Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Atmospheric Research 297 (2024) 107112

Contents lists available at ScienceDirect

Atmospheric Research
journal homepage: www.elsevier.com/locate/atmosres

Impact of the 2018 major sudden stratospheric warming on weather over


the midlatitude regions of Eastern Europe and East Asia
Yu Shi a, Oleksandr Evtushevsky b, Gennadi Milinevsky a, c, *, Xiaolong Wang d, **,
Andrew Klekociuk e, Wei Han a, Asen Grytsai b, Yuke Wang f, Lidong Wang d,
Bohdan Novosyadlyj a, g, Yulia Andrienko b
a
College of Physics, International Center of Future Science, Jilin University, 2699 Qianjin Str., Changchun 130012, China
b
Taras Shevchenko National University of Kyiv, 60, Volodymyrska Str., Kyiv 01601, Ukraine
c
State Institution National Antarctic Scientific Center, 16, Taras Shevchenko Boulevard, Kyiv 01601, Ukraine
d
State Key Laboratory of Integrated Optoelectronics, College of Electronic Science and Engineering, International Center of Future Science, Jilin University, 2699 Qianjin
Str., Changchun 130012, China
e
School of Physics, Chemistry and Earth Sciences, University of Adelaide, 10 Pulteney Str., Adelaide 5005, Australia
f
Chang Guang Satellite Technology Co., Ltd, 1299, Mingxi Road, Changchun 130000, China
g
Ivan Franko National University of Lviv, 8, Kyryla i Methodia Str., Lviv 79005, Ukraine

A R T I C L E I N F O A B S T R A C T

Keywords: The impact of the Arctic major sudden stratospheric warming (SSW) in 2018 on regional cold weather in the
Sudden stratospheric warming northern midlatitudes is analyzed. Data from temperature observations in East Asia (Changchun, Northern
Air temperature China) and Eastern Europe (Kharkiv, Northern Ukraine) and NCEP–NCAR and ERA5 reanalyses are used. The
Tropopause descent
variability of the vertical meridional profiles of temperature, temperature anomalies, vertical velocities and
Vertical velocity
Cooling
geopotential height during two strong surface cooling events is compared. Weather anomalies in Changchun and
Weather anomaly Kharkiv appeared under strong vortex conditions with wave 1 dominance (the pre-SSW period in January) and
during vortex perturbation and split due to intensification of wave 2 (major SSW in February), respectively. The
results show that the positive stratospheric temperature anomaly migrated both downward to the lowermost
stratosphere at 200–300 hPa, and equatorward, reaching midlatitudes. In both events, the change in the
lowermost stratosphere heating was associated with a lowering of the tropopause by 2–3 km in 5–8 days. The
descent of the tropopause over Changchun (Kharkiv) was accompanied by a downward air motion so that the
midtropospheric layer at altitudes 4–5 km (3–4 km) and temperature − 20 ◦ C to − 30 ◦ C (− 10 ◦ C to − 20 ◦ C) was
pushed down to the surface. Observations showed that the surface temperatures during these events dropped by
18 ◦ C in Changchun and by 14 ◦ C in Kharkiv. Besides the altitudinal (3–5 km), the results indicate latitudinal
(near the climatological mean edge of the stratospheric polar vortex at ~60◦ N) location of the cold air source for
the occurrence of extreme low temperatures in East Asia and Eastern Europe. Preliminary analysis of cooling
events suggests a relatively smaller contribution of meridional surface transport of cold air to the midlatitudes
compared to its downwelling from the midtropospheric level. Possible mechanisms of strong surface cooling in
midlatitudes are discussed.

1. Introduction horizontally into mid- and low latitudes (Chandran and Collins, 2014).
While the onset of SSW may develop over a few days, the main phase of
Rapid changes in zonal circulation and temperature in the polar warming and the recovery phase can last for weeks and months,
stratosphere occur during sudden stratospheric warming (SSW) events. respectively (Baldwin et al., 2021; Li et al., 2023). An SSW is caused by a
The impact of these dramatic changes extends both vertically, down into sharp increase in the amplitude of planetary-scale waves that propagate
the troposphere and up into the mesosphere (Baldwin et al., 2021), and upward from the troposphere and perturb the stratospheric polar vortex

* Corresponding author at: College of Physics, International Center of Future Science, Jilin University, 2699 Qianjin Str., Changchun 130012, China.
** Corresponding author.
E-mail addresses: milinevskyi@jlu.edu.cn (G. Milinevsky), brucewang@jlu.edu.cn (X. Wang).

https://doi.org/10.1016/j.atmosres.2023.107112
Received 5 April 2023; Received in revised form 9 November 2023; Accepted 9 November 2023
Available online 13 November 2023
0169-8095/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
Y. Shi et al. Atmospheric Research 297 (2024) 107112

(Butler et al., 2015). The main influences come from waves with zonal downward impact, as compared to the observed event in January 2019,
wavenumbers m = 1 (wave 1) and m = 2 (wave 2), which act to shift the while the dominant wave number (the wave-2/wave-1 amplitude) is not
vortex away from the pole or split it into two parts (Baldwin et al., the primary factor determining the downward propagation of SSWs in
2021). these events (Rao et al., 2020). In contrast to SSW occurred in early
In the Northern Hemisphere (NH), it is known that the upper- January 2019, the strong surface cooling was observed in late January
stratospheric and lower-mesospheric anomalies are driven immedi­ 2019 across much of the United States Midwest, which was explained by
ately following an SSW (Hitchcock and Shepherd, 2013), whereas the impact of the stratosphere through downward transport (Xiong
anomalies can descend downward into the troposphere on time scales of et al., 2022). An exceptionally strong and persistent vortex resulted in
weeks to months (Baldwin and Dunkerton, 2001; Thompson et al., 2002; the early February 2020 minor SSW (Manney et al., 2022). Due to weak
Wang et al., 2020). The stratosphere–troposphere coupling is the effect wave activity in the stratosphere associated with a quasi-simultaneous
of the extension of dynamically forced stratospheric wind anomalies reflection of both wave 1 and wave 2, a strong and zonally symmetric
downwards into the troposphere, however, a complete understanding of polar vortex led to extreme Eurasian weather with anomalously high
the mechanism has yet to be developed (Kodera et al., 1990; Kidston temperatures in early 2020 (Rupp et al., 2022). This indicated the
et al., 2015; Baldwin et al., 2021). The tropospheric response to an generally robust dynamical coupling between stratosphere and tropo­
Arctic SSW is associated with anomalies in the Northern Annular Mode sphere in both directions (Rupp et al., 2022). During the SSW onset,
(NAM) and the North Atlantic Oscillation (NAO) and surface weather wave 2 split the vortex in early January 2021 (Lu et al., 2021). The SSW
anomalies, in particular, with the events of extremely low temperatures 2021 was attributed also to a mixture of the two event types, with the
described in many studies (Baldwin and Dunkerton, 2001; Kidston et al., vortex displaced long enough but not far enough south to be a
2015; Nath et al., 2016; Yu et al., 2018; Karpechko et al., 2018; Charl­ displacement event (Wright et al., 2021). Zhang et al. (2022) emphasize
ton-Perez et al., 2018; Domeisen et al., 2020; Lü et al., 2020; Xie et al., the unique role of downward propagation of the major SSW on the
2020; Zhang et al., 2022). extreme cold wave in East Asia in early January 2021. The results by
Regional changes in the perturbed stratospheric vortex can induce Cohen et al. (2021, 2023) highlight an important type of stratospher­
zonally asymmetric effects in the tropopause and the troposphere and e–troposphere coupling, stratospheric polar vortex stretching, that in­
regionally asymmetric stratosphere–troposphere coupling (Thompson volves wave reflection and stretching of the vortex in the winter of 2021
and Wallace, 2000; Kodera et al., 2016; Hall et al., 2021; Kim and Kim, and is linked with extreme cold across parts of Asia and North America,
2021). SSWs can also influence the troposphere by causing planetary including the February 2021 Texas cold wave. Wright et al. (2021) show
wave reflection (Kodera et al., 2016; Nath et al., 2016; Rupp et al., that the early January SSW in 2021 may have acted as either a trigger or
2022). As shown in Nath et al. (2016), planetary wave packets propa­ an intensifier for several extreme winter weather events affecting
gated upward and were then reflected back down during the SSW 2013, densely populated regions of the NH over the next 2 months. The au­
and led to the formation of extreme cold weather over the central Eur­ thors, however, conclude, that their analysis is not able to explain why
asia, covering Kazakhstan and Northwestern China. different regions may have been impacted at different times during the
It has previously been shown that displacement of the stratospheric SSW evolution.
polar vortex can significantly affect troposphere–stratosphere coupling Another circumstance that plays an important role in the connection
and influence the occurrence of extreme cold events in East Asia (Lu between the troposphere and stratosphere is their coupled thermal state,
et al., 2022; Zhong and Wu, 2023). In the case of vortex splitting events, which is usually realized through the tropopause. On synoptic and
these have promoted tropospheric circulation changes with a negative monthly timescales, tropopause height variations are positively corre­
Arctic Oscillation or NAO pattern that have produced extreme cold lated with tropospheric temperature variations and anticorrelated with
events over Europe (Seviour et al., 2013; Lü et al., 2020). stratospheric temperature variations (Seidel and Randel, 2006). In
Statistical studies show that about a third of observed major SSWs general, the tropopause ascent (descent) is related to a warming (cool­
can be unambiguously classified as split events and another third as ing) in the upper troposphere and a cooling (warming) in the lower
displacement events; for the rest of the events, the polar vortex both stratosphere, as well as to anticyclonic (cyclonic) systems in the tropo­
displaces and splits within a period of several days (Baldwin et al., 2021, sphere (Thompson and Wallace, 2000; Seidel and Randel, 2006). The
and refences therein). The surface weather responses related to the 35 tropopause pressure is closely related to the temperature difference
major SSWs over the period 1958–2010 are stronger following split between lower stratosphere (100 hPa) and middle troposphere (500
events (Lehtonen and Karpechko, 2016). On average, split events (under hPa) (Zängl and Hoinka, 2001).
amplification of wave 2) account for about 50% of SSWs and a surface In this work, we analyze the 2018 Arctic SSW and the short-term
response can first appear within 2–3 days after their onset (Kidston et al., effects in the stratosphere–troposphere coupling that occurred during
2015). the week following the event. As noted above, the SSW 2018 was
It was shown that the stratospheric anomaly can lead the surface attributed to split event (Rao et al., 2020), however, the polar vortex
signal with a lag of around +3 days and surface anomalies can persist up asymmetry evolved from a displacement type before the SSW to a
to 60 days (Domeisen et al., 2020). White et al. (2021) found that on splitting type during the event (Xie et al., 2020). This occurred due to
seasonal timescales (i.e., approximately longer than 3–4 weeks), the alternate amplitude increases in wave 1 (vortex displacement) and wave
tropospheric responses to displacements (due to wave 1 amplification) 2 (vortex split) (Wang et al., 2019a). In February 2018 the largest
and splits are indistinguishable. However, at lags closer to the SSW remnant of the vortex moved to the Western Hemisphere and Europe (Lü
onset, there are clear differences, with splits being associated with a et al., 2020). This contributed to regional differences in the tropospheric
more barotropic NAM structure that extends deep into the troposphere, weather response to the SSW. To extend and update the results of Wang
whereas displacements are associated with a more gradual downward et al. (2019a, 2020), we compare the variability of vertical profiles of
propagation over ~ 10–15 days. Overall, vortex morphology is impor­ temperature and geopotential height anomalies and vertical velocity
tant for determining the tropospheric response (Kidston et al., 2015; omega over the East Asia (Northern China, the Changchun region) and
Lehtonen and Karpechko, 2016; Domeisen et al., 2020). the Eastern Europe (Northern Ukraine, the Kharkiv region) in February
In recent winters, particular variability has occurred in stratospheric 2018. The strong cooling observed in Changchun in January 2018, in the
dynamics and vortex evolution and their resulting surface weather re­ pre-SSW conditions, is considered in the context of stratosphere–­
sponses. As shown by Rao et al. (2020), the SSWs in 2018 (split event) troposphere coupling.
and 2019 (displacement followed by split) differed not only in their
morphology types but also in magnitudes (the former being stronger).
The observed strong SSW in February 2018 is more likely to have a

2
Y. Shi et al. Atmospheric Research 297 (2024) 107112

2. Data and method

We focus on the weather conditions in regions near Changchun,


China (43.9◦ N, 125.3◦ E) and Kharkiv, Ukraine (50.0◦ N, 36.3◦ E) during
January–March 2018. Both stations report to the World Meteorological
Organization (WMO) Global Telecommunications System (GTS), and are
thus available to be assimilated in reanalyses. The latitude/longitude
domains of 48–52◦ N/30–40◦ E (midlatitude Eastern Europe) and
42–46◦ N/120–130◦ E (midlatitude East Asia) for the regions of Kharkiv
and Changchun, respectively, are considered. Temperature and geo­
potential height anomalies in the stratosphere and troposphere are
analyzed based on the NCEP-NCAR reanalysis (NNR) data (Kalnay et al.,
1996) and the ERA5 reanalysis (Hersbach et al., 2020). Temperature
anomalies in the NNR and ERA5 time series and latitude–longitude grid
are presented relative to the 1991–2020 climatology. We first created
time–latitude, latitude–pressure and longitude–pressure cross sections
of the temperature anomalies with the climate plotting and analysis
tools available in NNR (https://psl.noaa.gov/data/getpage/). The re­
sults from the NNR data were then verified with the ERA5 data. In
addition, polar maps of the temperature in the stratosphere and tropo­
sphere are compared. Vertical velocities (omega) are also presented to
examine possible stratosphere–troposphere coupling. The tropopause
pressure data were obtained from the NCEP Global Data Assimilation
System GDAS at https://psl.noaa.gov/data/gridded/data.ncep.html.
The reanalysis datasets were downloaded as netCDF files and plotted in
the same graphical representation using the NASA Panoply Data Viewer
(https://www.giss.nasa.gov/tools/panoply/). The local temperature
variations are based on measurements at weather stations in Changchun
and Kharkiv shown by red curves in Fig. 1 (data from https://meteostat.
net/ for Changchun (WMO 54161) and Kharkiv (WMO 34300) weather
stations: https://meteostat.net/en/station/54161 and https://meteostat
.net/en/station/34300).
Fig. 1. Daily temperatures observed at weather stations (a) Changchun in East
Asia and (b) Kharkiv in Eastern Europe (red curves) in January–March 2018
3. Results compared with those obtained using the NNR (green curves) and ERA5 (blue
curves) reanalyses. Red vertical lines show the time of the SSW 2018 event.
3.1. Evolution of surface air temperatures These dates February 10 – March 01 were determined based on the reversal of
the 10 hPa zonal wind as used in (Wang et al., 2019a). Black (red) upward
The observed surface temperature is compared with temperature arrows 1–4 indicate significant cooling (warming). Downwards dashed arrows
from reanalyses NNR and ERA5 shown by green and blue curves, mark a temperature drop during strong cooling events. Black dashed curve
respectively, in Fig. 1. It is seen from Fig. 1 that the temperature vari­ shows daily climatology 1991–2020. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of
ations from reanalyses are very similar to the observed ones. Correla­
this article.)
tions between three datasets are r = 0.94–0.99 and differences between
them are shown in Table 1. Differences indicate that, in January–March
2018, NNR overestimated (underestimated) observed values by 1.2 ± Table 1
1.6 ◦ C (− 2.3 ± 1.5 ◦ C) at Kharkiv (Changchun). ERA continuously Difference between daily surface temperatures (in ◦ C), observed at Changchun
overestimated observations by 1.1 ± 1.0 ◦ C (Changchun) and 1.8 ± and Kharkiv, and reanalyses NNR and ERA5 in January–March 2018.
1.1 ◦ C (Kharkiv). These differences are to some extent influenced by the
Difference Changchun Kharkiv
topography over the broader area represented by the reanalyses. The
NNR minus obs − 2.3 ± 1.5 1.2 ± 1.6
mean invariant geopotential height in the reanalysis grid boxes closest
ERA5 minus obs 1.1 ± 1.0 1.8 ± 1.1
to the stations is within approximately <100 m of the altitude of each ERA5 minus NNR 3.4 ± 1.2 0.6 ± 1.1
site. So, the temperature lapse rate, which is on the order of − 1 ◦ C for
each 100 m of altitude gained (e.g., Marshall and Plumb, 2008), likely
accounts for at least some of each temperature difference when − 16 ◦ C was observed in Kharkiv on 24–27 February, during the SSW
compared with the WMO GTS data. We examine changes in temperature (cooling 3, red curve in Fig. 1b). The temperature drop was 18 ◦ C in 5
over time, which are equally well represented in both reanalyses and days in Changchun and 14 ◦ C in 8 days in Kharkiv (downwards dashed
strongly correlated with observations. arrows in Fig. 1). Cooling 4 at both sites occurred during the SSW re­
Since we pay the main attention to regional cooling events, four cases covery phase in March. The general tendency of seasonal temperature
of temperature decrease at each of the stations are indicated by black increase in Kharkiv was observed in late February – March, however the
upward arrows 1–4 in Fig. 1 (hereinafter referred to as cooling 1 – temperature remained noticeably below climatological values (black
cooling 4). The two cases of significant temperature increase in dashed curve in Fig. 1b). This anomalous cooling by about 5–10 ◦ C
Changchun are marked by red upward arrows (Fig. 1a). In general, the suggests the influence of SSW on seasonal time scale, which is not clearly
temperature variability at the stations was different. In Changchun, a visible in Changchun (Fig. 1a).
strong minimum of − 26 ◦ C was observed on 24–25 January, in the pre- In this work, we present the manifestations of temperature anomaly
SSW period (cooling 2, red curve in Fig. 1a). Then the temperature changes in vertical time/latitude/longitude representations to deter­
increased relatively monotonously following the annual cycle of air mine whether surface anomalies can be related to stratospheric ones. A
temperature (black dashed curve in Fig. 1a). In contrast, a minimum of comparative analysis of temperature, geopotential height, and vertical

3
Y. Shi et al. Atmospheric Research 297 (2024) 107112

velocity data is performed in terms of regional stratosphere–troposphere The polar vortex in January–February 2018 was generally shifted to
coupling in Eastern Europe and East Asia. the Western Hemisphere and Europe and split in two pieces with the
SSW onset: one centered over Canada, and the other migrating westward
from central Eurasia to western Europe between 11 and 15 February (Lü
3.2. Time–latitude temperature anomalies in January–March 2018
et al., 2020). Due to the asymmetry of the vortex relative to the pole and
the 6◦ latitude difference between the stations, the warm anomaly
Fig. 2 shows latitudinal changes of regional temperature anomalies
contour at 8 ◦ C in the lowermost stratosphere (T200; “lowermost
in January–March 2018. The longitude segments 30–40◦ E (Kharkiv re­
stratosphere” in terms of Holton et al. (1995)) extends south of 50◦ N
gion, left column in Fig. 2) and 120–130◦ E (Changchun region, right
(Fig. 2c) but does not quite reach 44◦ N (Fig. 2h). This suggests a possible
column) are compared. Anomalies at four pressure levels of 10, 100, 200
difference in the expected weather impact of the SSW at the two
and 500 hPa (panels T10, T100, T200 and T500, respectively) and at the
observation sites. In the troposphere, synoptic scales dominate in both
surface (panel T-surf) are presented.
the lifetime and the latitudinal extent of positive and negative anoma­
The SSW event is clearly seen in T10, T100 and T200 (Fig. 2) as the
lies, which thus appear more fragmented (from a few days to a week,
abrupt stratosphere warming between about 10 February and 1 March
panels T500 and T-surf) than in the stratosphere. All four coolings 1–4 in
2018 (red vertical lines), in correspondence with earlier studies (Rao
surface temperature, shown in Fig. 1, can be identified in time–latitude
et al., 2018; Wang et al., 2019a; Butler et al., 2020; Rao et al., 2020; Lu
cross sections as negative temperature anomalies in T-surf at station
et al., 2021; Rao et al., 2021). The strongest anomaly of >18 ◦ C appears
latitudes (upward arrows 1–4 and rectangles in Fig. 2e and j). It is also
in the middle stratosphere (panel T10) and decreases to 12 ◦ C and 8 ◦ C in
seen from Fig. 2e and j that while coolings 1–2 in January (pre-SSW
the lower stratosphere (panels T100 and T200, respectively). It is seen
period) are approximately close in time at both stations, coolings 3–4
that time–latitude extension of warm anomaly in the Changchun
differ by a few weeks. Only cooling 3 in the Kharkiv region, which is the
segment (Fig. 2f–h) is less than in the Kharkiv region (Fig. 2a–c).

Fig. 2. The NH time–latitude air temperature anomalies at the pressure levels 10, 100, 200, 500 hPa and at the surface (panels T10, T100, T200, T500 and T-surf,
respectively) in January–March 2018, averaged in longitude segments (a–e) 30–40◦ E centered in Kharkiv and (f–j) 120–130◦ E centered in Changchun. The latitude
range covers the Northern Hemisphere, and station latitudes of 50◦ N (Kharkiv, left) and 44◦ N (Changchun, right) are marked by dashed horizontal lines. The red
vertical lines indicate the duration of SSW 2018. The black (red) upward arrows in (e) and (j) are placed at the time of cold (warm) surface anomalies at the
corresponding latitude. The strongest cold (warm) anomalies are shown by thick solid (dashed) contours marked in ◦ C for T10, T100 and T200; for T500 and T-surf,
contours are shown for ±5 ◦ C anomalies. Temperature anomalies are based on the NNR climatology 1991–2020. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

4
Y. Shi et al. Atmospheric Research 297 (2024) 107112

strongest (up to − 12 ◦ C, white rectangle in Fig. 2e; compare to − 14 ◦ C propagates downward and equatorward, which is visible also over the
by local observation in Fig. 1b) among eight cases, coincides with the Kharkiv region (Fig. 3, upper panel). In both cases, stratospheric warm
SSW, falling at the end of the event in late February. The strongest anomalies of 4–8 ◦ C extend to midlatitudes across station latitudes
cooling 2 in the Changchun region was between − 5 ◦ C and − 10 ◦ C (upper panels in Figs. 3 and 4), but they are concentrated at the higher
(white rectangle in Fig. 2j) and occurred in the second half of January latitudes over the Changchun longitude segment (upper panel in Fig. 4).
(pre-SSW period). It was associated with the intrusion of cold polar air as The sensitivity of the latitudinal profiles of the tropopause to thermal
seen from latitudinal and temporal shift of corresponding negative anomalies in the lowermost stratosphere is quite noticeable (dashed
anomaly. The latter was located at polar cap latitudes during the first curves in Figs. 3 and 4, upper panels). As the lower edge of the warm
half of January and moved equatorward to the polar vortex edge and anomaly descends, the tropopause sags and reaches its lowest height. In
into the midlatitudes during the cooling 2 event (Fig. 2j). Note that the the Kharkiv segment at latitudes near 50◦ N, tropopause height varies
climatological mean latitude of the vortex edge defined by the westerly between pressure levels of 300 hPa and 250 hPa, or about 9 km and 10
maximum is about 60◦ N (Waugh et al., 2017). Similar evolution is also km on 20 and 22 February (Fig. 3a and b). It decreases then to the 400
observed in the appearance of coolings 1 and 3 at 44◦ N (Fig. 2j) and hPa (~7 km) pressure level on 24 and 26 February (Fig. 3c and d),
cooling 4 at 50◦ N (Fig. 2e). Therefore, cooling 3 in the Kharkiv region lowering by 2–3 km during a week. Note that the typical height of the
does not appear associated with polar intrusion and looks like mostly as winter tropopause in this region is 9–10 km (Zängl and Hoinka, 2001;
a midlatitude effect directly related to the SSW in February 2018. Seidel and Randel, 2006). On the same dates, the cold anomaly at the
Cooling 4 in the Changchun region arose immediately after the SSW surface intensifies from − 4 ◦ C to − 12 ◦ C. A similar interaction of the
event (Fig. 2j), did not show an obvious connection with the polar air positive stratospheric anomaly with the tropopause descent and surface
mass and can be considered to have a connection with the regional temperature decrease can be observed in the Changchun segment
stratospheric influence. (Fig. 4a–e). However, the cooling of the troposphere near the latitude of
Note that the two positive T-surf anomalies at Changchun latitude in the station was less strong, between − 4 ◦ C and − 8 ◦ C on 20, 22 and 24
March (red upward arrows in Fig. 2j), are associated with two warmings February (Fig. 4a–c). The cooling here was even weaker near the surface
of up to 7 ◦ C and 16 ◦ C in Changchun in March (red upward arrows and (0 ◦ C to − 4 ◦ C; Fig. 4, upper panel). This difference between regions
red curve in Fig. 1b). during the SSW corresponds to the difference in T-surf anomalies at
station latitudes (Fig. 2e and j, time interval between red vertical lines).
3.3. Temperature vertical latitude and longitude variability during SSW The stratosphere–tropopause–troposphere coupling, which is visible
from the longitudinal profiles, was the strongest in the Kharkiv region on
Variability of vertical profiles of temperature anomalies in the lat­ 26 February (Fig. 3i). On this date, the surface negative anomaly was as
itudinal and longitudinal directions is presented in Fig. 3 and Fig. 4 for strong as in the latitudinal profile: − 8 ◦ C to − 12 ◦ C. The longitude
the Kharkiv and Changchun regions, respectively. The SSW period is distribution of the surface anomaly in the Changchun region shows
shown every other day from 20 to 28 February, which corresponds to the weaker cooling, 0 ◦ C to − 4 ◦ C (Fig. 4, lower panel), consistent with the
surface cooling 3 in Fig. 2e (white rectangle). Upper panels show that results for latitudinal changes in Fig. 4 (upper panel) and T-surf in
the latitudinal gradient in the positive stratospheric anomaly is steeper Fig. 2j, as well as with the weak temperature oscillations observed in
over the Kharkiv region (Fig. 3a–e) than over the Changchun region Changchun during SSW (Fig. 1a, red curve between red vertical lines).
(Fig. 4a–e). In the latter case, this is the result not only of a shift of the Longitudinally, vertical cross sections also show weaker positive
polar vortex to the Western Hemisphere noted in Section 3.1, but also of anomalies in the stratosphere (Fig. 4, lower panel) than in the Kharkiv
the diagonal pattern in the T-anomaly distribution in Fig. 4 (upper region (Fig. 3, lower panel) due to the vortex dynamics. As noted in
panel) with a poleward tilt between the lower and middle stratosphere. Section 3.1, the polar vortex in February 2018 was generally shifted to
This pattern indicates that the positive stratospheric anomaly the Western Hemisphere and Europe and this may explain the difference

Fig. 3. Daily vertical profiles of air temperature anomalies in the troposphere–stratosphere in (a–e) latitude–pressure cross section (0–90◦ N, 1000–10 hPa), averaged
over the segment 30–40◦ E, and (f–j) longitude–pressure cross section (0–180◦ E, 1000–10 hPa), averaged in the zone 48–52◦ N; both ranges are centered on Kharkiv
coordinates (50◦ N, 35◦ E), indicated by vertical lines and open circles on the horizontal axis. Contour intervals are 4 ◦ C. The thick dashed curve shows the thermal
tropopause pressure. Temperature anomalies are based on the NNR climatology 1991–2020.

5
Y. Shi et al. Atmospheric Research 297 (2024) 107112

Fig. 4. Daily vertical profiles of air temperature anomalies in the troposphere–stratosphere in (a–e) latitude–pressure cross section (0–90◦ N, 1000–10 hPa), averaged
over the segment 120–130◦ E, and (f–j) longitude–pressure cross section (0–180◦ E, 1000–10 hPa), averaged in the zone 42–46◦ N; both ranges are centered on
Changchun coordinates (44◦ N, 125◦ E), indicated by vertical lines and open circles on the horizontal axis. Contours for positive (negative) anomalies are shown by
solid (dashed) lines with a step of 4 ◦ C. The thick dashed curve shows the thermal tropopause pressure. Temperature anomalies are based on the NNR clima­
tology 1991–2020.

in the stratospheric influence on the tropopause level and surface line in Fig. 5), the tropopause shifted downward from 230 hPa (Fig. 5a)
weather in the Northern Ukraine than in the Northern China. to 400 hPa (Fig. 5b), or from ~10 km to ~7 km over five days. At the
Note that the tropopause elevates to about 100 hPa (16 km) over the same time, the surface T-anomaly decreased from 0 ◦ C (that is, tem­
warm tropics and descends to about 300 hPa (9 km) in the polar region, perature was close to the climatological mean) on 19 January to − 12 ◦ C
which is clearly visible from the meridional profiles in Figs. 3 and 4 on 24 January (Fig. 5a and b, respectively). Cooling 2 in Fig. 2j was
(dashed curves in upper panels) and is consistent with the climatology associated with the intrusion of cold polar air into midlatitudes, and the
(Seidel and Randel, 2006). surface negative anomaly in Fig. 5 also shows similar equatorward
Positive stratospheric T-anomalies in Figs. 3 and 4 are significantly displacement from about 60◦ N (Fig. 5a) to 40–50◦ N (Fig. 5b). It is
larger in magnitude and size than the tropospheric ones and appear interesting that the positive stratospheric anomaly crosses the tropo­
dynamically more active. They cause anomalous descent of the tropo­ pause and penetrates into the subtropical troposphere (Fig. 5b) on the
pause, which is accompanied by troposphere cooling. Analyzing the pre- date of a strong temperature minimum of − 26 ◦ C in Changchun (cooling
SSW period, it was found that stratospheric disturbances could also 2 in Fig. 1a, red curve).
contribute to strong surface cooling observed in Changchun in January It is necessary to make some clarifications regarding estimates of the
2018 (cooling 2 in Fig. 1a). Fig. 5 demonstrates how the sharp move­ degree of surface cooling. On the date of extreme cooling, the negative
ment of the positive stratospheric anomaly downward (between the surface anomaly reached − 12 ◦ C in the Changchun region (Fig. 5b) and
middle and lowermost stratosphere) and equatorward (between the in the Kharkiv region (Fig. 3d and i), both covering the 4◦ × 10◦ lat­
polar and midlatitudes) from 19 to 24 January is accompanied by itude–longitude domains. The observed temperature drop was between
changes in the height of the midlatitude tropopause and the temperature − 8 ◦ C and − 26 ◦ C (by 18 ◦ C in 5 days) in Changchun in January and
of the troposphere. At the Changchun latitude (open circle and vertical between − 2 ◦ C and − 16 ◦ C (by 14 ◦ C in 8 days) in Kharkiv in February
(downwards dashed arrows in Fig. 1a and b, respectively). Therefore, a
rapid decrease in the height of the tropopause by 2–3 km in 5–8 days is
associated with a decrease in the local surface temperature by 18 ◦ C in
Changchun and by 14 ◦ C in Kharkiv. According to the larger extent and
magnitude of the stratospheric anomalies, as well as their higher dy­
namic activity, the strong winter coolings may be contributed by the
downward forcing caused by the rapid heating of the lowermost
stratosphere with a significant lowering of the tropopause above each of
the two stations (see Sections 3.4–3.6 for more details and Section 4 for
discussion). The noticeable strengthening of the negative tropospheric
anomaly towards the surface is also noteworthy (Fig. 3d and i and
Fig. 5b). This effect could be related to the so-called “surface amplifi­
cation” of the stratospheric signal observed in pressure over the polar
cap (Baldwin et al., 2019, 2021). Such a relation with respect to mid­
latitude tropospheric temperature can be analyzed separately in more
detail.
The results of Figs. 3–5 based on the NNR data are well reproduced
Fig. 5. The NH latitude–pressure cross sections of air temperature anomalies in
using the ERA5 data, both qualitatively and quantitatively (Supple­
the Changchun segment (120–130◦ E) on (a) 19 January and (b) 24 January
2018, which illustrate stratosphere–troposphere coupling in a case of cooling 2
mental Figs. S1–S3), in accordance with the general consistency of
in Fig. 1a. Open circle and vertical line indicate Changchun latitude 44◦ N. reanalyses in Fig. 1 and Table 1.
Temperature anomalies are based on the NNR climatology 1991–2020.

6
Y. Shi et al. Atmospheric Research 297 (2024) 107112

3.4. Temperature extremes on polar maps anomalies in the troposphere and at the surface in January in Fig. 2i and
j and Fig. 5. In contrast to the pre-SSW period, during SSW there is a
Next, we compare the geographic distribution of stratospheric and clear spatial anticorrelation between the extremes in the T-fields in the
tropospheric temperatures in pre-SSW and SSW cooling events using lowermost stratosphere and middle troposphere (Fig. 6c and d). The
polar maps based on the NNR data. Surface temperature minimum in temperature maximum of − 48 ◦ C in T200 (Fig. 6c), reproduced on the
Changchun on 24 January (cooling 2 in Fig. 1a) appeared under zonally polar map for T500 (dashed contour in Fig. 6d), coincides quite closely
asymmetric temperature extremes in the lowermost stratosphere (T200 with the temperature minimum of − 40 ◦ C. Both extremes are zonally
in Fig. 6a). The areas of lowest and highest temperature (red contour orientated and this stratosphere–troposphere interaction during SSW
− 60 ◦ C and black contours − 48 ◦ C, respectively) are located on opposite can contribute to the temperature minimum in both the troposphere
sides of the pole, approximately around meridians 0◦ E–180◦ E. One of over Kharkiv (white circle in Fig. 6d) and at the surface (cooling 3 in
the localized temperature maxima is just above Changchun (white circle Fig. 1b). As emphasized above in Subsection 3.2 (Fig. 2), cooling 3 in the
inside black contour − 48 ◦ C). In the middle troposphere (T500 in Kharkiv region does not appear associated with polar air intrusion, but is
Fig. 6b), the cold area enclosed by the red contour − 30 ◦ C covers the closely coupled with the cold air mass in the vortex edge region near
whole polar region and undulates over the middle latitudes. The tem­ 60◦ N. This conclusion is supported by the limited latitudinal extent of
perature minimum inside the contour of − 40 ◦ C stretches in a strip the elongated T500 minimum of − 40 ◦ C, which does not cover to polar
across the pole along meridians 120◦ E–60◦ W (Fig. 6b) falling roughly region (Fig. 6d).
between the temperature extremes in T200 (Fig. 6a). The main property that distinguishes T-minima in the troposphere is
Geographically, the T200 and T500 temperature anomalies in the their meridional and zonal orientation in pre-SSW and SSW conditions
pre-SSW period appear to be formed independently. Nevertheless, as seen from the T-fields in Fig. 6b and d, respectively, with an
Changchun is locally falls under two opposite regional extremes in the approximately mirror image of the stratospheric T-maximum in the
lowermost stratosphere (temperature maximum) and middle tropo­ latter case. Although the large-scale polar-to-midlatitude linkage is
sphere (temperature minimum). This is consistent with altitudinal different, the regional stratosphere–troposphere coupling in both cold
alternation of temperature anomalies at the station latitude (contours weather events is generally similar. It is further confirmed by the
±8 ◦ C around vertical line in Fig. 5b). On the other hand, meridional structure of vertical air movements and the distribution of geopotential
elongation of the T500 minimum − 40 ◦ C in Fig. 6b reaching Changchun height anomalies in Subsection 3.5.
reflects the intrusion of cold polar air into the midlatitudes Similar effect
was observed from the latitudinal migration of negative temperature 3.5. Vertical movements during cooling events

Fig. 7 illustrates stratosphere–troposphere coupling in the two


cooling events using the pressure vertical velocity and air temperature
cross sections in the NH meridional plane averaged in the corresponding
segments. To compare changes in vertical profiles depending on weather
conditions, pairs of days with the coldest and previous warmest tem­
perature were selected based on the observations (Fig. 1). The compared
dates are January 15 and 24, 2018 for the Changchun region (Fig. 7, left)
and February 15 and 26, 2018 for the Kharkiv region (Fig. 7, right). If
approximately zero omega values at the station latitudes are observed in
the warm days (Fig. 7a and e), then positive values of 0.08–0.15 Pa s− 1
(Fig. 7b) and 0.05–0.10 Pa s− 1 (Fig. 7f) are presented by the omega
contours in the cold days at corresponding vertical lines. Increasing
pressure with time indicates descending movements between the
lowermost stratosphere (200–300 hPa, just above tropopause shown by
dashed curve) and the troposphere throughout its depth. This descent
can favor the penetration of air particles that are colder than in the
surrounding areas (Fig. 6b and d) into the lower tropospheric layers all
the way to the surface, causing cool weather.
The midlatitude omega maxima in their latitudinal location closely
coincide with the positive temperature anomalies in the lowermost
stratosphere (Section 3.3). Here, the tropopause shows a sharp descent
with a steeper meridional gradient just near the latitude of each station
(dashed curves in Fig. 7b and f). The lowered tropopause heights,
accompanied by downward motion of air (Fig. 7b and f), resemble the
conditions of stratospheric intrusions or tropopause folds (Holton et al.,
1995; Kim et al., 2001; Škerlak et al., 2015; Barnes et al., 2022; Xiong
et al., 2022). However, the stratospheric intrusions usually penetrate
into the middle troposphere and do not reach the surface, except for
Fig. 6. The NH temperature on (a, b) 24 January and (c, d) 26 February 2018 extremely cold air outbreak (Xiong et al., 2022) or deep tropopause fold
in (a, c) the lowermost stratosphere (T200) and (b, d) the middle troposphere (Škerlak et al., 2015) events. Circulation in the vertical plane includes
(T500) in polar projection. Temperature fields compare conditions for the time neighboring ascending branches located further south in the tropics or
of appearance of local temperature minimum in Changchun (pre-SSW period,
subtropics (blue area in Fig. 7b and f, respectively). This pattern looks
left) and Kharkiv (during SSW, right). Red and black contours outline the areas
like the midlatitude Ferrel cell in the tropospheric circulation (Liu,
of temperature minima and maxima, respectively. Black dashed contour in (d)
reproduces temperature maximum − 48 ◦ C in (c). White circles indicate location 2021), but demonstrates opposite direction. It is possible that the cool­
of the stations and the thick solid circle shows the latitude 60◦ N. Temperature ing events are associated with the formation of a characteristic circu­
fields are derived from the NNR data. (For interpretation of the references to lation pattern in the meridional plane, as suggested by the results below.
colour in this figure legend, the reader is referred to the web version of The tropopause descent and air downwelling in the NH midlatitudes
this article.) are closely related to the vertical thermal transformation of the

7
Y. Shi et al. Atmospheric Research 297 (2024) 107112

Fig. 7. The NH latitude–pressure cross sections of (a, b, e, f) vertical velocity omega (Pa s− 1) and (c, d, g, h) air temperature in (left) the Changchun segment
(120–130◦ E) and (right) the Kharkiv segment (30–40◦ E). The dates of the warmer and coldest weather conditions based on Fig. 1 are compared. Open circles and
vertical lines indicate Changchun (44◦ N) and Kharkiv (50◦ N) latitudes. The dashed curve shows the thermal tropopause pressure. The tropospheric layers associated
with the surface cooling events are outlined by thin white contours: − 20 ◦ C to − 30 ◦ C in (c, d) and − 10 ◦ C to − 20 ◦ C in (g, h) and thick black vertical and dashed
horizontal lines show the change in altitudinal location of these layers over Changchun and Kharkiv; temperature extremes are highlighted by signed contours. The
yellow curves in (g, h) indicate the latitude–altitude change in the position of the temperature minimum. The NNR data were used. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

atmospheric layers, which contributes to regional cooling at the surface. surface temperature was warmer, about − 5 ◦ C (Fig. 1b, red curve).
Meridional temperature profiles between the dates of warmer and colder Thus, regardless of stratospheric vortex conditions, strong regional
surface weather show a 3–5 km descent of midtropospheric air to the cooling in the troposphere is accompanied by the descent of cold mid­
surface (Fig. 7c, d, g, h). In the Changchun segment, the temperature tropospheric air to the surface. At the Changchun latitude, the vertical
layer of − 20 ◦ C to − 30 ◦ C (thin white contours in Fig. 7d) was respon­ velocity estimated from the 9-day altitude decrease of the selected
sible for the strong temperature minimum of − 26 ◦ C observed in troposphere layer (black vertical lines in Fig. 7c and d) was approxi­
Changchun on 24 January 2018 (cooling 2 on red curve in Fig. 1a). This mately 0.09 Pa s − 1. This value corresponds to the omega values at the
layer was much higher at Changchun on 15 January 2018 (500–600 hPa, Changchun latitude on 24 January (0.08–0.15 Pa s − 1, Fig. 7b), on the
or about 4–5 km altitude; thin white contours − 20 ◦ C and − 30 ◦ C, black day of intensified air downwelling. The average estimated velocity for
vertical and dashed horizontal lines in Fig. 7c), when warmer temper­ the 11-day interval at the Kharkiv latitude was approximately 0.06 Pa s
− 1
ature − 10 ◦ C was observed (Fig. 1a, red curve). Similarly, the tem­ (Fig. 7g and h) and this value falls within the range of 0.05–0.10 Pa s
− 1
perature layer of − 10 ◦ C to − 20 ◦ C in the Kharkiv segment (30–40◦ E, on 26 February (Fig. 7f). The linear downward velocity of cold air
thin white contours in Fig. 7g and h) was responsible for the cold tem­ between the midtroposphere and the surface occurred at a velocity of
perature of − 16 ◦ C in Kharkiv on 26 February 2018 (cooling 3 on red about 0.5 km day− 1 and 0.3 km day− 1 over the Changchun and Kharkiv
curve in Fig. 1b). It was at pressure levels 600–700 hPa (about 3–4 km) regions, respectively.
on 15 February2018 (thin white contours − 10 ◦ C and − 20 ◦ C, black Meridional temperature profiles in Fig. 7d and h are in close agree­
vertical and dashed horizontal lines in Fig. 7g), when the observed ment with those in (Xiong et al., 2022, their Fig. 9, left) for the extreme

8
Y. Shi et al. Atmospheric Research 297 (2024) 107112

cooling event in late January 2019 based on the satellite and reanalysis larger during February 2018 SSW event (Fig. 8c and d) than in January
data. Xiong et al. (2022) provided direct evidence from satellite obser­ 2018 pre-SSW conditions (Fig. 8a and b). This relationship corresponds
vations to confirm the impact of the downward transport on the cold air to larger positive T-anomalies in the stratosphere in Figs. 3 and 4 (upper
outbreak process. In addition, our results in Fig. 7 allow to identify the panels) compared to Fig. 5. The downward and equatorward motions of
altitudinal location of the midtropospheric layer, the lowering of which positive and negative Z-anomalies in both events are similar, but they
down to the surface contributes to the extreme surface cooling. are a more noticeable in the Changcun segment (Fig. 8, upper panel)
Lowering this layer looks like pushing it down to the surface in condi­ than in the Kharkiv segment (Fig. 8, lower panel). This difference ap­
tions of consistent patterns in downward movement of the warm pears to be related to a change in the spatial temperature distributions.
stratospheric anomaly, the tropopause descent and general air mass In January 2018, Changchun fell under a positive anomaly that rapidly
downwelling through the troposphere (Figs. 3–5 and Fig. 7). The change descended from the polar stratosphere to the midlatitude tropopause
in the shape of the isolines in Fig. 7c, d, g and h can be considered as a (Fig. 5), with vertical and meridional components of motion. A merid­
large-scale deformation field observed during stratospheric intrusions ionally extended temperature minimum was observed in the tropo­
and tropopause folds (Holton et al., 1995). The dynamics of the strato­ sphere (Fig. 6b). The zonally oriented temperature anomalies over
spheric polar vortex in winter appears to be a determining factor Kharkiv in February 2018 (Fig. 6c and d) were probably less favored for
influencing surface weather through a chain of stratosphere–tropo­ the meridional component of movement in the lowermost stratosphere
sphere coupling processes, which are generally known from much pre­ and midtroposphere.
vious works (Thompson et al., 2002; Haigh et al., 2005; Kidston et al., Negative Z-anomaly in the Changchun region (Fig. 8a and b) reflects
2015; Charlton-Perez et al., 2018; Domeisen and Butler, 2020; Xiong a deepening of the East Asia major trough, which is a climatological
et al., 2022). In our analysis, we tried to quantify some elements of this midtropospheric cyclonic structure of low pressure (e.g. Wang et al.,
chain. 2010). On the western side, this structure causes northerly winds, which
promote the penetration of cold air masses from higher latitudes into the
middle latitudes and are accompanied by surface cooling (Wang et al.,
3.6. Geopotential height anomalies and surface winds 2010). Below the negative Z-anomaly, the surface low-pressure system
can be identified by counterclockwise circulation around its center L
Fig. 8 shows the latitude–pressure cross sections of geopotential marked on the surface vector wind map (Fig. 9a). The wind direction
height anomalies in the two cooling events occurred in East Asia (top) indicates the presence of northwesterlies (NW) in the Changchun region
and Eastern Europe (bottom). In general, the magnitude and lat­ on 24 January 2018 (thick curved arrow in Fig. 9a). The NW wind in this
itude–altitude extents of positive Z-anomalies in the stratosphere are region is on the western side of the low-pressure system L (Fig. 9a),
which is consistent with the winter climatology for the East Asia region

Fig. 9. (left) The surface vector wind and (right) surface temperature in (a, b)
East Asia (20–70◦ N, 100–160◦ E) on 24 January 2018 and (c, d) Eastern Europe
Fig. 8. Latitude–pressure cross sections of geopotential height anomalies (Z- (20–70◦ N, 10–70◦ E) on 26 February 2018. Black vertical and horizontal lines
anomalies, in m) in (a, b) the Changchun segment (120–130◦ E) on 19 and 24 indicate the coordinates of (top) Changchun and (bottom) Kharkiv, and the
January 2018 (for the dates in Fig. 5a and b, respectively) and (c, d) the Kharkiv curved arrows show the prevailing NW and NE wind directions over the station
segment (30–40◦ E) on 22 and 26 February 2018 (for the dates in Fig. 3b and d, regions, respectively. The letter “L” in (a, c) indicates the center of the low-
respectively). Open circles and vertical lines indicate Changchun latitude 44◦ N pressure system, with its characteristic counterclockwise circulation, as seen
in (a, b) and Kharkiv latitude 50◦ N in (c, d), and contour intervals are ±82.5 m from the vector directions. Temperature minima (maxima) in (b, d) are outlined
and ± 75.0 m, respectively. The dashed curve shows the thermal tropopause by signed white (black) contours and zero temperature is indicated by a
pressure. Z-anomalies are based on the NNR climatology 1991–2020. dashed contour.

9
Y. Shi et al. Atmospheric Research 297 (2024) 107112

(Wang et al., 2010, their Fig. 1). This wind direction contributes to the the same time the observed surface temperature decreased by 18 ◦ C
advancement of the region of negative temperatures into the middle (cooling 2 in Changchun, Fig. 1a) and 14 ◦ C (cooling 3 in Kharkiv,
latitudes, as can be seen from the contour of zero temperature (dashed Fig. 1b). The tropopause descent to about 350–400 hPa (7–8 km) has
curve in Fig. 9b). It is possible to conclude about the regional influence been observed in the events of stratospheric intrusions and tropopause
of the surface temperature minimum in the vortex edge region (~60◦ N) folds (Li et al., 2015; Škerlak et al., 2015; Xiong et al., 2022). The
on cooling in the midlatitude East Asia (Fig. 9b). Likewise, the north­ stratosphere warming during the SSW events generally decrease the
easterly (NE) winds in the Kharkiv region (thick curved arrow around tropopause height in the polar regions by 1.5–2 km and the stronger the
the low-pressure center L in Fig. 9c) contribute to surface cooling on 26 event is, the larger the influences of SSW events on the tropopause layer
February 2018 (Fig. 9d). The minimum temperature in the Changchun are (Wang et al., 2023a). In this work, we described the vertical
segment is about 10◦ lower, and negative temperatures are generally sequence of processes that connect anomalous warming of the strato­
colder than in the Kharkiv segment (white contours and dark and light sphere with the tropopause descent, downward air movements in the
blue in Fig. 9b and 9d, respectively). This explains the stronger observed stratosphere–troposphere and the extreme surface cooling.
cooling 2 in Changchun (Fig. 1a) compared to cooling 3 in Kharkiv In both regional cooling events, the clear vertical air movements
(Fig. 1b). Note that the location of the cyclone centers L and the struc­ between the lowermost stratosphere throughout the troposphere down
ture of the surface vector winds in Fig. 9a and 9c generally correspond to to the surface occurred (Fig. 7b and f). A large contribution to the surface
the location and shape of the low-pressure systems on the sea level cooling is made by the downward displacement of the air of the middle
pressure maps (not shown). troposphere (Fig. 7c, d, g, h). In undisturbed conditions, this atmo­
As can be estimated from Fig. 7d and h, the latitudinal minimum spheric layer is several tens of degrees colder than at the surface due to
temperature near the surface at ~60◦ N was about 5 ◦ C colder than the the negative vertical thermal gradient. Typically observed lapse rate is
temperature around each of the stations. At the same time, the meridi­ about 6.5 ◦ C km− 1 (Stone and Carlson, 1979), and the troposphere at an
onal component of the surface wind (v-wind) near the stations on the altitude of 3–5 km usually contains air with a temperature of about
extreme cold days ranged from − 1 m s− 1 to − 5 m s− 1 (not shown). This − 20 ◦ C to − 30 ◦ C. Being pushed down to the surface (Fig. 7d and h), this
results in a combination of relatively low temperature gradients with layer is capable of causing extreme cold weather. A nearly vertically
weak northerlies, suggesting a comparatively smaller contribution of aligned temperature minimum is in the troposphere (yellow curves in
meridional transport of cold air to surface cooling than the much colder Fig. 7d and h) and the T-contours here significantly deformed and
air downwelling from the 3–5 km tropospheric layer (thin white con­ deflected downward towards the surface. Such the change in the shape
tours in Fig. 7d and h). Recall, that the extreme cooling in these indi­ of the isolines in Fig. 7 indicates a significant role of vertical thermal
vidual events was by 14 ◦ C at Kharkiv and 18 ◦ C at Changchun (Section coupling in the formation of extreme surface cooling in the midlatitudes.
3.1; downwards dashed arrows in Fig. 1). However, a more reliable This process, to quote Holton et al. (1995, p. 421), “may also to some
quantification of the relationship between the contribution of vertical extent be regarded as the result of the systematic effect of the large-scale
and horizontal thermal advection in extremely cold weather needs to be deformation field”, inherent in the formation of stratospheric intrusions
made in future studies. and tropopause folds.
During cooling 2 in Changchun (Fig. 1a), which occurred in the pre-
4. Discussion SSW conditions, such a downward influence was especially noticeable
(Fig. 5, Fig. 7b, c, d). Rapid downward/equatorward propagation of the
4.1. Downward motion positive temperature anomaly occurred in a large latitudinal range from
the polar middle stratosphere into the subtropical middle troposphere
We have described the regional manifestations of stratospheric (Fig. 5). It was accompanied by significant lowering of the midlatitude
forcing in individual surface cooling events. Temperature variations in tropopause by about 3 km, air descent between the lowermost strato­
East Asia (Changchun, Northern China) in January 2018 (the pre-SSW sphere and troposphere up to the surface with omega velocity of about
period) and in Eastern Europe (Kharkiv, Northern Ukraine) in 0.1 Pa s− 1 (Fig. 7b) and strong surface cooling with negative anomaly of
February 2018 (the SSW period) appear to be largely influenced by − 12 ◦ C relative to the climatological mean (Fig. 1a and Fig. 5b).
dynamical changes in stratospheric temperature patterns. According to Downward motions with omega values of the order of ~0.1 Pa s− 1 are
our results, the downward stratospheric forcing in the two considered generally consistent with those observed during the formation of
events played a greater role in the occurrence of cold weather (Sections stratospheric intrusions and tropopause folds (Meloen et al., 2003;
3.2–3.5) compared to the circulation of air masses at the surface (Section Škerlak et al., 2015; Tao et al., 2017; Wang et al., 2023b).
3.6). There was a noticeable decrease in the tropopause height under the The downward transport of stratospheric air was found by the tem­
influence of a positive stratospheric temperature anomaly that moved perature, wind and relative humidity measurements using the Cross-
equatorward (between the polar and midlatitudes to subtropics) and track Infrared Sounder (CrIS) onboard Suomi National Polar-Orbiting
downward (between the middle and lowermost stratosphere). Rean­ Partnership (SNPP) in late January 2019 cooling (Xiong et al., 2022).
alysis ERA5 shows the anomaly patterns and magnitudes (Supplemental It has been demonstrated that along the path of cold air transport,
Figs. S1–S3) very close to those from the NNR in Figs. 3–5. The tropo­ particularly near the mid-latitude coldest surface center, there exists a
pause height is typically positively (negatively) correlated with the deep tropopause folding, lowest geopotential heights (from ERA-5
tropospheric (stratospheric) temperature (Seidel and Randel, 2006). The reanalysis), significant downward transport of stratospheric dry air,
positive stratospheric T-anomalies in Figs. 3–5 are much more intense, and a warm center above the tropopause. Unlike the New Year’s SSW on
spatially extended and dynamically active than tropospheric ones of 1–2 January 2019, which was classified as a non-downward propagating
both signs. This became the ground for an association of a sharp event (Rao et al., 2020) with only minor surface effects (Butler et al.,
decrease in the tropopause height precisely with the influence of the 2020), extreme cold weather in late January 2019 across a large region
warm anomaly in the lowermost stratosphere. In turn, the dynamics of in the midwestern United States and Eastern Canada was associated with
the stratospheric polar vortex is a determining factor in changes in the stratosphere-to-troposphere transport (Xiong et al., 2022). Along the
warm stratospheric anomaly above the tropopause (Tomassini et al., path of stratospheric intrusion, ozone concentration was enhanced
2012; Baldwin et al., 2019; Xiong et al., 2022), as was confirmed by (Xiong et al., 2022), which could be associated with an increase in air
zonally asymmetric patterns of the stratospheric temperature (Fig. 6a temperature. However, the downward transport of cold stratospheric air
and c). in this region ensured a lower temperature than in adjacent areas (Xiong
Over 5–8 days, the tropopause descended by 2–3 km (Fig. 3c and et al., 2022). These atmospheric circulation conditions are broadly
d and Fig. 5) to the 7 km altitude, or pressure level of ~400 hPa, and at consistent with our results for i) anomalous regional descent of the

10
Y. Shi et al. Atmospheric Research 297 (2024) 107112

tropopause beneath the warm stratospheric anomaly, ii) downward and descent of cold air from the tropopause and middle troposphere
migration of negative T- and Z-anomalies, iii) downward vertical ve­ regions. This process may also be involved in the phenomena of cold air
locity omega, and iv) descent of colder midtropospheric layers up to the outbreaks, stratospheric intrusions and deep tropopause folds (Škerlak
surface, leading to cold weather anomalies (Figs. 3–5 and Figs. 7 and 8). et al., 2015; Xiong et al., 2022).
Many atmospheric events show signs of intrusion of stratospheric air In term of ‘split–displacement’, the stratospheric polar vortex
into the troposphere and tropopause folding, which, however, are not evolved from a displacement type before SSW 2018 to a splitting type
usually associated with surface cooling effects. Most of works on during the event, and finally back to the displacement type (Xie et al.,
stratospheric intrusions and tropopause folds analyze aspects of ozone, 2020; Rao et al., 2021; Liang et al., 2022). This was a result of alternate
water vapor, wind, cyclogenesis, precipitation (Holton et al., 1995; Kim amplitude increase in planetary wave 1 (vortex displacement) and wave
et al., 2001; Škerlak et al., 2015; Barnes et al., 2022; Xiong et al., 2022) 2 (vortex split) (Wang et al., 2019a). In SSW 2018, the stratospheric
and do not highlight possible thermal effects. We emphasize that vortex was shifted to the Western Hemisphere and Europe (Lü et al.,
although the downward movements occur through the stratosphere–­ 2020), However, the regional stratosphere–troposphere coupling in this
troposphere, the extreme surface cooling can be directly related to the event was not associated with cold polar air, but with warm midlatitude
cold air in the midtroposphere. Namely, our results indicate the location anomaly above the tropopause (Fig. 3 and Fig. 6a and c). The longitu­
of the midtropospheric layer at an altitude of 3–5 km, the descent of dinal location of the maximum temperature in the lowermost strato­
which to the surface contributes to cold weather in both East Asia and sphere (Fig. 6c) determined a weaker weather response in the
Eastern Europe. The overall quantitative and qualitative coincidence of Changchun region than in Kharkiv region during the SSW 2018. It was
the results (Figs. 3–7) indicates in favor of fairly reliable establishment found earlier that wave 1 causes vortex asymmetry with more persistent
of the process of rapid downwelling of atmospheric air during the shift to Eurasia in recent decades (Zhang et al., 2016). The SSW onset
extreme cold weather. At the same time, there may be a meridional and evolution in the conditions of the wave 1 amplification were
(equatorward) component of air movement at the surface, which can accompanied by vortex shift to northern Eurasia and the adjacent polar
also contribute to the regional cooling (Section 3.6). However, accord­ ocean in more than three-quarters of the events (Li et al., 2023) and
ing to a preliminary estimate, this pathway appears to be less significant strongest response to SSW is observed in the North Atlantic Basin
compared to the downward stratospheric forcing. (Baldwin et al., 2021). The wave reflection during SSW 2013 led to the
formation of extreme cold weather over the central Eurasia (Nath et al.,
4.2. Surface cooling mechanisms 2016), whereas the vortex shift towards northern East Asia during SSW
2021 resulted in extreme low surface temperatures (Zhang et al., 2022).
The tropospheric response to SSW events is highly variable and These results are evidence of the diversity of regional manifestations
although well quantified statistically, is not completely understood in of SSW effects in weather depending on various factors influencing the
detail (Kidston et al., 2015; Baldwin et al., 2021; Wright et al., 2021; mechanisms of the downward forcing of the stratosphere (Domeisen
Rupp et al., 2022; Hall et al., 2023). Extreme cold weather conditions in et al., 2020; Baldwin et al., 2021; Kolstad et al., 2022; Li et al., 2023).
the boreal winter can occur due to tropospheric circulation anomalies Surface cooling is often associated with midlatitude cold air outbreaks
without any stratospheric connection (Wang et al., 2010; Li et al., 2019; driven by horizontal thermal advection of cold Arctic air (Tomassini
Wang et al., 2019b; Terpstra et al., 2021) and with a clear downward et al., 2012; Wang et al., 2019b; Zhang et al., 2022). Our study highlights
influence of the stratospheric circulation on the surface weather the contribution of downward air motion in the occurrence of cold
throughout much of the NH mid–high latitudes (Baldwin and Dunker­ weather phenomena, which is consistent with the impact of the strato­
ton, 1999; Thompson et al., 2002; Haigh et al., 2005; Kretschmer et al., sphere to the cold air outbreak process through the downward transport
2018; Rao et al., 2020; White et al., 2021; Lü et al., 2020; Lu et al., observed by the satellite measurements (Xiong et al., 2022). In our re­
2022). sults, the northerly wind component (Fig. 9) is associated with a nega­
Analyzing the occurrence of cold spells over Northern Europe, tive Z-anomaly (Fig. 8) and cyclonic circulation (Fig. 9a and 9c). The
Tomassini et al. (2012) show that the lowered tropopause height is a regions considered are located on the western side of low-pressure sys­
consequence of the warming in the lower stratosphere leading to a tems that favors the northerlies to contribute to the cold air advection
compression of the tropospheric column below, appearance of wind into the midlatitudes (Wang et al., 2010; Xiong et al., 2022). A
anomalies and advection of cold air from the north-east. Authors distinctive feature of our finding is that this southward transport path,
conclude that initially vertically propagating planetary waves disturb however, connects the midlatitude regions not with polar region, but
the stratospheric circulation in the polar region, causing strong strato­ with the vortex edge region with a temperature minimum near 60◦ N
spheric warmings, and the impact of the stratosphere on the troposphere (yellow curve in Fig. 7d and h and arrows in Fig. 9b and 9d). Never­
can be understood as dynamical response to these heating anomalies. A theless, preliminary analysis suggests a comparatively smaller contri­
decrease in the height of the tropospheric column as a result of strato­ bution of meridional surface transport of cold air from this region to the
spheric warming and a decrease in the height of the tropopause, as also midlatitudes in comparison with the much colder air downwelling from
suggested in (Kim and Kim, 2021), may be related to tropospheric the 3–5 km tropospheric layer (Section 3.6). Our study highlights the
anomalies. significant contribution of downward air motion in the occurrence of
The tropopause locates in the transition layer between the tropo­ cold weather phenomena, which is consistent with the impact of the
sphere and stratosphere with the lowest air temperature, so called stratosphere to the cold air outbreak process through the downward
tropopause cold-point region (Holton et al., 1995; Kim et al., 2013). transport observed using satellite measurements (Xiong et al., 2022).
Figs. 7 and 9 demonstrate possibility of the coldest air, which resides in Many aspects of planetary wave propagation, stratosphere heating,
the lowermost stratosphere and tropopause cold-point region, to tropopause lowering and weather anomalies, which have been thor­
descend into the troposphere pushing the midtropospheric layers of air oughly studied in last decades (Kodera et al., 1990; Baldwin and Dun­
down all the way to the surface and causing a cold weather anomaly. kerton, 1999; Thompson et al., 2002; Haigh et al., 2005; Škerlak et al.,
Our results, thus, indicate the presence of a chain of vertical air move­ 2015; Yu et al., 2018; Domeisen and Butler, 2020; Wang et al., 2023a),
ments, which is associated with stratosphere–tropopause–troposphere are still open to researchers. Negative tropospheric anomalies tend to
thermal interaction and the occurrence of cold weather events on a deepen near the surface (Figs. 3d and 5b), so it is worth taking into
regional scale not previously reported in the literature. The evolution of account the extent to which the surface amplification of the strato­
the temperature anomalies and air masses in both pre-SSW and SSW spheric signal (Baldwin et al., 2019, 2021) may act regardless of SSW
conditions appears to have occurred as a dynamical response to rapid and polar region, say, in the pre-SSW winter conditions over the NH
lowering of the tropopause during heating of the lowermost stratosphere midlatitudes. The contribution of the descent of cold midtropospheric

11
Y. Shi et al. Atmospheric Research 297 (2024) 107112

air (Fig. 7) may take place in suggested phenomenon. It also deserves to Annular Mode, zonal wind, Eliassen–Palm flux, tropospheric jet streams,
be clarified how effect of compression of the tropospheric column below associated with SSW, remained beyond the scope of this work and
the descended tropopause (Tomassini et al., 2012; Kim and Kim, 2021) require further analysis.
can manifest itself not only in the polar region but also in the mid­
latitudes. According to the results of this work, regionality and the chain Author contributions
of cause-and-effect relationships in stratosphere–troposphere coupling
are important areas of research. Conceptualization, O.E., Y.S. and G.M.; methodology, O.E. and Y.W.;
data acquisition, O.E., Y.A., B.N. and Y.W.; software and related figures,
5. Conclusions O.E., Y.A., and Y.S.; validation, O.E., Y.S., B.N. and G.M.; investigation
and interpretation, O.E., A.K., G.M., Y.A., L.W. and X.W.; writing-
We compare the evolution of temperature anomalies in the strato­ original draft preparation, O.E., A.G., A.K. and G.M.; writing-review
sphere and troposphere in two strong surface cooling events occurred in and editing, O.E., A.K., Y.S., G.M., W.H. and A.G.; visualization, O.E.,
the winter 2018 in East Asia (the Changchun region, Northern China) Y.W. and Y.A.; supervision, G.M. and X.W.; project administration, G.M.
and Eastern Europe (the Kharkiv region, Northern Ukraine). They and W.H. Each author contributed to the interpretation and discussion
developed under strong vortex conditions and wave 1 dominance (the of the results and edited the manuscript. All authors have read and
pre-SSW period in January) and during vortex perturbation and split due agreed to the published version of the manuscript.
to intensification of wave 2 (major SSW in February), respectively.
Analysis of vertical variability in temperature anomaly profiles in time, Funding
as well as by latitude and longitude, show that in both cases surface
cooling appears mainly as a response to rapid tropopause descent under This research has been supported in part by the Jilin University,
influence of heating in the lowermost stratosphere. China (Grant Numbers: G2022129014L, 2020L024J00170).
In both cases, the positive anomaly in the stratospheric temperature
rapidly migrates between polar and midlatitudes and penetrates into the
Declaration of Competing Interest
lowermost stratosphere at 200–300 hPa. This change in the lowermost
stratosphere heating is associated with a lowering of the midlatitude
The authors declare no conflict of interest.
tropopause by 2–3 km in 5–8 days and appearance of negative surface
anomaly of about − 12 ◦ C. The maximum tropopause pressure increased
Data availability
to 400 hPa (tropopause height decreased to ~7 km, while climatologi­
cally it is 9–10 km in midlatitudes). Below the lowered tropopause,
Data will be made available on request.
meridional temperature profiles demonstrate significant deformation
and deflection of the T-contours down to the surface. The cold mid­
tropospheric layer from altitudes of 3–5 km was pushed down to the Acknowledgments
surface moving with a downward velocity of about 0.3–0.5 km day− 1.
Observations showed that the surface temperatures during these events The authors thank the three anonymous reviewers for their careful
dropped by 18 ◦ C in Changchun and by 14 ◦ C in Kharkiv. To a large reading of our manuscript and their many important comments and
extent, the cold air descent can be responsible for the thermal coupling suggestions. This work was partially supported by the College of Physics,
between the lowermost stratosphere – tropopause cold-point regions International Center of Future Science, Jilin University, China, and by
and the surface on a synoptic time scale. Such cold weather events can the Ministry of Education and Science of Ukraine in the framework of
develop in both the pre-SSW conditions and the SSW conditions. project BF30-2021 at Taras Shevchenko National University of Kyiv.
In both events, a sharp lowering of the tropopause was accompanied This work contributed to Project 4293 of the Australian Antarctic Pro­
by a negative regional geopotential height anomaly in the altitude range gram, and to the National Antarctic Scientific Center of Ukraine research
of the lowermost stratosphere, tropopause and middle troposphere. The objectives. NCEP–NCAR reanalysis data were provided by the NOAA/
Z-anomaly minimum in the meridional cross-section was located near OAR/ESRL PSD, Boulder, Colorado, USA, from their website at
the climatological latitudes of the polar vortex edge (~60◦ N). It was http://www.esrl.noaa.gov/psd/. The tropopause pressure data were
associated with the surface low-pressure system with cyclonic circula­ obtained from the NCEP Global Data Assimilation System GDAS at http
tion responsible for the northerly winds on its western side. The north­ s://psl.noaa.gov/data/gridded/data.ncep.html. ERA5 data produced by
erlies contributed to the cold air advection near the surface into the the Copernicus Climate Change Service (C3S) were obtained from the
midlatitudes. An important feature of our finding is that this southward ECMWF reanalysis at https://cds.climate.copernicus.eu/cdsapp#!/
transport path, however, connected the midlatitude regions not with dataset/reanalysis-era5-pressure-levels?tab=overview.
polar region, but with the vortex edge region with a temperature min­
imum near 60◦ N, which served as a direct source of extremly cold air, Appendix A. Supplementary data
entering the midlatitudes both at the surface and at the midtroposphere
level. Preliminary analysis suggests a relatively smaller contribution of Supplementary data to this article can be found online at https://doi.
meridional surface transport compared to the midtropospheric layer org/10.1016/j.atmosres.2023.107112.
downwelling. A detailed quantification of the relative roles of vertical
and horizontal transport of cold air masses in the emergence of References
extremely low temperatures over East Asia and Eastern Europe in winter
Baldwin, M.P., Dunkerton, T.J., 1999. Propagation of the Arctic Oscillation from the
2018 is a task for further study. stratosphere to the troposphere. J. Geophys. Res. 104, 30937–30946. https://doi.
Comparisons of the NNR and ERA5 reanalyses demonstrate close org/10.1029/1999JD900445.
agreement between the temperature timeseries and vertical anomaly Baldwin, M.P., Dunkerton, T.J., 2001. Stratospheric harbingers of anomalous weather
regimes. Science 294, 581–584. https://doi.org/10.1126/science.1063315.
profiles in their variability, patterns and magnitudes. The results show
Baldwin, M.P., Birner, T., Brasseur, G., Burrows, J., Butchart, N., Garcia, R., Geller, M.,
general consistency with the concepts of compression of the tropo­ Gray, L., Hamilton, K., Harnik, N., Hegglin, M.I., Langematz, U., Robock, A., Sato, K.,
spheric column below the lowered tropopause, the surface amplification Scaife, A.A., 2019. 100 years of progress in understanding the stratosphere and
of the stratospheric signal, the large-scale deformation field, which are mesosphere. In: Chapter 27 in Meteorological Monographs, 59. https://doi.org/
10.1175/AMSMONOGRAPHS-D-19-0003.1, 27.1–27.62.
considered in the literature, but their possible effects in midlatitude Baldwin, M.P., Ayarzagüena, B., Birner, T., Butchart, N., Butler, A.H., Charlton-Perez, A.
cooling events are less studied. Aspects of changes in the Northern J., Domeisen, D.I.V., Garfinkel, C.I., Garny, H., Gerber, E.P., Hegglin, M.I.,

12
Y. Shi et al. Atmospheric Research 297 (2024) 107112

Langematz, U., Pedatella, N.M., 2021. Sudden stratospheric warmings. Rev. Kretschmer, M., Cohen, J., Matthias, V., Runge, J., Coumou, D., 2018. The different
Geophys. 59, e2020RG000708 https://doi.org/10.1029/2020RG000708. stratospheric influence on cold-extremes in Eurasia and North America. npj Clim.
Barnes, M.A., Ndarana, T., Sprenger, M., Landman, W.A., 2022. Stratospheric intrusion Atmos. Sci. 44 https://doi.org/10.1038/s41612-018-0054-4.
depth and its effect on surface cyclogenetic forcing: an idealized potential vorticity Lehtonen, I., Karpechko, A.Yu., 2016. Observed and modeled tropospheric cold
(PV) inversion experiment. Weather Clim. Dynam. 3, 1291–1309. https://doi.org/ anomalies associated with sudden stratospheric warmings. J. Geophys. Res. Atmos.
10.5194/wcd-3-1291-2022. 121, 1591–1610. https://doi.org/10.1002/2015JD023860.
Butler, A.H., Seidel, D.J., Hardiman, S.C., Butchart, N., Birner, T., Match, A., 2015. Li, D., Bian, J.C., Fan, Q.J., 2015. A deep stratospheric intrusion associated with an
Defining sudden stratospheric warmings. Bull. Am. Meteorol. Soc. 96, 1913–1928. intense cut-off low event over East Asia. Sci. China Earth Sci. 58, 116–128. https://
https://doi.org/10.1175/BAMS-D-13-00173.1. doi.org/10.1007/s11430-014-4977-2.
Butler, A.H., Lawrence, Z.D., Lee, S.H., Lillo, S.P., Long, C.S., 2020. Differences between Li, Y., Zhang, J., Lu, Y., Zhu, J., Feng, J., 2019. Characteristics of transient eddy fluxes
the 2018 and 2019 stratospheric polar vortex split events. Q. J. R. Meteorol. Soc. during blocking highs associated with two cold events in China. Atmosphere 10, 235.
146, 3503–3521. https://doi.org/10.1002/qj.3858. https://doi.org/10.3390/atmos10050235.
Chandran, A., Collins, R.L., 2014. Stratospheric sudden warming effects on winds and Li, Y., Kirchengast, G., Schwaerz, M., Yuan, Y., 2023. Monitoring sudden stratospheric
temperature in the middle atmosphere at middle and low latitudes: a study using warmings under climate change since 1980 based on reanalysis data verified by
WACCM. Ann. Geophys. 32, 859–874. https://doi.org/10.5194/angeo-32-859-2014. radio occultation. Atmos. Chem. Phys. 23, 1259–1284. https://doi.org/10.5194/
Charlton-Perez, A.J., Ferranti, L., Lee, R.W., 2018. The influence of the stratospheric acp-23-1259-2023.
state on North Atlantic weather regimes. Q. J. R. Meteorol. Soc. 144, 1140–1151. Liang, Z., Rao, J., Guo, D., Lu, Q., 2022. Simulation and projection of the sudden
https://doi.org/10.1002/qj.3280. stratospheric warming events in different scenarios by CESM2-WACCM. Clim. Dyn.
Cohen, J., Agel, L., Barlow, M., Garfinkel, C.I., White, I., 2021. Linking Arctic variability 59, 3741–3761. https://doi.org/10.1007/s00382-022-06293-2.
and change with extreme winter weather in the United States. Science 373, Liu, R., 2021. Latest trends of atmospheric cells under global warming. IOP Conf. Ser.:
1116–1121. https://doi.org/10.1126/science.abi9167. Earth Environ. Sci. 658, 012003 https://doi.org/10.1088/1755-1315/658/1/
Cohen, J., Agel, L., Barlow, M., Garfinkel, C.I., White, I., 2023. Response to Limited 012003.
surface impacts of the January 2021 sudden stratospheric warming. Nat. Commun. Lü, Z., Li, F., Orsolini, Y.J., Gao, Y., He, S., 2020. Understanding of European cold
14, 3289. https://doi.org/10.1038/s41467-023-38772-3. extremes, sudden stratospheric warming, and Siberian snow accumulation in the
Domeisen, D.I.V., Butler, A.H., 2020. Stratospheric drivers of extreme events at the winter of 2017/18. J. Clim. 33, 527–545. https://doi.org/10.1175/JCLI-D-18-
Earth’s surface. Commun. Earth Environ. 1, 59. https://doi.org/10.1038/s43247- 0861.1.
020-00060-z. Lu, Q., Rao, J., Liang, Z., Guo, D., Luo, J., Liu, S., Wang, C., Wang, T., 2021. The sudden
Domeisen, D.I.V., Butler, A.H., Charlton-Perez, A.J., Ayarzagüena, B., Baldwin, M.P., stratospheric warming in January 2021. Environ. Res. Lett. 16 (8), 084029 https://
Dunn-Sigouin, E., Furtado, J.C., Garfinkel, C.I., Hitchcock, P., Karpechko, A.Yu., doi.org/10.1088/1748-9326/ac12f4.
Kim, H., Knight, J., Lang, A.L., Lim, E.-P., Marshall, A., Roff, G., Schwartz, C., Lu, Q., Rao, J., Shi, C., Guo, D., Fu, G., Wang, J., Liang, Z., 2022. Possible influence of
Simpson, I.R., Son, S.-W., Taguchi, M., 2020. The role of the stratosphere in sudden stratospheric warmings on the atmospheric environment in the
subseasonal to seasonal prediction: 2. Predictability arising from stratosphere- Beijing–Tianjin–Hebei region. Atmos. Chem. Phys. 22, 13087–13102. https://doi.
troposphere coupling. J. Geophys. Res. Atmos. 125, e2019JD030923 https://doi. org/10.5194/acp-22-13087-2022.
org/10.1029/2019JD030923. Manney, G.L., Millán, L.F., Santee, M.L., Wargan, K., Lambert, A., Neu, J.L., Werner, F.,
Haigh, J.D., Blackburn, M., Day, R., 2005. The response of tropospheric circulation to Lawrence, Z.D., Schwartz, M.J., Livesey, N.J., Read, W.G., 2022. Signatures of
perturbations in lower-stratospheric temperature. J. Clim. 18, 3672–3685. https:// anomalous transport in the 2019/2020 Arctic stratospheric polar vortex. J. Geophys.
doi.org/10.1175/JCLI3472.1. Res. Atmos. 127, e2022JD037407 https://doi.org/10.1029/2022JD037407.
Hall, R.J., Mitchell, D.M., Seviour, W.J.M., Wright, C.J., 2021. Tracking the stratosphere- Marshall, J., Plumb, R.A., 2008. Atmosphere, ocean, and climate dynamics: an
to-surface impact of Sudden Stratospheric Warmings. J. Geophys. Res. Atmos. 126, introductory text. In: International Geophysics Series, 93. Elsevier Academic Press,
e2020JD033881 https://doi.org/10.1029/2020JD033881. Burlington, MA 01803, USA, p. 319. ISBN 978-0-12-558691-7.
Hall, R.J., Mitchell, D.M., Seviour, W.J.M., Wright, C.J., 2023. Surface hazards in North- Meloen, J., Siegmund, P., van Velthoven, P., Kelder, H., Sprenger, M., Wernli, H.,
west Europe following sudden stratospheric warming events. Environ. Res. Lett. 18, Kentarchos, A., Roelofs, G., Feichter, J., Land, C., Forster, C., James, P., Stohl, A.,
064002 https://doi.org/10.1088/1748-9326/acd0c3. Collins, W., Cristofanelli, P., 2003. Stratosphere-troposphere exchange: a model and
Hersbach, H., Bell, B., Berrisford, P., Hirahara, S., Horányi, A., Muñoz-Sabater, J., method intercomparison. J. Geophys. Res. 108, 8526. https://doi.org/10.1029/
Nicolas, J., Peubey, C., Radu, R., Schepers, D., et al., 2020. The ERA5 global 2002JD002274.
reanalysis. Q. J. R. Meteorol. Soc. 146, 1999–2049. https://doi.org/10.1002/ Nath, D., Chen, W., Cai, Z., Zelin, C., Pogoreltsev, A.I., Wei, K., 2016. Dynamics of 2013
qj.3803. Sudden Stratospheric Warming event and its impact on cold weather over Eurasia:
Hitchcock, P., Shepherd, T.G., 2013. Zonal-mean dynamics of extended recoveries from Role of planetary wave reflection. Sci. Rep. 6, 1–12. https://doi.org/10.1038/
stratospheric sudden warmings. J. Atmos. Sci. 70, 688–707. https://doi.org/ srep24174.
10.1175/JAS-D-12-0111.1. Rao, J., Ren, R., Chen, H., Yu, Y., Zhou, Y., 2018. The stratospheric sudden warming
Holton, J.R., Haynes, P.H., McIntyre, M.E., Douglass, A.R., Rood, R.B., Pfister, L., 1995. event in February 2018 and its prediction by a climate system model. J. Geophys.
Stratosphere-troposphere exchange. Rev. Geophys. 33, 403–439. https://doi.org/ Res. Atmos. 123 https://doi.org/10.1029/2018JD028908, 13,332–13,345.
10.1029/95RG02097. Rao, J., Garfinkel, C.I., White, I.P., 2020. Predicting the downward and surface influence
Kalnay, E., Kanamitsu, M., Kistler, R., Collins, W., Deaven, D., Gandin, L., Iredell, M., of the February 2018 and January 2019 sudden stratospheric warming events in
Saha, S., White, G., Woollen, J., Zhu, Y., Chelliah, M., Ebisuzaki, W., Higgins, W., subseasonal to seasonal (S2S) models. J. Geophys. Res. Atmos. 125, e2019JD031919
Janowiak, J., Mo, K.C., Ropelewski, C., Wang, J., Leetmaa, A., Reynolds, R., https://doi.org/10.1029/2019JD031919.
Jenne, R., Joseph, D., 1996. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Rao, J., Liu, S., Chen, Y., 2021. Northern Hemisphere sudden stratospheric warming and
Meteorol. Soc. 77, 437–472. https://doi.org/10.1175/1520-0477(1996)077<0437: its downward impact in four chinese CMIP6 models. Adv. Atmos. Sci. 38, 187–202.
TNYRP>2.0.CO;2. https://doi.org/10.1007/s00376-020-0250-0.
Karpechko, A.Y., Charlton-Perez, A., Balmaseda, M., Tyrrell, N., Vitart, F., 2018. Rupp, P., Loeffel, S., Garny, H., Chen, X., Pinto, J.G., Birner, T., 2022. Potential links
Predicting sudden stratospheric warming 2018 and its climate impacts with a between tropospheric and stratospheric circulation extremes during early 2020.
multimodel ensemble. Geophys. Res. Lett. 45 https://doi.org/10.1029/ J. Geophys. Res. Atmos. 127, e2021JD035667 https://doi.org/10.1029/
2018GL081091, 13,538–13,546. 2021JD035667.
Kidston, J., Scaife, A.A., Hardiman, S.C., Mitchell, M.D., Butchart, N., Baldwin, M.P., Seidel, D.J., Randel, W.J., 2006. Variability and trends in the global tropopause
Gray, L.J., 2015. Stratospheric influence on tropospheric jet streams, storm tracks estimated from radiosonde data. J. Geophys. Res. 111, D21101. https://doi.org/
and surface weather. Nat. Geosci. 8, 433–440. https://doi.org/10.1038/NGEO2424. 10.1029/2006JD007363.
Kim, J., Kim, K.-Y., 2021. The leading modes of NH extratropical tropopause variability Seviour, W.J.M., Mitchell, D.M., Gray, L.J., 2013. A practical method to identify
and their connection with stratosphere-troposphere variability. Clim. Dyn. 56, displaced and split stratospheric polar vortex events. Geophys. Res. Lett. 40,
2413–2430. https://doi.org/10.1007/s00382-020-05595-7. 5268–5273. https://doi.org/10.1002/grl.50927.
Kim, K.-E., Jung, E.-S., Campistron, B., Heo, B.-H., 2001. A physical examination of Škerlak, B., Sprenger, M., Pfahl, S., Tyrlis, E., Wernli, H., 2015. Tropopause folds in ERA-
tropopause height and stratospheric air intrusion – a case study. J. Meteorol. Soc. Interim: global climatology and relation to extreme weather events. J. Geophys. Res.
Jap. Ser. II. 79, 1093–1103. https://doi.org/10.2151/jmsj.79.1093. Atmos. 120, 4860–4877. https://doi.org/10.1002/2014JD022787.
Kim, J., Grise, K.M., Son, S.-W., 2013. Thermal characteristics of the cold-point Stone, P.H., Carlson, J.H., 1979. Atmospheric lapse rate regimes and their
tropopause region in CMIP5 models. J. Geophys. Res. Atmos. 118 https://doi.org/ parameterization. J. Atmos. Sci. 36, 415–423. https://doi.org/10.1175/1520-0469
10.1002/jgrd.50649. (1979)036<0415:ALRRAT>2.0.CO;2.
Kodera, K., Yamazaki, K., Chiba, M., Shibata, K., 1990. Downward propagation of upper Tao, W., Zhang, J., Zhang, X., 2017. The role of stratosphere vortex downward intrusion
stratospheric mean zonal wind perturbation to the troposphere. Geophys. Res. Lett. in a long-lasting late-summer Arctic storm. Q. J. R. Meteorol. Soc. 143, 1953–1966.
17 https://doi.org/10.1029/GL017i009p01263, 1263–1266. 90GL01408. https://doi.org/10.1002/qj.3055.
Kodera, K., Mukougawa, M., Maury, P., Ueda, M., Claud, C., 2016. Absorbing and Terpstra, A., Renfrew, I.A., Sergeev, D.E., 2021. Characteristics of cold-air outbreak
reflecting sudden stratospheric warming events and their relationship with events and associated polar mesoscale cyclogenesis over the North Atlantic region.
tropospheric circulation. J. Geophys. Res. Atmos. 121, 80–94. https://doi.org/ J. Clim. 34, 4567–4584. https://doi.org/10.1175/JCLI-D-20-0595.1.
10.1002/2015JD023359. Thompson, D.W., Wallace, J.M., 2000. Annular modes in the extratropical circulation.
Kolstad, E.W., Lee, S.H., Butler, A.H., Domeisen, D.I.V., Wulff, C.O., 2022. Diverse Part I: month-to-month variability. J. Clim. 13, 1000–1016. https://doi.org/
surface signatures of stratospheric polar vortex anomalies. J. Geophys. Res. Atmos. 10.1175/1520-0442(2000)013<1000:AMITEC>2.0.CO;2.
127 https://doi.org/10.1029/2022JD037422. Thompson, D.J., Baldwin, M.P., Wallace, J.M., 2002. Stratospheric connection to
Northern Hemisphere wintertime weather: Implications for prediction. J. Clim. 15,

13
Y. Shi et al. Atmospheric Research 297 (2024) 107112

1421–1428. https://doi.org/10.1175/1520-0442(2002)015<1421:SCTNHW>2.0. White, I.P., Garfinkel, C.I., Cohen, J., Jucker, M., Rao, J., 2021. The impact of split and
CO;2. displacement sudden stratospheric warmings on the troposphere. J. Geophys. Res.
Tomassini, L., Gerber, E.P., Baldwin, M.P., Bunzel, F., Giorgetta, M., 2012. The role of Atmos. 126, e2020JD033989 https://doi.org/10.1029/2020JD033989.
stratosphere–troposphere coupling in the occurrence of extreme winter cold spells Wright, C.J., Hall, R.J., Banyard, T.P., Hindley, N.P., Krisch, I., Mitchell, D.M.,
over northern Europe. J. Adv. Model. Earth Syst. 4, M00A03. https://doi.org/ Seviour, W.J.M., 2021. Dynamical and surface impacts of the January 2021 sudden
10.1029/2012MS000177. stratospheric warming in novel Aeolus wind observations, MLS and ERA5. Weather
Wang, B., Wu, Z., Chang, C.-P., Liu, J., Li, J., Zhou, T., 2010. Another look at interannual- Clim. Dynam. 2, 1283–1301. https://doi.org/10.5194/wcd-2-1283-2021.
to-interdecadal variations of the East Asian winter monsoon: the northern and Xie, J., Hu, J., Xu, H., Liu, S., He, H., 2020. Dynamic diagnosis of stratospheric sudden
southern temperature modes. J. Clim. 23, 1495–1512. https://doi.org/10.1175/ warming event in the boreal winter of 2018 and its possible impact on weather over
2009JCLI3243.1. North America. Atmosphere 11, 438. https://doi.org/10.3390/atmos11050438.
Wang, Y., Shulga, V., Milinevsky, G., Patoka, A., Evtushevsky, O., Klekociuk, A., Han, W., Xiong, X., Liu, X., Wu, W., Knowland, K.E., Yang, F., Yang, Q., Zhou, D.K., 2022. Impact
Grytsai, A., Shulga, D., Myshenko, V., Antyufeyev, V., 2019a. Winter 2018 major of stratosphere on cold air outbreak: Observed evidence by CrIS on SNPP and its
sudden stratospheric warming impact on midlatitude mesosphere from microwave comparison with models. Atmosphere 13, 876. https://doi.org/10.3390/
radiometer measurements. Atmos. Chem. Phys. 19, 10303–10317. https://doi.org/ atmos13060876.
10.5194/acp-19-10303-2019. Yu, Y., Cai, M., Shi, C., Ren, R., 2018. On the linkage among strong stratospheric mass
Wang, F., Vavrus, S.J., Francis, J.A., Martin, J.E., 2019b. The role of horizontal thermal circulation, stratospheric sudden warming, and cold weather events. Mon. Weather
advection in regulating wintertime mean and extreme temperatures over interior Rev. 146, 2717–2739. https://doi.org/10.1175/MWR-D-18-0110.1.
North America during the past and future. Clim. Dyn. 53, 6125–6144. https://doi. Zängl, G., Hoinka, K.P., 2001. The tropopause in the polar regions. J. Clim. 14,
org/10.1007/s00382-019-04917-8. 3117–3139. https://doi.org/10.1175/1520-0442(2001)014<3117:TTITPR>2.0.CO;
Wang, Y., Evtushevsky, O., Milinevsky, G., Shulga, V., Yukhymchuk, Y., Han, W., 2.
Shulga, D., Grytsai, A., 2020. The major sudden stratospheric warming impact on Zhang, J., Tian, W., Chipperfield, M.P., Xie, F., Huang, J., 2016. Persistent shift of the
mid-latitude surface weather. EPJ Web Conf. 237, 04007. https://doi.org/10.1051/ Arctic polar vortex towards the Eurasian continent in recent decades. Nat. Clim.
epjconf/202023704007. Chang. 6, 1094–1099. https://doi.org/10.1038/nclimate3136.
Wang, Y., Li, Y., Wang, G., Yuan, Y., Geng, H., 2023a. Influences of Sudden Stratospheric Zhang, Y.X., Si, D., Ding, Y.H., Jiang, D.B., Li, Q.Q., Wang, G.F., 2022. Influence of major
Warming events on tropopause based on GNSS Radio Occultation data. Atmosphere stratospheric sudden warming on the unprecedented cold wave in East Asia in
14, 1553. https://doi.org/10.3390/atmos14101553. January 2021. Adv. Atmos. Sci. 39, 576–590. https://doi.org/10.1007/s00376-022-
Wang, H., Wang, W., Shangguan, M., Wang, T., Hong, J., Zhao, S., Zhu, J., 2023b. The 1318-9.
stratosphere-to-troposphere transport related to Rossby wave breaking and its Zhong, W.G., Wu, Z.W., 2023. Interannual variability of the wintertime Asian-Bering-
impact on summertime ground-level ozone in Eastern China. Remote Sens. 15, 2647. north American teleconnection linked to Eurasian snow cover and Maritime
https://doi.org/10.3390/rs15102647. Continent Sea surface temperature. J. Clim. 36, 2815–2831. https://doi.org/
Waugh, D.W., Sobel, A.H., Polvani, L.M., 2017. What is the polar vortex and how does it 10.1175/JCLI-D-22-0367.1.
influence weather? Bull. Am. Meteorol. Soc. 98, 37–44. https://doi.org/10.1175/
BAMS-D-15-00212.1.

14

You might also like