1. Introduction
Feedbacks involving wind and sea surface temperature (SST), including the wind–evaporation–SST (WES) feedback, are important to the maintenance of aspects of the mean tropical climate such as the northward-biased intertropical convergence zone (ITCZ) and other equatorial asymmetries (Emanuel 1987; Neelin et al. 1987; Xie and Philander 1994; Maloney and Sobel 2004; Xie 2004). The WES feedback has also been argued to be an important process in modes of internal climate variability including the Madden–Julian oscillation (MJO; DeMott et al. 2016), the Indian Ocean dipole mode (Li et al. 2003), meridional modes (Chiang and Vimont 2004), El Niño–Southern Oscillation (ENSO; Wang et al. 1999; Zhu et al. 2016), decadal variability (Chang et al. 1997; Xie and Tanimoto 1998), and tropical–extratropical interactions (Stuecker 2018; Amaya 2019), and in shaping the response of the mean climate to radiative forcing (Xie et al. 2010).
As commonly posed (e.g., Xie and Philander 1994), the WES feedback is positive (Fig. 1a); a positive meridional SST gradient induces a negative meridional sea level pressure (SLP) gradient, which induces a northward surface wind that is deflected differentially along the SLP gradient due to the dependence of the Coriolis parameter on latitude. The subsequent zonal wind anomalies superimpose differentially upon the mean prevailing winds, reducing (enhancing) latent heat flux over the warm (cold) SST anomaly, thereby amplifying the initial anomalous SST gradient. This mechanism depends critically on the surface wind response to the SST anomalies being governed by the hydrostatic adjustment of SLP in a well-mixed boundary layer (Lindzen and Nigam 1987; Back and Bretherton 2009); that is, warm (cold) SST anomalies force low (high) SLP anomalies, and the resulting gradient therein propels the low-level wind.
Schematic illustrations of the WES feedback in which the surface wind response to an anomalous meridional SST gradient is governed by (a) the LN87 mechanism (hydrostatic pressure, i.e., the common WES feedback) and (b) the WH89 mechanism (vertical momentum mixing). In both diagrams, time progresses from left to right following steps A through F (see main text). Note that the final state (step F) in (a) is an amplified version of the initial state (step A), thus representing a positive feedback, and vice versa for (b).
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
Omitted from this formulation of the WES feedback is an alternative mechanism by which the surface wind field can respond to SST anomalies. Specifically, the potential for surface wind speed to be modulated by SST through vertical stratification and momentum mixing as proposed by Wallace et al. (1989) and Hayes et al. (1989) is not considered in the common WES feedback. Such a mechanism alone would also give rise to a feedback involving wind, evaporation, and SST—but a negative one (Fig. 1b). In that case, the warm (cold) SST anomaly would destabilize (stabilize) the atmospheric boundary layer by steepening (relaxing) its temperature lapse rate and therefore increase (decrease) the rate of vertical mixing of horizontal momentum. The result would be an increase (decrease) in surface wind speed and latent heat flux over the warm (cold) SST anomaly, thereby damping the initial anomalous meridional SST gradient.
Recognizing that the Lindzen and Nigam (1987) argument (hereafter the LN87 mechanism) and the Wallace et al. (1989) and Hayes et al. (1989) mechanism (hereafter the WH89 mechanism) need not be mutually exclusive, this study seeks to develop a simple yet useful model of the WES feedback that accounts for both. The intent is not to declare a winner, but to explore the degree to which the WES feedback as commonly posed may be improved upon when also considering the WH89 mechanism. The model is developed in the following section, with the results of some idealized experiments presented in section 3. The dependence of the magnitude and latitudinal structure of the WES feedback on low-frequency variations in mean climate is also explored, before offering some concluding remarks in section 4.
2. Model
The following section presents several experiments conducted with this model. The model equations (8)–(11) are integrated with a one-day time step for two locations (denoted by subscripts N for north and S for south) separated by 10° latitude and initialized with SST anomalies
3. Results
a. Solutions with a constant and uniform background state
The first experiments to examine are idealized ones in which only the LN87 mechanism is enabled, only the WH89 mechanism is enabled, and both mechanisms are enabled. Each experiment is initialized with an anomalous meridional SST gradient of 1 K over 10° latitude antisymmetric about the equator (as described in section 2), the parameter d is set to 0.25 K−1, and the prevailing zonal wind
Solutions to the WES model in which (a) the LN87 mechanism is enabled, (b) the WH89 mechanism is enabled, and (c) both the LN87 and WH89 mechanisms are enabled. Shown for each solution is the SST anomaly (T′; K), sea level pressure anomaly (P′; mb; 1 mb = 1 hPa), zonal wind anomaly (U′; m s−1), and latent heat flux anomaly (Q′; W m−2) for the northern box (thin solid line), southern box (dashed line), and their difference (ΔT′; heavy solid line).
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
Solutions to the WES model for SST anomaly (T′; K) in (a) the northern box, (b) the southern box, and (c) their difference (ΔT′; K), in which the LN87 mechanism is enabled (thin solid line), the WH89 mechanism is enabled (dashed line), and both the LN87 and WH89 mechanisms are enabled (heavy solid line).
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
The behavior of the model in which both the LN87 and WH89 mechanisms are enabled (Fig. 2c) is quite intriguing over course of the first year of integration. The value of ΔT′ is roughly quadrupled at 4 months and is then reduced to 1% of its initial value by the end of one year. Comparing the results of this experiment (Fig. 2c) with those of the LN87-only experiment (Fig. 2a) suggests that enabling the WH89 mechanism facilitates the eventual damping of an otherwise obviously unrealistic (ceaseless) growth of SST anomalies—strictly through the dynamics of the coupled system (i.e., without requiring a Newtonian damping term). In reality, of course, there is a thermal damping of this coupled system through adjustment of surface heat flux (Frankignoul and Kestenare 2002; Myers and Mechoso 2020) including variations in the air–sea specific humidity difference Δq (which has thus far been held constant). If the model is rephrased such that Δq is allowed to adjust during model integration (Fig. 4), a steady-state solution for ΔT′ is reached by about day 120 (Fig. 4c), and that final anomaly ΔT′ is about 11% smaller in the model that includes the WH89 mechanism in addition to LN87.
(a),(b) As in the left columns of Figs. 2a and 2c, i.e., combining SST and zonal wind anomaly results from the LN87-only model (red lines) and the model with both LN87 and WH89 mechanisms enabled (black lines). (c),(d) As in (a) and (b), but for a version of the model with interactive air–sea specific humidity difference Δq. In the latter version, it is necessary to specify additional parameters defining the background state including a mean SST (26°C) and sea level pressure (1013.25 mb) (to calculate the saturation specific humidity at the sea surface), and the mean near-surface specific humidity (0.016 kg kg−1); all other model parameters and initial conditions remain unchanged.
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
The sensitivity of this model solution to varying all of the parameters and initial conditions by 20% reveals some dependencies to be expected based on climate physics (e.g., increasing the mixed layer depth h slows SST change without changing the amplitude of the feedback), but a general insensitivity of the overall character of the solution as described above (Fig. 5). An additional sensitivity of the model solution with the LN87 mechanism enabled is on the meridional distance between the two points at which
Solutions to the WES model (for ΔT′) in which both the LN87 and WH89 mechanisms are enabled (i.e., as in Fig. 2c), but with each model parameter and initial condition changed by 20%. The thick gray line is for nominal values as indicated in the main text. The solid red line is for the prevailing easterlies
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
(a) Solutions to the WES model for anomalous meridional SST gradient (ΔT′; K) at day 60 as a function of central latitude with uniform prevailing easterlies (
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
Variation of the Coriolis parameter f is the root cause of the latitudinally dependent results shown in Fig. 6; all other model parameters are equal at different latitudes, and the simulations are run separately at each latitude using the same initial SST anomalies T′ including their meridional spacing Δy. Since the Coriolis parameter f (and its meridional gradient ∂/∂φ) is a nonlinear function of latitude φ, and ∂f/∂φ is maximum in the tropics, the extent to which the LN87-driven positive WES feedback in the presence of an initial positive ΔT′ (i.e., ∂T′/∂φ > 0) is strongly a function of latitude, which introduces latitudinally varying competition to the WH89 mechanism that would attempt to damp it. These latitudinally dependent evolving SST anomalies then modulate the efficiency of the WH89 mechanism by way of T′ in the second term of Eq. (3) (
As a means to avoid the singularity at ±5° evident in Fig. 6a, and thus shed further light on the latitudinal structure of the WES feedback, a Rayleigh damping term may be added to Eq. (3) [or term c as defined in Eq. (7)]. In that case, 1/f is simply replaced by
b. The role of variations in the background state
The remainder of this paper examines the potential influence of variations in background climate at seasonal, decadal, and centennial time scales on the magnitude and latitudinal structure of the WES feedback, particularly with both the LN87 and WH89 mechanisms enabled (d = 0.25 K−1). All subsequent experiments do not invoke Rayleigh damping, although it was checked that the overall conclusions of this paper (including retroactively; i.e., the solutions shown in Figs. 2 and 3) do not depend critically on whether term c as defined in Eq. (7) is treated with Rayleigh damping. For these experiments, realistic profiles of
In the idealized experiments described above,
(a) Profiles of global zonal mean zonal surface wind (m s−1) from the NCEP–NCAR reanalysis (averaged 1948–2019) for boreal winter (DJF; blue line) and boreal summer (JJA; red line). (b) Solutions to the WES model for anomalous meridional SST gradient (ΔT′; K) at day 60 as a function of central latitude, with the prevailing zonal wind
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
The bulk of the seasonal results are fundamentally driven by the LN87 mechanism (i.e., the WES feedback as commonly posed) and are not qualitatively sensitive to whether the WH89 mechanism is enabled in the WES model. However, the introduction of the WH89 mechanism to the WES model does render the WES feedback near the equator slightly negative during August and September (Fig. 7c), whereas it is positive all year with only the LN87 mechanism enabled (Fig. 7d), so that the WH89 mechanism serves to amplify the annual cycle of the strength of the WES feedback by ∼0.25 K in ΔT′ on the equator at day 60 (Fig. 7e). The prevailing zonal wind
Changes to the prevailing zonal wind
(a) Profiles of zonal mean (eastern Pacific, 135°–125°W) zonal surface wind (m s−1) from the NCEP–NCAR Reanalysis (1948–2019) for boreal winters with PDO > 0.5 (heavy line) and PDO < −0.5 (thin line). (b) Solutions to the WES model for anomalous meridional SST gradient (ΔT′; K) at day 60 as a function of central latitude, with the prevailing zonal wind
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
Finally, the potential change in the magnitude and structure of the WES feedback due to anthropogenic radiative forcing can be examined using the projected change in
(a) Profiles of zonal mean (eastern Pacific, 135°–125°W) zonal surface wind (m s−1) from the NCEP–NCAR reanalysis (averaged 1948–2019) for boreal winter (thin line), and plus the centennial trend predicted by one CMIP6 climate model (heavy line). (b) Solutions to the WES model for anomalous meridional SST gradient (ΔT′; K) at day 60 as a function of central latitude, with the prevailing zonal wind
Citation: Journal of Climate 35, 7; 10.1175/JCLI-D-20-0895.1
4. Summary and discussion
In light of the multiple and generally competing mechanisms by which SST anomalies can alter surface winds and hence guide the wind–evaporation–SST feedback, a simple coupled model was developed that enables exploration of such processes simultaneously. The atmospheric component of the model developed here accounts for both the hydrostatic adjustment of surface pressure to SST anomalies and the influence of SST anomalies on the stability of the boundary layer and hence vertical mixing of horizontal momentum from the free troposphere. Inclusion of the latter effect prevents the WES feedback from growing unstable beyond a few months, and damps its magnitude on the equator by order 10% for reasonable model parameters and cases [including with or without allowing for the (negative) heat flux feedback via adjustments in Δq]. This is not the first time the aperture on the WES feedback has been widened. A recent study, for example, uncovered an important role for a variable ocean mixed layer depth in shaping such feedbacks (Kataoka et al. 2019), and the WH89 mechanism could have interesting implications in that context as well.
Several other studies have employed a similar formulation of the ocean aspect of the coupled system to investigate the WES feedback (i.e., a slab ocean mixed layer with bulk formula for latent heat flux), but with substantially greater complexity in the atmosphere (e.g., Gill–Matsuno type), which of course leads to atmospheric Kelvin and Rossby wave propagation and consideration of horizontal mass convergence (e.g., Vimont 2010; Martinez-Villalobos and Vimont 2017)—processes that may well be necessary to fully resolve all of the implications of the WH89 mechanism for the WES feedback. Interestingly, the results of Takatama et al. (2012) suggest that the two mechanisms considered here (LN87 and WH89) may play different roles in terms of rendering surface convergence versus divergence—at least in the context of narrow frontal zones such as along western boundary currents. Additionally, the suitability of the WH89 mechanism to explain the correlation between SST and surface wind speed (positive in this model, by construction) may depend on proximity to, and the wind’s orientation relative to, the frontal transition between the warm and cold SST anomalies (Samelson et al. 2006; Spall 2007).
Hemispheric asymmetries in the strength of the WES feedback predicted by the simple model, even under uniform zonal winds, are consistent with previous studies that show that WES is more active in the Northern Hemisphere for the Atlantic meridional mode (Chang et al. 2000; Amaya et al. 2017). Model solutions for realistic cases, constrained by observed profiles of prevailing zonal wind, reveal the leading influence of seasonality on the strength and latitudinal structure of the WES feedback such that it is predicted to be much stronger near the equator during boreal winter than boreal summer, but active over a wider range of latitudes during boreal summer. Inclusion of the stability mechanism serves to amplify the annual cycle of the strength of the WES feedback by about 10%. Such seasonality of the efficiency of the WES feedback near the equator may contribute to its potential to play a role in ENSO development, and in particular ENSO’s observed phase locking to the annual cycle as previously argued by Wang et al. (1999).
This framework may be applied to lower-frequency variations in the background climate, including those arising due to anthropogenic forcing. It is important to be mindful that this is not a rigorous estimate of the forced change—nor is it intended to be. It is a demonstration using a single run by one climate model for a particular forcing scenario, and some nontrivial low-frequency natural variability may contaminate the trend. Liguori and Di Lorenzo (2018) analyzed output from a large ensemble of full complexity coupled GCM simulations driven by similar radiative forcing and calculated the so-called WES parameter to estimate the strength of the WES feedback post hoc using the model output. They concluded that the WES feedback in the PMM region (central northern tropical Pacific) exhibits an exponential increase in amplitude because of the nonlinear relationship between SST and evaporation in a warming mean climate. The simple model, when only driven by projected changes in zonal wind, predicts a more modest (∼3%) enhancement of the WES feedback in the east Pacific on the equator. North of the equator, however, the increase in WES feedback strength is quite stronger than this locally due to a northward shift in the latitude band in which WES is positive and effective. Despite very different models and with a large number of other caveats, one may conclude that these results are not inconsistent. The overarching conclusion is that temporal variability in the background climate state—forced and unforced alike—can alter the strength and structure of the WES feedback [as also shown by Okajima et al. (2003) by altering land distribution in a full-physics coupled model], and these insights may facilitate interpretation of coupled climate behavior in observations and more complex models.
Acknowledgments.
The author acknowledges support from NOAA Climate Program Office, Climate Variability and Predictability (CVP) Program (NA18OAR4310406) and a helpful discussion with Dr. Lei Zhang. The author also thanks three anonymous reviewers for insightful suggestions, and the editor for patience.
REFERENCES
Amaya, D. J., 2019: The Pacific meridional mode and ENSO: A review. Curr. Climate Change Rep., 5, 296–307, https://doi.org/10.1007/s40641-019-00142-x.
Amaya, D. J., M. J. DeFlorio, A. J. Miller, and S.-P. Xie, 2017: WES feedback and the Atlantic meridional mode: Observations and CMIP5 comparisons. Climate Dyn., 49, 1665–1679, https://doi.org/10.1007/s00382-016-3411-1.
Back, L. E., and C. S. Bretherton, 2009: On the relationship between SST gradients, boundary layer winds, and convergence over the tropical oceans. J. Climate, 22, 4182–4196, https://doi.org/10.1175/2009JCLI2392.1.
Boucher, O., and Coauthors, 2020: Presentation and evaluation of the IPSL‐CM6A‐LR climate model. J. Adv. Model. Earth Syst., 12, e2019MS002010, https://doi.org/10.1029/2019MS002010.
Chang, P., L. Ji, and H. Li, 1997: A decadal climate variation in the tropical Atlantic Ocean from thermodynamic air–sea interactions. Nature, 385, 516–518, https://doi.org/10.1038/385516a0.
Chang, P., R. Saravanan, L. Ji, and G. C. Hegerl, 2000: The effect of local sea surface temperatures on atmospheric circulation over the tropical Atlantic sector. J. Climate, 13, 2195–2216, https://doi.org/10.1175/1520-0442(2000)013<2195:TEOLSS>2.0.CO;2.
Chiang, J. C. H., and D. J. Vimont, 2004: Analogous Pacific and Atlantic meridional modes of tropical atmosphere–ocean variability. J. Climate, 17, 4143–4158, https://doi.org/10.1175/JCLI4953.1.
de Boyer Montégut, C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone, 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile‐based climatology. J. Geophys. Res., 109, C12003, https://doi.org/10.1029/2004JC002378.
DeMott, C. A., J. J. Benedict, N. P. Klingaman, S. J. Woolnough, and D. A. Randall, 2016: Diagnosing ocean feedbacks to the MJO: SST‐modulated surface fluxes and the moist static energy budget. J. Geophys. Res. Atmos., 121, 8350–8373, https://doi.org/10.1002/2016JD025098.
Emanuel, K. A., 1987: An air–sea interaction model of intraseasonal oscillations in the tropics. J. Atmos. Sci., 44, 2324–2340, https://doi.org/10.1175/1520-0469(1987)044<2324:AASIMO>2.0.CO;2.
Eyring, V., S. Bony, G. A. Meehl, C. A. Senior, B. Stevens, R. J. Stouffer, and K. E. Taylor, 2016: Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev., 9, 1937–1958, https://doi.org/10.5194/gmd-9-1937-2016.
Frankignoul, C., and E. Kestenare, 2002: The surface heat flux feedback. Part I: Estimates from observations in the Atlantic and the North Pacific. Climate Dyn., 19, 633–647, https://doi.org/10.1007/s00382-002-0252-x.
Grise, K. M., and Coauthors, 2019: Recent tropical expansion: Natural variability or forced response? J. Climate, 32, 1551–1571, https://doi.org/10.1175/JCLI-D-18-0444.1.
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.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437–471, https://doi.org/10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2.
Kataoka, T., M. Kimoto, M. Watanabe, and H. Tatebe, 2019: Wind–mixed layer–SST feedbacks in a tropical air–sea coupled system: Application to the Atlantic. J. Climate, 32, 3865–3881, https://doi.org/10.1175/JCLI-D-18-0728.1.
Li, T., B. Wang, C. Chang, and Y. Zhang, 2003: A theory for the Indian Ocean dipole–zonal mode. J. Atmos. Sci., 60, 2119–2135, https://doi.org/10.1175/1520-0469(2003)060<2119:ATFTIO>2.0.CO;2.
Liguori, G., and E. Di Lorenzo, 2018: Meridional modes and increasing Pacific decadal variability under anthropogenic forcing. Geophys. Res. Lett., 45, 983–991, https://doi.org/10.1002/2017GL076548.
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, Z., and S.-P. Xie, 1994: Equatorward propagation of coupled air–sea disturbances with application to the annual cycle of the eastern tropical Pacific. J. Atmos. Sci., 51, 3807–3822, https://doi.org/10.1175/1520-0469(1994)051<3807:EPOCAD>2.0.CO;2.
Maloney, E. D., and A. H. Sobel, 2004: Surface fluxes and ocean coupling in the tropical intraseasonal oscillation. J. Climate, 17, 4368–4386, https://doi.org/10.1175/JCLI-3212.1.
Mantua, N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis, 1997: A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Amer. Meteor. Soc., 78, 1069–1079, https://doi.org/10.1175/1520-0477(1997)078<1069:APICOW>2.0.CO;2.
Martinez-Villalobos, C., and D. J. Vimont, 2016: The role of the mean state in meridional mode structure and growth. J. Climate, 29, 3907–3921, https://doi.org/10.1175/JCLI-D-15-0542.1.
Martinez-Villalobos, C., and D. J. Vimont, 2017: An analytical framework for understanding tropical meridional modes. J. Climate, 30, 3303–3323, https://doi.org/10.1175/JCLI-D-16-0450.1.
Myers, T. A., and C. R. Mechoso, 2020: Relative contributions of atmospheric, oceanic, and coupled processes to North Pacific and North Atlantic variability. Geophys. Res. Lett., 47, e2019GL086321, https://doi.org/10.1029/2019GL086321.
Neelin, J. D., I. M. Held, and K. H. Cook, 1987: Evaporation–wind feedback and low‐frequency variability in the tropical atmosphere. J. Atmos. Sci., 44, 2341–2348, https://doi.org/10.1175/1520-0469(1987)044<2341:EWFALF>2.0.CO;2.
Newman, M., and Coauthors, 2016: The Pacific decadal oscillation, revisited. J. Climate, 29, 4399–4427, https://doi.org/10.1175/JCLI-D-15-0508.1.
Okajima, H., S.-P. Xie, and A. Numaguti, 2003: Interhemispheric coherence of tropical climate variability: Effect of the climatological ITCZ. J. Meteor. Soc. Japan, 81, 1371–1386, https://doi.org/10.2151/jmsj.81.1371.
Samelson, R. M., E. D. Skyllingstad, D. B. Chelton, S. K. Esbensen, L. W. 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.
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.
Staten, P. W., K. M. Grise, S. M. Davis, N. A. Davis, and K. B. Karnauskas, 2019: Regional widening of tropical overturning–forced change, natural variability, and recent trends. J. Geophys. Res. Atmos., 124, 6104–6119, https://doi.org/10.1029/2018JD030100.
Stuecker, M. F., 2018: Revisiting the Pacific meridional mode. Sci. Rep., 8, 3216, https://doi.org/10.1038/s41598-018-21537-0.
Takatama, K., S. Minobe, M. Inatsu, and R. J. Small, 2012: Diagnostics for near-surface wind convergence/divergence response to the Gulf Stream in a regional atmospheric model. Atmos. Sci. Lett., 13, 16–21, https://doi.org/10.1002/asl.355.
Vecchi, G. A., and B. J. Soden, 2007a: Global warming and the weakening of the tropical circulation. J. Climate, 20, 4316–4340, https://doi.org/10.1175/JCLI4258.1.
Vecchi, G. A., and B. J. Soden, 2007b: Increased tropical Atlantic wind shear in model projections of global warming. Geophys. Res. Lett., 34, L08702, https://doi.org/10.1029/2006GL028905.
Vimont, D. J., 2010: Transient growth of thermodynamically coupled variations in the tropics under an equatorially symmetric mean state. J. Climate, 23, 5771–5789, https://doi.org/10.1175/2010JCLI3532.1.
Wallace, J., 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.
Wang, C., R. H. Weisberg, and H. Yang, 1999: Effects of the wind speed–evaporation–SST feedback on the El Niño–Southern Oscillation. J. Atmos. Sci., 56, 1391–1403, https://doi.org/10.1175/1520-0469(1999)056<1391:EOTWSE>2.0.CO;2.
Wang, F., 2010: Thermodynamic coupled modes in the tropical atmosphere–ocean: An analytical solution. J. Atmos. Sci., 67, 1667–1677, https://doi.org/10.1175/2009JAS3262.1.
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.
Xie, S.-P., and S. G. H. Philander, 1994: A coupled ocean–atmosphere model of relevance to the ITCZ in the eastern Pacific. Tellus, 46A, 340–350, https://doi.org/10.3402/tellusa.v46i4.15484.
Xie, S.-P., and Y. Tanimoto, 1998: A pan-Atlantic decadal climate oscillation. Geophys. Res. Lett., 25, 2185–2188, https://doi.org/10.1029/98GL01525.
Xie, S.-P., C. Deser, G. A. Vecchi, J. Ma, H. Teng, and A. T. Wittenberg, 2010: Global warming pattern formation: Sea surface temperature and rainfall. J. Climate, 23, 966–986, https://doi.org/10.1175/2009JCLI3329.1.
Zhang, H., C. Deser, A. Clement, and R. Tomas, 2014: Equatorial signatures of the Pacific meridional modes: Dependence on mean climate state. Geophys. Res. Lett., 41, 568–574, https://doi.org/10.1002/2013GL058842.
Zhu, J., A. Kumar, B. Huang, M. A. Balmaseda, Z.-Z. Hu, L. Marx, and J. L. Kinter III, 2016: The role of off-equatorial surface temperature anomalies in the 2014 El Niño prediction. Sci. Rep., 6, 19677, https://doi.org/10.1038/srep19677.