1. Introduction
A change in the strength of the stratospheric polar vortex can have an appreciable influence on the position of the tropospheric midlatitude eddy-driven jet (e.g., Baldwin and Dunkerton 2001; Polvani and Kushner 2002; Kidston et al. 2015). In particular, there is considerable evidence in observations and models that a weakening of the polar vortex gives rise to a persistent equatorward shift of the lower-tropospheric jet, whereas a strengthening of the vortex, such as that which occurs under ozone depletion, yields a poleward-shifted jet (e.g., Thompson and Solomon 2002). One of the most striking examples of this downward coupling occurs during a sudden stratospheric warming (SSW), wherein the polar vortex weakens and warms in the space of a few days (Scherhag 1952). Following an SSW, the equatorward tropospheric jet shift can persist for four or more weeks, substantially longer than the tropospheric decorrelation time scale in the absence of such an event (e.g., Baldwin and Dunkerton 2001; Gerber et al. 2010; Simpson et al. 2011). Extreme vortex events such as SSWs can thus provide a potential source of skill for extratropical weather forecasts on subseasonal to seasonal time scales (e.g., Sigmond et al. 2013).
It is implicit in a number of studies that the tropospheric response to SSWs can be separated into two approximate stages: 1) the mechanism by which the stratospheric anomalies are initially communicated downward to the troposphere, and 2) the subsequent amplification and persistence of the tropospheric jet shift (e.g., Song and Robinson 2004; Thompson et al. 2006). In terms of the former, the mechanisms are not well understood and many have been proposed, including “downward control” via the wave-induced zonally symmetric meridional circulation (Haynes et al. 1991; Thompson et al. 2006), a balanced nonlocal response to a stratospheric potential vorticity anomaly (Hartley et al. 1998; Ambaum and Hoskins 2002; Black and McDaniel 2004), as well as changes in planetary wave propagation, breaking and reflection either directly or indirectly in both the stratosphere and troposphere (e.g., Matsuno 1971; Chen and Robinson 1992; Perlwitz and Harnik 2003; Shaw et al. 2010; Hitchcock and Haynes 2016; Hitchcock and Simpson 2016; Smith and Scott 2016).
To explain the second stage (i.e., the persistent jet shift at longer lags), the general consensus is that synoptic wave feedbacks are necessary (Limpasuvan et al. 2004; Kushner and Polvani 2004; Song and Robinson 2004; Garfinkel et al. 2013; Hitchcock and Simpson 2014). Indeed, Domeisen et al. (2013) employed a dry dynamical core, to show that in the absence of synoptic wave feedbacks in the troposphere, the tropospheric response to an SSW would be a poleward-shifted jet, opposite to what is observed. To our knowledge, no study has explicitly tried to separate the short- and long-lag response. It is the latter upon which we focus in this study.
To understand how changes in stratospheric temperature (such as those found during an SSW) influence the troposphere, many studies have imposed temperature perturbations to the stratosphere (e.g., Williams 2006; Lorenz and DeWeaver 2007). For instance, Polvani and Kushner (2002) and Kushner and Polvani (2004) developed a modification of the Held and Suarez (1994) forcing where tropospheric and stratospheric temperatures were relaxed to a chosen equilibrium state, to explore the impact of a high-latitude cooling on the troposphere. They demonstrated that the tropospheric response to a colder (stronger) polar vortex is a poleward-shifted jet stream. However, as they also relaxed the tropospheric temperatures, the downward impact was very sensitive to the details of the tropospheric climatology (e.g., Gerber and Polvani 2009). In fact, the magnitude of the tropospheric response to an identical stratospheric perturbation can differ by more than a factor of 3 depending on the tropospheric state (Garfinkel et al. 2013).
In another set of experiments, Haigh et al. (2005) and Simpson et al. (2009) imposed a steady stratospheric warming at high latitudes and found an equatorward tropospheric jet shift (although the main aim of their work was to understand the tropospheric response to tropical stratospheric warming). All of these studies found that changes in tropospheric eddy momentum fluxes and their feedbacks with the tropospheric circulation are crucial for the obtained response. Further, Simpson et al. (2009) found that the changes in the quasigeostrophic refractive index (Matsuno 1970) could explain the tropospheric eddy changes.
While many studies have imposed thermal perturbations to the stratosphere to explore changes in stratospheric variability (e.g., Taguchi et al. 2001; Jucker et al. 2013), the focus has been on the climatological (steady or seasonally evolving) modifications by applying the heating continuously. As SSWs are associated with a sudden onset of a high-latitude warming, we take a novel approach in this study by imposing a warming for only a few days to initiate an SSW, before switching it off and examining the coupled stratosphere–troposphere response. To do this, we perform a number of integrations with varying-magnitude heating profiles, using the Model of an Idealized Moist Atmosphere (MiMA; Jucker and Gerber 2017) and compare the evolution of the forced SSWs with SSWs taken from a free-running control integration. We also perform one additional experiment with an imposed high-latitude cooling in order to generalize the results to all extreme vortex events.
By triggering an SSW using a heating perturbation rather than by a modulation of the momentum budget, our experiments allow us to explicitly isolate the part of the downward influence that is attributable to changes in the polar vortex (e.g., subsequent changes in planetary and synoptic wave propagation in response to the weakened vortex), as opposed to the downward influence that is associated with the preceding planetary wave activity that drives a naturally occurring SSW, or with tropospheric precursors (as found to be important by a number of studies, e.g., Black and McDaniel 2004; Nakagawa and Yamazaki 2006; Karpechko et al. 2017; White et al. 2019).
Indeed, Plumb and Semeniuk (2003) found that upward-propagating planetary waves emanating from the troposphere can drive wind anomalies at successively lower levels akin to that observed during SSWs. In this case the downward migration occurs as a passive response to upward-propagating waves, such that downward migration during SSWs does not necessarily indicate any stratospheric influence on the troposphere. We will show that the tropospheric response to SSWs at longer lags is somewhat generic, insomuch that the evolution during the thermally triggered SSWs and the free-running SSWs (i.e., those initiated by momentum torques) are almost indistinguishable. We conclude that the persistent equatorward shift of the tropospheric jet at longer lags is independent of the wave fluxes that force an SSW, and that there is a genuine downward propagation of anomalies from the stratosphere (e.g., Hitchcock and Haynes 2016).
Section 2 provides a description of our model and experiments. Section 3 presents the results of our study, comparing SSWs in a free-running control integration (which are necessarily forced by momentum torques) with those that are thermally triggered. Finally, in section 4, a summary and discussion is provided.
2. Model and experimental setup
In this study we utilize the recently developed Model of an Idealized Moist Atmosphere (referred to hereafter as MiMA; Jucker and Gerber 2017). The most important features of MiMA that distinguish it from dry dynamical cores used in the studies aforementioned are its explicit treatment of moisture and radiation. These two features are important for simulating a realistic stratosphere and hence for stratosphere–troposphere coupling, which is the focus of this study.
a. MiMA
MiMA is an intermediate complexity atmospheric model with a dynamical core that has a variety of other well-motivated physical processes. Following Frierson et al. (2006), it includes a representation of large-scale moisture transport, latent heat release, a mixed layer ocean, a subgrid-scale convection scheme (Betts 1986; Betts and Miller 1986), and a Monin–Obukhov similarity boundary layer scheme. Also incorporated is a more realistic representation of radiation, namely the Rapid Radiative Transfer Model (RRTM) radiation scheme (Mlawer et al. 1997; Iacono et al. 2000), which replaces the gray radiation scheme of Frierson et al. (2006). The RRTM scheme allows for representation of the radiative impacts of both ozone and water vapor.
Neither a sponge layer nor Rayleigh damping scheme is utilized; instead, the gravity wave scheme of Alexander and Dunkerton (1999) is used to represent gravity wave momentum deposition, following Cohen et al. (2014). The gravity wave scheme is also modified to ensure that any gravity wave momentum fluxes that do reach close to the model lid are deposited in the top three layers so as to avoid possible sponge layer feedbacks and spurious meridional circulations associated with imposing heating perturbations (Shepherd et al. 1996; Shepherd and Shaw 2004). Full details regarding the model can be found in Jucker and Gerber (2017).
To generate a relatively realistic climatology (see Fig. S1 in the online supplemental material) on which our runs will be based, a number of parameters have been updated from the original version provided by Jucker and Gerber (2017). We follow Garfinkel et al. (2020), who modified the lower boundary conditions of the model to generate as realistic a stationary wave pattern as possible. There are differences between our study and theirs and these are documented in section 1 of the supplemental material, although these differences do not affect our results quantitatively. Another difference from Jucker and Gerber (2017) and Garfinkel et al. (2020) is the use of a monthly climatology zonal-mean input ozone file, taken from the preindustrial era CMIP5 forcing, as opposed to an annual-mean ozone input file. We note that this does not change the results qualitatively, although the SSW frequency is slightly higher using the latter. We refer readers to Garfinkel et al. (2020) for details on the exact model setup.
b. Experimental setup
A series of runs are performed at T42 horizontal resolution (2.8° × 2.8°) and with 40 vertical levels spanning from the surface to ~0.01 hPa (i.e., close to 70 km). We start by running the model freely for 50 years after discarding the first 10 years to allow the mixed layer ocean to reach an equilibrium state. This 50-yr control integration is herein referred to as the CTRL run.
In CTRL, 22 SSWs are found using the WMO criterion (McInturff 1978) that the zonal-mean zonal wind at 60°N and 10 hPa must reverse, along with the extra conditions that the SSW must occur during November–April, returning to westerly winds for at least 10 consecutive days (to avoid counting final warmings), and that two consecutive SSW events must be separated by at least 20 consecutive days of westerly winds (following Charlton and Polvani 2007). The ratio of SSWs in CTRL is 0.44 yr−1, which is a bit less than in observations (e.g., ~0.65 yr−1 in the latest ERA5 reanalysis). This may be due to the fact that in the climatology, the vortex is somewhat too strong and cold (by approximately 10–20 m s−1 and 5–10 K, respectively; see Fig. S1a) compared to in observations.
In total, five PTRB warming experiments are presented here, along with one PTRB cooling experiment, each with 50 ensemble members and with varying magnitude warmings and Nd = 3 days; the maximum thermal forcing is Q = 25 K day−1, incrementally decreasing by 5 K down to Q = 5 K day−1. The PTRB cooling experiment has a thermal forcing of Q = −10 K day−1. For example, in the 15-K PTRB, a forcing of Q = 15 K day−1 is switched on for 3 days, after which it is switched off and subsequently the model is allowed to run freely. Figure 1b shows the change in vortex strength (i.e., the ensemble mean zonal-mean zonal wind
Note that PTRB experiments where the duration of the thermal forcing has been varied have also been conducted (with Nd = 5 and 10 days). The key difference at longer lags is that the tropospheric response lasts for longer in conjunction with the thermal forcing duration. At shorter lags, a forcing that is imposed for longer, drives a stronger tropospheric response directly associated with the forcing itself that also lasts for a longer period (see the anomalous tropospheric westerlies in the bottom row of Fig. 2b). Hence, to avoid such direct tropospheric impacts that are not typical of observed SSWs, we limit the thermal forcing to Nd = 3 days.
Experiments have also been conducted wherein the vertical extent of the heating is modified. In particular, imposing the forcing only above pb = 70 hPa still yielded a clear near-surface response (~30% weaker in magnitude), whereas restricting the heating to above pb = 30 hPa, gave a much weaker near-surface response (along with a less barotropic structure). These results are in general agreement with Butler et al. (2010) insomuch that raising the lowest level of forcing does influence the magnitude of the near-surface response, although the details are different as their forcing was substantially weaker. We further note that the results presented herein are insensitive to different horizontal and vertical resolutions (T85 horizontal and using 60 levels rather than 40).
The initial stratospheric and tropospheric states for each ensemble member are not the same and are essentially random. This is indicated by the spread of the individual ensemble members for the 15-K PTRB (gray shading) before 1 January in Fig. 1b. Hence, any signal in the PTRB anomaly composites in relation to CTRL represents the deterministic response to the thermally forced stratospheric anomalies, which are thus independent of the initial stratospheric and tropospheric states. Also, note that there are two years in CTRL for which
3. Results: Zonal-mean circulation and wave evolution during free-running and thermally forced SSWs
We compare the evolution of the zonal-mean circulation and wave propagation/forcing between the 22 SSWs identified in CTRL (hereafter CTRL SSWs) and the thermally forced SSWs in PTRB. We focus primarily on the 15-K PTRB experiment although note that both the 10- and 15-K PTRB experiments provide similar results that are most similar to the CTRL SSWs. Nevertheless, we also make interexperiment comparisons to examine the tropospheric response sensitivity to the various magnitude thermal forcings.
The anomalies in this section are all deviations away from the unfiltered daily climatology in CTRL. For example, anomalies averaged over lags 1–3 in PTRB are calculated as the deviations away from the daily climatology in CTRL averaged over 1–3 January. By construction therefore, in PTRB, the ensemble-mean anomalies are identically zero.
a. Zonal wind, NAM, and temperature evolution
Composites of zonal-mean zonal wind
Lags 1–3 (Fig. 2b) represent the early onset in CTRL SSWs and the forcing stage in PTRB SSWs. In CTRL, there is a clear intensification of the
As the lags progress, the development of the stratospheric anomalies in both CTRL and PTRB are rather similar. There is a poleward and downward movement of the
In the troposphere, the
To further highlight the downward propagation to the troposphere, Fig. 3 shows height–time composites of the northern annular mode (NAM) index (shading) and
After the onset, the general structure of the stratospheric anomalies is similar between CTRL and PTRB with a sudden enhancement of negative NAM anomalies close to the onset date followed by recovery first aloft, and persistence in the lower stratosphere. However, the PTRB experiments in Figs. 3b–d have NAM anomalies that persist for much longer than in CTRL. In particular, the anomalies associated with the PTRB SSWs last until ~90 days after the switch-on forcing (i.e., up to 1 April) whereas in CTRL they last for ~65–70 days in the lower stratosphere. The second negative-NAM anomaly peak in April–May is associated with an earlier onset of the date of the final warming in all PTRB runs. In CTRL, the average final warming date over all 50 years is 12 May, whereas in PTRB, the average final warming date ranges from 28 April to 6 May.
The NAM index for our T85 15-K PTRB run (Fig. 3e) appears to explain the more persistent NAM in our T42 PTRB runs compared to in CTRL. The NAM in T85 persists for a similar period to the CTRL SSWs (although note the stronger recovery in T85 compared to in T42). Previous studies have found that coarser-resolution models tend to have more persistent annular mode variability (e.g., Gerber et al. 2008) and comparison of Figs. 3c and 3e confirm this in MiMA. Nevertheless, our T42 and T85 runs are qualitatively similar and the essential dynamics at play are the same (not shown). Further, the fluctuation dissipation theorem (Leith 1975) indicates that the extratropical jet response can be overly sensitive to external forcing if the intrinsic annular-mode time scales are too long, but a comparison of Figs. 3a and 3c (as well as Fig. 5 below) suggests that the PTRB and CTRL have tropospheric responses of similar magnitude and hence the response to external forcing is not exaggerated in our experiments.
In terms of the downward influence on the troposphere, the CTRL SSWs, 25- and 15-K PTRB experiments exhibit the classical “dripping-paint” pattern found by Baldwin and Dunkerton (2001). This is in contrast to the 5-K PTRB experiment that does not show any statistically significant downward propagation below ~200 hPa aside from a weakly negative tropospheric NAM in March. In particular, in the 15- and 25-K PTRB, the NAM and
In observations, the tropospheric
For the CTRL SSWs (Fig. 4, left), a dipole in
For the 15-K PTRB experiment (Fig. 4, right), the
A natural question arising from Figs. 2–4 is how the strength of the initial stratospheric warming relates to the subsequent strength and persistence of the tropospheric response. Hence, in Fig. 5a, the variability of the strength of the tropospheric response for all ensemble members for all PTRB experiments is shown as a scatterplot of the lower-stratospheric (100 hPa)
Overall, it is clear that the PTRB heating experiments give rise to an equatorward-shifted near-surface jet, whereas the PTRB cooling gives rise to a poleward-shifted jet (with the exception of a few ensemble members). There is a clear linear relationship, with a more negative lower-stratospheric
The regression slopes (shown at the top right in Figs. 5a,b) allow us to approximately quantify the magnitude of the downward impact. For instance, the near-surface
To further show that a stronger thermal perturbation yields a more-negative tropospheric NAM response, Fig. 5c shows histograms of the 850-hPa daily NAM indices at positive lags for the 25- and 5-K PTRB experiments (colored vertical lines indicate the ensemble means for the other three PTRB heating experiments, the PTRB cooling, and for CTRL). The main feature is that the 25-K PTRB leads to an overall shift of the tropospheric NAM toward more negative values in comparison to the 5-K PTRB rather than there being large changes in the skewness or kurtosis of the respective histograms (see values in top right). This is in agreement with Simpson et al. (2011), Sigmond et al. (2013), and Hitchcock and Simpson (2014) who also found that the main stratospheric influence during SSWs is to bias the troposphere to a more negative NAM-like state. We note that the 15-K PTRB produces a near-surface response of very similar magnitude to in CTRL (cf. pink and gray vertical lines).
In summary, the evolution of
Herein, the lag stages 4–10 and 11–20 are averaged into one (4–20). This is because the aim of this paper is to examine the long-lag (i.e., ≳3 week) tropospheric response to SSWs. The mechanisms behind the initial downward impact (i.e., the short-lag response), are beyond the scope of this paper.
b. Planetary and synoptic wave evolution
Figure 6 shows latitude–height composites of the EP flux divergence term Π = ∇ ⋅ F/ρ0acosφ (shading), EP fluxes F (arrows), and
At lags 4+ (Figs. 6c, d) planetary wave F anomalies are generally oriented poleward and downward along with anomalous Π > 0 in the high-latitude stratosphere, although the magnitudes of F and Π for planetary waves decrease at lags 21–90. This suppression following an SSW is the expected response to the weakened polar vortex (e.g., Limpasuvan et al. 2004). The presence of tropospheric precursors makes it difficult to separate the anomalies that are associated with the downward propagation from the preexisting tropospheric anomalies. The region of anomalous planetary wave Π < 0 near 55°–60°N in the middle troposphere contributes to the maintenance of the negative high-latitude
Tropospheric poleward-propagating synoptic waves are present at all lags straddling the
We now compare the anomalies in the CTRL SSWs with those for the 15-K PTRB in Fig. 7, which shows the same as Fig. 6 except without panels at negative lags. At lags 1–3 (Fig. 7a), a vertical dipole in Π for planetary waves is evident, which straddles the lowest level of maximum forcing at ~60 hPa, with anomalous divergence aloft, and convergence extending down to ~200 hPa. This dipole is associated with anomalous downward-propagating planetary waves and occurs as a direct response to the weakened vortex. In particular, the weakening vortex lowers the critical lines and hence prevents Rossby waves from propagating freely. The increase in static stability associated with the thermal forcing may also play a role in reducing the upward propagation of planetary waves [see Eq. (5b) and Chen and Robinson 1992]. This will also be explained by refractive index arguments in section 3c. In the region of anomalous tropospheric
At lags 4–20 (i.e., after the forcing has been switched off; Fig. 7b), the planetary wave anomalies are more widespread with an anomalous poleward and downward propagation extending from the stratospheric subtropics down to the high-latitude troposphere and with divergence aloft and convergence in the lower stratosphere–upper troposphere. In particular, the F anomalies extend down to 700 hPa in conjunction with the
At lags 21–90 (Fig. 7c), both the planetary wave and synoptic wave anomalies are similar to those in CTRL (Fig. 6). The planetary wave anomalies are essentially the same as at earlier lags, but with weaker magnitude as the vortex recovers. In terms of synoptic waves, there are clear poleward-propagating anomalies straddling the tropospheric
We next investigate the source of the tropospheric poleward-propagating synoptic waves. In Fig. 8a, a latitudinal profile of the Eady growth rate (σ = 0.31|f||∂u(φ, z, t)/∂z|/N) anomalies (Hoskins and Valdes 1990, blue line) at 400 hPa, averaged over lags 21–90 is shown for the 15-K PTRB. Also shown are the corresponding 400-hPa
It is next shown that the magnitude of the lower-stratospheric anomalies influence the strength of the tropospheric synoptic-wave anomalies. Figure 8c shows a scatter graph of the 100-hPa high-latitude
Overall, it appears that poleward-propagating synoptic waves play a key role in the maintenance of the equatorward-shifted tropospheric jet at longer lags in both the CTRL and PTRB SSWs (in contrast to equatorward-propagating synoptic waves in the PTRB cooling experiment). Such waves appear to be generated by the enhanced baroclinicity at midlatitudes, and propagate poleward where they break in the region of easterly anomalies (see Fig. S3c for evidence of EP flux convergence at high latitudes). Planetary waves on the other hand, are suppressed throughout the stratosphere and troposphere and may play a key role at short lags in initially bringing the polar vortex anomalies to the troposphere; however, examination of the initial downward communication is left to a future study.
c. Waveguide evolution
To calculate n2 for the CTRL SSWs, we first average
Figure 9 shows composites of n2 and
A developing feature at positive lags is a region of n2 < 0 in the midlatitude–subpolar upper troposphere–lower stratosphere, which intensifies as the lags progress. Upon comparison with the December–February climatology of n2 (see Fig. S1b), it appears that this feature extends the subtropical–midlatitude minimum of n2 to higher latitudes, and hence, may act to shield the stratosphere from subsequent upward wave propagation. Nevertheless, we note that tunneling of planetary waves through a region of n2 < 0 is still possible (e.g., Harnik 2002). Above ~50 hPa, n2 becomes positive after lags 1–3, as the vortex starts to recover (i.e.,
We now examine the PTRB SSWs (Fig. 9, bottom). First note that at negative lags (Fig. 9a), the presence of n2 anomalies close to the zero-wind line represent floating point errors due to the very small
At lags 4+ (Figs. 9c,d), n2 and and
Overall, as was the case in sections 3a and 3b, after ~3 weeks, the n2 anomalies in the thermally forced SSWs become similar to those in the CTRL SSWs. In particular, the mid- to high-latitude lower-tropospheric n2 > 0, the midlatitude lower-stratospheric n2 < 0, and the large positive n2 above ~50 hPa, are all common features to CTRL and PTRB. The EP fluxes in section 3b agree dynamically with the n2 anomalies here, and in particular, the high-latitude tropospheric region of n2 > 0 first develops in response to the downward migration of the stratospheric
d. Meridional circulation evolution
Figure 10 (top) shows composites of Ψ* at various lag stages for the CTRL SSWs. At negative lags (Fig. 10a), Ψ* is everywhere positive indicative of a strengthened Brewer–Dobson circulation during the lead up to an SSW. This is driven by an imbalance between the enhanced upward-propagating planetary wave activity (
The tropospheric Ψ* response at positive lags is an extratropical tripole with Ψ* > 0 at midlatitudes flanked at low and high latitudes by Ψ* < 0 (although the high-latitude cell is much weaker at lags 4+). This tripole corresponds to changes in the width of the polar, Ferrel, and Hadley cells (e.g., Martineau et al. 2018). Indeed, this tripole is the response associated with general stratospheric NAM variability rather than variability solely attributed to the tropospheric NAM (see Fig. S6). We note that the Ψ* < 0 anomalies at ~30°–45°N and the Ψ* > 0 anomalies at ~45°–65°N straddle the nodal line in
The bottom row of Fig. 10 shows Ψ* anomalies for the 15-K PTRB experiment. Note that Ψ* is qualitatively similar for all of our experiments. At lags 1–3 (i.e., during the forcing stage; Fig. 10b), Ψ* is everywhere negative, with largest magnitudes at ~55°N, ~50 hPa, and a second peak at ~45°N, 500 hPa. The
At lags 4–20 (Fig. 10c), the Ψ* anomalies are noticeably different to those in CTRL. For instance, the anomalous meridional circulation between ~400 and ~50 hPa completely reverses to Ψ* > 0. This occurs due to a slight imbalance between
However, by lags 21–90, the Ψ* anomalies appear to be very similar to those in CTRL, with an extratropical tripole in the troposphere and with weakly negative stratospheric anomalies. The tripole is the response to general stratospheric NAM variability and gives rise to changes in the width of the Ferrel cell, whereas the weakly negative Ψ* aloft is the response to the reduced upward-propagating planetary waves into the stratosphere (Fig. 7). Hence, after ~3 weeks, the circulation following the CTRL SSWs and that following the thermally forced SSWs in PTRB become very similar to one another.
In summary, there are large differences in Ψ* between the CTRL SSWs and the thermally forced SSWs at lags of less than ~3 weeks. However, at longer lags, the Ψ* anomalies evolve very similarly with a tropospheric tripole associated with the shifted jet, and a weakly negative stratospheric Ψ* associated with the suppressed planetary waves following the SSW onset (see section 3b).
4. Summary and discussion
We have examined the tropospheric response to varying magnitude high-latitude stratospheric heating perturbations in order to examine the downward influence of SSWs. To capture the sudden nature of an SSW, the heating perturbation was only switched on for a few days (spun off from a free-running control integration, CTRL), which, depending on the magnitude of the imposed heating, either gave rise to a weakened, or completely reversed vortex. The evolution of the thermally forced SSWs was then compared with naturally occurring SSWs identified in CTRL. Our novel approach has allowed us to isolate the tropospheric response associated with the weakened polar vortex, as opposed to the response associated with the original planetary waves (and hence momentum torques) that initiated the SSW. We have focused in particular on understanding the long-lag (i.e., >2–3 weeks) tropospheric response as opposed to the initial communication of the stratospheric anomalies to the troposphere at shorter lags.
Our results confirm a downward influence from the stratosphere following an SSW event (e.g., Baldwin and Dunkerton 2001). This is evidenced by the strong tropospheric signal following the thermally forced SSWs (Figs. 2–5) despite the fact that there are no momentum torques associated with preceding planetary waves that initiate the SSW (as is the case in the free-running CTRL SSWs). Plumb and Semeniuk (2003) demonstrated that the tropospheric zonal-wind anomalies following a SSW could occur passively in response to the upward-propagating planetary waves that initiated the SSW, and hence concluded that a downward migration of wind anomalies is not necessarily indicative of a downward stratospheric influence. Our results unambiguously confirm that a weakening of the stratospheric polar vortex drives a tropospheric circulation response.
Another key result is that at longer lags, the stratospheric and tropospheric evolution in the free-running CTRL SSWs and the thermally forced SSWs are remarkably similar, both in terms of the zonal-mean circulation and the eddy fluxes (Figs. 2, 6, 7, 9, and 10). This indicates that at longer lags the tropospheric response is somewhat generic and the initial formation of an SSW does not play a large role. Instead, the strength of the warming in the lower stratosphere determines the magnitude of the tropospheric response (Fig. 5, and in agreement with, e.g., Maycock and Hitchcock 2015). In particular, our results indicate a robust linear relationship between the strength of the lower-stratospheric warming and the tropospheric response, with the linearity also extending to sudden stratospheric cooling events (Figs. 5 and 8). The linear response rules out the presence of any regime-like tropospheric behavior (at least in MiMA). Nevertheless, at shorter lags, the particulars associated with the initial SSW formation may play a potentially important role, given the difference in evolution between the CTRL SSWs and PTRB SSWs.
In maintaining the tropospheric jet shift at longer lags, synoptic waves play a key role (see Figs. 6–8), in agreement with a number of studies (e.g., Limpasuvan et al. 2004; Polvani and Waugh 2004; Song and Robinson 2004; Domeisen et al. 2013). The collocation of upward-propagating synoptic waves and the peak Eady growth rate in the region of midlatitude westerly anomalies suggests that synoptic waves may be forced due to the enhanced baroclinicity (see Fig. 8 and e.g., Robinson 2000). The poleward propagation of these synoptic waves then appears to generate a positive feedback in concert with the region of enhanced high-latitude tropospheric refractive index that develops in response to the descending polar vortex anomalies and intensifies as the lags progress (Fig. 9). In particular, the poleward-propagating synoptic waves flux momentum equatorward [see Eq. (5a)] and thus weaken the winds further at high latitudes, which in turn enhances the ambient refractive index [due to
The initial 3-week period after 1 January in the PTRB experiments during which the polar-vortex anomalies migrate downward to the surface, requires further investigation. The circulation anomalies gradually propagate down to ~300 hPa over the first ~2 weeks, before they barotropically extend downward to the high-latitude lower troposphere (Fig. 3). The suppression of planetary waves appears to correlate with this downward propagation (Fig. 7) in agreement with Hitchcock and Haynes (2016) and Hitchcock and Simpson (2016). Once the mean-state anomalies reach the lower troposphere, they subsequently migrate equatorward before stalling at midlatitudes where they straddle the midlatitude jet (Figs. 2 and 4). The exact mechanisms for this downward and subsequently equatorward migration of the winds is beyond the scope of this paper, although we note that both the CTRL and PTRB SSWs exhibit anomalous wave convergence into the polar cap at lags of 4–20 (Figs. 6 and 7), which may play a role in the initial downward communication.
Unlike in our CTRL run (as well as in observations), for which the near-surface response following an SSW projects almost entirely onto the first EOF, the near-surface response following the PTRB SSWs projects onto both the first and second EOFs (Fig. 4), although with a larger projection onto EOF1. Parallels can therefore be drawn between the PTRB SSWs and the observed response during final warmings, which have been found in observations to project onto both leading EOFs (e.g., Black et al. 2006; Sheshadri et al. 2017), although we note that in our PTRB runs, the winds do reverse back to westerly. Hence, our experiments may be useful for examining the tropospheric response to a wide range of polar vortex variability.
It should be noted that the mechanisms for downward propagation discussed here are based on the evolution during thermally triggered SSWs, which, by construction, lack the vital ingredient of planetary-scale momentum torques that are ultimately responsible for observed SSWs. The meridional circulation anomalies associated with heating and momentum torques can be very different (e.g., Shepherd et al. 1996) and hence could conceivably have different effects on the troposphere. Nevertheless, given the similar evolution of the thermally forced SSWs to the CTRL SSWs at longer lags, these initial momentum torques seemingly do not play a large role in the tropospheric response at subseasonal to seasonal time scales.
One of the advantages of MiMA used here is that it has a realistic stratosphere and annual cycle due to the incorporation of a full radiation scheme (Jucker and Gerber 2017). It is therefore a more realistic setup than that used in previous studies (e.g., Polvani and Kushner 2002; Kushner and Polvani 2004) that have utilized dry dynamical cores with Newtonian cooling. Nevertheless, we note that the annular mode time scales in our presented T42 PTRB experiments are too long compared to our CTRL SSWs (Fig. 3). However, as the results are qualitatively similar to T85 experiments (which have similar annular-mode time scales to in CTRL), our conclusions are unchanged and the essential dynamics are the same.
It has been suggested that the strength of the original wave driving can be important for the tropospheric response to some SSWs (e.g., Nakagawa and Yamazaki 2006; White et al. 2019). This is somewhat similar to the strength of the lower-stratospheric warming in our study. It has also been suggested that the troposphere may need to be in a state to “receive” the stratospheric influence (e.g., Black and McDaniel 2004). We agree that the details of an SSW are important for the evolution of an SSW, as well as for the initial downward impact on the troposphere, but argue that the long-lag response of the tropospheric jet is a generic response to a weakened polar vortex.
Acknowledgments
We thank Hua Lu for useful discussion. We acknowledge the support of a European Research Council starting grant under the European Union Horizon 2020 research and innovation programme (Grant 677756). EPG also acknowledges support from the U.S. NSF through Grant AGS-1852727. MJ is supported by the ARC Centre of Excellence for Climate Extremes under Grant CE170100023 and ARC Grant FL150100035. JR also acknowledges support from the National Natural Science Foundation of China (41705024).
REFERENCES
Alexander, M., and T. J. Dunkerton, 1999: A spectral parameterization of mean-flow forcing due to breaking gravity waves. J. Atmos. Sci., 56, 4167–4182, https://doi.org/10.1175/1520-0469(1999)056<4167:ASPOMF>2.0.CO;2.
Ambaum, M. H. P., and B. J. Hoskins, 2002: The NAO troposphere–stratosphere connection. J. Climate, 15, 1969–1978, https://doi.org/10.1175/1520-0442(2002)015%3C1969:TNTSC%3E2.0.CO;2.
Andrews, D. G., J. R. Holton, and C. B. Leovy, 1987: Middle Atmosphere Dynamics. Academic Press, 489 pp.
Baldwin, M. P., and T. J. Dunkerton, 1999: Propagation of the Arctic oscillation from the stratosphere to the troposphere. J. Geophys. Res., 104, 30 937–30 946, https://doi.org/10.1029/1999JD900445.
Baldwin, M. P., and T. J. Dunkerton, 2001: Stratospheric harbingers of anomalous weather regimes. Science, 294, 581–584, https://doi.org/10.1126/science.1063315.
Baldwin, M. P., and D. W. J. Thompson, 2009: A critical comparison of stratosphere-troposphere coupling indices. Quart. J. Roy. Meteor. Soc., 135, 1661–1672, https://doi.org/10.1002/qj.479.
Betts, A., 1986: A new convective adjustment scheme. Part I: Observational and theoretical basis. Quart. J. Roy. Meteor. Soc., 112, 677–691, https://doi.org/10.1002/QJ.49711247307.
Betts, A., and M. Miller, 1986: A new convective adjustment scheme. Part II: Single column tests using GATE wave, BOMEX, ATEX and Arctic air-mass data sets. Quart. J. Roy. Meteor. Soc., 112, 693–709, https://doi.org/10.1002/QJ.49711247308.
Black, R. X., and B. A. McDaniel, 2004: Diagnostic case studies of the northern annular mode. J. Climate, 17, 3990–4004, https://doi.org/10.1175/1520-0442(2004)017<3990:DCSOTN>2.0.CO;2.
Black, R. X., B. A. McDaniel, and W. A. Robinson, 2006: Stratosphere–troposphere coupling during spring onset. J. Climate, 19, 4891–4901, https://doi.org/10.1175/JCLI3907.1.
Butler, A. H., D. W. J. Thompson, and R. Heikes, 2010: The steady-state atmospheric circulation response to climate change–like thermal forcings in a simple general circulation model. J. Climate, 23, 3474–3496, https://doi.org/10.1175/2010JCLI3228.1.
Charlton, A. J., and L. M. Polvani, 2007: A new look at stratospheric sudden warmings. Part I: Climatology and modeling benchmarks. J. Climate, 20, 449–469, https://doi.org/10.1175/JCLI3996.1.
Charney, J. G., and P. G. Drazin, 1961: Propagation of planetary-scale disturbances from the lower into the upper atmosphere. J. Geophys. Res., 66, 83–109, https://doi.org/10.1029/JZ066i001p00083.
Chen, P., and W. A. Robinson, 1992: Propagation of planetary waves between the troposphere and stratosphere. J. Atmos. Sci., 49, 2533–2545, https://doi.org/10.1175/1520-0469(1992)049<2533:POPWBT>2.0.CO;2.
Cohen, J., and J. Jones, 2011: Tropospheric precursors and stratospheric warmings. J. Climate, 24, 6562–6572, https://doi.org/10.1175/2011JCLI4160.1.
Cohen, N. Y., E. P. Gerber, and O. Buhler, 2014: What drives the Brewer–Dobson circulation? J. Atmos. Sci., 71, 3837–3855, https://doi.org/10.1175/JAS-D-14-0021.1.
de la Camara, A., M. Abalos, and P. Hitchcock, 2018: Changes in stratospheric transport and mixing during sudden stratospheric warmings. J. Geophys. Res. Atmos., 123, 3356–3373, https://doi.org/10.1002/2017JD028007.
Domeisen, D. I. V., L. Sun, and G. Chen, 2013: The role of synoptic eddies in the tropospheric response to stratospheric variability. Geophys. Res. Lett., 40, 4933–4937, https://doi.org/10.1002/grl.50943.
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.
Garfinkel, C. I., D. L. Hartmann, and F. Sassi, 2010: Tropospheric precursors of anomalous Northern Hemisphere stratospheric polar vortices. J. Climate, 23, 3282–3299, https://doi.org/10.1175/2010JCLI3010.1.
Garfinkel, C. I., T. A. Shaw, D. L. Hartmann, and D. W. Waugh, 2012: Does the Holton–Tan mechanism explain how the quasi-biennial oscillation modulates the Arctic polar vortex? J. Atmos. Sci., 69, 1713–1733, https://doi.org/10.1175/JAS-D-11-0209.1.
Garfinkel, C. I., D. W. Waugh, and E. P. Gerber, 2013: The effect of tropospheric jet latitude on coupling between the stratospheric polar vortex and the troposphere. J. Climate, 26, 2077–2095, https://doi.org/10.1175/JCLI-D-12-00301.1.
Garfinkel, C. I., I. P. White, E. P. Gerber, M. Jucker, and M. Erez, 2020: The building blocks of Northern Hemisphere wintertime stationary waves. J. Climate, 33, 5611–5633, https://doi.org/10.1175/JCLI-D-19-0181.1.
Gerber, E. P., and L. P. Polvani, 2009: Stratosphere–troposphere coupling in a relatively simple AGCM: The importance of stratospheric variability. J. Climate, 22, 1920–1933, https://doi.org/10.1175/2008JCLI2548.1.
Gerber, E. P., S. Voronin, and L. P. Polvani, 2008: Testing the annular mode autocorrelation time scale in simple atmospheric general circulation models. Mon. Wea. Rev., 136, 1523–1536, https://doi.org/10.1175/2007MWR2211.1.
Gerber, E. P., and Coauthors, 2010: Stratosphere-troposphere coupling and annular mode variability in chemistry-climate models. J. Geophys. Res., 115, D00M06, https://doi.org/10.1029/2009JD013770.
Haigh, J. D., M. Blackburn, and R. Day, 2005: The response of tropospheric circulation to perturbations in lower-stratospheric temperature. J. Climate, 18, 3672–3685, https://doi.org/10.1175/JCLI3472.1.
Harnik, N., 2002: The evolution of a stratospheric wave packet. J. Atmos. Sci., 59, 202–217, https://doi.org/10.1175/1520-0469(2002)059<0202:TEOASW>2.0.CO;2.
Hartley, D. E., J. T. Villarin, R. X. Black, and C. A. Davis, 1998: A new perspective on the dynamical link between the stratosphere and troposphere. Nature, 391, 471–474, https://doi.org/10.1038/35112.
Haynes, P. H., C. J. Marks, M. E. McIntyre, T. G. Shepherd, and K. P. Shine, 1991: On the “downward control” of extratropical diabatic circulations by eddy-induced mean zonal forces. J. Atmos. Sci., 48, 651–678, https://doi.org/10.1175/1520-0469(1991)048<0651:OTCOED>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.
Hitchcock, P., and I. R. Simpson, 2014: The downward influence of stratospheric sudden warmings. J. Atmos. Sci., 71, 3856–3876, https://doi.org/10.1175/JAS-D-14-0012.1.
Hitchcock, P., and P. H. Haynes, 2016: Stratospheric control of planetary waves. Geophys. Res. Lett., 43, 11 884–11 892, https://doi.org/10.1002/2016GL071372.
Hitchcock, P., and I. R. Simpson, 2016: Quantifying eddy feedbacks and forcings in the tropospheric response to stratospheric sudden warmings. J. Atmos. Sci., 73, 3641–3657, https://doi.org/10.1175/JAS-D-16-0056.1.
Hitchcock, P., T. G. Shepherd, and G. L. Manney, 2013: Statistical characterization of Arctic polar-night jet oscillation events. J. Climate, 26, 2096–2116, https://doi.org/10.1175/JCLI-D-12-00202.1.
Holton, J. R., and H.-C. Tan, 1980: The influence of the equatorial quasi-biennial oscillation on the global circulation at 50 mb. J. Atmos. Sci., 37, 2200–2208, https://doi.org/10.1175/1520-0469(1980)037<2200:TIOTEQ>2.0.CO;2.
Hoskins, B. J., and D. J. Karoly, 1981: The steady linear response of a spherical atmosphere to thermal and orographic forcing. J. Atmos. Sci., 38, 1179–1196, https://doi.org/10.1175/1520-0469(1981)038<1179:TSLROA>2.0.CO;2.
Hoskins, B. J., and P. J. Valdes, 1990: On the existence of storm-tracks. J. Atmos. Sci., 47, 1854–1864, https://doi.org/10.1175/1520-0469(1990)047<1854:OTEOST>2.0.CO;2.
Iacono, M. J., E. J. Mlawer, S. A. Clough, and J. J. Morcrette, 2000: Impact of an improved longwave radiation model, RRTM, on the energy budget and thermodynamic properties of the NCAR Community Climate Model, CCM3. J. Geophys. Res., 105, 14 873–14 890, https://doi.org/10.1029/2000JD900091.
Jucker, M., and E. P. Gerber, 2017: Untangling the annual cycle of the tropical tropopause layer with an idealized moist model. J. Climate, 30, 7339–7358, https://doi.org/10.1175/JCLI-D-17-0127.1.
Jucker, M., S. Fueglistaler, and G. K. Vallis, 2013: Maintenance of the stratospheric structure in an idealized general circulation model. J. Atmos. Sci., 70, 3341–3358, https://doi.org/10.1175/JAS-D-12-0305.1.
Karpechko, A. Y., P. Hitchcock, D. H. W. Peters, and A. Schneidereit, 2017: Predictability of downward propagation of major sudden stratospheric warmings. Quart. J. Roy. Meteor. Soc., 143, 1459–1470, https://doi.org/10.1002/qj.3017.
Kidston, J., A. A. Scaife, S. C. Hardiman, D. M. Mitchell, N. Butchart, M. P. Baldwin, and L. J. Gray, 2015: Stratospheric influence on tropospheric jet streams, storm tracks and surface weather. Nat. Geosci., 8, 433–440, https://doi.org/10.1038/ngeo2424.
Kuroda, Y., and K. Kodera, 2001: Variability of the polar night jet in the Northern and Southern Hemispheres. J. Geophys. Res., 106, 20 703–20 713, https://doi.org/10.1029/2001JD900226.
Kushner, P. J., and L. M. Polvani, 2004: Stratosphere–troposphere coupling in a relatively simple AGCM: The role of eddies. J. Climate, 17, 629–639, https://doi.org/10.1175/1520-0442(2004)017<0629:SCIARS>2.0.CO;2.
Leith, C. E., 1975: Climate response and fluctuation dissipation. J. Atmos. Sci., 32, 2022–2026, https://doi.org/10.1175/1520-0469(1975)032<2022:CRAFD>2.0.CO;2.
Limpasuvan, V., D. W. Thompson, and D. L. Hartmann, 2004: The life cycle of the Northern Hemisphere sudden stratospheric warmings. J. Climate, 17, 2584–2596, https://doi.org/10.1175/1520-0442(2004)017<2584:TLCOTN>2.0.CO;2.
Lorenz, D. J., and E. T. DeWeaver, 2007: Tropopause height and zonal wind response to global warming in the IPCC scenario integrations. J. Geophys. Res., 112, D10119, https://doi.org/10.1029/2006JD008087.
Martineau, P., S.-W. Son, M. Taguchi, and A. H. Butler, 2018: A comparison of the momentum budget in reanalysis datasets during sudden stratospheric warming events. Atmos. Chem. Phys., 18, 7169–7187, https://doi.org/10.5194/acp-18-7169-2018.
Matsuno, T., 1970: Vertical propagation of stationary planetary waves in winter Northern Hemisphere. J. Atmos. Sci., 27, 871–883, https://doi.org/10.1175/1520-0469(1970)027<0871:VPOSPW>2.0.CO;2.
Matsuno, T., 1971: A dynamical model of the stratospheric sudden warming. J. Atmos. Sci., 28, 1479–1494, https://doi.org/10.1175/1520-0469(1971)028<1479:ADMOTS>2.0.CO;2.
Maycock, A. C., and P. Hitchcock, 2015: Do split and displacement sudden stratospheric warmings have different annular mode signatures? Geophys. Res. Lett., 42, 10 943–10 951, https://doi.org/10.1002/2015GL066754.
McInturff, R. M. E., 1978: Stratospheric warmings: Synoptic, dynamic and general-circulation aspects. NASA Reference Publ. NASA-RP-1017, 174 pp., http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/19780010687.pdf.
Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res., 102, 16 663–16 682, https://doi.org/10.1029/97JD00237.
Nakagawa, K. I., and K. Yamazaki, 2006: What kind of stratospheric sudden warming propagates to the troposphere? Geophys. Res. Lett., 33, L04801, https://doi.org/10.1029/2005GL024784.
Perlwitz, J., and N. Harnik, 2003: Observational evidence of a stratospheric influence on the troposphere by planetary wave reflection. J. Climate, 16, 3011–3026, https://doi.org/10.1175/1520-0442(2003)016<3011:OEOASI>2.0.CO;2.
Plumb, R. A., and K. Semeniuk, 2003: Downward migration of extratropical zonal wind anomalies. J. Geophys. Res., 108, 4223, https://doi.org/10.1029/2002JD002773.
Polvani, L. M., and P. Kushner, 2002: Tropospheric response to stratospheric perturbations in a relatively simple general circulation model. Geophys. Res. Lett., 29, 1114, https://doi.org/10.1029/2001GL014284.
Polvani, L. M., and D. W. Waugh, 2004: Upward wave activity flux as a precursor to extreme stratospheric events and subsequent anomalous surface weather regimes. J. Climate, 17, 3548–3554, https://doi.org/10.1175/1520-0442(2004)017<3548:UWAFAA>2.0.CO;2.
Robinson, W. A., 2000: A baroclinic mechanism for the eddy feedback on the zonal index. J. Atmos. Sci., 57, 415–422, https://doi.org/10.1175/1520-0469(2000)057<0415:ABMFTE>2.0.CO;2.
Scherhag, R., 1952: Die explosionsartigen stratosphärenerwärmungen des spätwinter 1951/1952 (the explosive warmings in the stratosphere of the late winter 1951/1952). Ber. Dtsch. Wetterdienstes, 38, 51–63.
Shaw, T. A., J. Perlwitz, and N. Harnik, 2010: Downward wave coupling between the stratosphere and troposphere: The importance of meridional wave guiding and comparison with zonal–mean coupling. J. Climate, 23, 6365–6381, https://doi.org/10.1175/2010JCLI3804.1.
Shepherd, T. G., and T. A. Shaw, 2004: The angular momentum constraint on climate sensitivity and downward influence in the middle atmosphere. J. Atmos. Sci., 61, 2899–2908, https://doi.org/10.1175/JAS-3295.1.
Shepherd, T. G., K. Semeniuk, and J. N. Koshyk, 1996: Sponge layer feedbacks in middle-atmosphere models. J. Geophys. Res., 101, 23 447–23 464, https://doi.org/10.1029/96JD01994.
Sheshadri, A., R. A. Plumb, and E. P. Gerber, 2017: Propagating annular modes: Empirical orthogonal functions, principal oscillation patterns, and time scales. J. Atmos. Sci., 74, 1345–1361, https://doi.org/10.1175/JAS-D-16-0291.1.
Sigmond, M., J. F. Scinocca, V. V. Kharin, and T. G. Shepherd, 2013: Enhanced seasonal forecast skill following stratospheric sudden warmings. Nat. Geosci., 6, 98–102, https://doi.org/10.1038/ngeo1698.
Simpson, I. R., M. Blackburn, and J. D. Haigh, 2009: The role of eddies in driving the tropospheric response to stratospheric heating perturbations. J. Atmos. Sci., 66, 1347–1365, https://doi.org/10.1175/2008JAS2758.1.
Simpson, I. R., P. Hitchcock, T. G. Shepherd, and J. F. Scinocca, 2011: Stratospheric variability and tropospheric annular-mode timescales. Geophys. Res. Lett., 38, L20806, https://doi.org/10.1029/2011GL049304.
Smith, K. L., and R. K. Scott, 2016: The role of planetary waves in the tropospheric jet response to stratospheric cooling. Geophys. Res. Lett., 43, 2904–2911, https://doi.org/10.1002/2016GL067849.
Song, Y., and W. Robinson, 2004: Dynamical mechanisms for stratospheric influences on the troposphere. J. Atmos. Sci., 61, 1711–1725, https://doi.org/10.1175/1520-0469(2004)061<1711:DMFSIO>2.0.CO;2.
Taguchi, M., T. Yamaga, and S. Yoden, 2001: Internal variability of the troposphere–stratosphere coupled system simulated in a simple global circulation model. J. Atmos. Sci., 58, 3184–3203, https://doi.org/10.1175/1520-0469(2001)058<3184:IVOTTS>2.0.CO;2.
Thompson, D. W. J., and J. M. Wallace, 2000: Annular modes in the extratropical circulation. Part I: Month-to-month variability. J. Climate, 13, 1000–1016, https://doi.org/10.1175/1520-0442(2000)013<1000:AMITEC>2.0.CO;2.
Thompson, D. W. J., and S. Solomon, 2002: Interpretation of recent Southern Hemisphere climate change. Science, 296, 895–899, https://doi.org/10.1126/science.1069270.
Thompson, D. W. J., J. C. Furtado, and T. G. Shepherd, 2006: On the tropospheric response to anomalous stratospheric wave drag and radiative heating. J. Atmos. Sci., 63, 2616–2629, https://doi.org/10.1175/JAS3771.1.
White, I. P., C. G. Garfinkel, E. P. Gerber, M. Jucker, V. A. Aquila, and L. D. Oman, 2019: The downward influence of sudden stratospheric warmings: Association with tropospheric precursors. J. Climate, 32, 85–108, https://doi.org/10.1175/JCLI-D-18-0053.1.
Williams, G. P., 2006: Circulation sensitivity to tropopause height. J. Atmos. Sci., 63, 1954–1961, https://doi.org/10.1175/JAS3762.1.
Note that upon including c > 0, the K* peak is evident at subpolar latitudes (see below), but north of ~60°N, K* becomes imaginary (and hence represents wave evanescence). To better highlight the peak at subpolar latitudes therefore, we use a value of c = 0.