0% found this document useful (0 votes)
14 views11 pages

Flash

This study investigates the upper boundary of the Asian Tropopause Aerosol Layer (ATAL) during the Asian Summer Monsoon, revealing that it can extend up to the layer of maximum stability (LmaxS) at approximately 442 K potential temperature. Utilizing in-situ and multi-satellite observations, the research highlights that the ATAL's top is lower than previously predicted, constrained by a temperature inversion that limits aerosol transport to higher altitudes. The findings contribute to understanding the dynamics of the ATAL and its impact on atmospheric composition and radiative forcing.

Uploaded by

Mohd Iliyas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views11 pages

Flash

This study investigates the upper boundary of the Asian Tropopause Aerosol Layer (ATAL) during the Asian Summer Monsoon, revealing that it can extend up to the layer of maximum stability (LmaxS) at approximately 442 K potential temperature. Utilizing in-situ and multi-satellite observations, the research highlights that the ATAL's top is lower than previously predicted, constrained by a temperature inversion that limits aerosol transport to higher altitudes. The findings contribute to understanding the dynamics of the ATAL and its impact on atmospheric composition and radiative forcing.

Uploaded by

Mohd Iliyas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Atmospheric Pollution Research 13 (2022) 101451

Contents lists available at ScienceDirect

Atmospheric Pollution Research


journal homepage: www.elsevier.com/locate/apr

Defining the upper boundary of the Asian Tropopause Aerosol Layer


(ATAL) using the static stability
S.T. Akhil Raj a, *, M. Venkat Ratnam a, J.P. Vernier b, c, A.K. Pandit b, Frank G. Wienhold d
a
National Atmospheric Research Laboratory, Gadanki, India
b
National Institute of Aerospace, Hampton, VA, USA
c
NASA Langley Research Center, USA
d
Institute of Atmospheric and Climate Science, Universitaetstrasse 16, Zurich, Switzerland

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

Keywords: The Asian Tropopause Aerosol Layer (ATAL) is located in the Upper Troposphere and Lower Stratosphere (UTLS)
ATAL during the Asian Summer Monsoon. However, what dynamical feature separates the ATAL from the well-known
UTLS region stratospheric ‘Junge layer’ is not yet clear. In this study, using the in-situ (Radiosonde, Ozonesonde, backscatter
BATAL
sonde and cryogenic frost-point hygrometer) observations from multiple locations in India (Gadanki (13.45◦ N,
Asian summer monsoon anticyclone
79.18◦ E), Hyderabad (17.47◦ N, 78.58◦ E) and Varanasi (25.27◦ N, 82.99◦ E)) and multi-satellite observations
((Cloud-Aerosol Lidar and Infrared Pathfinder Observation, (CALIPSO), Atmospheric Chemistry Experiment
(ACE) Fourier Transform Spectrometer (FTS) and Constellation Observation System for Meteorology, Ionosphere
and Climate (COSMIC) Global Position System (GPS) Radio Occultation (RO) (COSMIC GPS-RO)) we show that
the ATAL can exist up to the layer of maximum stability (LmaxS), located a few kilometers above the tropopause,
determined using the square of Brunt Väisäla frequency. These in-situ observations over Indian stations collected
during the ISRO-NASA Balloon Measurement Campaigns of the Asian Tropopause Aerosol Layer (BATAL) show
that the ATAL top can reach up to ~442 K potential temperature level over the Indian region. The LmaxS
delineated from COSMIC GPSRO observations over the Asian Summer Monsoon Anticyclone (ASMA) region
indicates that the top of ATAL can reach up to 454 K potential temperature level, which is lower than the earlier
Lagrangian transport model predicted 460 K. The temperature inversion at LmaxS acts as a lid and constrains the
direct transport of aerosols to higher altitudes.

1. Introduction the ASMA region was discovered using Cloud-Aerosol Lidar and Infrared
Pathfinder Satellite Observations (CALIPSO) mission a few years later
The Asian Summer Monsoon Anticyclone (ASMA), a major dynam­ (Vernier et al., 2011, 2015). Further, the existence of ATAL from 1999
ical system extending from the upper troposphere to the lower strato­ onwards is observed in the Stratospheric Aerosols and Gas Experiments
sphere during the boreal summer is known to contain an enhanced (SAGE) II measurements (Thomason and Vernier, 2013). However, the
concentration of tropospheric pollutants, either lifted by the associated spatially resolved Ammonium Nitrate (NH4NO3, (AN)) observations
deep convection over the Indian subcontinent (Hoskins and Rodwell, with Cryogenic Infrared Spectrometers and Telescopes for the Atmo­
1995) or through long-range transport (Brunamonti et al., 2018; Vogel sphere (CRISTA) showed enhanced concentrations of solid AN in the
et al., 2019; Basha et al., 2020, 2021). The quasi-persistence of anti­ upper troposphere within the ASMA as early as 1997 (Höpfner et al.,
cyclone in the UTLS region results in the accumulation of tropospheric 2019). The ATAL occur between ~360 K and ~440 K potential tem­
trace gases (H2O, CH4, CO etc.) and a reduction in the stratospheric trace perature (~13–18 km) from the eastern Mediterranean Sea to western
gases (in O3, HNO3, HCl, etc.) within the ASMA (Park et al., 2008, 2013; China and India with a thickness of 3–4 km around 30–40◦ N and thinner
Randel et al., 2010; Garny and Randel, 2016; Basha et al., 2021). near the equator (Vernier et al., 2011, 2015, 2018; Hanumanthu et al.,
However, the existence of the Asian Tropopause Aerosol Layer (ATAL) in 2020; Bian et al., 2020). The UTLS aerosols long-term observations

Peer review under responsibility of Turkish National Committee for Air Pollution Research and Control.
* Corresponding author.
E-mail address: [email protected] (S.T.A. Raj).

https://doi.org/10.1016/j.apr.2022.101451
Received 21 January 2022; Received in revised form 6 May 2022; Accepted 6 May 2022
Available online 12 May 2022
1309-1042/© 2022 Turkish National Committee for Air Pollution Research and Control. Production and hosting by Elsevier B.V. All rights reserved.
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

suggest that the ATAL optical depth has increased 2–3 times since the monsoon circulation, which impacts the trace gases and aerosol con­
late 1990 (Vernier et al., 2018). The satellite observations of ATAL were centration in the ASMA (Santee et al., 2017, Hemanth Kumar and Rat­
consistent with the early balloon-borne observations made from Lhasa nam, 2021; Yuan et al., 2019). Hanumanthu et al. (2020) found
(29.7◦ N, 91.1◦ E) in August 1999 (Kim et al., 2003; Hanumanthu et al., substantial day-to-day variability of ATAL in both backscatter signal
2020). The first offline chemical analysis during the Balloon Measure­ intensity and altitude range. They have also performed Lagrangian
ment Campaigns of the Asian Tropopause Aerosol Layer (BATAL) using back-trajectory analysis to identify the air mass origin in the model
ion chromatography (IC) suggests that the nitrate is a more important boundary layer and its transport pathways to the ATAL over Nainital in
chemical component of the ATAL than the sulphate aerosols, which were August 2016.
lower than the IC detection limit (Vernier et al., 2018). This observation There have been many studies exists on ASMA composition, dy­
was consistent with the model results of Gu et al. (2016). These results namics and ATAL variability in the recent years (Park et al., 2008, 2013;
are further supported by the later observations of AN from CRISTA and Pan et al., 2016; Santee et al., 2017; Vernier et al., 2015, 2018; Hemanth
ammonia (NH3) observed using Michelson Interferometer for Passive Kumar and Ratnam, 2021; Yuan et al., 2019; Basha et al., 2020; Hanu­
Atmospheric Sounding (MIPAS) (Höpfner et al., 2019). The advanced manthu et al., 2020; Zhang et al., 2002; Vogel et al., 2015, 2019; He
balloon-borne observations of aerosol over the Indian region using the et al., 2020). With Lagrangian models trajectory simulations, few of
Compact Optical Backscatter AerosoL Detector (COBALD), Optical Par­ these studies have shown a slow ascend of air parcels within the anti­
ticle Counter (OPC) and offline chemical analysis (IC) shown that the cyclone and reaches an upper boundary of 400–460 K potential tem­
ATAL is mostly composed of relatively small (radius <0.25 μm) liquid perature (Müller et al., 2016; Vogel et al., 2019; Brunamonti et al.,
(80%–95%) aerosols with relatively low scattering ratio (SR) at 532 nm 2018). However, what dynamical aspect separates the ATAL from the
which is consistent with the CALIPSO observations (Vernier et al., well known Jungi layer is not clear. In the present study, we define the
2018). Independent observations over the ASMA region with upper boundary of the ATAL based on the atmospheric stability, which is
balloon-borne aircraft and satellite platforms indicate that the ATAL derived from the balloon-borne in situ and spaceborne measurements.
extends above the cold point tropopause (Vernier et al., 2018; Bruna­ The rest of the paper is organized as follows. Section 2 outlines the data
monti et al., 2018; Hanumanthu et al., 2020; Höpfner et al., 2019; sets used for the present study. Section 3 comprises the methodology
Mahnke et al., 2021; Yu et al., 2017). A relatively high concentration of adopted to identify the ATAL top boundary, followed by the description
aerosol is found near the tropopause and declines in the higher altitudes. of the results. Summary and discussion are provided in Section 4, and
Particle concentration for radius (r) > 0.094 dominated in the tropo­ finally, major conclusions for the present study are drawn in Section 5.
pause altitude, with its concentration comparable to that of the
boundary layer concentration (up to about 25 particles per cm− 3). 2. Data and methodology
Particle concentration for r > 0.15 and r > 0.30 μm is smaller by a factor
of 30 and 300, respectively. Brunamonti et al. (2018) referred to this In this paper, we used multiple in-situ observations from the BATAL
extended layer above the cold point tropopause as ‘confined lower campaign conducted during 2014–2019 (Vernier et al., 2018) and sat­
stratosphere’ (CLS) using the 2-week LAGRANTO backward trajectories ellite observations to derive the top of the ATAL from the basic atmo­
and water vapor measurements over Nainital, India (29.35◦ N, 79.46◦ E) spheric parameters. The top of the ATAL is delineated based on the static
and Dhulikhel, Nepal (27.62◦ N, 85.54◦ E). The top of the confined layer stability criteria (Sunilkumar et al., 2017; Gettelman and Wang, 2015;
is located around 18.6 km (421.5 K) and 19.5 km (441 K) over Nainital Birner et al., 2002) using radiosonde observations over Gadanki (13.45◦
and Dhulikhel, respectively. Ma et al. (2022) also observed a similarly N, 79.18◦ E), Hyderabad (17.47◦ N, 78.58◦ E) and Varanasi (25.27◦ N,
mixed layer around 1–1.5 km above the tropopause over ASMA region 82.99◦ E) along with supporting balloon-borne aerosol backscatter,
from the tracer-tracer relationship of O3 and H2O. The concept of the water vapor and ozone observations. The geographical location of the
confined layer is consistent with the concept of “upward spiralling launching sites is shown in Fig. 1, and the details of the sondes used in
range” by Vogel et al. (2019). They found that the slow diabatic uplift
within the ASMA can transport the airmass up to 460 K from the top of
the convective outflow level (~360 K) within a few months. Later these
uplifted air masses are transported to a higher altitude via tropical pipe
associated with large-scale Brewer-Dobson circulation. The HALO
aircraft in-situ observations over extratropical UTLS (15 - 75◦ N, 25◦ W-
15◦ E) and 50 days Lagrangian back trajectory analysis also shows a slow
diabatic ascent of air mass within the anticyclone up to potential tem­
perature >400 K and subsequent transport to the extratropical lower­
most stratosphere (Müller et al., 2016). The analysis of model
simulations also shows eddies detached from the anticyclone contribute
to transporting the trace gases and aerosols to the western Pacific and
western Africa from the Asian region (Fadnavis et al., 2018). The
descending motion in the western part (approximately 70◦ E) is also
found to play an important role in the dissipation of the ATAL, in
addition to the large scale ascending circulation inside ASMA and the
eddy shedding at the east and western edges of the ASMA (He et al.,
2020). Aerosol near the tropopause could impact the Earth’s radiative
balance and the cirrus cloud properties. Based on CALIOP observations
between 1995 and 2013, the summertime aerosol optical depth increase
within the ATAL resulted in a − 1 Wm2 radiative forcing at the top of the
atmosphere. This correspond to the one-third of the total radiative
forcing due to increased CO2 over the same period (Vernier et al., 2015).
ASMA is not a stable, persistent unimodal circulation, it can be
bimodal (Tibetan and Iranian modes) with substantial day-to-day vari­
ability (Zhang et al., 2002; Vogel et al., 2015, Hemanth Kumar and Fig. 1. Map showing the balloon launching locations over India during
Ratnam, 2021). There is also considerable interannual variability in the 2014–2019 BATAL Campaigns.

2
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

the present study are provided in Table 1. Tropopause parameters and has an optical power of around 700 mW. A silicon photodiode phase-
their detection criteria are discussed in section 2.7. sensitive detects the light scattered back from air molecules, aerosols
or ice particles with an uncertainty of 5% and precision better than 1%
2.1. Radiosonde and ozonesonde in the UTLS region (Vernier et al., 2015, 2018). In the present study, we
used the blue backscatter ratio (BSR455) to discern aerosol (ATAL) and
We used both Meisei (RS-11 G) and iMet radiosonde temperature and cloud, following Hanumanthu et al. (2020). BSR455 < 1.12 is considered
pressure data as a function of altitude. These radiosondes measure as aerosol, and larger values are treated as cloud/ice particles. We didn’t
temperature from − 90 ◦ C to 50 ◦ C with a resolution of 0.1 ◦ C and ac­ use COBALD observations after August 15, 2017 and 2019 in the present
curacy of 0.5 ◦ C at a time resolution of 1s. iMet employs a piezo-resistor study to avoid the effects of smoke from the Canadian wildfire and
to measure atmospheric pressure with an accuracy of ~1–2 hPa between Raikoke eruption, respectively.
2 and 1070 hPa. Since the Meisei Radiosonde does not have a pressure
sensor, it derives pressure from the observed temperature and GPS 2.4. Atmospheric Chemistry Experiment (ACE) Fourier Transform
altitude with the hypsometric equation. The ozone profile from the Spectrometer (FTS)
surface to the balloon burst altitude (30–35 km) was obtained by the EN-
SCI Electrochemical concentration cell (ECC) ozonesondes (Komhyr We used monthly mean ACE-FTS observations on-board SCISAT
et al., 1995). Ambient air is bubbled through a cathode chamber filled between 2014 and 2018 over the ASMA region (10–40◦ N, 0–120◦ E).
with potassium iodide solution by a 12 V pump, and the subsequent ACE-FTS measure solar spectra from 2.2 to 13.3 μm wavelength using a
reaction generates two electrons per ozone molecule. The current Michelson interferometer during solar occultations (Bernath et al.,
measured using an external circuit board later converts into ozone 2005). During each occultation, the vertical sampling of each constitu­
partial pressure with an accuracy of around 5–10%. More details on ent is made with a vertical resolution of 2–6 km depending upon the
radiosonde and ozonesonde are available in Ratnam et al. (2014), and occultation angle, and the final product is available at a 1 km vertical
Akhil Raj et al. (2015). resolution (Boone et al., 2005). The satellite provides 30 measurements
per day for over 30 chemical species from the 5 km–150 km range.
Version 3.6 ozone (O3), water vapor (H2O), hydrogen cyanide (HCN)
2.2. Cryogenic frostpoint hygrometer (CFH)
and carbon monoxide (CO) observations are used in the present study.
The ACE-FTS provides observations primarily over high latitudes with
The CFH is a compact, microprocessor-controlled instrument that
limited measurements over the tropics. Therefore, we have used
operates based on the chilled mirror principle. It accurately measures
June-July-August mean profiles within the ASMA region to study the
the frost-point/dew point temperature from the surface to the 28 km
vertical structure of the trace gases and their variation near the tropo­
altitude range. The CFH operates by precisely controlling the tempera­
pause from 2014 to 2018. Though the ACE-FTS provides temperature
ture on a small gold-plated oxygen-free high-density copper mirror to
measurements, we have used high-resolution temperature observations
generate a thin layer of dew or frost in thermodynamic equilibrium with
from Constellation Observation System for Meteorology, Ionosphere and
the ambient air. The thin layer of condensate is maintained by fast
Climate (COSMIC) Global Position System (GPS) Radio Occultation (RO)
heating and cooling with the help of an electrical heater and a cryogenic
to derive tropopause parameters.
liquid (Trifluoromethane). A detailed description of the instrument is
presented by Vömel et al. (2007). The observed frost point temperature
2.5. GPSRO on-board COSMIC satellite
is used to calculate the water vapor mixing ratio by estimating the water
vapor partial pressure following the Goff Gratch equation (Goff and
We used temperature profiles retrieved from COSMIC GPSRO ob­
Gratch, 1946). CFH measures water vapor with an accuracy of 4% in the
tained during 2014–2018 to derive tropopause parameters over the
lower tropical troposphere, 9% in the tropopause region and 10% in the
ASMA region. The temperature and relative humidity profiles are
lower stratosphere. The relative humidity with respect to ice (RHice) is
derived from the refractivity profile obtained from the RO method. The
calculated from the frost point temperature from CFH, and the ambient
details of the GPS RO method and retrieval technique are provided by
air temperature measured using iMet radiosonde. RHice accuracy is
Kursinski et al. (1997). The estimated precision of COSMIC temperature
5–7% in the upper troposphere and better than 10% in the lower
measurements is 0.1% (Alexander et al., 2014). The root mean square
stratosphere (Sunilkumar et al., 2016).
difference in the temperature between radiosonde and COSMIC GPSRO
is found around 0.64 K between 10 and 27 km (Rao et al., 2009). In the
2.3. Compact Optical Backscatter AerosoL detector (COBALD) present study, we used Level 2 dry profiles from the COSMIC GPS RO
data products at 2◦ x 2◦ spatial resolution.
COBALD is a lightweight balloon-borne backscatter sonde developed
by ETH Zurich designed to fly with weather balloons operating at optical 2.6. CALIOP on-board CALIPSO satellite
wavelengths 455 nm (blue) and 940 nm (infrared). Each of these LEDs
For investigating the spatial distribution of the ATAL and its upper
Table 1 boundary, we used Cloud-Aerosol LIdar with Orthogonal Polarization
BATAL data used in the present study with a minimum altitude coverage of 22 (CALIOP) on-board Cloud-Aerosol Lidar and Infrared Pathfinder
km. O3: Ozonesonde, COBALD: Compact Optical Backscatter AerosoL Detector, Observation (CALIPSO) (Winker et al., 2009) during 2014–2018 which
CFH: Cryogenic Frost-point Hygrometer. operates at 532 nm. CALIOP is a nadir viewing, sun synchronised active
Year Location Payloads remote sensing instrument with two optical wavelengths (1064 nm and
2014 Gadanki (13.45◦ N, 79.18◦ E) O3 (4), COBALD (2) 532 nm) produced simultaneously by the frequency doubling method.
2015 Gadanki (13.45◦ N, 79.18◦ E) O3 (1) These lasers produce a peak power of 100 mJ at a pulse repetition rate of
2015 Hyderabad (17.47◦ N, 78.58◦ E) O3 (5), COBALD (6), CFH (5) 20.16 Hz. Initially, we used the level 3 (L3) CALIOP extinction profile to
2015 Varanasi (25.27◦ N, 82.99◦ E) O3 (3), COBALD (3), CFH (1)
illustrate the spatial distribution of the ATAL during the study period.
2016 Gadanki (13.45◦ N, 79.18◦ E) O3 (6), COBALD (2)
2016 Varanasi (25.27◦ N, 82.99◦ E) COBALD (3) Since the CALIOP L3 product does not provide aerosol information in the
2017 Gadanki (13.45◦ N, 79.18◦ E) O3 (1), COBALD (2), CFH (1) upper troposphere (<16 km) over the ASMA region, we used the level 1
2017 Hyderabad (17.47◦ N, 78.58◦ E) O3 (6) (L1) CALIOP data product to obtain the backscatter ratio at 532 nm
2018 Hyderabad (17.47◦ N, 78.58◦ E) O3 (8), COBALD (5), CFH (3) between 10 and 22 km. To overcome the low signal-to-noise ratio of the
2019 Hyderabad (17.47◦ N, 78.58◦ E) O3 (9), CFH (7)
CALIOP observation in the UTLS region, Vernier et al. (2009, 2015,

3
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

2018) have developed a specific treatment for the level 1 product to dilation, respectively. An appropriate selection of a is the main challenge
retrieve the backscatter profiles in the UTLS region. The data sets used in in accurate detection of the cloud layer. After going through different
the present study is obtained from the same algorithm at 5◦ x 2◦ (lon x profiles and experimenting with various values, a = n × Δz = 200 m is
lat) spatial resolution. In the present study, we didn’t use CALOP ob­ chosen with n = 8 and altitude resolution, Δz = 25 m. The dilation a is
( )
servations after August 15, 2017 and 2019 to avoid the effect of smoke the extent of the step function, h z−a b and translation b determines the
from the Canadian wildfire and Raikoke eruption, respectively. location of the step. Negative and positive peaks are used to identify the
cloud base and top with respect to a pre-defined threshold value. An
2.7. ERA5 reanalysis appropriate threshold for the detection of cloud is fixed after inspecting
multiple profiles and further verified the cloud detection by checking
In the present study, we also used European Centre for Medium- manually. We fixed the threshold in such a way that the BSR455 > 1.12
Range Weather forecasts (ECMWF) Re-Analysis-5 (ERA-5) datasets. (Hanumanthu et al., 2020) will be identified as a cloud layer around the
ERA-5 is the fifth generation of the ECMWF atmospheric reanalysis of UTLS (10–22 km) region. By making use of threshold values that linearly
the global climate data set with a higher time resolution than the pre­ varies as a function of altitude it is possible to remove multiple clouds as
vious version of ERA-Interim (Hersbach et al., 2020). We used ERA-5 well as, geometrically and optically thin clouds (Pandit et al., 2014)
temperature, zonal wind, and potential vorticity at 18:00 UTC at 1◦ x from the backscatter signal. However, the prior knowledge of the
1◦ spatial resolution at 37 pressure levels. backscatter ratio that represents the cloud layer helps here to fix the
threshold to a single value.

2.8. Identification of tropopause parameters and cloud removal using


3. Results
wavelet covariance transform (WCT)
3.1. Defining the top of the ATAL
In the present study, we used altitude profiles of temperature (T),
potential temperature (θ) and square of Brunt-Väisälä frequency (N2)
In this study, we derived all the tropopause parameters from indi­
from radiosonde observations over three locations to identify the
vidual observations which have a minimum altitude coverage of 22 km.
different tropopause parameters such as cold point tropopause height
Fig. 2(a–c) shows the altitude profiles T, the gradient of θ (dθ/dz) and N2
(CPH), lapse rate tropopause height (LRH), the convective overflow
obtained from radiosonde observations on August 01, 2017,17:30 UTC
height (COH) and the layer of maximum stability (LmaxS) by following
over Gadanki, respectively. The tropopause parameters identified from
Sunilkumar et al. (2017). The CPH and LRH are derived from temper­
these profiles are shown with labelled dash lines. The CPH is located
ature profiles. The coldest point in the temperature profile below 20 km
around 17.0 km with a potential temperature (temperature) of 374.30 K
is considered as CPH. The traditional WMO tropopause (LRH) is defined
(190.94 K). The LRH is found at 16.80 km, 200 m below the CPH with a
as ‘the lowest level at which the lapse rate becomes less than 2 K/km,
potential temperature (temperature) of 370.86 K (191.12 K). The COH is
provided that the average lapse rate from this level to the next 2 km
located around 12.5 km and the potential temperature (temperature) is
remains less than 2 K/km (WMO, 1957)’. The COH is identified from the
found to be 351.97 K (218.42 K). N2 peaked around 18.4 km, 1.4 km
minimum gradient of θ below CPH after smoothening the nine-point
above CPH and marked the LmaxS with a potential temperature (tem­
running mean (Gettelman and Forster, 2002). If there are multiple
perature) of 428.43 K (203.74 K). The simultaneous observation of the
troughs, then whose value is less than 5 times the minimum value and is
backscatter ratio at 455 nm (BSR455) from the COBALD sonde is shown
less than 5 K/km and close to the CPH is considered as COH (Mehta
in Fig. 2(d) along with the tropopause parameters from the radiosonde.
et al., 2008). The altitude at which the N2 peaks is identified as the
Two high-level clouds have been noticed around 11 km and 15.7 km
LmaxS (Gettelman and Wang, 2015; Sunilkumar et al., 2017) after
with a thickness of 600 m and 250 m, respectively. An increase in BSR455
smoothening nine-point running mean. N2 is calculated from potential
above the tropopause is noticeable and it shows a minimum around the
temperature derived from radiosonde observations and defined as N2 = LmaxS. The LmaxS coincide with a thermal inversion and this inversion
( )( )
g/ dθ/ where, g is the acceleration due to gravity and dz is the limits the further vertical extend of the ATAL. Therefore, the top of the
θ dz
vertical interval. The maximum static stability is a result of the shape of ATAL can be determined by the upper-level inversion, a few meters/
the temperature profile. The thermal profile is a consequence of the kilometers above the tropopause. Fig. 3 shows individual profiles of
radiative (Randel et al., 2007) and/or dynamical processes (Son and BSR455 and T over Gadanki, Hyderabad and Varanasi along with
Polvani, 2007; Grise et al., 2010). tropopause parameters suggesting as well that LmaxS can be considered
We performed Wavelet Covariance Transform (WCT) (Brooks, 2003; as the upper boundary of the ATAL. The mean BSR455 is estimated by
Pandit et al., 2014) analysis to detect cloud layers in the observed removing the clouds from individual profiles using the wavelet covari­
backscatter ratio profile. Once the cloud top and bottom are identified, ance transform (WCT) method. The raw BSR455 profile, WCT corre­
the corresponding cloud layers are removed from the profile to create sponding to the raw BSR455 profile and cloud removed BSR455 on August
cloud-free backscatter ratio profiles. The WCT for the backscatter ratio 01, 2017, 17:30 UTC over Gadanki are assembled in Fig. S1. To avoid
profile is defined as, potential contamination, we have removed BSR455 of a few meters
∫zt ( ) below and above the cloud layer. The mean BSR455 with tropopause
Wp (a, b) =
1
P(z)h
z− b
dz (1) parameters estimated from the corresponding radiosonde observations
a a are shown in Fig. 4. The mean CPH over Gadanki, Hyderabad and
Varanasi are found at around 16.75 ± 0.68 km, 17.08 ± 0.42 km and
zb

With the Haar function, 17.33 ± 0.90 km, respectively with a potential temperature of 373.35 ±
⎧ a 10.45 K, 377.69 ± 5.65 and 378.95 ± 14 K, respectively. The mean
( ) ⎪

⎪ +1, b −
⎪ 2
≤ z ≤ b, LmaxS over these locations is seen around 18.98 ± 0.70 km, 18.84 ±

z− b 0.91 and 20.01 ± 1.0 km, respectively, with a potential temperature of
h = − 1, b ≤ z ≤ b + a (2)
a ⎪

⎪ 2 444.61 ± 20.81 K, 438.14 ± 27.43 and 450.18 ± 28.18 K, respectively.


0, elsewhere The LmaxS is found about 1.7–2.7 km above the CPH over these three
locations during the monsoon season (June-July-August (JJA)). The
P(z) in equation (1) is the backscatter signal from COBALD sonde, z is BSR455 shows a minimum at LmaxS and this minimum separate the
the altitude, zb and zt are the lower and upper limit of the profile, ATAL from the stratospheric aerosol layer (Junge layer).
respectively. In the Haar function, b and a are called translation and

4
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Fig. 2. Profiles of (a) temperature, (b) potential temperature gradient (dθ/dz) and (c) square of Brunt Väisäla frequency (N2) derived from August 01, 2017, 17:30
UTC radiosonde measurement over Gadanki. (d) Profile of backscatter ratio at 455 nm observed during the same launch using a COBALD sonde. The solid dash lines
indicate the tropopause parameters derived from the radiosonde observations and they are labelled in (a).

Fig. 3. Backscatter ratio at 455 nm and temperature profile over (a) Gadanki on August 01, 2017, (b) Hyderabad on August 26, 2018, and (c) Varanasi on August 21,
2015. The dashed lines indicate the respective day’s tropopause parameters derived from radiosonde observations.

3.2. Water vapor, ozone and cloud from balloon-borne observations small variation around the LmaxS, and the ozone shows a steady in­
crease above this stable layer. In Fig. S2 we have shown mean ozone
Our analysis of water vapor and ozone from CFH and ozonesonde profiles from all the available observations over Gadanki (14 profiles),
measurements reveals distinct features around the tropopause. Mean Hyderabad (28 profiles) and Varanasi (3 profiles) along with the mean
profiles of water vapor and ozone (8 profiles) from simultaneous CFH tropopause parameters estimated from individual profiles. The sharp
and ozonesonde measurements along with mean tropopause parameters gradient in the mean ozone profile around the LRH is clearly noticeable
calculated from individual observations over Hyderabad are shown in in all three station’s observations. However, the change in the ozone
Fig. 5. A confined layer of water vapor is observed between CPH and concentration around the LmaxS is not prominent, though a small
LmaxS with a localised maximum of ~ 5ppmv in the mean profile. The change is observed. In Figs. S3a and S3b, we plotted ozone relative
two vertical dashed lines indicate 4 ppmv (blue) and 5 ppmv (red) of difference and ozone mean absolute difference over Hyderabad along
water vapor mixing ratio. The minimum water vapor mixing ratio (~4 with tropopause parameters, respectively. The ozone relative difference
ppmv) is observed around CPH and LmaxS (within the standard devia­ is mostly less than 20 ppbv below COH and it gradually increases from
tion). Mean CPH (θ) from these observations is found at around 17.20 ± COH to LmaxS and beyond. Variation in ozone mixing ratio near the
0.31 km (377.06 ± 4.66 K) and LmaxS(θ) is located around 18.37 ± tropopause and LamxS is clearly depicted in the mean ozone absolute
0.28 (423.00 ± 7.12 K). The ozone mixing ratio profile shows higher difference than in the ozone relative difference profiles. The higher
variability between COH and LRH and it shows a sharp increase in its gradient in mean ozone absolute difference coincides with the COH,
concentration above LRH. However, the mean ozone profile shows a LRH/CPH and LmaxS (within the standard deviations). High altitude

5
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Fig. 4. Cloud removed mean backscatter ratio at 455 nm (blue) and mean temperature (red) along with their standard deviation observed over (a) Gadanki (6
profiles), (b) Hyderabad (11 profiles) and (c) Varanasi (6 profiles). The dashed straight lines show the mean tropopause parameters and its standard deviations are
indicated with vertical bars.

clouds were being detected by COBALD sonde during the BATAL CPH. Within TTL, the water vapor slowly ascends within the anticyclone
campaign. Convection frequently reaches the tropopause height during and increases the RHice with altitude. Once the RHice reaches the cloud
the Asian Summer Monsoon (ASM) (Basha et al., 2021), not necessarily nucleation level, ice crystals start to form in the upper TTL region. They
penetrating the tropopause, however. The overshooting clouds can hy­ fall through the TTL and grow in the saturation region by dehydrating
drate the lower stratosphere by injecting ice crystals directly into the the TTL and eventually sublimate below the TTL and hydrate the lower
lower stratosphere (~420 K) (Corti et al., 2008). Our observation on boundary of the TTL. This repetitive process enhances water vapor
August 17, 2018, 19:40 UTC using COBALD and CFH sondes detected an below the tropopause and around the COH, as observed. Higher RHice
overshooting cloud with cloud top (~385.22 K) above CPH (~374.64 K) values near the COH and around the CPH are prominent in Fig. 5(d).
and shown in Fig. 5(c). The BSR455 (>1.12) and RHice (>78%, >12 km) Above the tropopause, the RHice decreases since the temperature starts
(Narendra Reddy et al., 2018; Renju et al., 2021) clearly show a cloud to rise in the lower stratosphere. The meteorological condition on
extending from 15.5 km–17.6 km (~364.01 K–385.22 K). The approx­ August 17, 2018 is provided in Fig. S4. Higher tropopause over the
imate temperature at which all the liquid water will spontaneously monsoon region is noticed on August 17, 2018 compare to the global
condense into ice crystals is − 40 ◦ C (Pruppacher and Klett, 1980; Koop mean tropopause. PV anomaly shows higher PV on this day between
et al., 2000). The observed clouds base was found around 15.5 km with a sub-tropical jet (thick black line) and tropical easterly jet streams (thick
temperature of − 73.3 ◦ C and hence it is an ice cloud. On that day, the dashed black line) (20–40◦ N) near the tropopause.
CPH and LRH were found to be at the same altitude, ~17.2 km with a
minimum water vapor mixing ratio of ~3.40 ppmv. Above cloud top, an
enhancement in water vapor with a maximum of ~5.68 ppmv (RHice, 3.3. Aerosols from the satellite observations
~60%) is noticed, and the concentration decreases to typical strato­
spheric value around the LmaxS. An enlarged view of the tropopause ATAL is seen in both in-situ and satellite observations around the
region is shown in Fig. 5(d) along with tropopause parameters to portray cold point tropopause region. Stratospheric aerosol extinction from
the enhanced water vapor layer above the cloud top. Though the in­ CALIOP is shown in Fig. S5 over the Indian region (65–110◦ E) and a
crease in the water vapor is small, it is significant in the tropical lower high extinction coefficient is observed around the ATAL region. The
stratosphere at cold temperatures. Higher water vapor at this cold extinction coefficient shows the presence of ATAL well above the CPH
temperature can favour the in-situ production of aerosols via photo­ and minimum near LmaxS. However, the level 3 (L3) CALIOP extinction
chemistry (Höpfner et al., 2019) as well as the hygroscopic growth of the profile does not provide aerosol information in the upper troposphere
aerosols in the UTLS region (He et al., 2019). Simultaneously observed (~10–16 km). To overcome this limitation, we used the CALIOP L1 data
ozone from ozonesonde recorded low concentration in the upper product with a specific treatment (Vernier et al., 2009, 2015, 2018). The
troposphere and started to increase around CPH/LRH. The ozone profile zonal (65–110◦ E, Indian region) and meridional (15–25◦ N) average
has also shown a variation in its concentration around LmaxS, however, cloud-free aerosol backscatter ratio at 532 nm (BSR532) from CALIOP is
the change was small. A similar overshooting cloud is also detected over shown in Fig. 6(a) and (b), respectively. The ATAL is located above 5◦ N
Nainital, India during 2016 balloon-borne observations using COBALD latitude belt with maximum BSR532 between ~15 and 27.5◦ N. The
and CFH sondes (Hanumanthu et al., 2020, Fig. 2 and Fig. A1, average meridional BSR532 shows the ATAL between 0◦ – 150◦ E with
12-08-2016). maximum aerosols between 15 and 120◦ E. Similarly, higher BSR is seen
In addition to this, we also noticed an enhancement in water vapor around CPH, and above CPH the BSR532 gradually decreases with the
below the tropopause, ~16 km and around COH in the mean water altitude and forms a layer of low BSR532. This low BSR532 layer matches
vapor profile (Fig. 5(a)). This enhancement in water vapor may be ac­ with the COSMIC GPSRO derived LmaxS (within the standard devia­
cording to the hypothesis by Schoeberl et al. (2019). Most of the con­ tion). The BSR start to increase gradually above the LmaxS and it
vection detrained around the COH and are the primary source of water identifies the stratospheric aerosol layer called ‘Junge Layer’. The
vapor in the tropical tropopause layer (TTL), region between COH and gradual decrease and attaining a low BSR around LmaxS above CPH is
clearly noticed in both the COBALD and CALIOP observations.

6
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Fig. 6. (a) The zonal mean (65–110◦ E) and meridional mean (15–25◦ N)
backscatter ratio at 532 nm from CALIOP observation during 2014–2018 as a
function of altitude. The dashed straight lines with vertical bars indicate the
mean and standard deviation of the tropopause parameters derived from
COSMIC GPS-RO satellite’s dry temperature and pressure profiles during the
same period and they are labelled in the figures.

calculate the tropopause parameter within the anticyclone. The mean


CPH (θ) within the anticyclone is around 17.50 ± 0.29 km (392.30 ±
6.27 K). The LRH(θ) is found at around 17.00 ± 0.25 km (383.84 ± 4.72
K). The minimum potential temperature gradient is located around
13.10 ± 0.96 km with a θ of 355.78 ± 5.18 K. The LmaxS(θ) is found
around 19.63 ± 0.46 km (454.39 ± 13.89 K) within the anticyclone.

4. Summary and discussion

The existence of the lower stratospheric aerosol layer known as


‘Junge Layer’ and the elevated upper tropospheric aerosol layer called
ATAL are well established. However, what separates these two layers
was not yet clear so far since the ATAL is observed well above the
tropopause. In this study, we introduced the concept of static stability
(LmaxS) to define the top of the ATAL. The LmaxS derived from radio­
Fig. 5. Profiles of (a) mean water vapor mixing ratio and (b) ozone mixing ratio sonde along with the backscatter ratio from COBALD sonde observations
along with mean temperature profile over Hyderabad. The blue and red shades has clearly shown that the ATAL can vertically extend up to the LmaxS
show the standard deviation of the observations. The dashed straight lines in and beyond that, the Junge layer exists. This has been verified using
figure (a) and (b) indicates the mean tropopause parameters and the vertical balloon measurements from multiple locations over India and the sat­
bars show their standard deviation. (c) The simultaneously observed back­ ellite observations provided a similar but a regional picture. Now, we
scatter ratio at 455 nm (blue), water vapor mixing ratio (magenta), ozone will further discuss how the LmaxS is acting as an upper boundary for
mixing ratio (black), temperature (red) and relative humidity over ice (orange) the ATAL with the observed results and in light of the other supporting
on August 17, 2018 over Hyderabad. The dashed straight lines indicate the information.
tropopause parameters derived from the radiosonde observations and are
The higher water vapor mixing ratio found near the cold point
labelled accordingly. An enlarged view of the water profile is shown in (d)
tropopause suggests deep convection can reach up to the upper tropo­
along with tropopause parameters. The dotted vertical line in the enlarged plot
corresponds to the water vapor mixing ratio of ~3.40 ppmv (blue) and ~5.68
sphere and even penetrate the tropopause, which could also transport
ppmv (red). other trace gases, aerosols or its gas phase precursors (Fig. 5(a)). The
water vapor peak between CPH and LmaxS is consistent with the hy­
dration of the lower stratosphere by overshooting convective uplifts/
Therefore, the lower stratospheric temperature inversion act as a
anvils (Fig. 5(c)). Our balloon observations of overshooting cloud,
capping inversion for the ATAL and it restricts the further upward
higher water vapor above CPH and ATAL around tropopause are
transport of aerosols by the weak BDC during NH monsoon.
consistent with the earlier studies. In-situ observations from both
We have utilized the COSMIC-GPSRO satellite dry profiles to
balloon-borne and aircraft measurements have shown clear evidence of

7
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

direct transport of ice particles and water vapor into the lower strato­ Table 2
sphere (Corti et al., 2008; Lee et al., 2019; Khaykin et al., 2021). The Average tropopause parameters estimated over Gadanki, Hyderabad and Vara­
Geophysica campaign observations by Corti et al. (2008) has shown nasi from radiosonde measurements during the 2014–2019 BATAL campaigns.
convective system penetrated the stratosphere and deposits of ice par­ Gadanki Hyderabad Varanasi
ticle at altitude reaching 420 K potential temperature. In the present km Θ (K) km Θ (K) km Θ (K)
study, we have observed a cloud top reaching an altitude (17.6 km)
CPH 16.75 373.35 17.08 377.49 17.325 378.95
corresponding to 386.98 K potential temperature and above this level,
± 0.68 ± 10.45 ± 0.42 ± 5.65 ± 0.90 ± 14.04
evaporation of ice crystal leads to an increase in vapor (Fig. 5(d)). The LRH 16.26 366.69 16.62 371.49 16.5 ± 367.80
top of the enhanced water vapor layer matches with the LmaxS at a ± 0.43 ± 4.67 ± 0.59 ± 6.49 0.81 ± 4.10
potential temperature 425.5 K. The hydration of the lower stratosphere COH 13.05 352.02 13.30 354.92 13.92 ± 358.71
by direct injection of ice crystal and later evaporation provides favour­ ± 1.10 ± 3.47 ± 1.18 ± 4.20 1.69 ± 2.86
LmaxS 18.98 444.61 18.84 438.14 20.01 ± 450.18
able conditions for new particle formation at low temperature and ± 0.70 ± 20.81 ± 0.91 ± 27.43 1.00 ± 28.18
pressure in the presence of sunlight (Vernier et al., 2018; Höpfner et al.,
2019 and references therein). Apart from this, the higher water vapor
also facilitates hygroscopic growth of aerosols which are convectively high mixing ratio in the upper troposphere and its mixing ratio decreases
transported as well as in-situ generated in the UTLS (He et al., 2019). with altitude and reaches a typical stratospheric background value
Various model simulations have reported that the sulphur dioxide (SO2), around LmaxS. The CPH and LRH locate well below the level of typical
the sulphate aerosol precursors, can survive the convective uplift and stratospheric background mixing ratio and this indicates significant
reach up to the UTLS. The ammonia (NH3) observations from the MIPAS stratospheric in-mixing, which further implies neither CPH nor the LRH
satellite and AN observations from the CRISTA (Höpfner et al., 2019) represents the separation between the troposphere and stratosphere in
also agree with the hypothesis of new particle formation at UTLS. the ASMA region in both chemical and dynamical sense. Thus, the
Our analysis suggests that the LmaxS found ~1 km–2.7 km above the ‘tropospheric bubble’ (Pan et al., 2016) not only extends to the excep­
CPH corresponding to the potential temperature ~442.11 ± 25.64 K tionally high tropopause but also to the lower stratosphere over ASMA.
(454.39 ± 13.89 K) over India (ASMA) marks the upper boundary of the The carbon monoxide and ozone relationship for the ATAL and pacific
ATAL. The balloon observations over India show a mean BSR455 of (0–40◦ N, 180◦ W- 140◦ W) regions during 2014–2018 are shown in
~1.048 ± 0.016 at LmaxS, ~19.27 ± 0.87. The BSR455 observations Fig. S6 with corresponding potential temperatures (colored open cir­
from CALIOP over the ASMA region also show that the LmaxS identify cles). In the scatter plot, the filled diamond scatters show the CPH (red
the top of the ATAL with a BSR value of ~1.045 and above the LmaxS border) and LmaxS (black) potential temperatures, and the filled square
the BSR once again starts to increase with the altitude and it marks the boxes show COH (red) and LRH (black) potential temperatures,
Junge layer. The concept of an upward spiral region within the ASMA respectively. The transition of the trace gas mixing ratio in the UTLS
with its upper boundary around 460 K potential temperature (Vogel region is well represented in the CO–O3 tracer -tracer famous ‘L’ shape
et al., 2019) is in good agreement with our LmaxS. The mean altitude curve. CO mixing ratio shows a gradual reduction from its higher values
(potential temperature) of LmaxS is located over Gadanki, Hyderabad around COH altitude, where almost all convection ceases and from the
and Varanasi at 18.98 ± 0.70 km (444.61 ± 20.81 K), 18.84 ± 0.91 km same altitude O3 mixing ratio started to increase. As we showed in Fig. 7,
(438.14 ± 27.43 K) and 20.01 ± 1.00 km (450.18 ± 28.18 K), respec­ the CO mixing ratio approaches its stratospheric background mixing
tively. The LmaxS limit the ATAL, possibly because it is mainly ratio around LmaxS and O3 shows a steady increase in its mixing ratio
composed of small liquid aerosols (~80% - 95%) formed from gas to above LmaxS. However, the ‘L’ isn’t as sharp as in other regions such as
liquid conversion via photochemistry at low temperature and pressure the extra-tropics and the smoother transition between the tropospheric
in the presence of higher water vapor around the CPH (Vernier et al., and stratospheric branches could be an indication of a transition or
2018). The existence of a higher water vapor mixing ratio above the CPH missing layer. The observed higher mixing ratio of trace gases above
also implies the same. Further, these higher water vapor concentrations tropopause and approaching the mixing ratio to the typical stratospheric
could trigger the hygroscopic growth of the liquid or solid aerosols background value is in good agreement with the recent aircraft obser­
formed in-situ in the UTLS and the aerosols transported to the upper vations by Hobe et al. (2021). However, the poor vertical resolution of
troposphere via deep convection over south Asia (He et al., 2019). This ACE-FTS is inadequate to explain the possible photochemical loss of
will also limit the upward transport of aerosols via weak Brewer-Dobson trace gases such as CO and define the top of the ASMA as LmaxS. High
Circulation (BDC) during the northern hemisphere monsoon along with vertical resolution in-situ observations and model outputs are required
the temperature inversion at LmaxS. Therefore, a part of the aerosols to investigate the role of LmaxS as the top of ASMA since not only the
may sediment back to the upper troposphere under the action of gravity microphysical properties limit the transport of the trace gases to the
and limits its increase in the concentration in the lower stratosphere lower stratosphere but also the photochemical reactions.
alike the water vapor freeze-drying and ice crystal formation at tropical
tropopause. The large scale BDC transports the pollutants to a higher 5. Conclusions
altitude above 460 K via the tropical pipe (within ~1 year) (Vogel et al.,
2019). We analysed balloon measurements of aerosol backscatter ratio,
The observed changes in ozone in between the CPH and LmaxS is due ozone, water vapor, temperature and tropopause parameters over
to the injection of tropospheric air into the lower stratosphere. However, Gadanki, Hyderabad and Varanasi conducted during 2014–2019 as a
the strong gradient in ozone mixing ratio observed around the lapse rate part of Balloon Measurement Campaigns of the Asian Tropopause
tropopause suggests almost all updraft becomes weaker around LRH and Aerosol Layer (BATAL) to identify the top of the ATAL. We also make use
further uplift by spiralling range take a few days. The mean tropopause of multi-satellite observations to investigate its spatial behaviour. The
parameters and their corresponding potential temperature is provided in main conclusions drawn from the study are listed below.
Table 2. The bottom layer of the proposed spiralling range (~360 K) by
Vogel et al. (2019) is found between the COH and LRH. We also analysed 1. Our balloon-borne observations over India and CALIOP and COSMIC
ACE- FTS satellite trace gas products such as CO, H2O, HCN and O3 to GPSRO observations over Asian Summer Monsoon Anticyclone
further verify the role of LmaxS in defining the upper boundary of the (ASMA) region suggest that the layer of maximum stability (LmaxS)
ATAL and ASMA and shown in Fig. 7. It is clear from the figure that there derived from the square of Brunt Väisäla frequency (N2) can define as
is no sharp gradient exists in any of these trace gases around the LRH or the top of the Asian Tropopause Aerosol Layer (ATAL) and it sepa­
CPH. The tropospheric traces gases such as CO, H2O and HCN shows rates the ATAL from the well-known stratospheric ‘Junge layer’.

8
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Fig. 7. The average profile of (a) carbon monoxide (CO), (b) Water Vapor (H2O), (c) Hydrogen Cyanide and (d) ozone inside the anticyclone (10–40◦ N, 0–120◦ E)
during 2014–2018. The straight lines indicate the mean tropopause parameters inside the anticyclone during the same period derived from the COSMIC GPS-RO dry
temperature and pressure profiles. The vertical bar in the figures indicates the standard deviation of the tropopause parameters. The res vertical line in (b) represents
4 ppmv.

2. The LmaxS is found 1–2.7 km above the cold point tropopause cor­ interests or personal relationships that could have appeared to influence
responding to the potential temperature ~442.11 ± 25.64 K (454.39 the work reported in this paper.
± 13.89 K) over India (ASMA) and the COBALD observations over
Indian stations show that the typical value of BSR455 at LmaxS is Acknowledgements
~1.048 ± 0.016.
3. The top of the enhanced water vapor layer in the lower stratosphere The results presented in this paper are obtained from the ISRO-NASA
also coincides with the LmaxS with a secondary minimum in its joint BATAL campaign supported by the National Atmospheric Research
concentration. Laboratory (NARL), Department of Space (DoS) and NASA ROSES Upper
4. The trace gas observations from the ACE-FTS satellite do not show a Atmospheric Research and Atmospheric Composition Modelling and
sharp gradient in its concentration around tropopause. A gradual Analysis programs (UARP, ACMAP). We would like to thank technicians
reduction in tropospheric trace gas concentration is observed above and engineers from NARL, TIFR Balloon Facility and BHU Varanasi for
the CPH and it approaches the stratospheric background value their help during the BATAL campaigns. We would like to thank Dr. T. D.
around LmaxS. However, without considering the photochemical Fairlie and Prof. A. Jayaraman for their initial helps in the realization of
processes and mixing within the ASMA region during the slow ascent this project. The authors acknowledge UCAR/CDAAC and the ACE sat­
of air masses we could not conclude whether the LmaxS can also ellite mission funded by the Canadian Space Agency for providing
consider as the top of the ASMA tropospheric bubble. COSMIC GPSRO and ACE-FTS data sets through their websites, respec­
tively. We also thank CALIPSO, satellite teams for providing data sets
Trace gases are not affected by microphysics while crossing the through their website. We also thank Giovanni (https://giovanni.gsfc.
tropopause region. Hence, the photochemical reaction and lifespan of a nasa.gov/giovanni/) for providing AIRS tropopause data through their
trace gas are required to define the upper boundary of the confined website. We also like to thank the European Centre for Medium-Range
ASMA region. The ACE-FTS observations show no sharp gradient in Weather Forecasts ERA5 teams for providing the datasets through
trace gas concentration and a gradual reduction in its concentration. their website. Author Akhil Raj S T would like to thank Ms. Renju
Therefore, high vertical resolution in-situ observations and model Nandan for the help related to the WCT analysis.
simulation are required to verify further whether the LmaxS can
consider as the top of the confined ASMA region or ASMA tropospheric Appendix A. Supplementary data
bubble.
Supplementary data to this article can be found online at https://doi.
CRediT authorship contribution statement org/10.1016/j.apr.2022.101451.

S.T. Akhil Raj: Methodology, Visualization, Data curation, Writing – References


original draft. M. Venkat Ratnam: Visualization, Resources, Supervi­
sion, Writing – review & editing. J.P. Vernier: Resources, Data analysis, Akhil Raj, S.T., Venkat Ratnam, M., Narayana Rao, D., Krishna Murthy, B.V., 2015.
Vertical distribution of ozone over a tropical station: seasonal variation and
Writing – review & editing. A.K. Pandit: Writing – review & editing. comparison with satellite (MLS, SABER) and ERA-Interim products. Atmos. Environ.
Frank G. Wienhold: Data curation, Writing – review & editing. 116, 281–292. https://doi.org/10.1016/j.atmosenv.2015.06.047.
Alexander, P., de la Torre, A., Llamedo, P., Hierro, R., 2014. Precision estimation in
temperature and refractivity profiles retrieved by GPS radio occultations.
Declaration of competing interest J. Geophys. Res. 119, 8624–8638. https://doi.org/10.1002/2013JD021016.

The authors declare that they have no known competing financial

9
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Basha, G., Ratnam, M.V., Kishore, P., 2020. Asian summer monsoon anticyclone: trends Hobe, M. Von, Ploeger, F., Konopka, P., Kloss, C., Ulanowski, A., Yushkov, V.,
and variability. Atmos. Chem. Phys. 20, 6789–6801. https://doi.org/10.5194/acp- Ravegnani, F., Volk, C.M., Pan, L.L., Honomichl, S.B., Tilmes, S., Kinnison, D.E.,
20-6789-2020. Garcia, R.R., Wright, J.S., 2021. Upward transport into and within the Asian
Basha, G., Ratnam, M.V., Jiang, J.H., Kishore, P., Babu, S.R., 2021. Influence of indian monsoon anticyclone as inferred from StratoClim trace gas observations,
summer monsoon on tropopause, trace gases and aerosols in asian summer monsoon pp. 1267–1285.
anticyclone observed by cosmic, mls and calipso. Rem. Sens. 13, 3984–4003. Höpfner, M., Ungermann, J., Borrmann, S., Wagner, R., Spang, R., Riese, M., Stiller, G.,
https://doi.org/10.3390/rs13173486. Appel, O., Batenburg, A.M., Bucci, S., Cairo, F., Dragoneas, A., Friedl-Vallon, F.,
Bernath, P.F., McElroy, C.T., Abrams, M.C., Boone, C.D., Butler, M., Camy-Peyret, C., Hünig, A., Johansson, S., Krasauskas, L., Legras, B., Leisner, T., Mahnke, C.,
Carleer, M., Clerbaux, C., Coheur, P.F., Colin, R., DeCola, P., DeMazière, M., Möhler, O., Molleker, S., Müller, R., Neubert, T., Orphal, J., Preusse, P., Rex, M.,
Drummond, J.R., Dufour, D., Evans, W.F.J., Fast, H., Fussen, D., Gilbert, K., Saathoff, H., Stroh, F., Weigel, R., Wohltmann, I., 2019. Ammonium nitrate particles
Jennings, D.E., Llewellyn, E.J., Lowe, R.P., Mahieu, E., McConnell, J.C., formed in upper troposphere from ground ammonia sources during Asian monsoons.
McHugh, M., McLeod, S.D., Michaud, R., Midwinter, C., Nassar, R., Nichitiu, F., Nat. Geosci. 12, 608–612. https://doi.org/10.1038/s41561-019-0385-8.
Nowlan, C., Rinsland, C.P., Rochon, Y.J., Rowlands, N., Semeniuk, K., Simon, P., Hoskins, B.J., Rodwell, M.J., 1995. A model of the asian summer monsoon. Part I: the
Skelton, R., Sloan, J.J., Soucy, M.A., Strong, K., Tremblay, P., Turnbull, D., global scale. J. Atmos. Sci. 52, 1329–1340.
Walker, K.A., Walkty, I., Wardle, D.A., Wehrle, V., Zander, R., Zou, J., 2005. Khaykin, S.M., Moyer, E., Krämer, M., Clouser, B., Bucci, S., Lykov, A., Afchine, A.,
Atmospheric chemistry experiment (ACE): mission overview. Geophys. Res. Lett. Cairo, F., Formanyuk, I., Mitev, V., Rolf, C., Singer, C., Spelten, N., Volkov, V.,
https://doi.org/10.1029/2005GL022386. Yushkov, V., 2021. Persistence of moist plumes from overshooting convection in the
Bian, J., Li, D., Bai, Z., Li, Q., Lyu, D., Zhou, X., 2020. Transport of Asian surface Asian monsoon anticyclone. Atmos. Chem. Phys. Discuss. 1–27.
pollutants to the global stratosphere from the Tibetan Plateau region during the Kim, Y.-S., Shibata, T., Iwasaka, Y., Shi, G., Zhou, X., Tamura, K., Ohashi, T., 2003.
Asian summer monsoon. Natl. Sci. Rev. 7, 516–533. https://doi.org/10.1093/nsr/ Enhancement of aerosols near the cold tropopause in summer over Tibetan Plateau:
nwaa005. lidar and balloonborne measurements in 1999 at Lhasa, Tibet, China. Lidar Remote
Birner, T., Dörnbrack, A., Schumann, U., 2002. How sharp is the tropopause at Sens. Ind. Environ. Monit. III 4893, 496. https://doi.org/10.1117/12.466090.
midlatitudes? Geophys. Res. Lett. 29, 1–4. https://doi.org/10.1029/2002GL015142. Komhyr, W.D., Barnes, R.A., Brothers, G.B., Lathrop, J.A., Opperman, D.P., 1995.
Boone, C.D., Nassar, R., Walker, K.A., Rochon, Y., McLeod, S.D., Rinsland, C.P., Electrochemical concentration cell ozonesonde performance evaluation during
Bernath, P.F., 2005. Retrievals for the atmospheric chemistry experiment Fourier- STOIC 1989. J. Geophys. Res. 100, 9231–9244. https://doi.org/10.1029/
transform spectrometer. Appl. Opt. 44, 7218–7231. https://doi.org/10.1364/ 94JD02175.
AO.44.007218. Koop, T., Luo, B., Tsias, A., Peter, T., 2000. Water Activity as the Determinant for
Brunamonti, S., Jorge, T., Oelsner, P., Hanumanthu, S., Singh, B.B., Ravi Kumar, K., Homogeneous Ice Nucleation in Aqueous Solutions, pp. 611–614.
Sonbawne, S., Meier, S., Singh, D., Wienhold, F.G., Ping Luo, B., Boettcher, M., Kursinski, E.R., Hajj, G.A., Schofield, J.T., Linfield, R.P., Hardy, K.R., 1997. Observing
Poltera, Y., Jauhiainen, H., Kayastha, R., Karmacharya, J., DIrksen, R., Naja, M., Earth’s atmosphere with radio occultation measurements using the global
Rex, M., Fadnavis, S., Peter, T., 2018. Balloon-borne measurements of temperature, positioning system. J. Geophys. Res. Atmos. 102, 23429–23465. https://doi.org/
water vapor, ozone and aerosol backscatter on the southern slopes of the Himalayas 10.1029/97jd01569.
during StratoClim 2016-2017. Atmos. Chem. Phys. 18, 15937–15957. https://doi. Lee, K.O., Dauhut, T., Chaboureau, J.P., Khaykin, S., Krämer, M., Rolf, C., 2019.
org/10.5194/acp-18-15937-2018. Convective hydration in the tropical tropopause layer during the StratoClim aircraft
Corti, T., Luo, B.P., de Reus, M., Brunner, D., Cairo, F., Mahoney, M.J., Martucci, G., campaign: pathway of an observed hydration patch. Atmos. Chem. Phys. 19,
Matthey, R., Mitev, V., dos Santos, F.H., Schiller, C., Shur, G., Sitnikov, N.M., 11803–11820. https://doi.org/10.5194/acp-19-11803-2019.
Spelten, N., Vössing, H.J., Borrmann, S., Peter, T., 2008. Unprecedented evidence for Ma, D., Bian, J., Li, D., Bai, Z., Li, Q., Zhang, J., Wang, H., Zheng, X., Hurst, D.F.,
deep convection hydrating the tropical stratosphere. Geophys. Res. Lett. 35, 1–5. Vömel, H., 2022. Mixing characteristics within the tropopause transition layer over
https://doi.org/10.1029/2008GL033641. the Asian summer monsoon region based on ozone and water vapor sounding data.
Fadnavis, S., Roy, C., Chattopadhyay, R., Sioris, C.E., Rap, A., Müller, R., Ravi Kumar, K., Atmos. Res. 271 https://doi.org/10.1016/j.atmosres.2022.106093.
Krishnan, R., 2018. Transport of trace gases via eddy shedding from the Asian Mahnke, C., Weigel, R., Cairo, F., Vernier, J., Afchine, A., Krämer, M., Mitev, V.,
summer monsoon anticyclone and associated impacts on ozone heating rates. Atmos. Matthey, R., Viciani, S., D’Amato, F., Ploeger, F., Deshler, T., Borrmann, S., 2021.
Chem. Phys. 18, 11493–11506. https://doi.org/10.5194/acp-18-11493-2018. The ATAL within the 2017 Asian Monsoon Anticyclone: microphysical aerosol
Garny, H., Randel, W.J., 2016. Transport pathways from the Asian monsoon anticyclone properties derived from aircraft-borne in situ measurements. Atmos. Chem. Phys.
to the stratosphere. Atmos. Chem. Phys. 16, 2703–2718. https://doi.org/10.5194/ Discuss. 21, 15259–15282. https://doi.org/10.5194/acp-2020-1241.
acp-16-2703-2016. Mehta, S.K., Krishna Murthy, B.V., Narayana Rao, D., Venkat Ratnam, M.,
Gettelman, A., Forster, P.M. de F., 2002. A climatology of the tropical tropopause layer. Parameswaran, K., Rajeev, K., Suresh Raju, C., Rao, K.G., 2008. Identification of
J. Meteorol. Soc. Japan 80, 911–924. https://doi.org/10.2151/jmsj.80.911. tropical convective tropopause and its association with cold point tropopause.
Gettelman, A., Wang, T., 2015. Structural diagnostics of the tropopause inversion layer J. Geophys. Res. Atmos. 113, 2–9. https://doi.org/10.1029/2007JD009625.
and its evolution. J. Geophys. Res. 120, 46–62. https://doi.org/10.1002/ Nandan, Renju, VenkatRatnam, M., RaviKiran, V., Naik, Dinesh N., 2022. Retrieval of
2014JD021846. cloud liquid water path using radiosonde measurements: comparison with MODIS
Goff, J.A., Gratch, S., 1946. Low-pressure properties of water from -160 to 212. F. Trans. and ERA5. J. Atmos. Sol. Terr. Phys. 227 https://doi.org/10.1016/j.
Amer. Soc. Heat Vent. Eng 52, 95–122. jastp.2021.105799.
Grise, K.M., Thompson, D.W.J., Birner, T., 2010. A global survey of static stability in the Narendra Reddy, N., Venkat Ratnam, M., Basha, G., Ravikiran, V., 2018. Cloud vertical
stratosphere and upper troposphere. J. Clim. 23, 2275–2292. https://doi.org/ structure over a tropical station obtained using long-term high-resolution radiosonde
10.1175/2009JCLI3369.1. measurements. Atmos. Chem. Phys. 18, 11709–11727. https://doi.org/10.5194/
Gu, Y., Liao, H., Bian, J., 2016. Summertime nitrate aerosol in the upper troposphere and acp-18-11709-2018.
lower stratosphere over the Tibetan Plateau and the South Asian summer monsoon Pan, L.L., Honomichl, S.B., Kinnison, D.E., Abalos, M., Randel, W.J., Bergman, J.W.,
region. Atmos. Chem. Phys. 16, 6641–6663. https://doi.org/10.5194/acp-16-6641- Bian, J., 2016. Transport of chemical tracers from the boundary layer to stratosphere
2016. associated with the dynamics of the asian summer monsoon, 159-14 J. Geophys. Res.
Hanumanthu, S., Vogel, B., Müller, R., Brunamonti, S., Fadnavis, S., Li, D., Ölsner, P., 121 (14). https://doi.org/10.1002/2016JD025616, 174.
Naja, M., Singh, B.B., Kumar, K.R., Sonbawne, S., Jauhiainen, H., Vömel, H., Luo, B., Pandit, A.K., Gadhavi, H., Ratnam, M.V., Jayaraman, A., Raghunath, K., Rao, S.V.B.,
Jorge, T., Wienhold, F.G., Dirkson, R., Peter, T., 2020. Strong day-to-day variability 2014. Characteristics of cirrus clouds and tropical tropopause layer: seasonal
of the asian tropopause aerosol layer (ATAL) in August 2016 at the himalayan variation and long-term trends. J. Atmos. Sol. Terr. Phys. 121, 248–256. https://doi.
foothills. Atmos. Chem. Phys. 20, 14273–14302. https://doi.org/10.5194/acp-20- org/10.1016/j.jastp.2014.07.008.
14273-2020. Park, M., Randel, W.J., Emmons, L.K., Bernath, P.F., Walker, K.A., Boone, C.D., 2008.
He, Q., Ma, J., Zheng, X., Yan, X., Vömel, H., Wienhold, F.G., Gao, W., Liu, D., Shi, G., Chemical isolation in the Asian monsoon anticyclone observed in Atmospheric
Cheng, T., 2019. Observational evidence of particle hygroscopic growth in the upper Chemistry Experiment (ACE-FTS) data. Atmos. Chem. Phys. 8, 757–764. https://doi.
troposphere-lower stratosphere (UTLS) over the Tibetan Plateau. Atmos. Chem. org/10.5194/acp-8-757-2008.
Phys. 19, 8399–8406. https://doi.org/10.5194/acp-19-8399-2019. Park, M., Randel, W.J., Kinnison, D.E., Emmons, L.K., Bernath, P.F., Walker, K.A.,
He, Q., Ma, J., Zheng, X., Wang, Y., Wang, Y., Mu, H., Cheng, T., He, R., Huang, G., Boone, C.D., Livesey, N.J., 2013. Hydrocarbons in the upper troposphere and lower
Liu, D., Lelieveld, J., 2020. Formation and dissipation dynamics of the Asian stratosphere observed from ACE-FTS and comparisons with WACCM. J. Geophys.
tropopause aerosol layer. Environ. Res. Lett. 16 https://doi.org/10.1088/1748- Res. Atmos. 118, 1964. https://doi.org/10.1029/2012JD018327, 1980.
9326/abcd5d. Pruppacher, H.R., Klett, J.D., 1980. Microphysics of clouds and precipitation. Nature
Hemanth Kumar, A., Ratnam, M.V., 2021. Variability in the UTLS chemical composition 284, 88. https://doi.org/10.1038/284088b0.
during different modes of the asian summer monsoon anti-cyclone. Atmos. Res. 260, Randel, W.J., Wu, F., Forster, P., 2007. The extratropical tropopause inversion layer:
105700. https://doi.org/10.1016/j.atmosres.2021.105700. global observations with GPS data, and a radiative forcing mechanism. J. Atmos. Sci.
Hersbach, H., Bell, B., Berrisford, P., Hirahara, S., Horányi, A., Muñoz-Sabater, J., 64, 4489–4496. https://doi.org/10.1175/2007JAS2412.1.
Nicolas, J., Peubey, C., Radu, R., Schepers, D., Simmons, A., Soci, C., Abdalla, S., Randel, W.J., Park, M., Emmons, L., Kinnison, D., Bernath, P., Walker, K.A., Boone, C.,
Abellan, X., Balsamo, G., Bechtold, P., Biavati, G., Bidlot, J., Bonavita, M., De Pumphrey, H., 2010. Asian monsoon transport of pollution to the stratosphere.
Chiara, G., Dahlgren, P., Dee, D., Diamantakis, M., Dragani, R., Flemming, J., Science 328 (80), 611–614.
Forbes, R., Fuentes, M., Geer, A., Haimberger, L., Healy, S., Hogan, R.J., Hólm, E., Rao, D.N., Ratnam, M.V., Mehta, S., Nath, D., Basha, S.G., Rao, V.V.M.J., Murthy, B.V.K.,
Janisková, M., Keeley, S., Laloyaux, P., Lopez, P., Lupu, C., Radnoti, G., de Tsuda, T., Nakamura, K., 2009. Val i da tion of the COS MIC Ra dio Occultation Data
Rosnay, P., Rozum, I., Vamborg, F., Villaume, S., Thépaut, J.N., 2020. The ERA5 over Gadanki ( 13 . 48 ◦ N , 79 . 2 ◦ E): a Trop i cal Re gion. Terr. Atmos. Ocean Sci.
global reanalysis. Q. J. R. Meteorol. Soc. 146, 1999–2049. https://doi.org/10.1002/ 20, 59–70. https://doi.org/10.3319/TAO.2008.01.23.01(F3C)1.
qj.3803.

10
S.T.A. Raj et al. Atmospheric Pollution Research 13 (2022) 101451

Ratnam, M.V., Pravallika, N., Ravindra Babu, S., Basha, G., Pramitha, M., Krishna Vernier, J.-P., Fairlie, T.D., Deshler, T., Ratnam, M.V., Gadhavi, H., Kumar, S.,
Murthy, B.V., 2014. Assessment of GPS radiosonde descent data. Atmos. Meas. Tech. Natarajan, M., Pandit, A.K., Raj, S.T., Kumar, A.H., Jayaraman, A., Singh, A.K.,
7, 1011–1025. https://doi.org/10.5194/amt-7-1011-2014. Rastogi, N., Sinha, P.R., Kumar, S., Tiwari, S., Wegner, T., Baker, N., Vignelles, D.,
Santee, M.L., Manney, G.L., Livesey, N.J., Schwartz, M.J., Neu, J.L., Read, W.G., 2017. Stenchikov, G., Shevchenko, I., Smith, J., Bedka, K., Kesarkar, A., Singh, V.,
A comprehensive overview of the climatological composition of the Asian summer Bhate, J., Ravikiran, V., Rao, M.D., Ravindrababu, S., Patel, A., Vernier, H.,
monsoon anticyclone based on 10 years of Aura Microwave Limb Sounder Wienhold, F.G., Liu, H., Knepp, T., Thomason, L., Crawford, J., Ziemba, L., Moore, J.,
measurements. J. Geophys. Res. 122, 5491–5514. https://doi.org/10.1002/ Crumeyrolle, S., Williamson, M., Berthet, G., Jégou, F., Renard, J.B., 2018. BATAL:
2016JD026408. the balloon measurement campaigns of the asian tropopause aerosol layer. A. Bull.
Schoeberl, M.R., Jensen, E.J., Pfister, L., Ueyama, R., Wang, T., Selkirk, H., Avery, M., Am. Meteorol. Soc. 99, 955–973. https://doi.org/10.1175/BAMS-D-17-0014.1.
Thornberry, T., Dessler, A.E., 2019. Water vapor, clouds, and saturation in the Vogel, B., Günther, G., Müller, R., Grooß, J.U., Riese, M., 2015. Impact of different Asian
tropical tropopause layer. J. Geophys. Res. Atmos. 124, 3984–4003. https://doi.org/ source regions on the composition of the Asian monsoon anticyclone and of the
10.1029/2018JD029849. extratropical lowermost stratosphere. Atmos. Chem. Phys. 15, 13699–13716.
Son, Seok-Woo, Polvani, L.M., 2007. Dynamical formation of an extra-tropical https://doi.org/10.5194/acp-15-13699-2015.
tropopause inversion layer in a relatively simple general circulation model. Geophys. Vogel, B., Müller, R., Günther, G., Spang, R., Hanumanthu, S., Li, D., Riese, M., Stiller, G.,
Res. Lett. 34. 2019. Lagrangian simulations of the transport of young air masses to the top of the
Sunilkumar, S.V., Muhsin, M., Emmanuel, M., Ramkumar, G., Rajeev, K., Sijikumar, S., Asian monsoon anticyclone and into the tropical pipe. Atmos. Chem. Phys. 19,
2016. Balloon-borne cryogenic frost-point hygrometer observations of water vapour 6007–6034. https://doi.org/10.5194/acp-2018-724.
in the tropical upper troposphere and lower stratosphere over India: first results. Vömel, H., David, D.E., Smith, K., 2007. Accuracy of tropospheric and stratospheric
J. Atmos. Sol. Terr. Phys. 140, 86–93. https://doi.org/10.1016/j.jastp.2016.02.014. water vapor measurements by the cryogenic frost point hygrometer: instrumental
Sunilkumar, S.V., Muhsin, M., Venkat Ratnam, M., Parameswaran, K., Krishna Murthy, B. details and observations. J. Geophys. Res. Atmos. 112, 1–14. https://doi.org/
V., Emmanuel, M., 2017. Boundaries of tropical tropopause layer (TTL): a new 10.1029/2006JD007224.
perspective based on thermal and stability profiles. J. Geophys. Res. 122, 741–754. Winker, D.M., Vaughan, M.A., Omar, A., Hu, Y., Powell, K.A., Liu, Z., Hunt, W.H.,
https://doi.org/10.1002/2016JD025217. Young, S.A., 2009. Overview of the CALIPSO mission and CALIOP data processing
Thomason, L.W., Vernier, J.P., 2013. Improved SAGE II cloud/aerosol categorization and algorithms. J. Atmos. Ocean. Technol. 26, 2310–2323. https://doi.org/10.1175/
observations of the Asian tropopause aerosol layer: 1989-2005. Atmos. Chem. Phys. 2009JTECHA1281.1.
13, 4605–4616. https://doi.org/10.5194/acp-13-4605-2013. World Meteorological Organization, 1957. Definition of the tropopause. World Meteorol.
Vernier, J.P., Pommereau, J.P., Garnier, A., Pelon, J., Larsen, N., Nielsen, J., Organ. Bull. 6 (136).
Christensen, T., Cairo, F., Thomason, L.W., Leblanc, T., McDermid, I.S., 2009. Yu, P., Rosenlof, K.H., Liu, S., Telg, H., Thornberry, T.D., Rollins, A.W., Portmann, R.W.,
Tropical Stratospheric aerosol layer from CALIPSO Lidar observations. J. Geophys. Bai, Z., Ray, E.A., Duan, Y., Pan, L.L., Toon, O.B., Bian, J., Gao, R.S., 2017. Efficient
Res. Atmos. 114, 1–12. https://doi.org/10.1029/2009JD011946. transport of tropospheric aerosol into the stratosphere via the Asian summer
Vernier, J.P., Thomason, L.W., Kar, J., 2011. CALIPSO detection of an Asian tropopause monsoon anticyclone. Proc. Natl. Acad. Sci. U.S.A. 114, 6972–6977. https://doi.org/
aerosol layer. Geophys. Res. Lett. 38, 1–6. https://doi.org/10.1029/2010GL046614. 10.1073/pnas.1701170114.
Vernier, J.P., Fairlie, T.D., Natarajan, M., Wienhold, F.G., Bian, J., Martinsson, B.G., Zhang, Q., Wu, G., Qian, Y., 2002. The bimodality of the 100 hPa South Asia High and its
Crumeyrolle, S., Thomason, L.W., Bedka, K.M., 2015. Increase in upper tropospheric relationship to the climate anomaly over East Asia in summer. J. Meteorol. Soc.
and lower stratospheric aerosol levels and its potential connection with Asian Japan 80, 733–744. https://doi.org/10.2151/jmsj.80.733.
pollution. J. Geophys. Res. 120, 1608–1619. https://doi.org/10.1002/
2014JD022372.

11

You might also like