Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Atmospheric Research 301 (2024) 107293

Contents lists available at ScienceDirect

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

Three-dimensional structure and transport flux of springtime smoke


aerosols over the Indochina Peninsula
Yurun Liu a, b, Ke Gui b, *, Quanliang Chen a, Liangliang Feng c, Hongke Cai a, Xutao Zhang b,
Wenrui Yao b, Hengheng Zhao b, Nanxuan Shang b, Lei Li b, Yu Zheng b, Huizheng Che b, *
a
Plateau Atmospheric and Environment Key Laboratory of Sichuan Province, College of Atmosphere Sciences, Chengdu University of Information Technology, Chengdu
610225, China
b
State Key Laboratory of Severe Weather & Key Laboratory of Atmospheric Chemistry of CMA, Chinese Academy of Meteorological Sciences, Beijing 100081, China
c
Kunming Meteorological Bureau of Yunnan, Kunming 650501, China

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

Keywords: Springtime smoke aerosols attributable to biomass burning on the Indochina Peninsula (ICP) are a major source
CALIOP of absorbing aerosols globally, and play a nonnegligible perturbing role in climate change. Using smoke
Smoke aerosols extinction profiles provided by the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP) instrument, and
Three-dimensional structure
aerosol and meteorological reanalyses for the period 2007–2021, this study investigated the three-dimensional
Transport flux
Indochina Peninsula
structure and transport flux of springtime smoke aerosols over the ICP. Stratified analysis revealed that
CALIOP-derived peaks of smoke aerosol optical depth (SAOD) and smoke mass flux (SMF) at altitudes of 0–2 km
occur over the central ICP, which is the main source region. Driven by a southwesterly jet stream and updrafts,
massive amounts of smoke aerosols are transported toward the northeast of the ICP, resulting in the largest
stratified SAOD and SMF observed at altitudes of 2–4 km. Over the main smoke source regions, the estimated
springtime SMF was approximately 0.4 Tg accumulated through a distance of 2◦ latitude at altitudes of 2–4 km.
Meteorological analyses indicated a path for the transport of smoke aerosols to the mid-troposphere. Smoke
aerosols in the source region are first lifted thermally to the 600-hPa level. Lifting then continues to the 500-hPa
level via the combined effects of topography and frequent, deep, and dry convective updrafts over the moun­
tainous areas of the northern ICP. The smoke aerosols are subsequently transported to South China and the
western Pacific owing to strong westerly winds and weak vertical motion. Further analyses demonstrated that
differences in meteorological conditions between daytime and nighttime have substantial impact on the vertical
structure of smoke aerosols, particularly over the northeast of ICP. Compared to daytime, enhanced updrafts and
moisture transport at night favor further lifting of smoke aerosols and greater extinction intensity owing to
increased hygroscopic growth.

1. Introduction major impact on atmospheric chemistry (Donahue et al., 2009; Kalog­


ridis et al., 2018). Moreover, smoke aerosols can change the charac­
The primary sources of smoke aerosols categorized as carbonaceous teristics of clouds (Penner et al., 1992; Ackerman et al., 2000), and
are combustion of fossil fuels and biomass burning. The practice of absorb and scatter solar radiation (Hansen et al., 1997), which could
biomass burning dates back 2 million years, whereas fossil fuel com­ change the climate. Reid et al. (2005) reported that smoke aerosols are
bustion has played a notable role in smoke aerosol emissions only over primarily composed of two types of carbon: black carbon (BC) and
the past two centuries (Andreae and Merlet, 2001). Initially, research on organic carbon (OC). Ramanathan et al. (2005) demonstrated that BC
emissions of pollutants focused solely on those associated with fossil fuel and OC can substantially affect the intensity of the Indian monsoon by
combustion, and it was not until the 1970s that the first groundbreaking enhancing atmospheric stability. According to reports, enhanced crea­
studies on the effects of burning biomass on the atmosphere were pub­ tion of low clouds over Southeast Asia can be facilitated by the syner­
lished (e.g., Wong, 1978; Crutzen et al., 1979). Smoke aerosols can have gistic effects of aerosol, cloud, and boundary layer interactions with the

* Corresponding authors.
E-mail addresses: guik@cma.gov.cn (K. Gui), chehz@cma.gov.cn (H. Che).

https://doi.org/10.1016/j.atmosres.2024.107293
Received 14 November 2023; Received in revised form 25 January 2024; Accepted 10 February 2024
Available online 12 February 2024
0169-8095/© 2024 Elsevier B.V. All rights reserved.
Y. Liu et al. Atmospheric Research 301 (2024) 107293

monsoon (Ding et al., 2021). Additionally, because smoke aerosols are 18–25◦ N, 108–122◦ E) and Western Pacific (WP; 17–28◦ N, 122–150◦ E).
poisonous, they could be harmful to human health (Janssen et al., 2011; In addition, to study the regional diurnal differences of smoke aerosols,
Ho et al., 2018). we defined the approximate concentrated region “northeast of ICP”
The Indochina Peninsula (ICP) is a region characterized by frequent (19–22◦ N, 95–108◦ E) with the diurnal difference of SAOD above 0.1 in
fires and it is one of the global aerosol hotspots, similar to South springtime, according to the distribution of diurnal differences of SAOD
America, Africa, and Australia (van der Werf et al., 2017). Smoke from MERRA-2 (Fig. S2).
aerosol emissions from biomass burning on the ICP might be linked to
crop combustion and forest wildfires; crop combustion is most common
in the southern part of the ICP, whereas the northern part has complex 2.2. CALIOP observations
topography and high incidence of forest wildfires (Streets et al., 2003).
The aerosols in this area exhibit marked characteristics of both temporal Through acquisition of vertical observations, NASA’s Cloud-Aerosol
and spatial variation (Sahu and Sheel, 2014; Shi et al., 2014). The Lidar and Infrared Pathfinder Satellite Observations (CALIPSO) project
intertropical convergence zone and regional monsoon systems coordi­ offers fresh understanding of the function of aerosols and clouds in
nate to control the climate of the ICP, which has distinct seasons of rain relation to Earth’s climate and weather. The primary device on the
and drought (Dong and Fu, 2015). Consequently, the seasonal variation CALIPSO satellite is the CALIOP two-wavelength polarization-sensitive
in smoke aerosol emission over the ICP is considerable, with springtime lidar. The CALIOP instrument has provided the first global multiyear
emissions being most prominent. Smoke aerosols over the ICP are often dataset of aerosol and cloud profiles during both daytime and nighttime.
transported by the winds of the southwest monsoon and can cause Along the CALIOP observing track, it also offers vertically resolved
deterioration in the ambient environmental conditions over South China attenuated backscatter data collected from the Earth’s surface to the
(SC) (Huang et al., 2013). Distribution of smoke aerosols in vertical elevation at 40 km (Winker et al., 2009). The CALIPSO satellite uses the
direction is crucial considering it strongly affects atmospheric chemis­ cloud–aerosol discrimination technique to classify clouds and aerosols
try, their own transport, and climate impact (Kahn et al., 2008; Guan (Liu et al., 2009). In the recently released CALIPSO version 4 products,
et al., 2010; Bourgeois et al., 2018). The position of aerosols in the at­ seven categories of aerosols are defined: elevated smoke, dusty marine,
mospheric column largely affects their residence time during long- clean continental, polluted dust, dust, marine, and polluted continental/
distance transport (Bourgeois et al., 2015). To better comprehend the smoke aerosols (Kim et al., 2018).
long-distance transport of aerosols, it is crucial to examine their vertical This study used the CALIPSO Level 2 15-year aerosol extinction
structure characteristics. A few studies have revealed that during the profile dataset (05kmAPro, V4.10), which includes daytime and night­
Indian monsoon season, the elevated aerosols in the lower troposphere time “all-sky” circumstances spanning 2007–2021. The nominal altitude
are mainly formed through the long-range transport of the low level jet range of this dataset is between − 0.5 and 30.1 km (399 levels) above
and maintained by vertical shear in the horizontal winds (Ratnam et al., mean sea level (all subsequent heights are likewise higher than mean sea
2018; Prasad et al., 2019). VanCuren et al. (2005) reported that aerosols level) (Gui et al., 2021). The unified orbit-following spatial resolution of
in lower free troposphere of Asia can persist and be transported to the this dataset is 60 m vertically and 5 km horizontally. Because back­
northeast Pacific and western North America. Although smoke aerosol ground solar illumination in daytime reduces the signal-to-noise ratio,
transport above the ICP is regularly observed or simulated, quantitative CALIOP detections have higher sensitivity at night than during daytime.
studies of smoke aerosol transport remain limited. Therefore, it is This leads to the possibility that poorly scattered aerosol layers observed
necessary to assess the mass flux of smoke aerosols. at night might not be detected during daytime, thereby amplifying the
In this study, we comprehensively elucidated the long-distance uncertainty in daytime observation. We used the atmospheric volume
transport properties and the three-dimensional structure of smoke description dataset that includes a characteristic categorization sign for
aerosols from biomass combustion on the ICP. We used Cloud-Aerosol determining the aerosol category in each profile partition and in each
Lidar with Orthogonal Polarization (CALIOP) observations, obtained aerosol layer. Here, for simplicity, we used “smoke” to represent
during springtime of each year during 2007–2021, for the purpose of “elevated smoke.” In the CALIOP V4.2 classification algorithm, the
analyzing the spatial distribution characteristics of smoke aerosol opti­ smoke plumes raised from biomass burning, which are essentially
cal depth (SAOD) integrated at different heights, and to examine vertical composed of carbonaceous aerosols, are represented by the “smoke”
structure of smoke aerosol extinction coefficient (SAEC) over the ICP. category. We applied quality-screening methods to every all-sky curve to
Moreover, we used Modern-Era Retrospective analysis for Research and reduce the uncertainties related to the aerosol extinction curve, refer­
Applications version 2 (MERRA-2) reanalysis data to further investigate encing various individual studies (Tian et al., 2017; Liao et al., 2021)
the aerosols and we compared the results obtained from MERRA-2 with and the official Level-3 aerosol profile dataset quality-screening
those obtained from CALIOP. We further analyzed the meteorological methods (Tackett et al., 2018). In general, the fundamental input data
conditions that affect aerosol distribution using the fifth-generation required for quality-screening procedures are CALIOP level 2 aerosol
European Centre for Medium-Range Weather Forecasts atmospheric extinction profiles and layer classification information (aerosol, cloud,
reanalysis (ERA5) data. Additionally, we estimated the smoke mass flux and clear air). Here, the quality-screening procedures used is briefly
(SMF) and explored the transport characteristics of smoke aerosols. summarized as follows: (1) extinction quality control sign at 532 nm =
0, 1, 2, 16, or 18; (2) − 100 ≤ cloud aerosol identification fraction ≤
2. Data and methods − 20; (3) uncertainty of extinction coefficient at 532 nm ≤ 10 km− 1 ; (4)
0 ≤ extinction coefficient at 532 nm ≤ 1.25 km− 1 ; (5) aerosol extinction
2.1. Study area in “clear air” = 0 km− 1 ; (6) excluding extinction under 60 m elevation
(above ground surface). For detailed information on CALIOP Level 2
As shown in Fig. S1, the northern part of the ICP is a mountainous aerosol data products, including storage structure, variables, and quality
area with elevations of about 1000–3000 m. The main mountain ranges control flags, please refer to Kim et al. (2018). Following quality control,
are Arakan Yoma in the northwest and a series of mountains extending the aerosol extinction for quality screening were averaged over a 2◦
southward from the Shan Plateau in the north. Smoke aerosols emissions latitude × 2◦ longitude grid box. When calculating the monthly aver­
from the ICP may affect air quality over downstream areas from South ages, to obtain more reliable statistics, only 10 or more extinction values
China to the Pacific (Tsay et al., 2013; Dong and Fu, 2015; Zhang et al., for each stratified grid box were averaged for the aerosol extinction
2022). In order to study the three-dimensional structure and the trans­ values. Given that CALIPSO has constant equatorial crossing times,
port flux of smoke aerosol, the regions selected in this paper include the because it is a polar sun-synchronous satellite, we considered 13:30 and
Indochina Peninsula (ICP; 8–24◦ N, 92–108◦ E), South China (SC; 01:30 (local time) as representative of daytime and nighttime,

2
Y. Liu et al. Atmospheric Research 301 (2024) 107293

respectively, when exploring diurnal differences. for the MERRA-2 dataset.


We calculated the SAOD at different altitudes based on the extinction
of smoke aerosols at 532 nm, as shown below. 3. Results
∫ h2
SAOD = SAEC(h)dh (1) 3.1. Comparison of SAOD from CALIOP and MERRA-2
h1

We first contrast the 15-year seasonal average SAOD from CALIOP


where h1 and h2 represent the bottom and top elevations of SAOD, and
and MERRA-2 for the period 2007–2021. Fig. 1 depicts the spatial dis­
SAEC(h) is the smoke aerosol extinction coefficient at 532 nm (km− 1 ) at tribution of SAOD over the ICP, SC, and WP regions obtained from
height h. By dividing the distance from the ground to the altitude of 6 km CALIOP and MERRA-2 during spring (March–April–May; MAM), sum­
into three height segments, each comprising 2 km, we calculated SAOD mer (June–July–August; JJA), autumn (September–October–November;
at various heights. SON), and winter (December–January–February; DJF). Both CALIOP
The SMF were estimated using the following method. First, we and MERRA-2 observed that the ICP region has the highest smoke
assumed the mass extinction efficiency (MEE) of smoke to be 3.5 m2 g− 1 , loading during MAM, followed by that in DJF; the lowest smoke loading
thus converting the SAEC to smoke mass concentration M (g m− 3 ) is in JJA and SON. The smoke loading in MAM is considerably greater
(Hirsch and Koren, 2021; Wu et al., 2022). than that in the other three seasons. There is distinct seasonal pattern to
We calculated the zonal (east–west) SMF rate (FRu ; g s− 1 ) and the the regional biomass burning. Burning activities start in December of
meridional (north–south) SMF rate (FRv ; g s− 1 ) in each 2◦ × 2◦ grid cell, DJF and peak in March and April of MAM, which are associated with
as shown below: local crop burning, forest fires, and dry climate. In JJA, the burning
∫ h2 activities are greatly reduced as the Asian summer monsoon brings
FRu = M(h)U(h)L1 dh (2) sufficient rainfall to the ICP (Yin et al., 2019). The high smoke loading
h1
during MAM over the ICP can be attributed to the massive amount of
∫ h2 smoke aerosols emitted by large-scale biomass burning in this season
FRv = M(h)V(h)L2 dh (3) (Tsay et al., 2013). During DJF, SAOD is mainly distributed in the
h1
southern ICP. However, during MAM, SAOD is mainly distributed in the
northeastern ICP and the SC region, and the intensity of SAOD is the
where h1 and h2 represent the bottom and top elevations of FRu and FRv ,
largest. Given that CALIOP detection sensitivity is higher at night than
and the mass concentration at height h is represented by M(h). The 2◦
during daytime, we discuss smoke aerosols observed during daytime and
latitude segment along the longitude line is L1, and the 2◦ longitude
nighttime separately, but we consider only the differences in their dis­
segment along the latitude line is L2. Using the ERA5 data, U and V
tribution. As for the differences in magnitude, it is justified from the
represent the zonal wind (m s− 1 ) and the meridional wind (m s− 1 ),
MERRA-2 data. It should be noted that smoke loading during nighttime
respectively.
(Fig. 1e-2) from MERRA-2 is markedly higher than that during daytime
At 2-km intervals from the ground to the altitude of 6 km, three
(Fig. 1e-1).
layers of SMF were determined. The computation of monthly SMF
involved multiplying the average flux rate per month by the duration for
3.2. SAOD at different altitudes
that month. Because we discuss the diurnal and nocturnal durations of
SMF separately, the duration of SMF during daytime or nighttime was
The 15-year averages of SAOD during daytime and nighttime at
estimated to be 12 h. The SMF in springtime was determined through
various altitudes during the main period of smoke aerosol activity (i.e.,
summation of the SMF in springtime in each year. The multiyear average
MAM), as measured by the CALIOP instrument between 2007 and 2021,
SMF during springtime was determined by averaging the SMF during
are displayed in Fig. 2. The SAOD at an altitude of >6 km accounts for
springtime for the years 2007–2021. Negative flux was associated with
only 2.4% of the total gas column layer at night and only 1.2% during
transport toward the west and south, whereas positive flux was associ­
daytime. Therefore, as samples typical of the variation in smoke aerosols
ated with transport toward the east and north. The uncertainty in the
at various altitudes, we selected the SAOD at altitudes of 0–2, 2–4, and
estimations of SMF was associated with CALIOP extinction, smoke MEE,
4–6 km. The ICP is the main area of regional biomass burning, and the
vertical profile shape, and potential variation in the distribution of the
SAOD can be transported from the ICP downwind to SC and the WP,
size of smoke aerosols while undergoing transport.
forming the transport pathway of smoke aerosols along the middle
lower-latitude region of East Asia (Tsay et al., 2013; Dong and Fu, 2015;
2.3. MERRA-2 dataset Zhang et al., 2022). However, there are notable differences in SAOD
among the three selected altitude layers.
This study used the aerosol mixing ratio containing hydrophilic and In the Northern Hemisphere, at middle and low latitudes, SAOD
hydrophobic aerosols (product: inst3_3d_aer_Nv) and the aerosol optical enhancement mainly occurs at altitudes of <4 km near the origin of the
depth (product: tavg1_2d_aer_Nx) from daily datasets in springtime smoke aerosols. Within altitudes of 0–2 km, the SAOD over the areas
during 2007–2021 to demonstrate the three-dimensional distribution of (12◦ –18◦ N, 95◦ –108◦ E) can exceed 0.1, which is considered to be the
smoke aerosols (OC and BC), with 0.625◦ × 0.5◦ spatial resolution. For main source regions of smoke aerosols. In addition, the SAOD that ex­
maintaining consistency with CALIOP datasets on exploring diurnal ceeds 0.15 over the northwest and southeast of the source regions is
variations, we used 06:00 (UTC) and 18:00 (UTC) to represent daytime roughly consistent with the hot spots of biomass burning during
and nighttime, respectively, over the ICP, after time zone conversion. springtime as studied by Huang et al. (2016). Compared to the lower
Then, all daily data were processed into seasonally averaged data. layer, the SAOD in altitude range of 2–4 km increases, with the peak
value reaching 0.22. The increase in SAOD over SC is substantial. The
2.4. ERA5 dataset SAOD initially increases and then decreases with increasing altitude,
and the SAOD in range of 4–6 km is markedly reduced.
The ERA5 dataset applies 4D variational assimilation with a It is noteworthy that SAOD enhancement at altitudes of 0–2 km is
geographic resolution of 0.25◦ × 0.25◦ . Here, we used hourly datasets of mainly distributed over a large area in the middle of the ICP, whereas
the ERA5 uvw, divergence, eastward water vapor, northward water SAOD enhancement at altitudes of 2–4 km is shifted to the northeastern
vapor, and divergence of moisture in the springtime during 2007–2021, ICP and southern parts of SC. It indicates that some transport mecha­
with the daytime and nighttime selection consistent with that adopted nisms might exist between these areas. At the 0–2 km over the ICP,

3
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Fig. 1. Seasonal climatology of SAOD from CALIOP (a-1 to d-1) and MERRA-2 (e-1 to h-1) during daytime and from CALIOP (a-2 to d-2) and MERRA-2 (e-2 to h-2) at
nighttime during 2007–2021. Red box (19◦ –22◦ N, 95◦ –108◦ E) represents the northeast of ICP. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

under the influence from the southwesterly jet (Duncan et al., 2003; Ooi
et al., 2021). In the 16◦ –30◦ N region at altitudes of >2 km, the transport
pathway of smoke aerosol from ICP to SC and WP can be observed.

3.3. Vertical structure of SAEC

Fig. 3 displays the vertical structure of daytime and nighttime 15-


year averages of the SAEC along certain representative pressur­
e–longitude cross sections in springtime that characterize the spatial
transport of smoke aerosols. High values of the SAEC are evident over
the source areas (~15◦ –20◦ N), exceeding 0.1 km− 1 . The maximum
height of the smoke aerosol layer (MHSAL) (SAEC >0.01 km− 1 ) is
approximately at the 700-hPa level in the 10◦ N profile. Peak values of
smoke aerosols are mainly distributed in the 95◦ –108◦ E region along the
15◦ N and 20◦ N profiles. During transport from the ICP to the WP, the
smoke plumes continuously diminish, and the SAEC increases and then
decreases with altitude. The pressure–longitude profiles of 20◦ N and
25◦ N pass over the mountains in the northern ICP. Compared to that in
the 10◦ N and 15◦ N profiles, the MHSAL (SAEC >0.01 km− 1 ) increases in
the 20◦ N and 25◦ N profiles. The strong smoke aerosol peak in the 20◦ N
Fig. 2. The 15-year averaged daytime and nighttime distribution of SAOD at
profile is expanded eastward to approximately 116◦ E, and the SAEC
altitudes of 0–2 km (a, b), 2–4 km (c, d), and 4–6 km (e, f) in MAM 2007–2021. value at approximately 130◦ E is increased. The SAEC value at approxi­
mately 140◦ E is increased in the 25◦ N profile, which means that the
intensity of the eastward transport of smoke increases at higher lati­
SAOD enhancement is blocked near 18◦ N, possibly owing to the influ­
tudes. This might reflect that the low-latitude equatorial easterly wind is
ence of mountainous areas with elevation of 1000–3000 m in the
not conducive to eastward transport of smoke, whereas the relatively
northern ICP, resulting in reduced transport by the southerly wind.
enhanced low–mid-latitude westerly wind can result in smoke accu­
However, At the 2–4 km altitude, the influence of mountain topography
mulation and eastward downwind transport.
is greatly reduced. Consequently, SAOD can continue to be transported
It is important to note that the mountains in the northern ICP appear
northward. Additionally, the increased westerly wind at this altitude
to affect the MHSAL (SAEC >0.01 km− 1 ) in the transport pathway. As
provides support for eastward transport of SAOD. In the 2–4-km altitude
the mountainous topography extends from low-latitude areas of the ICP
layer, SAOD might be transported across most of SC and on to the WP
to higher latitudes, the MHSAL increases. In the 20◦ N and 25◦ N profiles,

4
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Fig. 3. The 15-year averaged daytime (a-1, b-1, c-1, d-1) and nighttime (a-2, b-2, c-2, d-2) SAEC of pressure–longitude cross sections at 10◦ N, 15◦ N, 20◦ N, and 25◦ N
in MAM 2007–2021. Corresponding location maps are displayed at the top of each plot. Black shading represents terrain.

the height of the smoke layer reaches a peak value over the northern aerosols from the ICP might be transported to SC and the WP by the
mountains. This might be attributable to frequent convective updrafts westerly wind.
forcing the smoke over the mountains and the smoke subsequently Normally, the MHSAL (SAEC >0.01 km− 1 ) can reach the 600-hPa
maintaining an upward trajectory. The smoke aerosol flow gradually level over the source areas. However, the MHSAL (SAEC >0.01 km− 1 )
diminishes over the WP, as the transport distance increases. Smoke in the mountainous region of the northern ICP might exceed 600-hPa

Fig. 4. The 15-year averaged daytime (a-1, b-1, c-1, d-1) and nighttime (a-2, b-2, c-2, d-2) vertical speed (10− 2 Pa s− 1 ; shading) and zonal circulation (arrows) at
10◦ N, 15◦ N, 20◦ N, and 25◦ N in MAM 2007–2021. Corresponding location maps are displayed at the top of each plot. Black, blue, and red shading represents terrain,
downward motion, and upward motion, respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

5
Y. Liu et al. Atmospheric Research 301 (2024) 107293

level. This raises a question regarding the mechanism that results in the topography and favorable wind patterns, the MHSAL (SAEC
smoke layer in the mountainous area of the northern ICP being higher >0.01 km− 1 ) of the nighttime is slightly higher compared to that in
than that in the source regions. daytime. The CALIOP-derived distribution reflects that the maximum
value of the SAEC is limited to a vertical range, with a lower limit and
3.4. Impact of circulation on the three-dimensional structure of the SAEC upper limit of 850 and 700 hPa respectively; conversely, the MERRA-2
distribution shows that the SAEC is very high below 700 hPa.
Smoke aerosol transport is influenced substantially by atmospheric It must be noted that even though the CALIOP data cannot account
circulations. Fig. 4 shows the 15-year averages of the spring zonal cir­ for the differences in magnitude between daytime and nighttime owing
culation during daytime and nighttime. We selected pressure–longitude to detector sensitivity, the maximum loading value of the smoke aerosol
profiles along 10◦ N, 15◦ N, 20◦ N, and 25◦ N to discuss the vertical at­ mixing ratio at night observed in the MERRA-2 data is higher than that
mospheric circulation above ICP. Compared to the 10◦ N and 15◦ N in daytime. It indicates that there might be differences in the meteoro­
profiles, the 20◦ N profile shows a stronger updraft in the region of logical fields between the daytime and nighttime periods in the north­
95◦ –108◦ E (northeast of ICP). Below 600-hPa level, the vertical velocity east of ICP that account for the greater maximum height (SAEC >0.01
averaged during springtime over the mountainous areas of the northern km− 1 ) and higher maximum loading value of smoke aerosol layer during
ICP can vary approximately from − 8 to − 40 10− 2 Pa s− 1 . In the 25◦ N nighttime compared to those in daytime.
profile, there is a sinking air current at approximately 100◦ E, whereas
strong updrafts to 400 hPa exist over the peripheral slopes. Wu et al. 3.6. Influence of meteorological conditions on differences between
(2012) speculated that this might be the result of slope heating. Previ­ daytime and nighttime
ously, Lau et al. (2006, 2008) proposed the elevated heat pump hy­
pothesis, whereby smoke aerosols deposited on mountain slopes might Fig. 6 shows the 15-year mean values of the meridional and zonal
absorb more solar radiation, thereby enhancing slope heating. Frequent circulation, water vapor flux and its divergence during MAM in the
updrafts over mountainous areas might facilitate smoke aerosols trans­ northeast of ICP. Vertical updrafts are evident during both daytime and
port into the mid–upper troposphere. nighttime in the northeast of ICP. However, it should be noted that in the
16◦ –23◦ N, the strength of the upward flow at night is greater than that in
3.5. Vertical structure of the SAEC in the northeast of ICP daytime, and that there is a vigorous updraft zone below 600 hPa at
night (Fig. 6a–c) that may be related to the mountainous terrain with
As previously observed, the values of smoke aerosols differ between high drop elevations in the north of ICP. At the high elevations of
daytime and nighttime within the range of 19◦ –22◦ N and 95◦ –108◦ E, i. 26◦ –30◦ N, updrafts are stronger during daytime than at nighttime, while
e., the northeast of ICP. To explore the specific underlying reasons, we at lower elevations of 16◦ –23◦ N, updrafts are stronger at nighttime
further analyzed this area. In this section, we examine profiles of the (Fig. 6c). This may be because during daytime, the high elevations heat
SAEC derived from CALIOP and smoke mixing ratios derived from up faster, which is conducive to the development of updrafts, while the
MERRA-2 to explain differences in the vertical structure of smoke over low elevations produce compensatory subsidence; at nighttime, the high
the northeast of ICP between daytime and nighttime. elevations radiate cold faster, which is conducive to the sinking move­
Fig. 5 illustrates the averaged daytime and nighttime distributions of ment, while the low elevations produce compensatory rise (Rampanelli
the SAEC and the smoke mixing ratio in the northeast of ICP, in both et al., 2004; Pan and Chen, 2019). To illustrate the causes of regional
meridional and zonal directions. Overall, owing to the nearby differences in vertical motion, Fig. S3 shows the nighttime-daytime

Fig. 5. The 15-year averaged meridional (average between 95◦ –108◦ E) (a-1, b-1) and zonal (average between 19◦ –22◦ N) (a-2, b-2) profiles of the SAEC from
CALIOP, and the meridional (c-1, d-1) and zonal (c-2, d-2) profiles of the smoke aerosol mixing ratio from MERRA-2 over the northeast of ICP (central area of the red
lines) in MAM 2007–2021. Black shading represents terrain. (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)

6
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Fig. 6. The 15-year averaged mean and difference (nighttime value minus the daytime value) of meridional circulation (average between 95◦ –108◦ E; arrows) and
zonal circulation (average between 19◦ –22◦ N; arrows) and vertical speed (10− 2 Pa s− 1 ; shading) (a, b, c, d, e, f), together with the water vapor flux (arrows) and
water vapor dispersion (10− 5 kg m− 2 s− 1 ; shading) (g, h, i) between daytime and nighttime in MAM 2007–2021. Black shading represents terrain.

differences in horizontal wind and divergence. In contrast to the day­ 3.7. SMF at different altitudes
time, there are significant convergences near the black terrain of the
925 hPa and 850 hPa layers during nighttime, which may lead to Atmospheric aerosols over East Asia can also have far-reaching
stronger upward movement that can transport more smoke from the environmental impacts on continents, hemispheres, and even the
source vertically upward to approximately the 600-hPa level. entire globe owing to their long-distance transport. To better compre­
Conversely, the weaker vertical updraft in daytime might be able to hend the origins and paths associated with pollution and the transport of
transport only a small amount of smoke from the source to the 600-hPa smoke, we conducted further research on SMF. The SMF represents a
level. Therefore, the smoke loading value and the peak smoke aerosol broad estimate of the specific tonnages of smoke transport, which might
layer height (SAEC >0.01 km− 1 ) in daytime are both lower than those at also be important for model simulation.
night. Additionally, the southerly wind at night is weaker than that in Fig. 7 illustrates the 15-year averages of the zonal (east-west) SMF
daytime between 700 and 500 hPa (Fig. 6b–c). accumulated through the distance of 2◦ latitude at altitudes of 0–2, 2–4,
In the northeast of ICP, water vapor is largely dispersed during and 4–6 km during MAM. Notable changes are evident in SMF from the
daytime (Fig. 6g), whereas it tends to be condensed at night (Fig. 6h). lower to upper atmosphere, consistent with the SAOD (Fig. 2). In the
The intensity of the southwest water vapor flux transport at night is main smoke source areas, SMF initially increases and then decreases
stronger than that during daytime (Fig. 6i). As shown in Fig. S4, the with increasing altitude. The strong SMF region is observed at altitudes
difference between nighttime and daytime of hydrophilic smoke aerosol of 2–4 km. Smoke aerosols transported from their areas of origin on the
is obviously higher than that of hydrophobic smoke aerosol. This might ICP toward SC and the WP might be accelerated by the establishment of
be because water vapor affects the hygroscopic growth of aerosols, lower subtropical westerlies. A smoke band at approximately 17◦ –25◦ N
thereby influencing their optical characteristics (Zieger et al., 2013). begins to appear at altitudes of between 2 and 4 km, and the maximum
The favorable water vapor environment during nighttime is instru­ SMF value over the northeast region of the ICP is 0.4 Tg. However, at
mental in influencing the magnitude of aerosol extinction coefficient altitudes of 4–6 km, the maximum SMF value decreases to 0.08 Tg. The
(Jung et al., 2009). The growth of aerosol optical depth is facilitated by smoke band not only expands northward to 30◦ N but also transports the
increase in the total column water vapor content (Smirnov et al., 2002). SMF eastward to 150◦ E. Greater SMF can also be observed near 120◦ E

7
Y. Liu et al. Atmospheric Research 301 (2024) 107293

averaged SMF. The second uncertainty is that the factors contributing to


diurnal and nocturnal differences in smoke in the ICP hotspots are not
limited to meteorological conditions, but may also be influenced by
ultraviolet and visible light from solar radiation, which can alter the
optical properties of aerosols by affecting their ageing time. Strong
photochemical reactions during the daytime cause aerosols to age more
rapidly than at nighttime (Chen et al., 2017). Furthermore, the vertical
resolution of the wind field cannot match that of CALIOP, which has 60-
m resolution. And, given the inability to separate the smoke class from
“polluted continental aerosol”, we only consider the “elevated smoke” in
CALIPSO as “smoke”, so the loading of smoke aerosols may be under­
estimated. We suggest that future work should utilize multisource data
fusion to form a more accurate aerosol extinction dataset, which would
reduce the uncertainty in the SMF. Additionally, we plan to undertake
bias correction for CALIOP extinction to rectify the aerosol extinction
and further mitigate the impact caused by the uncertainty of CALIOP’s
own detection. Furthermore, more refined high-resolution vertical wind
fields data (e.g., Aeolus data) might be required to support CALIOP’s
observations.

5. Conclusions

This study used smoke extinction profiles from the CALIOP instru­
ment, and aerosol and meteorological reanalyses from MERRA-2 and
Fig. 7. The 15-year averaged SMF (Tg; 1 Tg = 1012 g) accumulated through a ERA5 for the period 2007–2021, for the purpose of characterizing the
distance of 2◦ latitude at altitudes of 0–2, 2–4, and 4–6 km dominated by U in three-dimensional structure of springtime smoke aerosols over the ICP,
MAM 2007–2021. Positive (negative) SMF indicates eastward (west­
and estimating the stratified transport fluxes over the ICP, SC, and WP.
ward) transport.
Stratified analysis revealed that the spatial distribution of smoke aero­
sols has obvious differences; specifically, the position of the northeast of
and 130◦ E relative to that of the main source areas of smoke. It is worth ICP is vastly different between the altitudes of 0–2, 2–4, and 4–6 km. At
noting that there is distinct westward transport of SMF between the east altitudes of 0–2 km, the peak values of SAOD and SMF appeared in the
of the ICP and the region of the Philippines at altitudes of 0–2 km. This central region of the ICP, which might reflect that fact that this region is
might be attributable to the fact that the ICP is at the intersection of low- the primary source of smoke, whereas at 2–4 km the peaks observed
layer easterly and westerly winds. The smoke generated in the region of appeared over the northeastern ICP and SC. Additionally, along zonal
the Philippines is transported westward by the equatorial easterlies to transects, the highest smoke layer was typically seen in the mountainous
SC and eastern parts of the ICP, causing thick aerosol clouds to build on a area of the northern ICP. Within the height range of 0–2 km over the ICP
vast scale across the South China Sea (Zhang et al., 2012). Owing to the region and 2–4 km over SC, the SAOD magnitude reached 0.1, whereas it
effects of terrain, the vertically integrated SAEC over the ICP will be was only 0.01 at altitudes of 4–6 km. The estimated springtime SMF over
lower than its true value. In fact, the SMF transported westward near the the main smoke source area was approximately 0.4 Tg accumulated
Philippines should be weaker than that transported eastward from the through a distance of 2◦ latitude at 2–4 km, which can be transported to
ICP. Fig. 8 shows the average SMF for 15 years in both east–west and the northeast of the ICP under the action of the southwesterly jet stream.
north–south directions. Owing to terrain blocking, strong SMF at alti­ Although the SMF drops to 0.1 Tg at 4–6 km, it can be transported
tudes of 0–2 km above the ICP appears south of 20◦ N. The transport further to SC and the WP by strong westerly winds. In the region of
pathway of SMF at altitudes of 2–4 km moves northward, especially 20◦ –30◦ N at altitudes of >4 km, with strengthening of the westerlies,
reaches SC. At altitudes of 4–6 km, SMF decreases sharply, which might there is an obvious band of smoke aerosols transported to 150◦ E.
be attributable to both the weaker southerly wind at higher levels Atmospheric circulation patterns influence the transport of smoke
compared to that at lower levels, and the much lower value of high-level aerosols. Upward air currents can be clearly observed in springtime over
SAEC compared to that at mid–lower levels. the northern region of the ICP, with a seasonal mean vertical velocity
ranging from approximately − 8 to − 40 10− 2 Pa s− 1 . Meteorological
4. Discussion analyses revealed a pathway for smoke aerosol transport to the mid-
troposphere. Source area smoke aerosols are first lifted thermally to
We proposed a method to analyze the three-dimensional structure of the 600-hPa level. They are then lifted further to 500 hPa via a combi­
smoke aerosols and to calculate their transport flux using CALIOP ob­ nation of topographic effects and frequent, deep, dry convective up­
servations, and we integrated the MERRA-2 reanalysis dataset for drafts over the mountainous areas of the northern ICP. Further analysis
assisted analysis. However, the method has some inherent uncertainties. of the meteorological conditions showed that, in the northeast of ICP
The first is that background solar illumination in daytime reduces the with large diurnal differences in smoke aerosols, the peak SAEC is
signal-to-noise ratio, which leads to the possibility that poorly scattered located between 850 and 700 hPa, which is related to the local vertical
aerosol layers observed at nighttime might not be detected during circulation. The vertical updraft speed at 850–700 hPa is sufficiently
daytime, thereby amplifying the uncertainty in daytime observations. strong to continuously lift smoke aerosols from the surface of the source
Therefore, the method can be used to compare differences in the spatial region to the 700-hPa level. However, at 600 hPa and above, the vertical
distribution of smoke aerosols between daytime and nighttime, but it speed is substantially reduced, making it difficult for continued uplift of
cannot be used to compare differences in their magnitude. To remedy smoke aerosols. Additionally, differences in meteorological conditions
this deficiency, we considered the MERRA-2 reanalysis datasets to between daytime and nighttime have substantial impact on the vertical
demonstrate the distinctions in the magnitude and regional distribution distribution of SAEC, especially in the northeast of ICP. Compared to the
of smoke aerosols between day and night. However, this uncertainty in conditions in daytime, the weaker southerly wind at night makes it
the SMF calculation leads to imprecise estimates and uncertainties in the easier for smoke aerosols to accumulate in the area to the south of the

8
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Fig. 8. The 15-year averaged SMF (Tg) at altitudes of 0–2, 2–4, and 4–6 km dominated by U and V in MAM 2007–2021.

mountainous region in the northern ICP, and enhanced water vapor analysis. Yu Zheng: Data curation, Formal analysis. Huizheng Che:
concentration at night also increases the hygroscopic growth of aerosols, Conceptualization, Funding acquisition, Supervision.
resulting in a notably higher SAEC value. In contrast to daytime,
enhanced updrafts at night result in higher MHSAL (SAEC >0.01 km− 1 ). Declaration of competing interest
The findings of this study are extremely important both for eluci­
dating the long-distance transport of smoke aerosols over East Asia, and We affirm that the study presented here could not have been affected
for providing an analytical framework of smoke aerosol transport that by any of their financial or personal conflicts.
could help improve global climate models. In future studies, we hope to
compare CALIOP observation data with multiple sets of modeling data, Data availability
combine the advantages of multiple data sets, and analyze the transport
flux of springtime smoke aerosols by integrating a biomass burning in­ CALIPSO Level 2 aerosol profile data are available from
ventory. This integrated study will provide additional insight regarding https://subset.larc.nasa.gov/calipso/index.php. MERRA-2 datasets are
the impact of ICP biomass combustion on environment locally and on available from https://earthdata.nasa.gov/eosdis/daacs/gesdisc. The
other regions attributable to the transport of smoke aerosols. ERA5 data are available from https://cds.climate.copernicus.
eu/cdsapp#!/dataset/.
CRediT authorship contribution statement
Acknowledgments
Yurun Liu: Conceptualization, Data curation, Formal analysis,
Investigation, Methodology, Visualization, Writing – original draft. Ke This research was supported by grants from the National Key
Gui: Conceptualization, Data curation, Formal analysis, Investigation, Research and Development Program of China (2023YFC3706305), Na­
Methodology, Supervision, Writing – review & editing. Quanliang tional Natural Science Foundation of China project (42175153,
Chen: Conceptualization, Supervision. Liangliang Feng: Methodology, 42105138, 42375188, and 42030608), Young Elite Scientists Sponsor­
Software. Hongke Cai: Formal analysis, Supervision. Xutao Zhang: ship Program by BAST, and Basic Research Fund of CAMS (2023Z021).
Data curation, Formal analysis. Wenrui Yao: Data curation, Formal We express our gratitude to the science teams of MERRA-2, ERA5, and
analysis. Hengheng Zhao: Data curation, Formal analysis. Nanxuan CALIPSO for their outstanding and easily available data products.
Shang: Data curation, Formal analysis. Lei Li: Data curation, Formal

9
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Appendix A chemistry and aging of Siberian biomass burning emissions in a large aerosol
chamber. Atmos. Environ. 185, 15–28. https://doi.org/10.1016/j.
atmosenv.2018.04.033.
Supplementary data to this article can be found online at https://doi. Kim, M.-H., Omar, A.H., Tackett, J.L., Vaughan, M.A., Winker, D.M., Trepte, C.R., Hu, Y.,
org/10.1016/j.atmosres.2024.107293. Liu, Z., Poole, L.R., Pitts, M.C., Kar, J., Magill, B.E., 2018. The CALIPSO version 4
automated aerosol classification and lidar ratio selection algorithm. Atmos. Meas.
Tech. 11, 6107–6135. https://doi.org/10.5194/amt-11-6107-2018.
References Lau, K.M., Kim, M.K., Kim, K.M., 2006. Asian summer monsoon anomalies induced by
aerosol direct forcing: the role of the Tibetan Plateau. Clim. Dyn. 26, 855–864.
Ackerman, A.S., Toon, O.B., Stevens, D.E., Heymsfield, A.J., Ramanathan, V., Welton, E. https://doi.org/10.1007/s00382-006-0114-z.
J., 2000. Reduction of tropical cloudiness by soot. Science. 288, 1042–1047. https:// Lau, K.M., Ramanathan, V., Wu, G.X., Li, Z., Tsay, S.C., Hsu, C., Sikka, R., Holben, B.,
doi.org/10.1126/science.288.5468.1042. Lu, D., Tartari, G., Chin, M., Koudelova, P., Chen, H., Ma, Y., Huang, J.,
Andreae, M.O., Merlet, P., 2001. Emission of trace gases and aerosols from biomass Taniguchi, K., Zhang, R., 2008. The Joint Aerosol-Monsoon Experiment: a New
burning. Glob. Biogeochem. Cycles 15, 955–966. https://doi.org/10.1029/ Challenge for Monsoon climate Research. Bull. Am. Meteorol. Soc. 89, 369–384.
2000gb001382. https://doi.org/10.1175/bams-89-3-369.
Bourgeois, Q., Ekman, A.M.L., Krejci, R., 2015. Aerosol transport over the Andes from the Liao, T., Gui, K., Li, Y., Wang, X., Sun, Y., 2021. Seasonal distribution and vertical
Amazon Basin to the remote Pacific Ocean: a multiyear CALIOP assessment. structure of different types of aerosols in Southwest China observed from CALIOP.
J. Geophys. Res. Atmos. 120, 8411–8425. https://doi.org/10.1002/2015jd023254. Atmos. Environ. 246 https://doi.org/10.1016/j.atmosenv.2020.118145.
Bourgeois, Q., Ekman, A.M.L., Renard, J.-B., Krejci, R., Devasthale, A., Bender, F.A.M., Liu, Z., Vaughan, M., Winker, D., Kittaka, C., Getzewich, B., Kuehn, R., Omar, A.,
Riipinen, I., Berthet, G., Tackett, J.L., 2018. How much of the global aerosol optical Powell, K., Trepte, C., Hostetler, C., 2009. The CALIPSO lidar cloud and aerosol
depth is found in the boundary layer and free troposphere? Atmos. Chem. Phys. 18, discrimination: Version 2 algorithm and initial assessment of performance. J. Atmos.
7709–7720. https://doi.org/10.5194/acp-18-7709-2018. Ocean. Technol. 26, 1198–1213. https://doi.org/10.1175/2009jtecha1229.1.
Chen, X., Wang, Z., Yu, F., Pan, X., Li, J., Ge, B., Wang, Z., Hu, M., Yang, W., Chen, H., Ooi, M.C.-G., Chuang, M.-T., Fu, J.S., Kong, S.S., Huang, W.-S., Wang, S.-H.,
2017. Estimation of atmospheric aging time of black carbon particles in the polluted Pimonsree, S., Chan, A., Pani, S.K., Lin, N.-H., 2021. Improving prediction of trans-
atmosphere over Central-Eastern China using microphysical process analysis in boundary biomass burning plume dispersion: from northern peninsular Southeast
regional chemical transport model. Atmos. Environ. 163, 44–56. https://doi.org/ Asia to downwind western North Pacific Ocean. Atmos. Chem. Phys. 21,
10.1016/j.atmosenv.2017.05.016. 12521–12541. https://doi.org/10.5194/acp-21-12521-2021.
Crutzen, P.J., Heidt, L.E., Krasnec, J.P., Pollock, W.H., Seiler, W., 1979. Biomass burning Pan, H., Chen, G., 2019. Diurnal variations of precipitation over north china regulated by
as a source of atmospheric gases CO, H2, N2O, NO, CH3Cl, and COS. Nature. 282, the mountain-plains solenoid and boundary-layer inertial oscillation. Adv. Atmos.
253–256. https://doi.org/10.1007/978-3-319-27460-7_4. Sci. 36, 863–884. https://doi.org/10.1007/s00376-019-8238-3.
Ding, K., Huang, X., Ding, A., Wang, M., Su, H., Kerminen, V.-M., Petäjä, T., Tan, Z., Penner, J.E., Dickison, R.E., O’Neill, C.A., 1992. Effects of aerosol from biomass burning
Wang, Z., Zhou, D., Sun, J., Liao, H., Wang, H., Carslaw, K., Wood, R., Zuidema, P., on the global radiation budget. Science. 256, 1432–1434. https://doi.org/10.1126/
Rosenfeld, D., Kulmala, M., Fu, C., Pöschl, U., Cheng, Y., Andreae, M.O., 2021. science.256.5062.1432.
Aerosol-boundary-layer-monsoon interactions amplify semi-direct effect of biomass Prasad, P., Raman, M.R., Ratnam, M.V., Ravikiran, V., Madhavan, B.L., Bhaskara Rao, S.
smoke on low cloud formation in Southeast Asia. Nat. Commun. 12 https://doi.org/ V., 2019. Nocturnal, seasonal and intra-annual variability of tropospheric aerosols
10.1038/s41467-021-26728-4. observed using ground-based and space-borne lidars over a tropical location of India.
Donahue, N.M., Robinson, A.L., Pandis, S.N., 2009. Atmospheric organic particulate Atmos. Environ. 213, 185–198. https://doi.org/10.1016/j.atmosenv.2019.06.008.
matter: from smoke to secondary organic aerosol. Atmos. Environ. 43, 94–106. Ramanathan, V., Chung, C., Kim, D., Bettge, T., Buja, L., Kiehl, J.T., Washington, W.M.,
https://doi.org/10.1016/j.atmosenv.2008.09.055. Fu, Q., Sikka, D.R., Wild, M., 2005. Atmospheric brown clouds: impacts on south
Dong, X., Fu, J.S., 2015. Understanding interannual variations of biomass burning from Asian climate and hydrological cycle. Proc. Natl. Acad. Sci. 102, 5326–5333. https://
Peninsular Southeast Asia, part II: Variability and different influences in lower and doi.org/10.1073/pnas.0500656102.
higher atmosphere levels. Atmos. Environ. 115, 9–18. https://doi.org/10.1016/j. Rampanelli, G., Zardi, D., Rotunno, R., 2004. Mechanisms of up-valley winds. J. Atmos.
atmosenv.2015.05.052. Sci. 61, 3097–3111. https://doi.org/10.1175/JAS-3354.1.
Duncan, B.N., Martin, R.V., Staudt, A.C., Yevich, R., Logan, J.A., 2003. Interannual and Ratnam, M.V., Prasad, P., Roja Raman, M., Ravikiran, V., Bhaskara Rao, S.V., Krishna
seasonal variability of biomass burning emissions constrained by satellite Murthy, B.V., Jayaraman, A., 2018. Role of dynamics on the formation and
observations. J. Geophys. Res. Atmos. 108 https://doi.org/10.1029/2002jd002378. maintenance of the elevated aerosol layer during monsoon season over south-east
Guan, H., Esswein, R., Lopez, J., Bergstrom, R., Warnock, A., Follette-Cook, M., peninsular India. Atmos. Environ. 188, 43–49. https://doi.org/10.1016/j.
Fromm, M., Iraci, L.T., 2010. A multi-decadal history of biomass burning plume atmosenv.2018.06.023.
heights identified using aerosol index measurements. Atmos. Chem. Phys. 10, Reid, J.S., Koppmann, R., Eck, T.F., Eleuterio, D.P., 2005. A review of biomass burning
6461–6469. https://doi.org/10.5194/acp-10-6461-2010. emissions part II: intensive physical properties of biomass burning particles. Atmos.
Gui, K., Che, H., Zheng, Y., Zhao, H., Yao, W., Li, L., Zhang, L., Wang, H., Wang, Y., Chem. Phys. 5, 799–825. https://doi.org/10.5194/acp-5-799-2005.
Zhang, X., 2021. Three-dimensional climatology, trends, and meteorological drivers Sahu, L.K., Sheel, V., 2014. Spatio-temporal variation of biomass burning sources over
of global and regional tropospheric type-dependent aerosols: insights from 13 years South and Southeast Asia. J. Atmos. Chem. 71, 1–19. https://doi.org/10.1007/
(2007–2019) of CALIOP observations. Atmos. Chem. Phys. 21, 15309–15336. s10874-013-9275-4.
https://doi.org/10.5194/acp-21-15309-2021. Shi, Y., Sasai, T., Yamaguchi, Y., 2014. Spatio-temporal evaluation of carbon emissions
Hansen, J., Sato, M., Ruedy, R., 1997. Radiative forcing and climate response. from biomass burning in Southeast Asia during the period 2001–2010. Ecol. Model.
J. Geophys. Res. Atmos. 102, 6831–6864. https://doi.org/10.1029/96jd03436. 272, 98–115. https://doi.org/10.1016/j.ecolmodel.2013.09.021.
Hirsch, E., Koren, I., 2021. Record-breaking aerosol levels explained by smoke injection Smirnov, A., Holben, B.N., Kaufman, Y.J., Dubovik, O., Eck, T.F., Slutsker, I., Pietras, C.,
into the stratosphere. Science. 371, 1269–1274. https://doi.org/10.1126/science. Halthore, R.N., 2002. Optical properties of atmospheric aerosol in maritime
abe1415. environments. J. Atmos. Sci. 59, 501–523. https://doi.org/10.1175/1520-0469
Ho, H.C., Wong, M.S., Yang, L., Shi, W., Yang, J., Bilal, M., Chan, T.-C., 2018. (2002)059<0501:Opoaai>2.0.Co;2.
Spatiotemporal influence of temperature, air quality, and urban environment on Streets, D.G., Yarber, K.F., Woo, J.H., Carmichael, G.R., 2003. Biomass burning in Asia:
cause-specific mortality during hazy days. Environ. Int. 112, 10–22. https://doi.org/ annual and seasonal estimates and atmospheric emissions. Glob. Biogeochem. Cycles
10.1016/j.envint.2017.12.001. 17. https://doi.org/10.1029/2003gb002040.
Huang, K., Fu, J.S., Hsu, N.C., Gao, Y., Dong, X., Tsay, S.-C., Lam, Y.F., 2013. Impact Tackett, J.L., Winker, D.M., Getzewich, B.J., Vaughan, M.A., Young, S.A., Kar, J., 2018.
assessment of biomass burning on air quality in Southeast and East Asia during CALIPSO lidar level 3 aerosol profile product: version 3 algorithm design. Atmos.
BASE-ASIA. Atmos. Environ. 78, 291–302. https://doi.org/10.1016/j. Meas. Tech. 11, 4129–4152. https://doi.org/10.5194/amt-11-4129-2018.
atmosenv.2012.03.048. Tian, P., Cao, X., Zhang, L., Sun, N., Sun, L., Logan, T., Shi, J., Wang, Y., Ji, Y., Lin, Y.,
Huang, W.R., Wang, S.H., Yen, M.C., Lin, N.H., Promchote, P., 2016. Interannual Huang, Z., Zhou, T., Shi, Y., Zhang, R., 2017. Aerosol vertical distribution and
variation of springtime biomass burning in Indochina: Regional differences, optical properties over China from long-term satellite and ground-based remote
associated atmospheric dynamical changes, and downwind impacts. J. Geophys. Res. sensing. Atmos. Chem. Phys. 17, 2509–2523. https://doi.org/10.5194/acp-17-2509-
Atmos. 121 https://doi.org/10.1002/2016jd025286. 2017.
Janssen, N.A.H., Hoek, G., Simic-Lawson, M., Fischer, P., van Bree, L., ten Brink, H., Tsay, S.-C., Hsu, N.C., Lau, W.K.M., Li, C., Gabriel, P.M., Ji, Q., Holben, B.N., Judd
Keuken, M., Atkinson, R.W., Anderson, H.R., Brunekreef, B., Cassee, F.R., 2011. Welton, E., Nguyen, A.X., Janjai, S., Lin, N.-H., Reid, J.S., Boonjawat, J., Howell, S.
Black Carbon as an additional Indicator of the adverse Health Effects of Airborne G., Huebert, B.J., Fu, J.S., Hansell, R.A., Sayer, A.M., Gautam, R., Wang, S.-H.,
Particles compared with PM10 and PM2.5. Environ. Health Perspect. 119, Goodloe, C.S., Miko, L.R., Shu, P.K., Loftus, A.M., Huang, J., Kim, J.Y., Jeong, M.-J.,
1691–1699. https://doi.org/10.1289/ehp.1003369. Pantina, P., 2013. From BASE-ASIA toward 7-SEAS: a satellite-surface perspective of
Jung, J., Lee, H., Kim, Y.J., Liu, X., Zhang, Y., Gu, J., Fan, S., 2009. Aerosol chemistry and boreal spring biomass-burning aerosols and clouds in Southeast Asia. Atmos.
the effect of aerosol water content on visibility impairment and radiative forcing in Environ. 78, 20–34. https://doi.org/10.1016/j.atmosenv.2012.12.013.
Guangzhou during the 2006 Pearl River Delta campaign. J. Environ. Manag. 90, van der Werf, G.R., Randerson, J.T., Giglio, L., van Leeuwen, T.T., Chen, Y., Rogers, B.M.,
3231–3244. https://doi.org/10.1016/j.jenvman.2009.04.021. Mu, M., van Marle, M.J.E., Morton, D.C., Collatz, G.J., Yokelson, R.J., Kasibhatla, P.
Kahn, R.A., Chen, Y., Nelson, D.L., Leung, F.Y., Li, Q., Diner, D.J., Logan, J.A., 2008. S., 2017. Global fire emissions estimates during 1997–2016. Earth. Syst. Sci. Data. 9,
Wildfire smoke injection heights: two perspectives from space. Geophys. Res. Lett. 697–720. https://doi.org/10.5194/essd-9-697-2017.
35 https://doi.org/10.1029/2007gl032165. VanCuren, R.A., Cliff, S.S., Perry, K.D., Jimenez-Cruz, M., 2005. Asian continental
Kalogridis, A.C., Popovicheva, O.B., Engling, G., Diapouli, E., Kawamura, K., aerosol persistence above the marine boundary layer over the eastern North Pacific:
Tachibana, E., Ono, K., Kozlov, V.S., Eleftheriadis, K., 2018. Smoke aerosol Continuous aerosol measurements from Intercontinental Transport and Chemical

10
Y. Liu et al. Atmospheric Research 301 (2024) 107293

Transformation 2002 (ITCT 2K2). J. Geophys. Res. Atmos. 110 https://doi.org/ Yin, S., Wang, X., Zhang, X., Guo, M., Miura, M., Xiao, Y., 2019. Influence of biomass
10.1029/2004jd004973. burning on local air pollution in mainland Southeast Asia from 2001 to 2016.
Winker, D.M., Vaughan, M.A., Omar, A., Hu, Y., Powell, K.A., Liu, Z., Hunt, W.H., Environ. Pollut. 254 https://doi.org/10.1016/j.envpol.2019.07.117.
Young, S.A., 2009. Overview of the CALIPSO mission and CALIOP data processing Zhang, Y.-N., Zhang, Z.-S., Chan, C.-Y., Engling, G., Sang, X.-F., Shi, S., Wang, X.-M.,
algorithms. J. Atmos. Ocean. Technol. 26, 2310–2323. https://doi.org/10.1175/ 2012. Levoglucosan and carbonaceous species in the background aerosol of coastal
2009jtecha1281.1. Southeast China: case study on transport of biomass burning smoke from the
Wong, C.S., 1978. Atmospheric Input of Carbon Dioxide from burning Wood. Science. Philippines. Environ. Sci. Pollut. Res. 19, 244–255. https://doi.org/10.1007/
200, 197–200. https://doi.org/10.1126/science.200.4338.197. s11356-011-0548-7.
Wu, G., Liu, Y., He, B., Bao, Q., Duan, A., Jin, F.-F., 2012. Thermal controls on the Asian Zhang, L., Ding, S., Qian, W., Zhao, A., Zhao, S., Yang, Y., Weng, G., Tao, M., Chen, H.,
summer monsoon. Sci. Rep. 2 https://doi.org/10.1038/srep00404. Zhao, S., Wang, Z., 2022. The impact of long-range transport of biomass burning
Wu, D., Niu, X., Chen, Z., Chen, Y., Xing, Y., Cao, X., Liu, J., Wang, X., Pu, W., 2022. emissions in Southeast Asia on Southern China. Atmosphere. 13, 1029. https://doi.
Causes and effects of the long-range dispersion of carbonaceous aerosols from the org/10.3390/atmos13071029.
2019–2020 Australian wildfires. Geophys. Res. Lett. 49 https://doi.org/10.1029/ Zieger, P., Fierz-Schmidhauser, R., Weingartner, E., Baltensperger, U., 2013. Effects of
2022gl099840. relative humidity on aerosol light scattering: results from different European sites.
Atmos. Chem. Phys. 13, 10609–10631. https://doi.org/10.5194/acp-13-10609-
2013.

11

You might also like