42
1 The seasonality of convective events in the Labrador Sea 1 Hao Luo 1 , Annalisa Bracco 1 *, Fan Zhang 1 2 1 School of Earth and Atmospherics Sciences, 3 Georgia Institute of Technology, Atlanta, GA, 30332, USA 4 5 6 7 8 9 10 11 12 13 Corresponding author: Dr. Annalisa Bracco 14 School of Earth and Atmospheric Sciences 15 Georgia Institute of Technology 16 Email: [email protected] 17

1! The seasonality of convective events in the Labrador Sea · ! ! 1! 1! The seasonality of convective events in the Labrador Sea 2! Hao Luo1, Annalisa Bracco1*, Fan Zhang1 3! 1School

Embed Size (px)

Citation preview

    1  

The seasonality of convective events in the Labrador Sea 1  

Hao Luo1, Annalisa Bracco1*, Fan Zhang1 2  

1School of Earth and Atmospherics Sciences, 3  

Georgia Institute of Technology, Atlanta, GA, 30332, USA 4  

5  

6  

7  

8  

9  

10  

11  

12  

13  

Corresponding author: Dr. Annalisa Bracco 14  

School of Earth and Atmospheric Sciences 15  

Georgia Institute of Technology 16  

Email: [email protected] 17  

    2  

18  Abstract 19  

Modeling deep convection is a key challenge for climate science. Here two simulations of the 20  

Labrador Sea circulation obtained with the Regional Oceanic Modeling System (ROMS) run at a 21  

horizontal resolution of 7.5 km are used to characterize the response of convection to 22  

atmospheric forcing, and its seasonal variability over the period 1980-2009. The integrations 23  

compare well with the sparse observations available. The modeled convection varies in three key 24  

aspects over the 30 years considered. First, its magnitude changes greatly at decadal scales. This 25  

aspect is supported by the in-situ observations. Second, the initiation and peak of convection (i.e. 26  

initiation and maximum) shifts by two to three weeks between strong and weak convective years. 27  

Third, the duration of convection varies by approximately one month between strong and weak 28  

years. The last two changes are associated to the variability of the time integrated surface heat 29  

fluxes over the Labrador Sea during winter and spring, while the first results from changes in 30  

both atmospheric heat fluxes and oceanic conditions through the lateral inflow of warm Irminger 31  

Water from the boundary current system to the basin interior. Changes in surface heat fluxes 32  

over the convective region are linked to large scale modes of variability, the North Atlantic 33  

Oscillation and Arctic Oscillation. Implications for modeling the climate variability of the 34  

Labrador basin are discussed. 35  

36  

37  

38  

39  

40  

    3  

1. Introduction 41  

The Meridional Overturning Circulation (MOC) plays a critical role in transporting heat 42  

across latitudes in the ocean, and its variability has significant impacts on the global climate 43  

system (Bryden et al., 2005). The North Atlantic branch of the MOC is characterized by deep 44  

water formation in few specific locations, including the Nordic Seas and the Labrador Sea 45  

(Dickson et al. 1996; Marshall et al. 2001). Here in winter and early spring the ocean releases 46  

heat to the atmosphere and the surface waters become dense enough to mix by convective 47  

instability (Kuhlbrodt et al. 2007). 48  

In the Labrador Sea, deep convection takes places in the deeper portion of the basin, 49  

seaward of the western continental slope, and forms the Labrador Sea Water (LSW) that often 50  

reaches as deep as 2000 m, exits the Labrador Sea and becomes a distinct component of the 51  

North Atlantic Deep Water, feeding the MOC (Talley and McCartney 1982; Marshall et al. 1998; 52  

Lazier et al. 2002; Rhein et al. 2002; Yashayaev et al. 2003; Yashayaev 2007). The LSW 53  

variability therefore influences the MOC, and it has been suggested that an intensification of the 54  

convective activity in the Labrador Sea will lead to an intensification of the MOC, with an 55  

overall increase in poleward heat transport (Eden and Willebrand, 2001). 56  

The formation of LSW displays strong interannual variability that is influenced by the 57  

atmospheric fluxes, which are highly variable at the latitudes in consideration, and by the 58  

characteristics of the Irminger Current (Myers et al. 2007; Rattan et al. 2010). The Irminger 59  

Current flows along the Greenland coast and its warm and salty subsurface water is brought to 60  

the Labrador Sea interior through anticyclonic mesoscale eddies, the Irminger Rings (Lilly et al. 61  

2003; Katsman et al. 2004; Bracco et al. 2008;  Hátún et al. 2007; Chanut et al. 2008; Rykova et 62  

    4  

al. 2009; Gelderloos et al. 2011; Luo et al. 2012). Since the mid-1990s the Irminger Current has 63  

experienced continuous warming, which has been linked to the concurrent weakening of the 64  

subpolar gyre (Holland et al. 2008). Such weakening, in turn, may be the expression of the 65  

subpolar gyre decadal variability (Häkkinen and Rhines 2004; Böning et al. 2006). 66  

The Labrador Sea convection has been sampled by highly frequency hydrographic 67  

measurements the Ocean Weather Station Bravo (OWSB) from 1945 until 1974 with 24 years of 68  

continuous recording (e.g. Lazier, 1980; Sathiyamoorthy and Moore, 2002), by moorings and P-69  

ALACE floats in the second half of the 1990s (e.g. Marshall et al., 1998, Lavender et al., 2000; 70  

Avsic et al., 2006), by gliders (Frajka-Williams et al. 2014), by hydrographic surveys usually 71  

conducted in late spring or summer and at least yearly from 1990 by the Bedford Institute of 72  

Oceanography (BIO), and by ARGO profiling floats since June 2002. The surveys primarily 73  

sample the AR7W WOCE/CLIVAR line extending from Newfoundland (53.67° N, 55.5° W) to 74  

the west coast of Greenland (60.5° N, 48.25° W), and have been extensively analyzed (e.g. 75  

Lazier et al. 2002; Yashayaev et al. 2003; Lu et al. 2006; Yashayaev 2007; Yashayaev et al. 2007; 76  

Yashayaev and Loder 2009). ARGO floats, on the other hand, provide temperature and salinity 77  

measurements of the upper 2000 m of the water column, and represent a continuous in time but 78  

irregularly in space data set (Yashayaev and Loder 2009; Luo et al. 2012). 79  

Focusing on the last three decades (1980-2009) in-situ data show that the convection in the 80  

Labrador Sea intensified in the late 1980s and early 1990s, and weakened since 1995, with a 81  

limited recovery in the 2007-2008 winter (Våge et al. 2008; Yashayaev and Loder 2009). 82  

Various external causes have been proposed to influence the onset and intensity of convection, 83  

including local and remote atmospheric forcing, the Irminger Current conditions, and the state of 84  

the subpolar gyre (Marshall and Schott 1999; Lazier et al. 2002; Straneo 2006; Straneo et al. 85  

    5  

2010). The variability of convective events, other than their interannual intensity, however, have 86  

been – and can be - investigated only for limited time spans using observations (Straneo, 2006; 87  

Gelderloos et al., 2012) due to their sparseness in space and/or time. 88  

On the modeling side, until recently the attribution of variability of convective events in the 89  

Labrador Sea have been attempted only for decadal modulations (Mizoguchi et al., 2003) due to 90  

the generally poor representation of the details of LSW formation (Canuto et al., 2004; Tréguier 91  

et al., 2005). Luo et al. (2012, LBYD12 in the following) using a regional ocean model were able 92  

to reproduce the interannual component as well, and to quantify the relative importance of the 93  

atmospheric forcing and of the Irminger Current conditions. Here we build on two of the 94  

numerical simulations described in LBYD12 to zoom on the seasonal scales. We analyze how 95  

the time of initiation, peak, and seasonality of convection vary in years of intense versus weak 96  

activity, with immediate application to monitoring planning. Additionally, we explore the drivers 97  

of those changes, and we discuss the modeling challenges. Both ocean-only and coupled climate 98  

models display large biases and divergent behaviors in simulating the formation of deep water 99  

masses and their variability (Canuto et al., 2004; MacMartin et al., 2013). By analyzing the 100  

interplay of oceanic and atmospheric forcings in the representation of the seasonal cycle and of 101  

the interannual modulation of LSW in the North Atlantic in realistic simulations, we hope to 102  

provide a base to validate and test the representation of the Atlantic MOC in coupled climate 103  

models. 104  

2. Model configuration 105  

The Regional Ocean Modeling System (ROMS) is a free-surface, primitive equation model 106  

based on the Boussinesq approximation and hydrostatic balance (Shchepetkin and McWilliams 107  

    6  

2003, 2005). Here ROMS is configured in the Labrador Sea over the domain 35-65°W, 51-66°N 108  

(Fig. 1a), with horizontal resolution of 7.5 km and 30 vertical layers, 8 of which are confined in 109  

the upper 300 m. The nonlocal K-Profile Parameterization (KPP) scheme is used to parameterize 110  

vertical mixing (Large et al. 1994). The integration period begins in January 1980 and ends in 111  

December 2009. 112  

The bathymetry is derived from ETOPO2 (Smith and Sandwell 1997). A modified Shapiro 113  

smoother (Penven et al. 2008) is applied to the original bathymetric data to avoid pressure 114  

gradient errors. The smoother is applied everywhere except for three small regions around Cape 115  

Desolation. Retaining the original bathymetric details in these regions does not affect 116  

significantly the bottom velocities. However, these details are critical for a correct representation 117  

of the eddy-generation along the West Greenland coast, as shown by Bracco et al., (2008). A 118  

detailed discussion of this problem is contained in Appendix A of Luo et al. (2011, LBD11 in the 119  

following). 120  

Boundaries are open to the east, south and north sides of the domain, where the velocity, 121  

temperature and salinity fields are nudged to the Simple Ocean Data Assimilation (SODA) ocean 122  

reanalysis version 2.1.6 (Carton and Giese 2008). A modified radiation boundary condition is 123  

also applied following Marchesiello et al. (2001). SODA has been evaluated against other ocean 124  

reanalysis products in LBYD12, and it presents the advantage of reproducing quite realistically 125  

the lateral and vertical extent of the Irminger Current at the north-east corner of the domain. As 126  

mentioned, the time period investigated in this work extends from 1980 to 2009. Since SODA 127  

2.1.6 ends in 2008, boundary conditions of 2008 are repeated for the computation of 2009. 128  

Surface fluxes are from NCEP/NCAR (1980-2009). Over the period November 1999 – June 129  

2009 the surface wind data consist of monthly averages of the spatial blending of high-resolution 130  

    7  

satellite data (Seawinds instrument on the QuikSCAT satellite - QSCAT) and the NCEP re-131  

analysis (Milliff et al., 2004). To avoid long-term drift of SST linked to the errors in the 132  

NCEP/NCAR heat fluxes (Josey 2001), QNCEP are corrected by the NOAA extended SST 133  

(SSTNOAA) (Smith and Reynolds, 2004) on a monthly timescale and 2°×2° resolution, according 134  

to QMod = QNCEP + dQMod/dSSTMod × (SSTMod - SSTNOAA), where the subscript Mod indicates 135  

modeled quantities.1 Nudging to the monthly surface salinity climatology provided by the Word 136  

Ocean Atlas 2009 (WOA09, Antonov et al., 2009) partially account for the seasonal cycle of the 137  

fresh water fluxes associated to sea-ice melting and the Hudson River outflow. Those anomalies 138  

dominate the seasonal variability of the surface salinity field along the western coast of the basin. 139  

In this study, therefore, we do not account for the interannual variability of the surface fresh 140  

water fluxes except for the eastern and southern model boundaries. 141  

We consider two simulations forced by the same atmospheric forcings, consisting of 142  

monthly varying NCEP/NCAR heat and momentum fluxes, but by different ocean boundary 143  

conditions. In the first simulation (CLIMA), we nudge the open boundaries to monthly mean 144  

climatological values obtained by averaging the SODA monthly output from 1980 to 2009 145  

(where 2009 is identical to 2008). In the second simulation (VARY), we retain the interannual 146  

variability in the boundary conditions without any further averaging. Initial conditions for both 147  

integrations are derived from a spin-up run, forced by climatological monthly averaged 148  

NCEP/NCAR atmospheric fluxes and SODA boundary conditions, that extend for 50 years after 149  

a stationary state is reached. CLIMA and VARY are then initiated in 1976 and the first four 150  

years of simulations are discarded. 151  

                                                                                                                         1  Differences between QNCEP and QMod are very small and to all effects of analyses here negligible.  

    8  

Additionally, we investigate if using high frequency surface heat fluxes (daily) and winds (6 152  

hourly) to force ROMS influences its representation of convective events in the Labrador Sea 153  

performing a third run limited to the period November 1999 – December 2004. The set-up is 154  

identical to VARY but for the frequency of the atmospheric forcing products. No significant 155  

changes were detected in the representation of LSW formation. The outcome of this comparison 156  

is summarized in the Appendix. 157  

3. Validation of model output and potential temperature variability 158  

A detailed validation of the model representation of the surface circulation is provided in 159  

LBD11. It is found that ROMS reproduces accurately the Labrador Sea circulation and its 160  

variability compared to altimeter data, from seasonal to interannual scales, including the surface 161  

EKE along the Greenland coast and in the convective area (Fig. 1a). Independently of the forcing 162  

fields used, and although the horizontal resolution of the model is just below the Rossby radius 163  

of deformation (~13 km) of the basin, the model captures the observed eddy variability. In 164  

particular, most statistics related to the Irminger Ring population are in good agreement with 165  

observations, with the exception of the lifespan of large eddies (six to seven months in CLIMA 166  

and eight to eleven months in VARY, versus twelve to eighteen months in observations). 167  

Consequently, fewer than observed Irminger Rings migrate as far south as 58o S in the model. 168  

The interannual variability of potential temperature through the water column over the 169  

period 1980-2009 is discussed in LBYD12. Modeled PT values are compared with hydrographic 170  

surveys along the AR7W line (Lazier et al. 2002; Yashayaev 2007; Straneo 2006; van Aken et al. 171  

2011) and with ARGO data from the summer of 2002 in the so-called Central Labrador Sea 172  

(CLS), defined accordingly to Yashayaev and Loder (2009) as the region comprised by the 3250 173  

    9  

m isobaths and within 150 km of the AR7W hydrographic line. The CLS is not the region of 174  

greatest convective activity, but is close enough to it to be affected by the spreading of LSW 175  

within weeks from the events, and is where we have the most uniform - in time and space - in-176  

situ data coverage over the period considered. The model representation matches well the 177  

observations from the base of the mixed layer to 2500 m; below such depth ROMS is 178  

approximately half a degree too warm compared to in-situ data and less uniformly stratified than 179  

observed. This discrepancy is likely due to poor representation of vertical mixing at depth and/or 180  

to insufficient model vertical resolution; SODA boundary conditions may further contribute to 181  

the model bias. 182  

The PT in the first half of the record (1980-1994) is characterized by the alternation of 183  

warmer and colder periods, while the second half (1995-2009) is dominated by a warming trend 184  

beginning around 1995 in both CLIMA and VARY (Fig. 1b). The modeled interannual 185  

variations are consistent with hydrographic observations (LBYD12); for example, the model 186  

successfully captures the strong convection events in 1982-1984 and in the early 1990s’, the 187  

reduction in the convective activity after 1995 and the partial recovery in 2008. Independent of 188  

the integration considered, the intensity of convection is reduced significantly after 1995 in both 189  

CLIMA and VARY. LBYD12 showed that the warming in CLS results from the combined 190  

changes in local heat fluxes and warming of the Irminger Current, which has been reported 191  

continuously since 1995 (Böning et al. 2006; Straneo 2006) and is realistically represented in 192  

SODA. Irminger Current water is advected by Irminger eddies into the central portion of the 193  

basin, where it facilitates restratification (Katsman et al. 2004; LBD11). 194  

Moving to seasonal scales, LBYD12 compared the PT seasonal cycle in ROMS with the 195  

observed one by ARGO floats from mid-2002 (see their Fig. 6) and found very good agreement 196  

    10  

between the two. The ARGO period, however, is characterized by weak convective episodes in 197  

all years except 2008, and it is not representative of the interannual variability observed during 198  

the 30 years under investigation. Therefore, as further validation, in Fig. 2 we present a 199  

comparison of the seasonal cycle of heat content and salinity in ROMS and P-ALACE float data 200  

covering 1996-2000 in the top 200 m and in a deeper level, extending from 200 to 1300 m, 201  

following the analysis in Straneo (2006) for the CLS. This period includes four strong convective 202  

events. The general behavior is well represented both at the surface and at depth. The model 203  

slightly underestimates the amplitude of the heat content cycle and overestimates the salinity one 204  

in the surface layer. At depth the seasonal evolution of ROMS heat content is within the standard 205  

deviation of the observations, while the modeled salinity does not agree with the P-ALACE data 206  

and does not display evident seasonal variations. While a similar lack of seasonality was noticed 207  

by Straneo (2006) using OWSB data over the period 1964-1974 independently of the inclusion 208  

of Great Salinity Anomaly years (Dickson et al., 1988), we cannot exclude a model bias. Indeed 209  

the representation of salinity at high latitudes poses a challenge to ocean modeling, as previously 210  

noticed by Tréguier et al. (2005) analyzing the variability of the North Atlantic subpolar gyre in 211  

four high resolution models. Nonetheless, we can conclude that ROMS provides a good 212  

representation of seasonal changes in potential temperature also in periods of strong convection. 213  

4. Seasonality of convective events in the Labrador Sea 214  

4.1 Vertical velocity and convection 215  

Ocean deep convection takes place within vertical plumes of O(1) km in radius and vertical 216  

extent of about 2km (Marshall and Shott, 1999). Those plumes cause the water column to 217  

homogenize within the so-called mixed patch. The model adopted in this investigation, and more 218  

generally all ocean models used for climate studies, do not have sufficient horizontal resolution 219  

    11  

to represent directly the convective process. The integral effect of the convective plumes and 220  

their impact on vertical mixing, and temperature and salinity distributions, however, can be 221  

parameterized (Send and Marshall 1995). In ROMS they are accounted for by the nonlocal KPP 222  

scheme. KPP assumes that the vertical turbulent fluxes of momentum, heat and salinity (or any 223  

other tracer) can be expressed as the sum of a down-gradient flux and a non-local contribution. 224  

Convective events can then be characterized by any quantity directly affected by such fluxes, 225  

from the absolute value of the vertical velocity field, to temperature and/or salinity profiles, or 226  

mixed layer depth (MLD) and turbulent kinetic energy. Vertical velocities present the advantage 227  

over the other tracers in defining convective episodes, and in particular in defining initiation and 228  

duration of convection, by allowing a clear-cut threshold. This cannot be achieved as easily using 229  

temperature or density profiles because following strong convective events the water mixed at 230  

depth during the convective season remains in the central portion of the Labrador basin for long 231  

enough to influence the PT or density averages for the following year (see for example Fig. 13 in 232  

LBYD12). 233  

Fig. 3 shows the mean seasonal cycle of the absolute value of the vertical velocity field, |w|, 234  

averaged over depths comprised between 150 and 2000 m as modeled by ROMS during the 30 235  

years considered. The top 150 m are excluded as strongly influenced by Ekman pumping and 236  

surface momentum forcing (see e.g. Koszalka et al., 2009 for an analysis of relative contributions 237  

to w in an idealized wind driven domain). High values of |w| are found at all times along the 238  

boundary current system in correspondence of steep gradients of the continental slope, broadly 239  

covering the areas comprised between the 1000 and 2000 m isolines. In those areas eddy 240  

generation by baroclinic instability contributes to high level of surface EKE, and the eddies 241  

extend from the ocean surface to the slope (e.g. LBD11). In the basin interior, on the other hand, 242  

    12  

the largest vertical velocities are found only in late winter and early spring over a region 243  

encompassing the deeper portion of the basin east of the western continental slope as observed, 244  

among others, by Lavander et al. (2000), Pickart et al. (2002) and Frajka-Williams et al. (2014). 245  

This area, indicated as CR for convective region in the following, is approximated by the black 246  

box in Fig. 3 with coordinates (56.6 – 52o W, 57 - 59.5o N). We verified that in ROMS the same 247  

area identifies also the maximum MLD (Fig. 3e and 3f). Any MLD definition requires the use of 248  

an ad-hoc threshold in temperature or density. In the Labrador Sea the density differences 249  

between the surface and the base of the mixed layer during convective events can be very small 250  

(Fraika-Williams et al. 2014), smaller than 0.01 kg m−3, usually adopted in numerical studies 251  

(Lazier et al. 2002). A robust observational constrain is not available, and the ‘correct’ threshold 252  

is likely to be one that accounts for seasonal changes and interannual variability, particularly in 253  

presence of trends. In this work the MLD is defined as the depth at which density differences 254  

with the surface are equal to 0.008 kg m−3. As further test, we compared the modeled convective 255  

region from 2003 to 2009 with the one identified by the MLD in the ARGO data. In both 256  

datasets, convection is patchy, and approximately located over a smaller area than the CLS or the 257  

CR regions, and close to the intersection of the two. Whenever the average over the period is 258  

considered, the agreement between model and observations is very good (not shown). 259  

The time series of |w| and MLD in CLIMA and VARY and their differences are presented 260  

in Fig. 4. Compared to the few observational estimates the modeled MLD appears generally too 261  

deep. This bias is directly linked to the deep layers in ROMS being less uniformly stratified than 262  

observed. The correlation between |w| and MLD has coefficient R = 0.9. Increasing the density 263  

threshold in the MLD criterion to the more common 0.01 kg/m-3 (e.g. Tréguier et al., 2005) does 264  

not affect the correlation but further increase the maximum MLD in the last ten years of the 265  

    13  

integration, when a warming trend is apparent. If a threshold of 0.02 kg/m-3 or higher is used 266  

instead, the correlation decreases dramatically and the mixed layer reaches the ocean bottom 267  

during most events. 268  

From the |w| time series we separate years of strong and weak convection using a threshold 269  

of 0.85x10-3 m/s. This allows us to group 15 years (1982, 1983, 1984, 1985, 1988, 1989, 1990, 270  

1991, 1992, 1993, 1994, 1999, 2000, 2002, 2008) in the strong, and 15 years (1980, 1981, 1986, 271  

1987, 1995, 1996, 1997, 1998, 2001, 2003, 2004, 2005, 2006, 2007, 2009) in the weak 272  

convection categories, respectively. For six of the thirty years considered a change in the 273  

threshold by ± 10% could modify their classification, and we choose to include three of those 274  

normal events in each category. While the |w| criterion provides a differentiation similar to the 275  

one of MLD exceeding (or not) 1000 m in observational data (e.g. Pickart et al. 2002), the 276  

classification in VARY does not match perfectly the observations. For example the winter of 277  

1999 is commonly categorized as having weak convection, while 1997 is seen as a strong. This is 278  

captured in CLIMA but not in VARY, even if the two runs have similar Irminger Current inflow 279  

over the 1995-1999 period, and is associated to the (modeled) internal ocean variability. In our 280  

analysis, we retain the same groups in CLIMA to simplify the comparison and have an equal 281  

number of events in each group. The vertical velocity time series in Fig. 4a and 4b are highly 282  

correlated (correlation coefficient R = 0.95), but according to the chosen w threshold four events 283  

classified as weak in VARY would follow under the strong category in CLIMA. We verified that 284  

our conclusions do not depend on the details of the classification in CLIMA, or on the |w| 285  

threshold in VARY. 286  

The interannual variability of convective events that emerges from the vertical velocity field 287  

is consistent with the one from PT or potential density over the same region (Fig. 5 and Fig. 6). 288  

    14  

The classification of weak and strong years does not change with the variable used, but the 289  

differences between VARY and CLIMA, indicative of the influence of the Irminger Current on 290  

convective activity, appear amplified (Fig. 5c). For the first half of the record the potential 291  

temperature averaged between 150 and 2000 m depth over the CLS domain is greater in CLIMA 292  

by about 0.07 °C on average, and vertical velocities display large differences but of alternating 293  

sign, so that their average is close to zero. After 1995, PT (density) in VARY is consistently 294  

higher (lower) than in CLIMA, and the absolute vertical velocities are mostly higher in CLIMA. 295  

The PT warming trend is consistent with the one found in the incoming Irminger Current at the 296  

model boundary (see LBYD12, their Fig. 9). Overall, the boundary current contribution explains 297  

about half of the warming recorded in VARY in the CR or CLS regions after 1995. 298  

It is important to notice that the Irminger Current warming recorded after 1995 is not 299  

unprecedented, at least according to SODA, but is part of the decadal variability of the system. 300  

The period from 1958 to 1970 was indeed characterized by an Irminger Current as warm as 301  

during the first decade of the XXI century. 302  

4.2 Seasonality and strength of convection 303  

The seasonal cycle of convective events over the Labrador Sea from 1980 to 2002 cannot be 304  

easily quantified in the observations due to their limited time and/or space coverage except for 305  

the P-ALACE period (1996-2000) (Straneo 2006; Lavender et al., 2000). Shipboard and 306  

hydrographic surveys have sampled late spring and summer, moorings provides continuous time 307  

series but at one location, and floats have been deployed only for limited time intervals. Since 308  

mid-2002 ARGO data, complemented by the K1 mooring (deployed in 1995 near the former 309  

    15  

BRAVO station at about 25 km distance, e.g. Avsic et al., 2006) and at times by gliders, allowed 310  

to measure the seasonal cycle with greater confidence. 311  

Here we compare the seasonality of weak and strong convective events using the modeled 312  

|w|, and we find that they differ in the timing of their initiation, maxima and shutdown (Fig. 7a). 313  

On average, weak events initiate two (CLIMA) to three (VARY) weeks after the strong ones, 314  

they reach their peak of intensity two or three weeks later, and terminate approximately two 315  

weeks earlier. This behavior is common to all weak events, independent of the decade considered, 316  

and is further amplified if the six ‘normal’ years are excluded by the calculation. Consequently, 317  

the convective activity in weak years is approximately one month shorter than in strong ones and 318  

shifted further into spring. 319  

The reduced intensity, shifted seasonality and shortened duration of convection in the CR 320  

for weak events in both integrations are associated predominately to reduced heat fluxes and 321  

wind intensity in winter (Figs. 7b and 7c). It is noticeable that atmospheric changes in the 322  

reanalysis are found only between November and March, and both heat and momentum fluxes 323  

do not show any significant difference through the remaining of the year. Also, the atmospheric 324  

fluxes display a substantial change in intensity, but almost no shift of seasonality in their annual 325  

cycle. In weak years, it takes at least two weeks more for the convection to start because of the 326  

reduced heat flux and associated wind intensity. Fig. 7 reveals also that the convection period 327  

defined by |w| does not coincide with the period during which the heat flux to the atmosphere 328  

forces convective mixing. The offset at initiation is associated with the excess buoyancy 329  

accumulated in the upper 150 – 200 m between May and August, when the surface heat fluxes 330  

are, on average, positive. Such excess buoyancy needs to be removed (Bailer et al. 2005; 331  

Gelderloos et al., 2012; Frajka-Williams et al. 2014) before convection can take place. Thereafter, 332  

    16  

it is enough for the surface heat fluxes to remain negative, even if weakly, for the convection to 333  

continue in the top 200 - 300 m of the water column, and for |w| to remain large at those depths. 334  

The input of warm, salty Irminger water from the Western Greenland coast to the center of 335  

the Labrador basin increases the stratification, and consequently more heat has to be released to 336  

the atmosphere for the convective instability to begin, further delaying its initiation. The impact 337  

of the boundary current interannual variability on convection is quantified as the difference 338  

between VARY and CLIMA and it is more prominent in weak than in strong years, due to the 339  

significant warming of Irminger Current experienced since 1995 (Stein 2005; Myers et al. 2007). 340  

For instance, the maximum absolute value of the CR-averaged vertical velocities in VARY for 341  

weak years is 0.5×10-3 m/s, to be compared 0.7×10-3 m/s in CLIMA, and it is achieved about a 342  

week later (Fig. 7a). In contrast, mean |w| in VARY and CLIMA are almost identical in strong 343  

convective years. 344  

4.3 Strength of convection and heat fluxes 345  

Figure 8 shows the relation between the strength of convection, measured by the absolute 346  

value of w averaged over the period in which |w| > 0.2 ms-1, and the surface heat fluxes averaged 347  

over the CR in different seasons for all, or weak and strong years separately. A high coefficient 348  

(R) implies that the modeled convection intensity is regulated by the local surface heat fluxes. 349  

The Labrador Sea convection occurs in late winter and spring, and the correlation between 350  

vertical velocities in the CR and the atmospheric heat fluxes is the greatest when heat fluxes 351  

from December to April are considered. Coefficients are generally higher in CLIMA than in 352  

VARY (Figs. 8a and 8b), but are greater than 0.85 in both cases when all years are considered. 353  

The wintertime heat flux, particularly in January and February, is the most important (Figs. 8c-354  

    17  

8d) since the conditions for the initiation of the convective events are set. Moreover, the 355  

accumulated heat flux in winter is larger than in spring (Fig. 7a). On the other hand, the delayed 356  

onset in weak years requires considering early spring, March and April, together with (D)JF to 357  

obtain a statistical significant relation for weak events in VARY, and greatly improve the 358  

correlation in CLIMA.2 In contrast, the atmospheric forcing averaged over three, six and nine 359  

months prior of the inception of convection does not influence the convective activity, in 360  

agreement with the analysis by Straneo (2006) that showed that the positive heat fluxes into the 361  

ocean in late spring and summer (May to August) are absorbed in the upper 200 m and are then 362  

released in autumn and early winter (September to December). 363  

The strength of convection depends also on the inflow of Irminger Current waters: the 364  

maxima correlations achieved in VARY are all slightly smaller than in CLIMA, and the slopes 365  

describing the linear relation between vertical velocities and heat fluxes change more 366  

dramatically between strong and weak years. Notwithstanding the modulation at decadal scales 367  

associated with changes of the Irminger Current characteristics, the seasonal variability of the 368  

convective activity in the Labrador Sea appears controlled to a large extent by the local heat 369  

fluxes immediately before and during the convective season. Heat fluxes over the CR result, in 370  

part, from large scale anomalies and during the winter season are correlated with both the North 371  

Atlantic Oscillation (NAO) (Dickson et al., 1996), and the Arctic Oscillation (AO) (Thompson 372  

and Wallace, 1998) (see Table 1). 373  

5. Discussion and conclusions 374  

                                                                                                                         2  If only JFM are considered the correlation in VARY decreases to R = 0.8 when all years are considered and R = 0.4 for weak events.  

    18  

In this paper, two regional simulations of the Labrador Sea circulation in the period 1980-375  

2009 are used to investigate changes in the seasonal cycle of convective events. The numerical 376  

integrations, CLIMA and VARY, differ in their boundary conditions and allow for isolating the 377  

impact of atmospheric forcing, through heat and momentum fluxes, and oceanic forcing, through 378  

changes in the properties of the incoming currents in the basin, on the interannual and seasonal 379  

variability of convective events. Both CLIMA and VARY reproduce well the observed 380  

circulation, the surface eddy kinetic energy and its interannual variability. In particular ROMS 381  

simulates realistically the population of Irminger Rings, and the surface EKE agrees with 382  

observational estimates from altimeter data along the west coast of Greenland, where the Rings 383  

are formed, and in the central portion of the basin where convection occurs (Luo et al., 2011). 384  

Additionally, the evolution of potential temperature throughout the water column is well 385  

represented (Luo et al., 2012) up to 2500 m of depth. As expected, VARY reproduces satellite 386  

and hydrographic observations more realistically than CLIMA by accounting for the changes in 387  

the Irminger Current. In ROMS the region characterized by convective activity is found seaward 388  

of the western continental slope in agreement with observational (Lavander et al., 2000; Pickart 389  

et al., 2002) and numerical studies (Mizoguchi et al., 2003). A recent work by Zhu et al. (2014) 390  

related the ability of a model to properly simulate the localization of deep convection in the 391  

Labrador Sea to its representation of eddy-induced lateral fluxes, associated to the Irminger 392  

Rings and to the smaller eddies formed by baroclinic instability along the boundary current 393  

system. A number of previous studies have indeed shown limitations in this respect, likely due to 394  

coarse resolution (Willebrand et al., 2001), and/or parameterization choices (Canuto et al., 2004). 395  

The analysis presented indicate that the annual changes in the strength of convection in the 396  

Labrador Sea are predominately determined by local atmospheric forcing, consistent with 397  

    19  

previous studies (Delworth and Greatbatch 2000; Eden and Willebrand 2001; Bentsen et al. 398  

2004). Heat fluxes over the CR from December to April correlate highly with the oceanic 399  

vertical velocities averaged between 150 and 2000 m depth in both simulations, with the 400  

wintertime heat flux having larger impacts in years of strong convection (Dickson et al. 1996), 401  

and spring heat fluxes being as important as winter ones during weak convective episodes. In 402  

contrast, the atmospheric forcing averaged over three, and six months prior of the inception of 403  

convection does not influence the convective activity, in agreement with the analysis by Straneo 404  

(2006). The state of the Irminger Current contributes to the modulation of the convective activity 405  

through the Irminger Rings that carry warm and salty water to the convective area, increasing the 406  

stratification through lateral mixing. Its variability, however, plays only a secondary role in the 407  

thirty years considered, as manifested by the difference of vertical velocity between CLIMA and 408  

VARY (Fig. 4c). More quantitatively, the comparison between the seasonal cycle of |w| in 409  

CLIMA and VARY (Fig. 7a) indicates that the interannual modulation of the boundary current is 410  

responsible for a further reduction of about 25% in the maximum of the mixing intensity. 411  

The seasonal cycle of Labrador Sea convection has been observed since the deployment of 412  

ARGO floats in the basin, but we do not have observations that are continuous in time and space 413  

prior to the summer of 2002. Here we show that weak and strong convective events are 414  

characterized not only by different intensity, but also by a different seasonal cycle. Both 415  

initiation and peak of convection for weak events in VARY are delayed by about three weeks 416  

compared to strong ones. Additionally, the duration of convective activity is approximately one 417  

month shorter for weak episodes. Those characteristics are linked to the reduced atmospheric 418  

cooling observed between December and April – but not in other months – in the climatology of 419  

the heat fluxes during weak years. The seasonal cycle of the atmospheric heat and momentum 420  

    20  

fluxes is almost unaltered, but variations in surface cooling rates are translated in changes in 421  

intensity, seasonality and duration of ocean convection. Those changes are important for 422  

understanding the variability in the oxygen drawdown and carbon sequestration in the basin. 423  

Finally, local heat fluxes reflect atmospheric variability at broader scales, and in winter are 424  

significantly correlated with NAO/AO. 425  

In light of previous work, our analysis suggests that a realistic representation of the 426  

interannual variability of LSW formation can be achieved also in a nonhydrostatic model, but at 427  

least two ingredients are required. First, the ocean model resolution has to be below the Rossby 428  

deformation radius to allow for the direct representation of lateral eddy fluxes. Canuto et al. 429  

(2004) have shown that models where the eddies are parameterized display large biases 430  

independently of the vertical mixing scheme adopted. The recent analysis of glider data by 431  

Frajka-Williams et al. (2014) further pinpointed to the importance of horizontal variations in 432  

density over short distances (tens of kilometers) in the convective area for restratifying the region. 433  

Variations at those scales (1-2 grid points), and of about 1/3 of the observed value (~ 0.003 kg m-434  

3), are present in the 3-day average density fields used in Fig. 6. Model resolution plays also an 435  

important role in the representation of the boundary current system and its instabilities. Second, 436  

the seasonal cycle and the interannual variability of the surface heat fluxes are both important. 437  

The vast majority of coupled climate models, including the ones adopted in the Coupled Model 438  

Intercomparison Project Phase 5, have a biased seasonal cycle, with unrealistically large seasonal 439  

excursions, and spring (fall) surface atmospheric fields closely resembling the ones in winter 440  

(summer). On the positive side, the representation of the time integrated (monthly) effect of 441  

winter storms on the surface wind and heat fluxes is sufficient to achieve a realistic 442  

    21  

representation of LSW formation (see Appendix), suggesting that a detailed representation of 443  

extreme winter events is not required if their average impacts are accounted for. 444  

Given the biases found in the representation of the Atlantic meridional overturning 445  

circulation in ocean and coupled climate models, and of the divergent behavior of the latter in 446  

future projections (e.g. MacMartin et al., 2013), the high correlation between the intensity of 447  

convective activity in the Labrador Sea and both local and large scale atmospheric heat flux may 448  

provide an important test for validating climate models, while pointing to a simple way to 449  

parameterize and improve the representation of convection when coarser resolution is used. 450  

Acknowledgements 451  

The simulations used in this work were performed under NSF OCE-0751775. The analysis has 452  

been partially supported through the NSF grant OCE-1357373. We thank three anonymous 453  

reviewers whose thoughtful comments greatly improved this work. 454  

Appendix 455  

The results presented in this work are obtained using ROMS forced by monthly atmospheric 456  

forcing fields interpolated to the model time step. In the Labrador Sea, convection is driven, 457  

predominantly, by the integral of the surface heat flux. This is shown, for example, by 458  

Yashayaev and Loder (2009) using NCEP heat fluxes and ARGO data, and has been confirmed 459  

by our integrations (see Figure 10 in LBYD12). The prominent role of the heat fluxes is further 460  

supported by the investigation of the recovery mechanisms of convection after the shut down due 461  

to the Great Salinity Anomaly by Gelderloos et al. (2012), and by the analysis of the 2005 462  

convective season using ARGO data and sea-gliders performed by Frajka-Williams et al. (2014). 463  

This happens because over most of the ocean, including the convective region in the Labrador 464  

    22  

Sea as defined in the manuscript, wind mixing is limited to the top 200 m of the water column 465  

even during strong events. Consequently, the use of higher frequency winds (or/and higher 466  

frequency heat fluxes) does not substantially change the representation of deep convection in the 467  

center of the basin. The above discussion is supported by a 5-yr long simulation – from Jan 2000 468  

to Dec 2004 – identical in set-up to VARY but forced by NCEP/NCAR daily heat fluxes and 469  

QuikSCAT – NCEP blended 6-hour winds, VARY-HF in the following. Comparing the mean 470  

climatology of potential temperature in VARY-HF and VARY over the common period, it is 471  

found that the mean modeled temperatures associated with the high frequency atmospheric 472  

fluxes are ~ 0.3 oC warmer immediately below the surface and within the upper 100 m, and 473  

slightly cooler (~ 0.1 oC or less) in the deeper layers, in agreement with results by Ezer (2000) 474  

over the North Atlantic, and by Cardona and Bracco (2012) in the South China Sea. The latter 475  

has shown that the high frequency winds excite near inertial waves as ageostrophic expression of 476  

the eddy field, whenever a vigorous one is present, as in the case of the Labrador Sea. Enhanced 477  

mixing by near inertial waves causes, in turn, the subsurface warming and the deeper cooling 478  

(the frequency spectra analysis reported in Cardona and Bracco has been repeated for this 479  

domain confirming their conclusions once the different Coriolis frequency is accounted for). The 480  

seasonal cycle of the SST anomalies (Fig. A2) and the interannual variability (Fig. A3) of 481  

potential temperature in the CR, however, do not depend on the frequency of the atmospheric 482  

forcing used, being the eddy field analogous in the two integrations everywhere but at the tip of 483  

Greenland, where the EKE in VARY-HF is twice as higher as in VARY, and in better agreement 484  

with observations (Pickart et al., 2003). In conclusion, we exclude a significant role of the 485  

atmospheric forcing frequency in determining strength and timing of modeled convective events 486  

in the Labrador Sea. 487  

    23  

References 488  

Antonov, J. I., D. Seidov, T. P. Boyer, R. A. Locarnini, A. V. Mishonov, H. E. Garcia, O. K. 489  

Baranova, M. M. Zweng, and D. R. Johnson, 2010. World Ocean Atlas 2009, Volume 2: 490  

Salinity. S. Levitus, Ed. NOAA Atlas NESDIS 69, U.S. Government Printing Office, 491  

Washington, D.C., 184 pp. 492  

Avsic, T., J. Karstensen, U. Send, and J. Fisher, 2006: Interannual variability of newly formed 493  

Labrador Sea Water from 1994 to 2005. Geoph. Res. Lett., 33, L21S02, 494  

doi:10.1029/2006GLo26913. 495  

Bailer, D.A., P.R. Rhines and S. Häkkinen, 2005: Formation and pathways of North Atlantic 496  

Deep Water in a coupled ice-ocean model of the Arctic-North Atlantic Oceans. Clim. Dyn., 497  

25, 497-516. 498  

Bentsen, M., H. Drange, T. Furevik, and T. Zhou, 2004: Simulated variability of the Atlantic 499  

meridional overturning circulation. Clim. Dyn., 22, 701-720. 500  

Böning, C. W., M. Scheinert, J. Dengg, A. Biastoch, and A. Funk, 2006: Decadal variability of 501  

subpolar gyre transport and its reverberation in the North Atlantic overturning. Geophys. Res. 502  

Lett., 33, L21S01. 503  

Bracco, A., J. Pedlosky, and R. S. Pickart, 2008: Eddy formation near the west coast of 504  

Greenland. J. Phys. Oceanogr., 38, 1992-2002. 505  

Bryden, H. L., H. R. Longworth, and S. A. Cunningham, 2005: Slowing of the Atlantic 506  

meridional overturning circulation at 25 N. Nature, 438, 655-657. 507  

Canuto V.M., A. Howard, P. Hogan, Y. Cheng, M.S. Dubovikov, and L.M. Montenegro, 2004: 508  

Modeling ocean deep convection. Ocean Model., 7, 75-95, doi:10.1016/S1463-509  

5003(03)00038-6. 510  

Cardona, Y., and A. Bracco, 2012: Enhanced vertical mixing within mesoscale eddies due to 511  

high frequency winds in the South China Sea. Ocean Modeling, 42, 1-15. 512  

Carton, J. A., and B. S. Giese, 2008: A reanalysis of ocean climate using Simple Ocean Data 513  

Assimilation (SODA). Mon. Weather Rev., 136, 2999-3017. 514  

Chanut, J., B. Barnier, W. Large, L. Debreu, T. Penduff, J. M. Molines, and P. Mathiot, 2008: 515  

Mesoscale eddies in the Labrador Sea and their contribution to convection and restratification. 516  

J. Phys. Oceanogr., 38, 1617-1643. 517  

    24  

Delworth, T. L., and R. J. Greatbatch, 2000: Multidecadal thermohaline circulation variability 518  

driven by atmospheric surface flux forcing. J. Clim., 13, 1481-1495. 519  

Dickson, R. R., J. Meincke, S. A. Malmberg, and A. J. Lee, 1988: The “Great Salinity Anomaly” 520  

in the northern North Atlantic 1968–1982. Progr. Ocean., 20, 103–151. 521  

Dickson, R., J. Lazier, J. Meincke, P. B. Rhines, and J. Swift, 1996: Long-term coordinated 522  

changes in the convective activity of the North Atlantic. Prog. Oceanogr., 38, 241-295. 523  

Eden, C., and J. Willebrand, 2001: Mechanism of interannual to decadal variability of the North 524  

Atlantic circulation. J. Clim., 14, 2266-2280. 525  

Ezer, T., 2000: On the seasonal mixed layer simulated by a basin-scale ocean model and the 526  

Mellor-Yamada turbulence scheme. J. Geophys. Res., 105, 16843–16855. 527  

Frajka-Williams, E., P.B. Rhines, and C.C. Eriksen, 2014: Horizontal Stratification during Deep 528  

Convection in the Labrador Sea. J. Phys. Oceanogr., 44, 220-228. 529  

Gelderloos, R., C. A. Katsman, and S. S. Drijfhout, 2011: Assessing the roles of three eddy types 530  

in restratifying the Labrador Sea after deep convection. J. Phys. Oceanogr., 41, 2102-2119. 531  

Gelderloos, R., F. Straneo and C.A. Katsman, 2012: Mechanisms behind the Temporary 532  

Shutdown of Deep Convection in the Labrador Sea: Lessons from the Great Salinity 533  

Anomaly Years 1968-71. J. Climate, 25, 6743-6755. 534  

Häkkinen, S., and P. B. Rhines, 2004: Decline of subpolar North Atlantic circulation during the 535  

1990s. Science, 304, 555-559. 536  

Holland, D. M., R. H. Thomas, B. De Young, M. H. Ribergaard, and B. Lyberth, 2008: 537  

Acceleration of Jakobshavn Isbrae triggered by warm subsurface ocean waters. Nat. Geosci., 538  

1, 659-664. 539  

Josey, S., 2001: A comparison of ECMWF, NCEP-NCAR, and SOC surface heat fluxes with 540  

moored buoy measurements in the subduction region of the northeast Atlantic. J. Clim., 14, 541  

1780-1789. 542  

Katsman, C. A., M. A. Spall, and R. S. Pickart, 2004: Boundary current eddies and their role in 543  

the restratification of the Labrador Sea*. J. Phys. Oceanogr., 34, 1967-1983. 544  

Koszalka, I., A. Bracco, J.C. McWilliams, A. Provenzale, 2009: Dynamics of wind-forced 545  

coherent anticyclones in the open ocean, J. Geoph. Res., 114, C08011, 546  

doi:10.1029/2009JC005388. 547  

    25  

Kuhlbrodt, T., A. Griesel, M. Montoya, A. Levermann, M. Hofmann, and S. Rahmstorf, 2007: 548  

On the driving processes of the Atlantic meridional overturning circulation. Rev. Geophys., 549  

45, RG2001. 550  

Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and 551  

a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363-403. 552  

Lavender, K. L., R. E. Davis, and W. B. Owens, 2000: Mid-depth recirculation observed in the 553  

interior Labrador and Irminger seas by direct velocity measurements. Nature, 407, 66-69. 554  

Lazier, J. N. R., 1980: Oceanographic conditions at Ocean Weather Ship Bravo, 1964–1974. 555  

Atmos.–Ocean, 18, 227–238. 556  

Lazier, J. R. N., R. Hendry, A. Clarke, I. Yashayaev, and P. Rhines, 2002: Convection and 557  

restratification in the Labrador Sea, 1990–2000. Deep Sea Res. Part I, 49, 1819-1835. 558  

Lilly, J. M., P. B. Rhines, F. Schott, K. Lavender, J. Lazier, U. Send, and E. D’Asaro, 2003: 559  

Observations of the Labrador Sea eddy field. Prog. Oceanogr., 59, 75-176. 560  

Lu, Y., D. G. Wright, and I. Yashayaev, 2007: Modelling hydrographic changes in the Labrador 561  

sea over the past five decades, Prog. Oceanogr., 73, 406-426. 562  

Luo, H., A. Bracco, and E. Di Lorenzo, 2011: The interannual variability of the surface eddy 563  

kinetic energy in the Labrador Sea. Prog. Oceanogr., 91, 295-311. 564  

Luo, H., A. Bracco, I. Yashayaev, and E. Di Lorenzo, 2012: The interannual variability of 565  

potential temperature in the central Labrador Sea. J. Geophys. Res., 117, C10016, 566  

doi:10.1029/2012JC007988. 567  

MacMartin, D.G., E. Tzipperman, and L. Zanna, 2013: Frequency domain multimodel analysis 568  

of the response of Atlantic meridional overtuning circulation to surface forcing. J. Climate, 569  

26, 8323-8340. 570  

Marchesiello, P., J. C. McWilliams, and A. Shchepetkin, 2001: Open boundary conditions for 571  

long-term integration of regional oceanic models. Ocean Model., 3, 1-20. 572  

Marshall, J., and Coauthors, 1998: The Labrador Sea deep convection experiment. Bull. Am. 573  

Meteorol. Soc., 79, 2033-2058. 574  

Marshall, J., and F. Schott, 1999: Open-ocean convection: Observations, theory, and models. Rev. 575  

Geophys, 37, 1-64. 576  

Marshall, J., and Coauthors, 2001: North Atlantic climate variability: phenomena, impacts and 577  

mechanisms. Int. J. Climatol., 21, 1863-1898. 578  

    26  

Milliff, R. F., J. Morzel, D. B. Chelton, and M. H. Freilich, 2004: Wind Stress Curl and Wind 579  

Stress Divergence Biases from Rain Effects on QSCAT Surface Wind Retrievals. J. Atmos. 580  

Ocean Tech., 21, 1216-1231. 581  

Mizoguchi, K.-I., S. L. Morey, J. Zavala-Hidalgo, N. Suginohara, S. Häkkinen, and J. J. O’Brien, 582  

2003: Convective activity in the Labrador Sea: Preconditioning associated with decadal 583  

variability in subsurface ocean stratification, J. Geoph. Res., 108(C10), 3330, 1-28, 584  

doi:10.1029/2002JC001735 585  

Myers, P. G., N. Kulan, and M. H. Ribergaard, 2007: Irminger water variability in the West 586  

Greenland Current. Geophys. Res. Lett., 34, L17601. 587  

Penven, P., P. Marchesiello, L. Debreu, and J. Lefèvre, 2008: Software tools for pre-and post-588  

processing of oceanic regional simulations. Environ. Model. Software, 23, 660-662. 589  

Pickart, R. S., D. J. Torres, and R. A. Clarke, 2002: Hydrography of the Labrador Sea during 590  

Active Convection. J. Phys. Ocean., 32, 428-457. 591  

Pickart, R. S., M. A. Spall, M. H. Ribergaard, G. K. Moore, and R. F. Milliff, 2003: Deep 592  

convection in the Irminger Sea forced by the Greenland tip jet. Nature, 424, 152-156. 593  

Rattan, S., P. G. Myers, A.-M. Tréguier, S. Theetten, A. Biastoch, and C. Böning, 2010: Towards 594  

an understanding of Labrador Sea salinity drift in eddy-permitting simulations, Ocean Model., 595  

35, 77-88, doi:10.1016/j.ocemod.2010.06.007. 596  

Rhein, M., and Coauthors, 2002: Labrador Sea Water: Pathways, CFC inventory, and formation 597  

rates. J. Phys. Oceanogr., 32, 648-665. 598  

Rykova, T., F. Straneo, J. M. Lilly, and I. Yashayaev, 2009: Irminger current anticyclones in the 599  

Labrador Sea observed in the hydrographic record, 1990-2004, J. Mar. Res., 67, 361-384. 600  

Sathiyamoorthy, S., and G. W. K. Moore, 2002: Buoyancy flux at Ocean Weather Station Bravo. 601  

J. Phys. Oceanogr., 32, 458–474 602  

Shchepetkin, A. F., and J. C. McWilliams, 2003: A method for computing horizontal pressure-603  

gradient force in an oceanic model with a nonaligned vertical coordinate. J. Geophys. Res., 604  

108, 3090. 605  

——, 2005: The regional oceanic modeling system (ROMS): a split-explicit, free-surface, 606  

topography-following-coordinate oceanic model. Ocean Model., 9, 347-404. 607  

Smith, T. M., and R. W. Reynolds, 2004: Improved extended reconstruction of SST (1854-1997). 608  

J. Clim., 17, 2466-2477. 609  

    27  

Smith, W. H., and D. Sandwell, 1997: Marine gravity anomaly from Geosat and ERS 1 satellite 610  

altimetry. J. Geophys. Res., 102, 10039-10054. 611  

Stein, M., 2005: North Atlantic subpolar gyre warming–impacts on Greenland offshore waters. J. 612  

Northw. Atl. Fish. Sci., 36, 43-54. 613  

Straneo, F., 2006: Heat and freshwater transport through the central Labrador Sea, J. Phys. 614  

Oceanogr., 36, 606-628. 615  

Straneo, F., and Coauthors, 2010: Rapid circulation of warm subtropical waters in a major glacial 616  

fjord in East Greenland. Nat. Geosci., 3, 182-186. 617  

Talley, L. D., and M. S. McCartney, 1982: Distribution and circulation of Labrador Sea water. J. 618  

Phys. Oceanogr., 12, 1189-1205. 619  

Thompson, D. W., and J. M. Wallace, 1998: The Arctic Oscillation signature in the wintertime 620  

geopotential height and temperature fields, Geophys. Res. Lett., 25, 1297-1300. 621  

Tréguier, A.-M., S. Theetten, E. Chassignet, T. Penduff, R. Smith, L. Talley, J. O. Beismann, and 622  

C. Böning, 2005: The North Atlantic subpolar gyre in four high-resolution models. J. Phys. 623  

Oceanogr., 35, 757-774. 624  

Våge, K., and Coauthors, 2008: Surprising return of deep convection to the subpolar North 625  

Atlantic Ocean in winter 2007–2008. Nat. Geosci., 2, 67-72. 626  

van Aken, H. M., M. Femke de Jong, and I. Yashayaev, 2011: Decadal and multi-decadal 627  

variability of Labrador Sea Water in the north-western North Atlantic Ocean derived from 628  

tracer distributions: Heat budget, ventilation, and advection. Deep Sea Res. Part I, 58, 505-629  

523. 630  

Yashayaev, I., J. R. N. Lazier, and R. A. Clarke, 2003: Temperature and salinity in the central 631  

Labrador Sea, ICES Marine Symposia Series, 219, 32-39. 632  

Yashayaev, I., 2007: Hydrographic changes in the Labrador Sea, 1960–2005. Prog. Oceanogr., 633  

73, 242-276. 634  

Yashayaev, I., H. M. van Aken, N. P. Holliday, and M. Bersch, 2007: Transformation of the 635  

Labrador Sea Water in the subpolar North Atlantic, Geophys. Res. Lett., 34, L22605, 636  

doi:10.1029/2007GL031812. 637  

Yashayaev, I., and J. W. Loder, 2009: Enhanced production of Labrador Sea Water in 2008, 638  

Geophys. Res. Lett., 36, L01606, doi: 10.1029/2008GL036162. 639  

    28  

Willebrand, J., and Coauthors, 2001: Circulation characteristics in three eddy-permitting models 640  

of the North Atlantic. Progr. Oceanography, 48, 123-161 641  

Zhu, J. S., E. Demirov, Y. Zhang, and A. Polomska-Harlick, 2014: Model simulations of 642  

mesoscale eddies and deep convection in the Labrador Sea. Adv. Atmos. Sci., doi: 643  

10.1007/s00376-013-3107-y,in press. 644  

645  

646  

    29  

Table 1: Correlation coefficients between seasonal averages (DJF and DJFMA) of the heat 647  

fluxes into the ocean over the CR region and the AO/NAO indices for 15 strong convective 648  

years, 15 weak episodes, and all 30 years. Correlations using monthly values over the same 649  

seasons are statistically indistinguishable. Correlations significant at the 95% confidence 650  

level according to a t-test are in bold. 651  

652  

AO strong

AO weak

AO all

NAO strong

NAO weak

NAO all

Winter (DJF) -0.49 -0.27 -0.47 -0.53 -0.53 -0.55 WinterSpring (DJFMA) -0.40 -0.18 -0.37 -0.48 -0.39 -0.46

653  

654  

655  

656  

    30  

Figure captions 657  

Fig. 1 (a) Annual mean distribution of surface eddy kinetic energy cast as a speed !!"! =658  

2!"! (in color) and mean surface velocity (arrows). The locations of the AR7W WOCE 659  

hydrographic line (black line) and of the Ocean Weather Station BRAVO (red star) are marked. 660  

(b) Annual mean of potential temperature in the region comprised by the 3250 m isobaths 661  

following Yashayaev (2007). Potential temperature values are averaged over depths comprised 662  

between 150 m and 2000 m. 663  

Fig. 2 Seasonal variation in heat content (top) and salinity (bottom) in the central Labrador Sea 664  

for the top 200 m of the water column (left), and for a lower layer comprised between 200 and 665  

1300 m (right) in the model (blue and red lines) and in the observations presented in Straneo 666  

(2006) (black and gray lines). The model and the P-ALACE float data cover the period 1996-667  

2000. The gray lines represent the standard deviation (std) around the float data. The dashed 668  

lines indicate the std of the Ocean Weather Station Bravo data from 1964 to 1974 (also from 669  

Straneo, 2006). 670  

Fig. 3 Seasonal mean distribution of vertical velocity |w| averaged between 150 and 2000 m 671  

depth (a) at the peak of convection (February to April), (b) in late spring and early summer 672  

(May to July), (c) in late summer and early fall (August to October), and (d) in fall to early 673  

winter (November to January). The seasonal mean distribution of mixed-layer depth in (e) 674  

February to April, and (f) May to July, is added for comparison. The black box indicates the 675  

convective region (CR) as defined in the paper. 676  

Fig. 4 Time series of absolute vertical velocity (in m/s) averaged between 150 and 2000 m depth 677  

and of mixed-layer depth (in m) defined using a density criterion of 0.008 kg m-3 over the CR 678  

    31  

from 1980 to 2009. (a) CLIMA, (b) VARY and (c) the difference (CLIMA-VARY). The dashed 679  

line in (b) indicates the chosen threshold separating strong and weak convections years. The 680  

time-series are constructed using 3-day averages. 681  

Fig. 5 Evolution of potential temperature (PT, in oC) in the convective region (CR) from 1980 to 682  

2009. (a) CLIMA, (b) VARY and (c) difference between the two integrations (CLIMA-VARY) 683  

for time series obtained averaging from 150 to 2000 m depth. The plot is obtained using 3-day 684  

averages. 685  

Fig. 6 Evolution of potential density (σθ in kg m-3 ) in the convective region (CR) from 1980 to 686  

2009. (a) CLIMA, (b) VARY. The plot is obtained using 3-day averages. 687  

Fig. 7 (a) Annual cycle of |w| averaged between 150 m and 2000 m depth over the convective 688  

region CR. Dashed lines for CLIMA, and solid lines for VARY. The shading indicates the 689  

standard deviation around VARY averages, and the vertical lines the time when |w| reaches its 690  

maximum in VARY. (b) Annual cycle of downward surface heat flux QMod and (c) surface wind 691  

intensity averaged over CR. 692  

Fig. 8 Relation between downward surface heat flux as seen by the model (QMod) and absolute 693  

value of vertical velocities averaged over the CR region between 150 and 2000 m during 694  

convective events. (a) CLIMA, December to April, DJFMA; (b) VARY, DJFMA; (c) CLIMA 695  

winter only, DJF; (d) VARY, DJF. Blue, red and black lines show fits using strong, weak, and all 696  

years, respectively. The two-digit numbers indicate the year. 697  

698  

    32  

Fig. A1 Potential temperature difference between VARY-HF and VARY averaged over the 699  

period January 2000 – December 2004 along the AR7W line. 700  

Fig. A2 Seasonal cycle of potential temperature anomalies (in oC) in the convective region (CR) 701  

in the top 500 m over the period January 2000 – December 2004. (a) VARY, (b) VARY-HF. 702  

Fig. A3 Evolution of potential temperature (PT, in oC) in the convective region (CR) from 2000 703  

to 2004. (a) VARY, (b) VARY-HF. The plot is obtained using 3-day averages of PT. 704  

705  

    33  

706  

(a) (b) 707  

Fig. 1 (a) Annual mean distribution of surface eddy kinetic energy cast as a speed !!"! =708  

2!"! (in color) and mean surface velocity (arrows). The locations of the AR7W WOCE 709  

hydrographic line (black line) and of the Ocean Weather Station BRAVO (red star) are marked. 710  

(b) Annual mean of potential temperature in the region comprised by the 3250 m isobaths 711  

following Yashayaev (2007). Potential temperature values are averaged over depths comprised 712  

between 150 m and 2000 m. 713  

!"!#

!"!#

!"!#

!"!#

!"!# !"!# !"!#

!"#$%!"#" ! !"#"

!"!!

!"!#

!"$!

!"$#

!"%!

!"%#

!"&!

!"&#

'())*+,*-

.,/!"#$ !"#$ !""# !""# !""" !""#

!"#

!"#

!"#

!"#!!

.0/

!"#$%#&'()*$+,$-'#.-$

!"#$%&'()(*+

!"!#

!"!!

!"!#

    34  

714  

Fig. 2 Seasonal variation in heat content (top) and salinity (bottom) in the central Labrador Sea 715  

for the top 200 m of the water column (left), and for a lower layer comprised between 200 and 716  

1300 m (right) in the model (blue and red lines) and in the observations presented in Straneo 717  

(2006) (black and gray lines). The model and the P-ALACE float data cover the period 1996-718  

2000. The gray lines represent the standard deviation (std) around the float data. The dashed 719  

lines indicate the std of the Ocean Weather Station Bravo data from 1964 to 1974 (also from 720  

Straneo, 2006). 721  

722  

!"#$!"#$%&$%! !"#$%

&'

%&'(#)" *+,"'%#-$!"#$%

!"# !$#

!%# !&#

    35  

723  

Fig. 3 Seasonal mean distribution of vertical velocity |w| averaged between 150 and 2000 m 724  

depth (a) at the peak of convection (February to April), (b) in late spring and early summer 725  

(May to July), (c) in late summer and early fall (August to October), and (d) in fall to early 726  

winter (November to January). The seasonal mean distribution of mixed-layer depth in (e) 727  

February to April, and (f) May to July, is added for comparison. The black box indicates the 728  

convective region (CR) as defined in the paper. 729  

730  

    36  

731  

Fig. 4 Time series of absolute vertical velocity (in m/s) averaged between 150 and 2000 m depth 732  

and of mixed-layer depth (in m) defined using a density criterion of 0.008 kg m-3 over the CR 733  

from 1980 to 2009. (a) CLIMA, (b) VARY and (c) the difference (CLIMA-VARY). The dashed 734  

line in (b) indicates the chosen threshold separating strong and weak convections years. The 735  

time-series are constructed using 3-day averages. 736  

737  

!" !# !$ !% !& !' !( !) !! !* *" *# *$ *% *& *' *( *) *! ** "" "# "$ "% "& "' "( ") "! "*"

"+""#

"+""$

"+""%

"+""&

!" !# !$ !% !& !' !( !) !! !* *" *# *$ *% *& *' *( *) *! ** "" "# "$ "% "& "' "( ") "! "* "

#"""

$"""

%"""

&"""

!" !# !$ !% !& !' !( !) !! !* *" *# *$ *% *& *' *( *) *! ** "" "# "$ "% "& "' "( ") "! "*"

"+""#

"+""$

"+""%

"+""&

!" !# !$ !% !& !' !( !) !! !* *" *# *$ *% *& *' *( *) *! ** "" "# "$ "% "& "' "( ") "! "* "

#"""

$"""

%"""

&"""!"#$%

&'(

&)(

&*(

+%,-

!"#$%./.+%,-

!"#$%&'()$*&+$,-.

!

!"#$%&'()$*&+$,-.

!

!"#$%&'()$*&+$,-.

!

!!'+

    37  

738  

Fig. 5 Evolution of potential temperature (PT, in oC) in the convective region (CR) from 1980 to 739  

2009. (a) CLIMA, (b) VARY and (c) difference between the two integrations (CLIMA-VARY) 740  

for time series obtained averaging from 150 to 2000 m depth. The plot is obtained using 3-day 741  

averages. 742  

743  

!"#

!$#

!%#

&

'&&

(&&&

('&&

)&&&

)'&&

&

'&&

(&&&

('&&

)&&&

)'&&

*+&

*+,

*+-

*+)

)+.

*+&

*+,

*+-

*+)

)+.

/0123

4356

/01237874356

    38  

744  

Fig. 6 Evolution of potential density (σθ in kg m-3 ) in the convective region (CR) from 1980 to 745  

2009. (a) CLIMA, (b) VARY. The plot is obtained using 3-day averages. 746  

747  

    39  

748  

Fig. 7 (a) Annual cycle of |w| averaged between 150 m and 2000 m depth over the convective 749  

region CR. Dashed lines for CLIMA, and solid lines for VARY. The shading indicates the 750  

standard deviation around VARY averages, and the vertical lines the time when |w| reaches its 751  

maximum in VARY. (b) Annual cycle of downward surface heat flux QMod and (c) surface wind 752  

intensity averaged over CR. 753  

754  

755  

Oct NovDec Jan Feb Mar Apr May Jun Jul AugSep0

0.05

0.1

0.15

0.2

0.25

0.3

strongweak

Month

Hea

t flu

xes

Win

d in

tens

ity

(W/m

2)

(N/m

2)

|w| (103 m/s)

0.0

0.5

1.0

1.5

2.0(a)

(b)

(c)

    40  

756  

Fig. 8 Relation between downward surface heat flux as seen by the model (QMod) and absolute 757  

value of vertical velocities averaged over the CR region between 150 and 2000 m during 758  

convective events. (a) CLIMA, December to April, DJFMA; (b) VARY, DJFMA; (c) CLIMA 759  

winter only, DJF; (d) VARY, DJF. Blue, red and black lines show fits using strong, weak, and all 760  

years, respectively. The two-digit numbers indicate the year. 761  

762  

(c)

ww

    41  

763  

Fig. A1 Potential temperature difference between VARY-HF and VARY averaged over the 764  

period January 2000 – December 2004 along the AR7W line. 765  

766  

767  

Fig. A2 Seasonal cycle of potential temperature anomalies (in oC) in the convective region (CR) 768  

in the top 500 m over the period January 2000 – December 2004. (a) VARY, (b) VARY-HF. 769  

770  

    42  

771  

Fig. A3 Evolution of potential temperature (PT, in oC) in the convective region (CR) from 2000 772  

to 2004. (a) VARY, (b) VARY-HF. The plot is obtained using 3-day averages of PT. 773  

!"#

$"%

$"#

&"%!"#

$"%

$"#

&"%&### &##' &##& &##$ &##!

#%##'###'%##&###&%##

#%##'###'%##&###&%## !"#

!$#