1. Introduction
Satellite measurements have shown evidence of a local response of the atmospheric boundary layer to oceanic mesoscale structures (ranging from tens to hundreds of kilometers). It takes the form of a positive correlation between wind stress and sea surface temperature (SST) anomalies at all latitudes (Xie 2004). Equivalent relationships exist with correlation of divergence of the wind stress with along-wind SST gradient, or wind stress curl and across-wind SST gradient (Chelton et al. 2001, 2004; O’Neill et al. 2003). It was also revealed through the signature of ocean eddies in turbulent air–sea fluxes of sensible and latent heat (Bourras et al. 2004), or in cloud cover and rain rates (Frenger et al. 2013).
The coupling between the atmosphere and narrow oceanic structures has been explored through various analyses of the horizontal momentum budget in the boundary layer based on theoretical models (Samelson et al. 2006; Schneider and Qiu 2015) or idealized simulations (Spall 2007; Kilpatrick et al. 2014, 2016). The general setting of these analyses was a large-scale wind blowing across (or along) an SST gradient, potentially leading to a change in the stability of the boundary layer. In locally unstable conditions (i.e., winds blowing from cold to warm waters), an increase of the downward transfer of momentum explains the correlation of wind or wind stress with SST anomalies (Wallace et al. 1989; Hayes et al. 1989). The mechanism of downward momentum mixing (DMM) was proposed to explain the relation between the divergence of wind stress and downwind SST gradients (e.g., Chelton et al. 2001; O’Neill et al. 2003).
Another mechanism that is considered in the literature is related to surface pressure variations induced by SST structures. It was initially proposed as an important source of coupling at tropical latitudes (Lindzen and Nigam 1987), and more recently as an important forcing for surface wind convergence over midlatitude SST fronts (Feliks et al. 2004; Minobe et al. 2008). The mechanism is based on a thermal adjustment of the boundary layer to the underlying SST, which creates local variations of the hydrostatic pressure. Through a mechanism in terms of Ekman balance mass adjustment (EBMA), the divergence of the surface wind correlates with the Laplacian of sea level pressure. The latter is itself very close to the Laplacian of the atmospheric temperature if the boundary layer has adjusted to the underlying SST, which is more likely for weak winds (Brachet et al. 2012; Lambaerts et al. 2013).
At midlatitudes, the importance of the pressure term compared to vertical mixing still remains unclear, largely depending on the spatial scales (Small et al. 2008) but also on the region of interest (Shimada and Minobe 2011) or on the season that is considered (Takatama et al. 2015). Moreover, the two mechanisms can be active together to force a surface divergence response. For instance, in the Kuroshio Extension region, Putrasahan et al. (2013) show that the divergence of wind stress correlates with downwind SST gradients (see their Fig. 4). At the same time, divergence of surface wind correlates with the Laplacian of SST (see their Fig. 7).
Most past studies have examined the time-average response (at least weekly averages) or the transient response (a few hours) of the atmospheric boundary layer to SST anomalies. As pointed out by Liu and Zhang (2013), O’Neill et al. (2017), or Plougonven et al. (2018), the responses differ when considering averaged or transients fields. Here, our goal is to determine the nature of the surface divergence response to mesoscale SST perturbations separating between classes of different large-scale wind conditions. For that purpose, we use an idealized simulation of an atmospheric storm track above a frontal SST zone including a variety of oceanic structures of horizontal scales from 40 to 400 km.
Section 2 presents the configuration of the model with a brief description of the simulated storm track. We then document in section 3 the surface divergence response at the oceanic eddy scale by a composite analysis, and we show that the simulations are consistent with observational results such as those of Frenger et al. (2013). Section 4 describes the spatial organization of the boundary layer response, investigating how the response mechanisms change for different synoptic wind configurations. Differences between the responses in wind divergence and wind stress divergence are also investigated. Section 5 summarizes the results of the previous sections and compares them with previous studies.
2. Model description
a. General configuration
The 3.6.1 version of the WRF Model (Skamarock et al. 2008) is used to simulate a characteristic midlatitude storm track above a prescribed SST field. The model integrates the nonhydrostatic compressible moist Euler equations. Microphysics is represented with the Kessler (1969) scheme, and convection with the Kain and Fritsch (1993) scheme. The model uses the Yonsei University (YSU) parameterization (Hong et al. 2006) for the atmospheric boundary layer in conjunction with a Monin–Obukhov parameterization for surface layers (MM5 scheme). We do not include the effect of ocean surface currents in the wind stress calculation although it is known to affect the atmospheric boundary layer above oceanic eddies (Renault et al. 2016; Takatama and Schneider 2017).
The Cartesian domain, periodic in the zonal direction x, is of size Lx × Ly = 9216 km × 9216 km. Horizontal resolution is set to 18 km, and 50 η levels are used for the hydrostatic pressure vertical coordinate, equally spaced in pressure. Top pressure is set to 36 hPa, corresponding to an altitude of approximately 20 km, and 13 levels are below 2 km of altitude. Free-slip boundary conditions are used at the poleward and equatorial walls of the domain, and
The model is forced by using a gray radiation scheme with an atmosphere transparent to water vapor and clouds, as proposed by Frierson et al. (2006). This forcing allows us to mimic simple relaxation forcings on dry variables (e.g., Held and Suarez 1994), but with the sole dependence on the SST field. The details of the radiative scheme are described in appendix B.
b. Oceanic forcing
Common parameters.


The eddying component
Figure 1 shows the total SST field and the corresponding SST anomalies. The maximum value of

(a) Total SST and (b) eddy SST fields (K). In (b), black crosses mark centers of the eddies used to create composites following the method described in section 3.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

(a) Total SST and (b) eddy SST fields (K). In (b), black crosses mark centers of the eddies used to create composites following the method described in section 3.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
(a) Total SST and (b) eddy SST fields (K). In (b), black crosses mark centers of the eddies used to create composites following the method described in section 3.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
c. Mean state of the troposphere
A first simulation using

Time and zonal average of zonal wind (blue thick contours, m s−1), potential temperature (black contours, K), and meridional flux of potential temperature (red thick contours, K m s−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Time and zonal average of zonal wind (blue thick contours, m s−1), potential temperature (black contours, K), and meridional flux of potential temperature (red thick contours, K m s−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Time and zonal average of zonal wind (blue thick contours, m s−1), potential temperature (black contours, K), and meridional flux of potential temperature (red thick contours, K m s−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
A typical storm track forms as a response to the large-scale forcing: a tropospheric jet is located around y = 6000 km with a maximum speed larger than 25 m s−1 around p = 250 hPa. The height of the tropopause changes from 200 hPa on the equatorial side of the domain down to 400 hPa on the poleward side (not shown). The eddy poleward heat flux is maximum in the free troposphere at the center of the domain between y = 4000 and 6000 km, while the eddy kinetic energy has its maximum slightly poleward at y = 5500 km (not shown). The simulated storm track is weaker than the Southern Hemisphere storm track for which the zonal jet reaches values of 35 m s−1 but has realistic features of midlatitudes baroclinic zones. A more detailed analysis of the response of the storm track to the oceanic mesoscale SST field is carried out in Foussard et al. (2019).
3. Composite analysis at the oceanic eddy scale
To assess the consistency of our idealized simulations with the observed relation between surface variables and SST anomalies, we first discuss the main features of the response of the atmospheric boundary layer to a typical mesoscale eddy. To that end, composites for cold and warm eddies are computed in the line of Park et al. (2006) or Frenger et al. (2013). For the sole purpose of identifying the position of the eddies, we use a method based on a wavelet packet decomposition [see details in Lapeyre and Klein (2006) and Doglioli et al. (2007)]. The procedure is to decompose SSTeddy in elementary wavelets of compact support (using the Haar basis). Then wavelet coefficients smaller than a given value are filtered out. The field that is recomposed with the remaining wavelets is such that it is zero at a given point if it does not belong to an eddy. This allows us to determine the precise location of each structure in order to compute the composites. The amplitude of an eddy is defined as the spatial average of the SST anomaly over the set of grid points within the eddy. The coordinates of its center are defined as their averaged values, and the eddy radius is defined as
For each eddy and each instantaneous snapshot, the large-scale background wind is defined as the average of the 10-m wind within a square box of width equal to 10 radii centered on each eddy. This yields a direction (used for the rotation of different quantities) and an amplitude (used to separate strong- and weak-wind conditions). For presentation of the composites, all fields, including sea surface temperature, are rotated so that the large-scale wind blows toward

Composites of surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Composites of surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Composites of surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
We now turn to the analysis of surface wind divergence. Rather than computing the mean divergence, we choose to separate the response depending on the large-scale wind speed. To that end, we have selected conditions with large-scale winds larger than 10 m s−1 (to be called strong-wind conditions) and smaller than 3 m s−1 (to be called weak-wind conditions). These categories correspond to 33% and 7% of instantaneous snapshots, respectively. In the following, we only consider the response to warm eddies as the results with cold eddies are qualitatively similar, but with an opposite sign (not shown). Finally, we have tested that changing the thresholds does not change qualitatively the results.
The divergence of the surface wind reveals significant differences between strong- and weak-wind conditions (Figs. 4a and 4d). Strong-wind conditions (Fig. 4a) are characterized by a dipolar spatial pattern with a divergent wind field upwind of the eddy and a convergent wind field downwind, with a typical amplitude on the order of 10−5 s−1. This is consistent with accelerated wind speeds over warm eddies and is similar to observations (e.g., Park et al. 2006; Ma et al. 2015). Note also that the downwind convergence is twice as large as the upwind divergence, which is generally not observed when doing averages over all weather conditions (e.g., Frenger et al. 2013). For weak-wind conditions (Fig. 4d), the situation is different as a strong monopolar convergence pattern is located slightly downwind of the warm eddy.

Composites above warm eddies of (a),(d) divergence of surface wind, (b),(e) downwind SST gradient, and (c),(f) Laplacian of boundary layer temperature under (a)–(c) strong-wind conditions (wind speeds greater than 10 m s−1) and (d)–(f) weak-wind conditions (wind speeds less than 3 m s−1). Contours correspond to SST (K).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Composites above warm eddies of (a),(d) divergence of surface wind, (b),(e) downwind SST gradient, and (c),(f) Laplacian of boundary layer temperature under (a)–(c) strong-wind conditions (wind speeds greater than 10 m s−1) and (d)–(f) weak-wind conditions (wind speeds less than 3 m s−1). Contours correspond to SST (K).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Composites above warm eddies of (a),(d) divergence of surface wind, (b),(e) downwind SST gradient, and (c),(f) Laplacian of boundary layer temperature under (a)–(c) strong-wind conditions (wind speeds greater than 10 m s−1) and (d)–(f) weak-wind conditions (wind speeds less than 3 m s−1). Contours correspond to SST (K).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
To determine the importance of the DMM and EBMA, the surface divergence was compared with the downwind SST gradient and the Laplacian of atmospheric temperature in the boundary layer. The downwind SST gradient
For each wind condition, the Laplacian of boundary layer temperature
For weak-wind conditions, the temperature Laplacian is monopolar and negative above the SST anomaly because of weak temperature advection by the wind (Fig. 4f). Comparing Figs. 4d–f, we see that the surface divergence pattern is highly correlated with the temperature Laplacian, while it is not the case when compared to the downwind SST gradient. Actually, because of the weak temperature advection, the SST Laplacian is correlated with the temperature Laplacian as well as with the surface divergence (not shown). This is in agreement with the results of Lambaerts et al. (2013), who examined the fast adjustment of the boundary layer from rest to a turbulent eddy SST field. A possible interpretation of this result can rely on the EBMA mechanism: the warm SST anomaly creates a warm temperature anomaly in the boundary layer, which then creates a convergence field in the Ekman layer.
A last remark concerns moderate-wind conditions (i.e., winds between 3 and 10 m s−1). In such conditions, it was found that the wind divergence response is between those for the two other wind conditions (not shown).
The difference in terms of the atmospheric response between weak- and strong-wind conditions is reminiscent of results obtained by Chen et al. (2017) for eddies in the Kuroshio Extension region. In their study, they separated two different classes, one with a dipolar pattern in divergence of surface wind (corresponding to 60% of the oceanic eddies that were observed) and one with a monopolar pattern (corresponding to 10% of the eddies). The first class was attributed to DMM while the second class to EBMA. An inspection of their Fig. 3c shows that, for the first class of eddies, the convergence maximum extends farther downstream, a result consistent with our result for strong-wind conditions (Fig. 4a).
4. Atmospheric response to a turbulent field of mesoscale eddies
The previous section has characterized the response of the wind field at the oceanic eddy scale in a simulation forced by the ocean. It showed that our simulation with fixed SST compares well with observations for strong-wind conditions. We now turn to examine the spatial organization of the atmospheric response in relation with the oceanic turbulent field, that is, for scales smaller than 400 km. This contrasts with studies focusing on eddy composites or unidimensional fronts. To this end, we focus on a part of the spatial domain, of width 1400 km × 1400 km and centered at
In the following, we consider anomalies from the large-scale environment. These turbulent-scale anomalies, denoted as
The anomaly of time-mean surface wind speed

Anomaly of (a) time-mean surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Anomaly of (a) time-mean surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Anomaly of (a) time-mean surface wind speed
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
The time-mean response in the surface winds generally reflects convergence above warm eddies (such as eddies B or D in Fig. 5b) and divergence above cold eddies (eddies A or C). The divergence field does not bear resemblance with the downwind SST gradients (Fig. 5c) while there is a high correlation with the Laplacian of temperature in the boundary layer (cf. shading in Figs. 5b and 5d). As discussed in the previous section, such a comparison is not helpful to reveal in which wind conditions EBMA or DMM are important. This is probably due to the fact that, in this region, the time-mean wind is weak (not shown).
We propose below to contrast conditions of strong and weak winds as well as different wind directions to better assess the role of the background wind and of the stability of the boundary layer. Several effects are anticipated: The wind speed will influence both how turbulent the boundary layer is and how much advection decorrelates boundary layer temperature from SST. The direction of the wind will also play a role through the presence of the large-scale meridional SST gradient. For example, northerly winds will advect cold air above warm waters, inducing a larger temperature difference between ocean and atmosphere, and hence a more turbulent boundary layer.
a. Method
Composite atmospheric fields depending on large-scale wind conditions are built through the following steps. We consider the square box of size 900 km × 900 km, centered at
b. Surface wind divergence
We now examine the differences of spatial structures in surface wind divergence for different wind conditions. Three large-scale wind conditions are considered in details: northerly strong winds
Figures 6a and 6b present the surface divergence anomaly
Correlation coefficients between different parameters related to wind divergence for northerly, weak, and southerly large-scale winds. Each quantity was computed over the domain displayed in Fig. 5.



Composites of surface divergence (shading, 10−5 s−1) for days with (a),(b) northerly winds, (c),(d) weak winds, and (e),(f) southerly winds. In (a), (c), and (e), contours correspond to the downwind SST gradient (10−5 K m−1). In (b), (d), and (f), contours correspond to the Laplacian of the atmospheric boundary layer temperature (10−10 K m−2).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Composites of surface divergence (shading, 10−5 s−1) for days with (a),(b) northerly winds, (c),(d) weak winds, and (e),(f) southerly winds. In (a), (c), and (e), contours correspond to the downwind SST gradient (10−5 K m−1). In (b), (d), and (f), contours correspond to the Laplacian of the atmospheric boundary layer temperature (10−10 K m−2).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Composites of surface divergence (shading, 10−5 s−1) for days with (a),(b) northerly winds, (c),(d) weak winds, and (e),(f) southerly winds. In (a), (c), and (e), contours correspond to the downwind SST gradient (10−5 K m−1). In (b), (d), and (f), contours correspond to the Laplacian of the atmospheric boundary layer temperature (10−10 K m−2).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
The weak-wind case is represented in Figs. 6c and 6d. The surface divergence is found to be generally weaker than for northerly winds (cf. Figs. 6a and 6c). Looking at Fig. 6d, surface divergence and temperature Laplacian are well correlated (with a correlation coefficient of 0.63). Also, there is a fair correspondence between SST Laplacian and surface divergence (correlation coefficient of 0.39), because the temperature anomalies lie almost above the SST anomalies (not shown). At particular locations (near eddy A, or in some filamentary structures in the northern part of the domain), the surface divergence resembles the downwind SST gradient (Fig. 6c). However, in many other places (such as eddies B, C, D), the two fields do not coincide with each other. We conclude that, in these weak-wind conditions, there is a preferential response following EBMA.
The situation is different for a southerly wind (Figs. 6e and 6f) for which we see a clear correlation of the surface divergence with the downwind SST gradient (correlation coefficient of 0.83; see Table 2). This manifests in similar spatial structures not only for eddies B, C, and D but also for several filamentary structures between them. The response above eddies B and C shows a typical dipolar structure of convergence–divergence corresponding to a DMM response. On the contrary, the connection between the surface divergence and the temperature Laplacian is less obvious when comparing the spatial structures of the two fields (Fig. 6f), although the spatial correlation is still high (around 0.48).

Correlation coefficient of surface wind divergence with (a) downwind SST gradient and (b) Laplacian of atmospheric boundary layer temperature, as a function of the large-scale background wind at 10 m. (c) Correlation coefficient of downwind SST gradient with Laplacian of temperature. White squares denote an insufficient number of snapshots for averaging over wind conditions.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Correlation coefficient of surface wind divergence with (a) downwind SST gradient and (b) Laplacian of atmospheric boundary layer temperature, as a function of the large-scale background wind at 10 m. (c) Correlation coefficient of downwind SST gradient with Laplacian of temperature. White squares denote an insufficient number of snapshots for averaging over wind conditions.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Correlation coefficient of surface wind divergence with (a) downwind SST gradient and (b) Laplacian of atmospheric boundary layer temperature, as a function of the large-scale background wind at 10 m. (c) Correlation coefficient of downwind SST gradient with Laplacian of temperature. White squares denote an insufficient number of snapshots for averaging over wind conditions.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
To understand why surface divergence correlates with downwind SST gradient in some situations, and with the temperature Laplacian in others, we examine the dependence on the wind conditions of the boundary layer height and the air–sea temperature difference, both spatially averaged over the domain of Fig. 5. The result is displayed in Fig. 8 and is significant in the sense than the mean change of both quantities between different wind conditions is larger than their change across the SST front (for a given wind condition). Northerly winds tend to be associated with high boundary layers (Fig. 8a) and an atmospheric temperature much colder than the underlying SST (Fig. 8b). This can be explained by the advection of cold air from the north, tending to decrease stability over the region that is examined. This results in a typical situation of strong turbulence in the boundary layer associated with a deep boundary layer. Southerly winds are associated with warm air advected in the region creating a stable boundary layer (Fig. 8b), which results in shallow boundary layers (Fig. 8a). These differences can explain the different response in terms of wind divergence, as the surface pressure anomaly and the surface divergence tend to be proportional to the height of the boundary layer (Feliks et al. 2004). Conditions with higher boundary layers will result in stronger EBMA. This can be confirmed by examining the coupling coefficient, computed as the regression coefficient between wind speed anomalies and SST anomalies as a function of the background wind

(a) Boundary layer height (m) and (b) difference between 0 and 500 m vertically averaged temperature θ and SST (K). Both quantities are averaged in space and plotted as a function of the large-scale background wind at 10 m. (c) Regression coefficient of wind speed and SST anomalies as a function of the background wind (m s−1 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

(a) Boundary layer height (m) and (b) difference between 0 and 500 m vertically averaged temperature θ and SST (K). Both quantities are averaged in space and plotted as a function of the large-scale background wind at 10 m. (c) Regression coefficient of wind speed and SST anomalies as a function of the background wind (m s−1 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
(a) Boundary layer height (m) and (b) difference between 0 and 500 m vertically averaged temperature θ and SST (K). Both quantities are averaged in space and plotted as a function of the large-scale background wind at 10 m. (c) Regression coefficient of wind speed and SST anomalies as a function of the background wind (m s−1 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Note that we repeated the analysis and compared the vertical velocity at 500 m

Composites of vertical velocity at 500 m (mm s−1) for weak winds. Contours correspond to
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Composites of vertical velocity at 500 m (mm s−1) for weak winds. Contours correspond to
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Composites of vertical velocity at 500 m (mm s−1) for weak winds. Contours correspond to
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
c. Wind stress divergence
Several observational studies (O’Neill et al. 2003; Chelton et al. 2004) pointed out a robust relation between wind stress divergence and downwind SST gradient. We now try to relate this result with the response of the surface divergence that we analyzed above.
Figure 10a presents the divergence of the wind stress for northerly winds

(a),(d),(g) Composites of wind stress divergence
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

(a),(d),(g) Composites of wind stress divergence
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
(a),(d),(g) Composites of wind stress divergence
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Correlation coefficients between different parameters related to wind stress divergence for northerly, weak, and southerly large-scale winds.


For weak-winds conditions
We now consider the case of southerly winds; that is,
RMS values of different parameters related to wind stress divergence for northerly, weak, and southerly large-scale winds.


More generally, for all wind conditions of

Correlation coefficient of wind stress divergence with (a) downwind SST gradient and (b) temperature Laplacian. (c) Regression coefficient of wind stress divergence and downwind SST gradient anomalies as a function of the large-scale background wind at 10 m (10−2 N m−2 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

Correlation coefficient of wind stress divergence with (a) downwind SST gradient and (b) temperature Laplacian. (c) Regression coefficient of wind stress divergence and downwind SST gradient anomalies as a function of the large-scale background wind at 10 m (10−2 N m−2 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Correlation coefficient of wind stress divergence with (a) downwind SST gradient and (b) temperature Laplacian. (c) Regression coefficient of wind stress divergence and downwind SST gradient anomalies as a function of the large-scale background wind at 10 m (10−2 N m−2 K−1).
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
Figure 11c shows the value of the regression coefficient between wind stress divergence and downwind SST gradient for all wind conditions. Values are of the typical range of those found in the literature (Perlin et al. 2014). The first dependence of the regression coefficient is on the wind speed, consistent with observations (e.g., O’Neill et al. 2012). It is modulated by the direction of the large-scale background wind relative to the large-scale front, in agreement with the coupling coefficient between wind speed and SST (Fig. 8c).
We conclude that the response in wind stress divergence to SST anomalies depends both on the magnitude of the mean surface wind and on the stability of the atmospheric boundary layer. In strong-wind conditions, we essentially find a wind stress divergence proportional to the downwind SST gradient. This is true for stable as well as for unstable boundary layers and is in agreement with the results of O’Neill et al. (2003) and Chelton et al. (2004). It can be understood as the addition of two effects in the wind stress. The first one comes from the variation of the drag coefficient and the surface wind speed due to SST [in relation with
5. Conclusions
In the present study, the response of surface winds to SST anomalies associated with oceanic eddies has been explored in an idealized simulation of an atmospheric storm track. Two mechanisms are generally invoked to explain the response in terms of divergence of surface wind and wind stress. A first one is related to pressure adjustment (EBMA mechanism; Lindzen and Nigam 1987) while a second is related to downward momentum mixing (DMM mechanism; Wallace et al. 1989). It is expected that the surface wind divergence resembles the Laplacian of the atmospheric temperature in the first case and the downwind SST gradient in the latter case. Our study has documented in which large-scale wind conditions one of the mechanisms is more active than the other. One advantage of our idealized simulation approach is that we could directly inspect the response in surface winds, contrary to other studies that considered equivalent neutral winds or wind stress. Also, using instantaneous winds averaged in composites (grouping together similar large-scale wind conditions) allows us to separate the rapid response without a temporal filter, in a manner similar to Byrne et al. (2015).
We first examined the response at the oceanic eddy scale through a composite analysis. It revealed that the surface wind divergence projects both onto the downwind SST gradient and onto the Laplacian of the atmospheric temperature in the boundary layer. For weak winds, the divergence of surface wind is proportional to the Laplacian of the boundary layer temperature. On the other hand, for strong winds, the surface divergence has a main pattern similar to the downwind SST gradient, but with a downstream extension (related to the temperature Laplacian spatial extension).
The atmospheric response was then investigated over a large region including a field of mesoscale oceanic eddies and filaments of scales between 40 and 400 km. The analysis revealed a more complex response that depends on the wind conditions, and more generally on the mean stability of the boundary layer. For large-scale unstable conditions or for weak winds, the divergence of the surface wind is correlated with the temperature Laplacian (corresponding to EBMA), while for large-scale stable conditions, it is correlated with downwind SST gradient (corresponding to DMM). For strong winds, the correlation of the surface divergence with the SST Laplacian is found to be small, because of the effect of the mean-wind advection.
Concerning the response in terms of wind stress divergence, a different picture is obtained. For strong winds, the divergence of wind stress is proportional to downwind SST gradient, even in large-scale unstable conditions. For weak winds, wind stress divergence is proportional to some extent to the temperature Laplacian. These results are valid at the scales of oceanic eddies, as well as smaller filamentary scales. This discussion shows that wind stress and surface wind divergences may behave differently considering their response to SST anomalies. We point out that such a distinction is rarely made in the literature and should be given greater consideration. Actually wind stress is directly related to the stability of the boundary layer while horizontal velocities in the atmosphere are less so but have a strong dependence on gradients of boundary layer temperature.
Several studies have examined the relevant parameters that set the atmospheric response sensitivity to DMM (Spall 2007; Small et al. 2008; Schneider and Qiu 2015; Ayet and Redelsperger 2019). The first one is related to the magnitude of the mean wind speed. Our study confirms that the relative importance of DMM increases with wind speed. This is shown by a better correlation of surface divergence with downwind SST gradient than with temperature Laplacian for strong winds, except in situations of winds blowing from cold to warm waters. A second important parameter is the spatial scale of the SST field (Small et al. 2008). One would expect the smaller the length scale, the larger the sensitivity to DMM. However, for strong winds blowing from cold to warm waters, we found that EBMA still dominates with surface divergence proportional to temperature Laplacian down to 40 km. The dominance of EBMA over DMM (in terms of relation between surface wind divergence and temperature Laplacian or downwind SST gradient) was found to be related to the large-scale stability of the boundary layer. For unstable and deep boundary layers, an EBMA response is found, while DMM prevails for large-scale stable conditions. This may be related to the dependence of the pressure Laplacian to the mean height of the boundary layer (which favors EBMA) and to the dependence of the coupling coefficient (between wind and SST anomalies) on the stability (which favors DMM).
As the focus of the paper concerns the boundary layer and surface dynamics, we did a sensitivity study to the boundary layer parameterization scheme, using the Mellor–Yamada–Nakanishi–Niino (MYNN) scheme (Nakanishi and Niino 2004). We obtained qualitatively similar results (see Fig. 12), but with different intensities in agreement with results of Lambaerts et al. (2013) and Perlin et al. (2014). A sensitivity to the number of vertical levels within the first 1000 m showed a weak dependence of the results on vertical resolution as well.

As in Figs. 7a and 7b, but for the simulation with the MYNN parameterization.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1

As in Figs. 7a and 7b, but for the simulation with the MYNN parameterization.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
As in Figs. 7a and 7b, but for the simulation with the MYNN parameterization.
Citation: Journal of the Atmospheric Sciences 76, 7; 10.1175/JAS-D-18-0204.1
The present study has different limitations. The first one is the stationarity in time of the oceanic anomalies especially for scales below 50 km. However, because of the fast variability of the atmosphere, conditions of given wind do not occur over long time scales compared to the SST variability of the ocean. Therefore, the composite analysis focuses only on the rapid response of the atmosphere and not on its time average, which is tightly linked to fixed SST. Another limitation is the fact that ocean–atmosphere coupling was not considered although different feedbacks are known to modify the surface wind response to oceanic mesoscale anomalies. In particular, our parameterization of the surface atmospheric layer does not take into account ocean currents modulation on the wind work (Renault et al. 2016; Moulin and Wirth 2016; Takatama and Schneider 2017). Moreover the air–sea coupling tend to damp oceanic eddies through Ekman pumping at the scale of oceanic eddies (Stern 1965; Dewar and Flierl 1987) as well at the scale of a turbulent eddy field (Oerder et al. 2018). A full air–sea coupling could reduce the SST amplitude and modulate the atmospheric response. These different mechanisms need to be taken into account in future studies.
Acknowledgments
This work was granted access to the HPC resources of IDRIS under the allocation A0020106852 from Grand Equipement National de Calcul Intensif (GENCI). The authors want to thank Arnaud Czaja, Caroline Muller, and Gwendal Rivière for useful discussion about this work and three anonymous reviewers for their time and numerous comments that significantly improved the clarity of the results and discussion.
APPENDIX A
Coriolis Parameter
APPENDIX B
Radiative Scheme
To ensure that the boundary layer response does not depend on radiative parameterization choices, two other sensitivity runs were done. In the first one,
REFERENCES
Ayet, A., and J.-L. Redelsperger, 2019: An analytical study of the atmospheric boundary layer flow and divergence over a SST front. Quart. J. Roy. Meteor. Soc., https://doi.org/10.1002/qj.3578, in press.
Bourras, D., G. Reverdin, H. Giordani, and G. Caniaux, 2004: Response of the atmospheric boundary layer to a mesoscale oceanic eddy in the northeast Atlantic. J. Geophys. Res., 109, D18114, https://doi.org/10.1029/2004JD004799.
Brachet, S., F. Codron, Y. Feliks, M. Ghil, H. Le Treut, and E. Simonnet, 2012: Atmospheric circulations induced by a midlatitude SST front: A GCM study. J. Climate, 25, 1847–1853, https://doi.org/10.1175/JCLI-D-11-00329.1.
Byrne, D., L. Papritz, I. Frenger, M. Münnich, and N. Gruber, 2015: Atmospheric response to mesoscale sea surface temperature anomalies: Assessment of mechanisms and coupling strength in a high-resolution coupled model over the South Atlantic. J. Atmos. Sci., 72, 1872–1890, https://doi.org/10.1175/JAS-D-14-0195.1.
Chelton, D. B., and Coauthors, 2001: Observations of coupling between surface wind stress and sea surface temperature in the eastern tropical Pacific. J. Climate, 14, 1479–1498, https://doi.org/10.1175/1520-0442(2001)014<1479:OOCBSW>2.0.CO;2.
Chelton, D. B., M. G. Schlax, M. H. Freilich, and R. F. Milliff, 2004: Satellite measurements reveal persistent small-scale features in ocean winds. Science, 303, 978–983, https://doi.org/10.1126/science.1091901.
Chen, L., Y. Jia, and Q. Liu, 2017: Oceanic eddy-driven atmospheric secondary circulation in the winter Kuroshio Extension region. J. Oceanogr., 73, 295–307, https://doi.org/10.1007/s10872-016-0403-z.
Dewar, W. K., and G. R. Flierl, 1987: Some effects of the wind on rings. J. Phys. Oceanogr., 17, 1653–1667, https://doi.org/10.1175/1520-0485(1987)017<1653:SEOTWO>2.0.CO;2.
Doglioli, A. M., B. Blanke, S. Speich, and G. Lapeyre, 2007: Tracking coherent structures in a regional ocean model with wavelet analysis: Application to Cape Basin eddies. J. Geophys. Res., 112, C05043, https://doi.org/10.1029/2006JC003952.
Feliks, Y., M. Ghil, and E. Simonnet, 2004: Low-frequency variability in the midlatitude atmosphere induced by an oceanic thermal front. J. Atmos. Sci., 61, 961–981, https://doi.org/10.1175/1520-0469(2004)061<0961:LVITMA>2.0.CO;2.
Foussard, A., G. Lapeyre, and R. Plougonven, 2019: Storm tracks response to oceanic eddies in idealized atmospheric simulations. J. Climate, 32, 445–463, https://doi.org/10.1175/JCLI-D-18-0415.1.
Frenger, I., N. Gruber, R. Knutti, and M. Münnich, 2013: Imprint of Southern Ocean eddies on winds, clouds and rainfall. Nat. Geosci., 6, 608–612, https://doi.org/10.1038/ngeo1863.
Frierson, D. M., I. M. Held, and P. Zurita-Gotor, 2006: A gray-radiation aquaplanet moist GCM. Part I: Static stability and eddy scale. J. Atmos. Sci., 63, 2548–2566, https://doi.org/10.1175/JAS3753.1.
Hayes, S., M. McPhaden, and J. Wallace, 1989: The influence of sea-surface temperature on surface wind in the eastern equatorial Pacific: Weekly to monthly variability. J. Climate, 2, 1500–1506, https://doi.org/10.1175/1520-0442(1989)002<1500:TIOSST>2.0.CO;2.
Held, I. M., and M. J. Suarez, 1994: A proposal for the intercomparison of the dynamical cores of atmospheric general circulation models. Bull. Amer. Meteor. Soc., 75, 1825–1830, https://doi.org/10.1175/1520-0477(1994)075<1825:APFTIO>2.0.CO;2.
Hong, S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an explicit treatment of entrainment processes. Mon. Wea. Rev., 134, 2318–2341, https://doi.org/10.1175/MWR3199.1.
Kain, J. S., and J. M. Fritsch, 1993: Convective parameterization for mesoscale models: The Kain-Fritsch scheme. The Representation of Cumulus Convection in Numerical Models, K. A. Emanuel and D. J. Raymond, Eds., Amer. Meteor. Soc., 165–170.
Kessler, E., 1969: On the Distribution and Continuity of Water Substance in Atmospheric Circulation. Meteor. Monogr., No. 32, Amer. Meteor. Soc., 84 pp.
Kilpatrick, T., N. Schneider, and B. Qiu, 2014: Boundary layer convergence induced by strong winds across a midlatitude SST front. J. Climate, 27, 1698–1718, https://doi.org/10.1175/JCLI-D-13-00101.1.
Kilpatrick, T., N. Schneider, and B. Qiu, 2016: Atmospheric response to a midlatitude SST front: Alongfront winds. J. Atmos. Sci., 73, 3489–3509, https://doi.org/10.1175/JAS-D-15-0312.1.
Lambaerts, J., G. Lapeyre, R. Plougonven, and P. Klein, 2013: Atmospheric response to sea surface temperature mesoscale structures. J. Geophys. Res. Atmos., 118, 9611–9621, https://doi.org/10.1002/jgrd.50769.
Lapeyre, G., 2017: Surface quasi-geostrophy. Fluids, 2, 7, https://doi.org/10.3390/fluids2010007.
Lapeyre, G., and P. Klein, 2006: Impact of the small-scale elongated filaments on the oceanic vertical pump. J. Mar. Res., 64, 835–851, https://doi.org/10.1357/002224006779698369.
Lindzen, R. S., and S. Nigam, 1987: On the role of sea surface temperature gradients in forcing low-level winds and convergence in the tropics. J. Atmos. Sci., 44, 2418–2436, https://doi.org/10.1175/1520-0469(1987)044<2418:OTROSS>2.0.CO;2.
Liu, J.-W., and S.-P. Zhang, 2013: Two types of surface wind response to the East China Sea Kuroshio front. J. Climate, 26, 8616–8627, https://doi.org/10.1175/JCLI-D-12-00092.1.
Ma, J., H. Xu, C. Dong, P. Lin, and Y. Liu, 2015: Atmospheric responses to oceanic eddies in the Kuroshio Extension region. J. Geophys. Res. Atmos., 120, 6313–6330, https://doi.org/10.1002/2014JD022930.
Minobe, S., A. Kuwano-Yoshida, N. Komori, S.-P. Xie, and R. J. Small, 2008: Influence of the Gulf Stream on the troposphere. Nature, 452, 206–209, https://doi.org/10.1038/nature06690.
Moulin, A., and A. Wirth, 2016: Momentum transfer between an atmospheric and an oceanic layer at the synoptic and the mesoscale: An idealized numerical study. Bound.-Layer Meteor., 160, 551–568, https://doi.org/10.1007/s10546-016-0153-x.
Nakanishi, M., and H. Niino, 2004: An improved Mellor–Yamada level-3 model with condensation physics: Its design and verification. Bound.-Layer Meteor., 112, 1–31, https://doi.org/10.1023/B:BOUN.0000020164.04146.98.
O’Neill, L. W., D. B. Chelton, and S. K. Esbensen, 2003: Observations of SST-induced perturbations of the wind stress field over the Southern Ocean on seasonal timescales. J. Climate, 16, 2340–2354, https://doi.org/10.1175/2780.1.
O’Neill, L. W., D. B. Chelton, and S. K. Esbensen, 2012: Covariability of surface wind and stress responses to sea surface temperature fronts. J. Climate, 25, 5916–5942, https://doi.org/10.1175/JCLI-D-11-00230.1.
O’Neill, L. W., T. Haack, D. B. Chelton, and E. Skyllingstad, 2017: The Gulf Stream convergence zone in the time-mean winds. J. Atmos. Sci., 74, 2383–2412, https://doi.org/10.1175/JAS-D-16-0213.1.
Oerder, V., F. Colas, V. Echevin, S. Masson, and F. Lemarié, 2018: Impacts of the mesoscale ocean-atmosphere coupling on the Peru-Chile ocean dynamics: The current-induced wind stress modulation. J. Geophys. Res. Oceans, 123, 812–833, https://doi.org/10.1002/2017JC013294.
Park, K., P. Cornillon, and D. L. Codiga, 2006: Modification of surface winds near ocean fronts: Effects of Gulf Stream rings on scatterometer (QuikSCAT, NSCAT) wind observations. J. Geophys. Res., 111, C03021, https://doi.org/10.1029/2005JC003090.
Perlin, N., S. P. De Szoeke, D. B. Chelton, R. M. Samelson, E. D. Skyllingstad, and L. W. O’Neill, 2014: Modeling the atmospheric boundary layer wind response to mesoscale sea surface temperature perturbations. Mon. Wea. Rev., 142, 4284–4307, https://doi.org/10.1175/MWR-D-13-00332.1.
Plougonven, R., A. Foussard, and G. Lapeyre, 2018: Comments on “The Gulf Stream convergence zone in the time-mean winds.” J. Atmos. Sci., 75, 2139–2149, https://doi.org/10.1175/JAS-D-17-0369.1.
Putrasahan, D. A., A. J. Miller, and H. Seo, 2013: Isolating mesoscale coupled ocean–atmosphere interactions in the Kuroshio Extension region. Dyn. Atmos. Oceans, 63, 60–78, https://doi.org/10.1016/j.dynatmoce.2013.04.001.
Renault, L., M. J. Molemaker, J. C. McWilliams, A. F. Shchepetkin, F. Lemarié, D. Chelton, S. Illig, and A. Hall, 2016: Modulation of wind work by oceanic current interaction with the atmosphere. J. Phys. Oceanogr., 46, 1685–1704, https://doi.org/10.1175/JPO-D-15-0232.1.
Samelson, R., E. Skyllingstad, D. Chelton, S. Esbensen, L. O’Neill, and N. Thum, 2006: On the coupling of wind stress and sea surface temperature. J. Climate, 19, 1557–1566, https://doi.org/10.1175/JCLI3682.1.
Schneider, N., and B. Qiu, 2015: The atmospheric response to weak sea surface temperature fronts. J. Atmos. Sci., 72, 3356–3377, https://doi.org/10.1175/JAS-D-14-0212.1.
Shimada, T., and S. Minobe, 2011: Global analysis of the pressure adjustment mechanism over sea surface temperature fronts using AIRS/Aqua data. Geophys. Res. Lett., 38, L06704, https://doi.org/10.1029/2010GL046625.
Skamarock, W., and Coauthors, 2008: A description of the Advanced Research WRF version 3. NCAR Tech. Note NCAR/TN-475+STR, 113 pp., https://doi.org/10.5065/D68S4MVH.
Skyllingstad, E. D., D. Vickers, L. Mahrt, and R. Samelson, 2007: Effects of mesoscale sea-surface temperature fronts on the marine atmospheric boundary layer. Bound.-Layer Meteor., 123, 219–237, https://doi.org/10.1007/s10546-006-9127-8.
Small, R., and Coauthors, 2008: Air–sea interaction over ocean fronts and eddies. Dyn. Atmos. Oceans, 45, 274–319, https://doi.org/10.1016/j.dynatmoce.2008.01.001.
Song, Q., D. B. Chelton, S. K. Esbensen, N. Thum, and L. W. O’Neill, 2009: Coupling between sea surface temperature and low-level winds in mesoscale numerical models. J. Climate, 22, 146–164, https://doi.org/10.1175/2008JCLI2488.1.
Spall, M. A., 2007: Midlatitude wind stress–sea surface temperature coupling in the vicinity of oceanic fronts. J. Climate, 20, 3785–3801, https://doi.org/10.1175/JCLI4234.1.
Stern, M. E., 1965: Interaction of a uniform wind stress with a geostrophic vortex. Deep-Sea Res. Oceanogr. Abstr., 12, 355–367, https://doi.org/10.1016/0011-7471(65)90007-0.
Stull, R. B., 1989: An Introduction to Boundary Layer Meteorology. Kluwer Academic, 666 pp.
Takatama, K., and N. Schneider, 2017: The role of back pressure in the atmospheric response to surface stress induced by the Kuroshio. J. Atmos. Sci., 74, 597–615, https://doi.org/10.1175/JAS-D-16-0149.1.
Takatama, K., S. Minobe, M. Inatsu, and R. J. Small, 2015: Diagnostics for near-surface wind response to the Gulf Stream in a regional atmospheric model. J. Climate, 28, 238–255, https://doi.org/10.1175/JCLI-D-13-00668.1.
Wallace, J. M., T. Mitchell, and C. Deser, 1989: The influence of sea-surface temperature on surface wind in the eastern equatorial Pacific: Seasonal and interannual variability. J. Climate, 2, 1492–1499, https://doi.org/10.1175/1520-0442(1989)002<1492:TIOSST>2.0.CO;2.
Xie, S.-P., 2004: Satellite observations of cool ocean–atmosphere interaction. Bull. Amer. Meteor. Soc., 85, 195–208, https://doi.org/10.1175/BAMS-85-2-195.