899

with satellite-tracked drifters. J. Geophys. Res. 106(C10), 22,067�22,093.900

901

Fratantoni, D.M., T.K. McKee, B.A. Hodges, H.H. Furey, J.M. Lund, 2010. CLIMODE 902

Bobber Data Report: July 2005–May 2009. Woods Hole Oceanographic Institution 903

Technical Report WHOI-2010-03, 154 pp. 904

905

Hanawa, K., L.D. Talley, 2001. Mode Waters. Ocean Circulation and Climate, in: 906

Siedler, G., Church, J., (Eds.), International Geophysics Series, Academic Press, 373–907

386.908

909

Page 42: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

41

Jayne, S.R., 2006. Circulation of the North Atlantic Ocean from altimetry and the 910

Gravity Recovery and Climate Experiment geoid. J. Geophys. Res. 111, doi: 911

10.1029/2005JC003128.912

913

Joyce, T.M., 2012. New perspectives on eighteen-degree water formation in the North 914

Atlantic. J. Oceanography 68(1), 45�52, doi:10.1007/s10872-011-0029-0. 915

916

Joyce, T.M., Y.-O. Kwon, L. Yu., 2009. On the Relationship between Synoptic 917

Wintertime Atmospheric Variability and Path Shifts in the Gulf Stream and the Kuroshio 918

Extension. J. Climate 22(12), 3177–3192. 919

920

Joyce, T.M., J.R. Luyten, A. Kubryakov, F.B. Bahr, J.S. Pallant, 1998. Meso- to large-921

scale structure of subducting water in the subtropical gyre of the eastern North Atlantic 922

Ocean. J. Phys. Oceanog. 28(1), 4061. 923

924

Joyce, T.M., L.N. Thomas, W.K. Dewar, J.B. Girton, 2012. Eighteen degree water 925

formation within the Gulf Stream: a new paradigm arising from CLIMODE. Deep-Sea 926

Res. II, this issue. 927

928

Page 43: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

42

Klein, B., N. Hogg, 1996. On the variability of 18 Degree Water formation as observed 929

from moored instruments at 55˚ W. Deep-Sea Res. 43(1112), 1777–1806. 930

931

Kwon, Y.-O., 2003. Observation of general circulation and water mass variability in the 932

North Atlantic subtropical mode water region. Ph.D. thesis, University of Washington, 933

161 pp.934

935

Kwon, Y.-O., S.C. Riser, 2004. North Atlantic Subtropical Mode Water: A history of 936

ocean-atmosphere interaction 1961–2000. Geophys. Res. Lett. 31, L19307, 937

doi:10.1029/2004GL021116.938

939

Kwon, Y.-O., S.C. Riser, 2005. General circulation of the western subtropical North 940

Atlantic observed using profiling floats. J. Geophys. Res. 110, C10012, 941

doi:10.1029/2005JC002909.942

943

Laine, E.P., C.D. Hollister, 1981. Geological effects of the Gulf Stream system on the 944

northern Bermuda Rise. Marine Geology 39, 277–310. 945

946

Lavender, K.L., W.B. Owens, R.E. Davis, 2005. The mid-depth circulation of the 947

Page 44: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

43

subpolar North Atlantic Ocean as measured by subsurface floats. Deep-Sea Res. 52, 948

767–785.949

950

Lumpkin, R., A.-M. Treguier, K. Speer, 2002. Lagrangian Eddy Scales in the North 951

Atlantic Ocean. J. Phys. Oceanog. 32, 2425–2440. 952

953

Lund, S.P., L. Keigwin, 1994. Measurement of the degree of smoothing in sediment 954

paleomagnetic secular variation records: An example from late Quaternary deep-sea 955

sediments of the Bermuda Rise, western North Atlantic Ocean. Earth and Planetary 956

Science Letters 122, 317–330. 957

958

Marshall, J., A. Andersson, N. Bates, W. Dewar, S. Doney, J. Edson, R. Ferrari, G. 959

Forget, D. Fratantoni, M. Gregg, T. Joyce, K. Kelly, S. Lozier, R. Lumpkin, G. Maze, J. 960

Palter, R. Samelson, K. Silverthorne, E. Skyllingstad, F. Straneo, L. Talley, L. Thomas, J. 961

Toole, R. Weller, 2009. The Climode Field Campaign: Observing the Cycle of 962

Convection and Restratification over the Gulf Stream. Bull. Amer. Meteor. Soc. 90, 963

1337–1350.964

965

Maze, G., J. Deshayes, J. Marshall, A.-M. Tréguier, A. Chronis, L. Vollmer, 2012. 966

Surface vertical PV fluxes and subtropical mode water formation in an eddy-resolving 967

Page 45: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

44

numerical simulation. Deep-Sea Res. II, this issue. 968

969

McGrath, G.G., T. Rossby, J.T. Merrill, 2010. Drifters in the Gulf Stream. J. Mar. Res. 970

68, 699–721. 971

972

McWilliams, J.C., 1985. Submesoscale, coherent vortices in the ocean. Rev. Geophys. 973

23, 165–182. 974

975

Old, E., K. Haines, 2006. North Atlantic Subtropical Mode Waters and Ocean Memory 976

in HadCM3. J. Climate 19, 1126–1148. 977

978

Palter, J.B., M.S. Lozier, R.T. Barber, 2005. The effect of advection on the nutrient 979

reservoir in the North Atlantic subtropical gyre. Nature 437, 687–692.980

981

Park, J.J., D.M. Fratantoni, Y.-O. Kwon, M.S. Lozier, 2012. Patchiness Length Scale of 982

Eighteen Degree Water. Poster presentation at AGU Ocean Sciences Meeting, Salt Lake 983

City, UT. 984

985

Peng, G., E.P. Chassignet, Y.-O. Kwon, S.C. Riser, 2006. Investigation of the variability 986

Page 46: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

45

of the North Atlantic Subtropical Mode Water using profile float data and numerical 987

model output. Ocean Modeling 13, 65–85. 988

989

Reid, J., 1978. On the middepth circulation and salinity field in the North Atlantic 990

Ocean. J. Geophys. Res. 83(10), 5063, doi:10.1029/JC083iC10p05063. 991

992

Richardson, P.L., A. Bower, W. Zenk, 2000. A census of meddies tracked by floats. 993

Progress in Oceanography 45, Pergamon, 209–250. 994

995

Rio, M.-H., P. Schaeffer, J.-M. Lemoine, F. Hernandez, 2005. Estimation of the ocean 996

Mean Dynamic Topography through the combination of altimetric data, in-situ 997

measurements and GRACE geoid: From global to regional studies. Proceedings of the 998

GOCINA international workshop, Luxembourg. 999

1000

Roemmich, D., M. Belbeoch, H. Freeland, S.L. Garzoli, W.J. Gould, F. Grant, M. 1001

Ignaszewski, B. King, B. Klein, P.-Y. Le Traon, K.A. Mork, W.B. Owens, S. Pouliquen, 1002

M. Ravichandran, S. Riser, A. Sterl, T. Suga, M.-S. Suk, P. Sutton, V. Thierry, P.J. 1003

Velez-Belchi, S. Wijffels, J. Xu, 2009. Argo: The challenge of continuing 10 years of 1004

progress. Oceanography 22(3), 46–55. 1005

1006

Rossby, T., D. Dorson, J. Fontaine, 1986. The RAFOS System. J. Atmos. Oceanic 1007

Page 47: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

46

Technol 3, 672–679, http://dx.doi.org/10.1175/1520-1008

0426(1986)003<0672:TRS>2.0.CO;2. 1009

1010

Rossby, T., M.D. Prater, 2005. On variations in static stability along Lagrangian 1011

trajectories. Deep-Sea Res. II 52(3–4), 465–479, doi:10.1016/j.dsr2.2004.12.003. 1012

1013

Rossby, H.T., S.C. Riser, A.J. Mariano, 1983. The western North Atlantic – a 1014

Lagrangian view point, in: Eddies in marine science, A.R. Robinson, (Ed.), Springer-1015

Verlag, Berlin, pp. 66–91.1016

1017

Rossby, H.T., A.D. Voorhis, D. Webb, 1975. A Quasi-Lagrangian study of mid-ocean 1018

variability using long range SOFAR floats. J. Mar. Res. 33, 355–382.1019

1020

Rypina, I.I., L.J. Pratt, M.S. Lozier, 2011. Near-surface transport pathways in the North 1021

Atlantic Ocean: Looking for throughput from the Subtropical to the Subpolar Gyre. J. 1022

Phys. Oceanog. 41, 911–925.1023

1024

Siegel, D.A., S.C. Doney, J.A. Yoder, 2002. The North Atlantic Spring Phytoplankton 1025

Bloom and Sverdrup’s Critical Depth Hypothesis. Science 296(5568), 730–733, 1026

Page 48: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

47

doi:10.1126/science.1069174.1027

1028

Speer, K., E. Tziperman, 1992. Rates of Water Mass Formation in the North Atlantic 1029

Ocean. J. Phys. Oceanog. 22, 93–104. 1030

1031

Talley, L.D., M.E. Raymer, 1982. Eighteen degree water variability. J. Mar. Res. 40 1032

(Suppl.), 757–777. 1033

1034

Thomas, L.N., 2005. Destruction of potential vorticity by winds. J. Phys. Oceanog. 35, 1035

2457–2466.1036

1037

Thomas, L.N., T.M. Joyce, 2010. Subduction on the northern and southern flanks of the 1038

Gulf Stream. J. Phys. Oceanog. 40, 429–438. 1039

1040

Trenberth, K.E., W.G. Large, J.G. Olson, 1990. The Mean Annual Cycle in Global 1041

Ocean Wind Stress. J. Phys. Oceanog. 20, 1742–1760. 1042

1043

Wang, G., W.K. Dewar, 2003. Meddy–Seamount Interactions: Implications for the 1044

Page 49: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

48

Mediterranean Salt Tongue. J. Phys. Oceanog. 33, 2446–2461. 1045

1046

Warren, B.A., 1972. In sensitivity of subtropical mode water characteristics to 1047

meteorological fluctuations. Deep-Sea Res. 19, 1–19. 1048

1049

Wooding, C.M., H.H. Furey, M.A. Pacheco, 2005. RAFOS float processing at the 1050

Woods Hole Oceanographic Institution. Woods Hole Oceanographic Institution 1051

Technical Report WHOI-2005-02, 37 pp. 1052

1053

Worthington, L.V., 1959. The 18˚ water in the Sargasso Sea. Deep-Sea Res. 5, 297–305.1054

1055

Worthington, L.V., 1972. Negative Oceanic Heat Flux as a Cause of Water-Mass 1056

Formation. J. Phys. Oceanog. 2(3), 205–211. 1057

1058

Worthington, L.V., 1976. On the North Atlantic Circulation. The Johns Hopkins 1059

Oceanogr. Stud. 6, The John Hopkins Press, 110 pp. 1060

1061

1062

1063

Page 50: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

49

1064

Fig. 1. Sequence of monthly deep (0–1000 m) temperature profiles collected by CLIVAR Mode 1065Water Dynamics Experiment (CLIMODE) Bobber 474 between February and November 2007. 1066For clarity, all profiles are truncated at 800 m depth and profiles subsequent to the first are shifted 1067by +1˚ C relative to the preceding profile. The highlighted segments of each profile meet the 1068temperature and temperature gradient criteria for Eighteen Degree Water (EDW) as proposed by 1069Kwon and Riser (2004) and discussed in Appendix A. During this 10-month period the bobber 1070drifted southward from approximately 38˚ N to 28˚ N. Profile positions are labeled by month 1071number in the inset. 1072

1073

Page 51: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

1

111111

075

Fig. 2081tempe082Dynam083The A084period085isothe086

2. Visual comerature-depth mics Experim

Argo float is ld. The CLIMOermal drift an

mparison of (l(CTD) and d

ment (CLIMOarger, heavierODE bobber d rapid profil

eft) a convendissolved oxygODE) bobber wr, and designehardware and

ling at relative

50

ntional Argo fgen sensors, awith temperated for deep isd software is oely shallow d

float equippedand (right) a Cture sensor an

sobaric operatoptimized for

depths.

d with conducCLIVAR Mond a RAFOS tion over a lonr actively-ball

ctivity-de Water hydrophone. ng time lasted

Page 52: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

1111111111111

082083084

Fig. 3093salinit094(dotte095subtro096tempe097minim098deviat099EDW100(CTD101

094

3. (a) Temperty (T-S) relat

ed line). Dataopical gyre, sperature gradiemum near 18.tion from the

W stratificationD) temperature

rature-Salinityionship near t

a source: Statpanning roug

ent (˚ C/m) fro5˚ C associatemean at each

n minimum. De profile data

y diagram illuthe 18.5˚ C cetions 37–69 o

ghly 20–40˚ Nom the westered with EDWh depth illustrData source: since 1980 w

51

ustrating the aenter of the E

of 1997 WOCN. (b) Represern subtropical

W. The shadedrating the relaAll available

within the dom

approximatelyEighteen DegrCE A22 sectio

entation of thl gyre illustrad region encoative spatial/tee NODC condmain 30–38˚ N

y linear tempree Water (EDn in the weste

he mean verticating the localompasses one emporal stabiductivity-tempN, 50–65˚ W.

erature-DW) layer erncall stratificationstandard

ility of the perature-dept

n

th

Page 53: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

1111111111111

095096

Fig. 4105Dynam106CLIM107shown108invest109source110when111wall (112observ113

106107

4. Location omics Experim

MODE field can are additiontigators from es were occasappropriate.

(and a one-stavations since

f RAFOS soument (CLIMOampaign and nal sound southe U.S. (EPB

sionally heard In this and su

andard-deviat1966 (courte

und source moODE) bobbers

were active brces, active dB, EPC3), Frd by the bobbubsequent figtion uncertainesy Fisheries a

52

oorings used s. Source moobetween Noveduring the CLIance (AP3, A

bers and were gures the nomnty band) is shand Oceans C

for tracking torings A–D wember 2005–NIMODE expe

AP5), and Gerused during p

minal location hown based oCanada / MED

the CLIVAR were deployedNovember 20eriment, instarmany (IM8). post-experimof the Gulf Sn monthly-m

DS).

Mode Water d as part of th007. Also lled by These ent tracking

Stream north mean

he

Page 54: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

53

110711081109

Fig. 5. Timeline illustrating the lifetime of each CLIVAR Mode Water Dynamics Experiment 1110(CLIMODE) bobber. Open circles indicate receipt of an Argos-derived position report from each 1111bobber. The shaded portions of each bobber’s timeline indicates the period over which acoustic 1112tracking was successful. For testing purposes, two bobbers (518 and 476) were instructed to 1113surface every 12 days rather than the standard 30 days. 1114

111511161117

Page 55: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

1111111111

119120121

Fig. 6126two m127positi128valid129achiev130

127128

6. A valid RAmoored soundon to the 3 clposition/veloved within 50

AFOS positiond sources. Sholosest sourcesocity measurem00 km of the c

n measuremeown here is ths (closest sourment. A majoclosest sound

54

ent requires dehe distributionrce range indiority of the vasource.

etection of tran of range froicated by darkalid RAFOS p

ansmissions from each validkest bar) at thposition solut

from at least d bobber he time of eactions were

h

Page 56: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

55

112811291130

Fig. 7. Position of CLIVAR Mode Water Dynamics Experiment (CLIMODE) bobber 1131deployments and the moored sound source array. Also shown is the annual mean depth of the 113218.5˚ C isotherm as derived from historical hydrographic data (Curry, 1996). 1133

113411351136

Page 57: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

56

1137113811391140

Fig. 8. Histograms of (a) the number of RAFOS and profile measurements obtained by the 1141bobber array during each month of the CLIVAR Mode Water Dynamics Experiment 1142(CLIMODE) field campaign, and (b, c) the average monthly distribution of these data. 1143

11441145

Page 58: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

57

1146114711481149115011511152115311541155115611571158115911601161116211631164116511661167116811691170117111721173117411751176

Fig. 9. Histograms of the depth of the top (shallow limit), bottom (deep limit), and total vertical 1177extent of each 17˚–20˚ C bob performed by the bobber array. Where the 20˚ C isotherm did not 1178exist the bobber was allowed to profile to the surface – values corresponding to a depth of zero 1179have been suppressed in this figure. 1180

11811182

Page 59: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

58

11831184

Fig. 10. (a) Total lifetime displacement of each CLIVAR Mode Water Dynamics Experiment 1185(CLIMODE) bobber from its launch location. Final position is indicated with a filled circle. The 1186sound source array is indicated by the yellow circles. (b) Argos-derived trajectories showing the 1187drift of each bobber at approximately 30-day temporal resolution. In both panels the colors 1188assigned to trajectories are random and arbitrary. 1189

1190

Page 60: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

59

11911192

Fig. 11. (a) Position of each valid bob profile (blue dot) and deep (1000 m) profile (red dot) 1193reported by the bobber array. (b) Final record-length RAFOS-derived isothermal (18.5˚ C) 1194trajectories and fragments for all trackable bobbers. Discontinuities in several trajectories are due 1195to intermittent periods in which range, topography, or stratification issues resulted in inability to 1196compute accurate position estimates, or in which the bobber deviated from the target 18.5˚ C 1197parking isotherm. The colors assigned to trajectories are random and arbitrary. 1198

1199

Page 61: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

60

120012011202

Fig. 12. Composite (RAFOS + Argos) trajectories of 16 CLIVAR Mode Water Dynamics 1203Experiment (CLIMODE) bobbers on the 18.5˚ C temperature surface in the western North 1204Atlantic subtropical gyre. Bobber deployment location is indicated by a green triangle in each 1205panel. Small black dots are shown at 30-day intervals along each trajectory. 1206

1207

Page 62: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

61

120812091210

Fig. 13a. Composite trajectory, temperature stratification, drift temperature and pressure, 1211Eighteen Degree Water (EDW) and 17˚–19˚ C thickness, and velocity as measured by CLIVAR 1212Mode Water Dynamics Experiment (CLIMODE) bobbers 448 (left; a–e) and 721 (right; f–j). (a) 1213Composite (RAFOS + Argos) trajectory with monthly tickmarks and date (YYMM) labels.1214Initial/final bobber position indicated by the green/red triangle. (b) Temperature as a function of 1215time and depth. Contour interval is 1˚ C. The dark solid line indicates the actual drift depth of the 1216bobber. Temporal resolution at depths between 17˚–19˚ C is 3 days; below that it is 30 days. (c) 1217Daily measured bobber temperature (light line) and pressure (dark line). The target drift 1218temperature was 18.5˚ C +/- 0.1˚ C. (d) Thickness (see Appendix A) of water identified as EDW 1219using the Kwon and Riser (2004) temperature / temperature gradient criteria. Black line shown is 1220a 30-day smoothed representation of the individual 3-day estimates (dots). Gray line and dots are 1221corresponding estimates of 17˚–19˚ C layer thickness. (e) Meridional (red) and zonal (blue) 1222velocity (cm/s) computed from the RAFOS-tracked segments of the bobber trajectory. Stars 1223highlight instrument positions on 1 January of each year. 1224

12251226

Page 63: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

62

122712281229

Fig. 13b. As in Fig. 13a, but for bobbers 598 (left) and 474 (right). 12301231

Page 64: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

63

123212331234

Fig. 14. Over 4 months Bobber 474 was observed to complete at least 10 anticyclonic loops 1235while translating southwestward towards the Bahamas. Recovery of the moored sound source 1236array in November 2007 precluded direct observation of additional looping and the interaction 1237between this eddy and the islands of the Bahamas. The observed loops were 40–50 km in radius 1238with a 9 to 11-day rotational period. Maximum measured azimuthal velocity was approximately 123955 cm/s. During this period the depth of the bobber increased slightly from 300–330 m while the 1240local Eighteen Degree Water (EDW) thickness remained relatively constant near 300 m. 1241

Page 65: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

64

1242

1243Fig. 15. Estimated Lagrangian velocity, time, and space scales derived from historical float and 1244drifter measurements as originally reported by Lumpkin et al., (2002) compared with similar 1245estimates from the CLIVAR Mode Water Dynamics Experiment (CLIMODE) bobbers. Values 1246for each parameter are shown at several depths in the vicinity of the Gulf Stream extension 1247(circles) and the southwestern subtropical gyre (squares). Parameters derived from a subset of the 1248CLIMODE bobbers (squares; see text for details) are shown at the mean depth of valid RAFOS 1249velocity measurements (185 +/- 80 m). This figure adapted from Fig. 3 of Lumpkin et al., (2002). 1250The originally published error bars and all measurements below 2000 m depth have been 1251suppressed for clarity. 1252

1253

Page 66: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

65

1254Fig. 16. (a) Raw count of RAFOS-derived velocity measurements per 2˚ square grid cell 1255summed over the entire CLIVAR Mode Water Dynamics Experiment (CLIMODE) bobber 1256record. (b) Contoured representation of the estimated number of degrees of freedom (DOF) per 2˚1257grid cell for only those valid RAFOS-derived velocity measurements on the 18.5˚ C temperature 1258surface. The contour interval is 10. Binned values obtained in regions with DOF < 10 are not 1259contoured and are suppressed in subsequent fields. Also shown in both panels are the initial 1260deployment locations of the 40 CLIMODE bobbers (white squares).1261

1262

Page 67: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

66

12631264

Fig. 17. Record-length mean vector velocity on the 18.5˚ C temperature surface. Circles drawn 1265at the root of each vector represent one standard error of the 2˚ binned velocity data. The velocity 1266vector is suppressed in each grid cell with a mean velocity magnitude less than one standard error 1267or with less than 10 degrees of freedom (DOF). Also shown (contours; interval = 0.05 m) is the 1268mean dynamic topography derived from satellite sea surface height measurements and the 1269GRACE geoid (Jayne, 2006). 1270

12711272

Page 68: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

67

127312741275

Fig. 18. (a) Record-length eddy kinetic energy (EKE) on the 18.5˚ C temperature surface derived 1276from CLIVAR Mode Water Dynamics Experiment (CLIMODE) bobber RAFOS velocity 1277measurements. (b) Surface EKE derived from satellite-tracked surface drifter data as reported by 1278Fratantoni (2001), remapped to a grid consistent with the subsurface bobber observations. Both 1279fields have been spatially interpolated using a bi-cubic spline prior to contouring. 1280

12811282

Page 69: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

68

12831284

Fig. 19. The estimated array bias associated with the mean vector field shown in Fig. 17. The 1285magnitude of the array bias over the core of our study region is less than 1 cm/s. 1286

12871288128912901291

Page 70: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

69

1292129312941295129612971298129913001301130213031304130513061307130813091310131113121313131413151316131713181319132013211322132313241325132613271328132913301331133213331334133513361337

Fig. 20. Record-length mean fields of (a) 18.5˚ C isotherm depth, (b) 17˚–19˚ C layer thickness, 1338and (c) Eighteen Degree Water (EDW) thickness (see Appendix A) derived from the CLIVAR 1339Mode Water Dynamics Experiment (CLIMODE) bobber temperature profile measurements. 1340Values were binned in 2˚x2˚ cells. No data is shown in grid cells with fewer than 10 independent 1341measurements. 1342

Page 71: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

70

13431344

13451346

Fig. 21. As in Fig. 20, but seasonally averaged for boreal winter (left; January–April) and 1347summer (right; July–October). 1348

1349

Page 72: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

71

135013511352

Fig. 22. Record-length mean Eighteen Degree Water (EDW) thickness flux, <uh>, computed by 1353bin-averaging daily values of the product of EDW thickness (h) and RAFOS-derived velocity (u) 1354on the 18.5˚ C isotherm. The resulting field summarizes the mean horizontal (isothermal) 1355transport of EDW in units of volume transport/width. Vectors with magnitude less than the bin-1356averaged standard error are suppressed for clarity. 1357

13581359

Page 73: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

72

1360136113621363136413651366136713681369137013711372137313741375137613771378137913801381138213831384138513861387138813891390139113921393139413951396139713981399140014011402

Fig. 23. Record-length mean transport of Eighteen Degree Water (EDW) as a function of (a) 1403latitude (positive north) and (b) longitude (positive east) computed as per Fig. 22. Meridional 1404transport was computed between 72–42˚ W. Zonal transport was computed between 28–38˚ N. 1405

14061407140814091410

Page 74: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

73

14111412141314141415

Fig. 24. Histogram summarizing the observed elapsed time between subduction and reemergence 1416of Eighteen Degree Water (EDW) as inferred from bobber measurements. A total of 55 examples 1417were noted of a bobber, initially in the unstratified surface mixed layer, submerging beneath the 1418seasonal stratification. In all but a very few cases, each of these bobbers was subsequently 1419reexposed to the mixed layer within 24 months. Note that the number of examples exceeds the 1420number of bobbers: Those bobbers that recycled within the subtropical gyre (i.e. experienced 1421more than one cycle of subduction and reemergence) are included more than once in this tally. 1422

14231424

Page 75: Direct observation of subtropical mode water circulation in the western North Atlantic Ocean

74

142514261427

Fig. 25. Composite annual cycle of Eighteen Degree Water (EDW) thickness computed from all 1428bobber temperature measurements during the CLIVAR Mode Water Dynamics Experiment 1429(CLIMODE) field campaign. Colors represent the number of observations of EDW thickness in 1430bins defined by a 20 m vertical interval and a 10-day period. The black lines were drawn 1431subjectively to illustrate the annual evolution of EDW thickness in the western subtropical 1432Atlantic.1433

14341435