ArticlePDF Available

Analysis of a Gigantic Jet in Southern China: Morphology, Meteorology, Storm Evolution, Lightning, and Narrow Bipolar Events

Wiley
Journal of Geophysical Research: Atmospheres
Authors:

Abstract and Figures

At about 22:43:30 BJT (Beijing Time = UTC + 8) on 13 August 2016, two amateur astronomers in Shikengkong, Guangdong province, and Jiahe County, Hunan province, respectively, fortuitously captured a gigantic jet (GJ) event simultaneously, and the GJ exact location could be triangulated. The parent thunderstorm was in a very humid environment (Precipitable Water [PWAT] in excess of 60 mm), featuring high convective available potential energy (CAPE of 2,428 J/kg). The GJ occurred in the region with the coldest cloud top brightness temperature of −64 °C, suggesting the GJ was associated with strong vertical development of the thunderstorm. The vertical cross sections of radar reflectivity also show that the GJ occurred near the thunderstorm strong convection region (overshooting top). The negative cloud‐to‐ground flashes dominated during the thunderstorm evolution. Three positive narrow bipolar events (NBEs) were detected within 30 s before and after the GJ. It indicates that the NBEs were occurred in the upper and middle layers of the thunderstorm (altitude of 11–13 km) with radar reflectivity of 30–35 dBZ.
This content is subject to copyright. Terms and conditions apply.
Analysis of a Gigantic Jet in Southern China: Morphology,
Meteorology, Storm Evolution, Lightning, and Narrow
Bipolar Events
Jing Yang
1
, Xiushu Qie
1
, Lihua Zhong
1,2
, Qijia He
3
, Gaopeng Lu
4
, Zhichao Wang
5
,
Yu Wang
6
, Ningyu Liu
7
, Feifan Liu
4
, KangMing Peng
4
, Baoyou Zhu
4
, Anjin Huang
1
,
Mitsuteru Sato
8
, Huien Pan
9
, and Hualong Li
10
1
Key Laboratory of Middle Atmosphere and Global Environment Observation (LAGEO), Institute of Atmospheric
Physics, Chinese Academy of Sciences, Beijing, China,
2
School of Atmospheric Sciences, Chengdu University of
Information Technology, Chengdu, China,
3
Chongqing Air Trafc Control Branch of Southwest Regional Air Trafc
Administration of Civil Aviation of China, Chongqing, China,
4
School of Earth and Space Science, University of Science
and Technology of China, Hefei, China,
5
China Meteorological Administration, Beijing, China,
6
State Grid Electric Power
Research Institute, Wuhan, China,
7
Space Science Center, Department of Physics, University of New Hampshire,
Durham, NH, USA,
8
Department of Cosmosciences, Hokkaido University, Sapporo, Japan,
9
Xinfeng Culture, Broadcast,
Television, Tourism and Sports Bureau, Shaoguan, China,
10
Finance Bureau of Jiahe County, Hunan Province, Chenzhou,
China
Abstract At about 22:43:30 BJT (Beijing Time = UTC + 8) on 13 August 2016, two amateur astronomers
in Shikengkong, Guangdong province, and Jiahe County, Hunan province, respectively, fortuitously
captured a gigantic jet (GJ) event simultaneously, and the GJ exact location could be triangulated. The
parent thunderstorm was in a very humid environment (Precipitable Water [PWAT] in excess of 60 mm),
featuring high convective available potential energy (CAPE of 2,428 J/kg). The GJ occurred in the region
with the coldest cloud top brightness temperature of 64 °C, suggesting the GJ was associated with strong
vertical development of the thunderstorm. The vertical cross sections of radar reectivity also show that the
GJ occurred near the thunderstorm strong convection region (overshooting top). The negative cloudto
ground ashes dominated during the thunderstorm evolution. Three positive narrow bipolar events (NBEs)
were detected within 30 s before and after the GJ. It indicates that the NBEs were occurred in the upper and
middle layers of the thunderstorm (altitude of 1113 km) with radar reectivity of 3035 dBZ.
1. Introduction
The upward electrical discharges, as one type of transient luminous events (TLEs) (Liu et al., 2015), are clas-
sied by their top altitudes: blue starters (2030 km) (Wescott et al., 1996, 2001), blue jets (4050 km) (Lyons
et al., 2003; Wescott et al., 1995, 1998, 2001), and gigantic jets (GJs) (7090 km) (Boggs et al., 2018, 2019;
Cummer et al., 2009; He et al., 2019; Kuo et al., 2009; Peng et al., 2018; Su et al., 2003; Van der Velde et al.,
2007, 2010; Yang & Feng, 2012; Yang etal., 2018). GJs have the largest vertical extension in the upward elec-
trical discharge group that starts from the thunderclouds and propagate upward to the lower ionosphere.
The GJ established a direct path of an electrical connection between the cloud top and the ionosphere
(Cummer et al., 2009; Kuo et al., 2009), which has an important inuence on the electron density and poten-
tial of the ionosphere, electromagnetic environment of the near space, and radio communication. Since the
rst GJ was observed on 14 September 2001 by Pasko et al. (2002) in Puerto Rico, several GJs have been
observed by groundbased (Cummer et al., 2009; He et al., 2019; Lu et al., 2011; Peng et al., 2018; Soula
et al., 2011; Su et al., 2003; Van der Velde et al., 2007, 2010, 2019) and satellitebased experiments (Boggs
et al., 2019; Chen et al., 2008; Kuo et al., 2009). Optical observations show that most of the observed GJs have
treeor carrotshape (Liu et al., 2015; Pasko et al., 2002; Soula et al., 2011; Su et al., 2003) and there is a
color transition zone between the two areas where the bottom (altitude 2040 km) is blue and the top (alti-
tude over 65 km) is red (Peng et al., 2018; Soula et al., 2011).
Past studies show that GJ's parent thunderstorms present diversity and were associated with tall summer
thunderstorm cells (Meyer et al., 2013; Pasko et al., 2002; Su et al., 2003), high precipitation supercell
(Van der Velde et al., 2007), multicell thunderstorm (Van der Velde et al., 2007), tropical thunderstorm
©2020. American Geophysical Union.
All Rights Reserved.
RESEARCH ARTICLE
10.1029/2019JD031538
Key Points:
The location of a gigantic jet was
triangulated using photographs
The jet occurred just after the peak
of a convective pulse
Three narrow bipolar events
clustered around the jet time
Correspondence to:
X. Qie,
qiex@mail.iap.ac.cn
Citation:
Yang, J., Qie, X., Zhong, L., He, Q., Lu,
G., Wang, Z., et al. (2020). Analysis of a
gigantic jet in southern China:
Morphology, meteorology, storm
evolution, lightning, and narrow
bipolar events. Journal of Geophysical
Research: Atmospheres,125,
e2019JD031538. https://doi.org/
10.1029/2019JD031538
Received 21 AUG 2019
Accepted 24 JUN 2020
Accepted article online 29 JUN 2020
YANG ET AL. 1of14
(Boggs et al., 2019; Cummer et al., 2009; Lu et al., 2011). Additionally, GJs can also occur in the
thunderstorm belt of hurricanes or typhoons (Cummer et al., 2009), containing embedded mini
supercells. However, GJ can also be produced by winter thunderstorms with low CAPE of 300 J/kg and
the cloud top height of only 6.5 km as reported by Van der Velde et al. (2010). In addition, lightning
activity in the parent thunderstorms of the GJs was also diverse. Van der Velde et al. (2010) found that
the storm was positive cloudtoground (CG) ashes dominated but there was a very low rate from about
30 min before the time of the GJ. The results from Yang and Feng (2012) demonstrated that CG
(negative CG ashes) dominated during a time window containing the GJ in a summer type
thunderstorm. In contrast, Soula et al. (2011) and Peng et al. (2018) found that no CG ashes were
associated with the occurrence of the GJs. As for the formation mechanism of GJs, Lu et al. (2011) suggest
that narrow bipolar events (NBEs) may be the initial events in the process of GJs. The NBEs are a very
shortlasting, special kind of intracloud discharge. However, NBEs were only found in Lu et al. (2011) and
Liu et al. (2015).
Previous studies have provided ample insights into the GJs. However, the characteristics of the GJ, its asso-
ciated thunderstorm, and lightning activity are still hot issues due to the rarity of GJs (global incidence of
0.01 jets per minute) (Chen et al., 2008). Moreover, most groundbased observed GJs were recorded only
by a single station, and the exact location of the GJs cannot be determined. At about 22:43:30 BJT on 13
August 2016, a GJ event in southern China has been recorded by two amateur astronomers in
Shikengkong, Guangdong Province (24.895°N, 112.966°E), and Jiahe County, Hunan Province (25.595°N,
112.379°E), respectively. The cameras are not GPS synchronized, and the GJ occurrence time given in the
present paper is not very accurate. For the rarity of the GJ and the time of the event are close in
Figures 1a and 1b, the discharge image obtained by the two amateur astronomers is very likely the same,
and we assume that they are the same event. Since the GJ has been captured by two sites, its exact location
can be triangulated. The distance between the GJ and the camera in Guangdong province is the closest
(~30 km) so far. In the present paper, the morphology of the GJ has been analyzed in detail. Meanwhile, a
comprehensive analysis of the meteorological environment, the structure of the parent thunderstorm, and
Figure 1. (a) Image of the GJ, seen in Shikengkong in Ruyuan Yao autonomous county, Qingyuan City, Guangdong
Province, China, 13 August 2016 at approximately 22:43:30 BJT (UTC + 8). Equipment: Canon EOS 6D, Sigma 15 mm
sheye. Exposure parameters: 10 s, f/2.8, ISO1600. The eld view of the camera system is 135° horizontally and 90°
vertically. Image is printed with permission (Huien Pan). (b) Image of the GJ, seen in Jiahe County, Chenzhou City,
Hunan Province, China, 13 August 2016 at approximately 22:43:30 BJT (UTC + 8). Equipment: Nikon D60, Nikon 18
55 mm. Exposure parameters: 15 s, f/3.5, ISO1600. The image is printed with permission (Hualong Li). The stars Markab,
Enif, Homam, Suudalnujun, and Fomalhaut and Skat are used to establish the azimuth and elevations.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 2of14
lightning activity have also been performed by using data from Doppler radar, lightning detection network,
European Centre for MediumRange Weather Forecasts (ECMWF) reanalysis, infrared weather maps,
sounding, and magnetic eld data. What is interesting is that three NBEs were detected in a time window
containing the GJ, and the characteristics of NBEs were also analyzed. The above results enriched our under-
standing of the GJ and the associated thunderstorms.
2. Data and Measurements
Data used in the present paper include optical images, Doppler weather radar data, lightning detection net-
work data, magnetic eld, ECMWF reanalysis, sounding data, and the infrared weather maps of MTSAT
(MultiFunction Transport Satellite).
The optical images of the GJ are obtained from two amateur astronomers in Shikengkong, Guangdong
Province, and Jiahe County, Hunan Province, respectively. This GJ event has been preliminarily analyzed
by another amateur astronomer, and the results are shown in Figure 1 (https://card.weibo.com/article/m/
show/id/2309404009706455701119). They used the following method to triangulate the GJ: First, analyze
the background stars in Figure 1, and nd the azimuth and vertical angles of the GJ relative to
Shikengkong and Jiahe. The intersection point (the location of the GJ) of the line of sight of two amateur
astronomers on the map is calculated by the azimuth angles. The distances between the GJ location and
the observers can be obtained by simple calculations. The eld of view (FOV), the location and altitude of
the camera, the radius of the Earth (6,378 km) are all known. With all the information above are known
and assuming that the GJ is vertical and perpendicular to the ground, the GJ vertical extension can be cal-
culated by using a geometric method, for example, in Yang et al. (2008). We will use this result for further
analysis.
The ECMWF reanalysis is the global grid point data obtained by assimilating ground observations, sounding
balloon observations, satellite inversion, and other data. It is considered to be basically credible and widely
used all over the world. In this paper, ECMWF reanalysis data (http://apps.ecmwf.int/datasets/data/
interim-full-daily/levtype=pl/) are used to analyze the meteorological environment of the GJ parent thun-
derstorm. The resolution is 0.125° × 0.125°.
The atmospheric sounding data have been downloaded from the University of Wyoming (weather.uwyo.
edu/upperair/sounding.html). The data are provided twice daily at 00:00 and 12:00 UTC
(BJT = UTC + 8), and the time used in the present paper is BJT except noted. The sounding station that
is close to the GJ location is Qingyuan Station (23.6820°N, 113.0561°E, Station Number 59280). The cloud
top brightness temperature is obtained from the MTSAT (MultiFunction Transport Satellite) satellite obser-
vations, and the data are downloaded online (from http://weather.is.kochi-u.ac.jp/archive-e.html). The data
were updated every hour, and the spatial resolution is 0.05° × 0.05°. Data from the Sband WSR98D fully
coherent Doppler weather radar located in Guangzhou (24.7908°N, 113.5658°E) are used to obtain more
detailed information of the thunderstorm. The radar data are updated every 6 min with a scanning range
of 230 km and its spatial resolution of 1 km. The location and peak current of CG ashes are given by
Guangdong lightning detection network operated by the State Grid Electric Power Research Institute. The
detection efciency and location error of the network are 92% and 760 m, respectively (Zhang et al., 2014),
which is obtained by using articially triggering lightning with known ground striking points. The electric
eld data are provided by Jianghuai Area Sferic Array (JASA). JASA is a lightning location network, which
consists of six stations. Each station is equipped with a very low frequency (VLF)/low frequency (LF) detec-
tor that consists of a vertical Eeld antenna and two orthogonal magnetic eld frame antennas (Liu
et al., 2018).
3. Analysis of GJ Images
The optical images of the GJ obtained in Shikengkong (Ruyuan Yao autonomous county) and Jiahe county
are shown in Figures 1a and 1b, respectively. Figure 1a is the clearest optical image of the GJ recorded from
groundbased observations so far. Although the upward propagating feature could not be obtained from the
static image, it shows clearly that the GJ connected with the thundercloud top, conrming the observations
of previous studies (Soula et al., 2011). Figure 1a indicates there is a color transition region between the
lower and upper part of the GJ, and the lower part is in blue while the upper part is in red. The top part
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 3of14
of the GJ in Figure 1a was much darker than the lower part; thus, it was not obvious on the image as the
lower part. Only the part connected with the cloud top at the bottom of the GJ can be clearly seen in
Figure 1b. Van der Velde et al. (2007) explained this is due to the much shorter duration of the upper jet
compared to the lower jet. The relative durations can be seen very well in Van der Velde et al. (2019).
Another reason is likely that at f/3.5, ISO1600, and 15 s exposure time, the background light overwhelms
the little light emitted by the GJ during just a few hundred milliseconds.
Due to the long exposure time, the ne structure of GJ cannot be seen. The GJs reported by Peng et al. (2018)
and Soula et al. (2011) have low luminosity and color image; the ne structure of GJ can be seen in low
luminosity image but not in color image. From the analysis results in Figure 1a, it can be seen that the initial
altitude of GJ is about 20 km, but according to the initiating point of the GJ images taken in Guangdong, it is
blocked by clouds due to the close distance and high elevation angle. Therefore, we will use the radar data for
further analysis in Part 5.
4. Meteorological Environment
Figures 2a and 2b show 500 hPa geopotential height, wind, and temperature and 850 hPa geopotential
height, wind, and water vapor ux at 20:00 BJT on August 13, 2016, respectively. The results show that there
was an upper trough on 500 hPa near 117°E and Guangdong Province was located in the front of the upper
trough. The PseudoEquivalent potential temperature was high from the ground to 800 hPa at 2228°N in
Figure 2c, and a large amount of unstable energy was accumulated in the lower atmosphere. The results
Figure 2. (a) 500 hPa geopotential height (black contours, unit: gpdam), wind (black arrow, unit: M/s), and temperature (shaded, unit: °C); (b) 850 hPa
geopotential height (black contours, unit: Gpdam), wind (black arrow, unit: M/s), and vapor ux (shaded, unit: g·cm
1
·hPa
1
·s
1
); (c) vertical prole of
pseudoequivalent potential temperature along 113.324°E of GJ (unit: K); (d) vertical prole of divergence along 24.835°N of GJ (unit:10
5
s
1
) at 20:00 BJT on 13
August 2016. The square box in (a) and (b) represents the region of interest, and the red +represents the location of GJ.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 4of14
of Qingyuan sounding at 20:00 BJT (Table 1) also show the Lifted
Index (LI) was 5.78 °C, indicating that a large amount of unstable
energy was accumulated in the region of interest.
Figure 2b shows there is a distinct transport belt of water vapor on
850 hPa and the center of the large value of vapor ux was controlled
by the westerly wind, which is conducive to continuous water trans-
port to Guangdong Province located in the northeast of the belt.
Distribution of vapor ux divergence at 850 hPa shows that most
regions in Guangdong Province were located in the convergence zone of water vapor ux, which provides
favorable water vapor conditions for the occurrence and development of the thunderstorm. The vertical pro-
le of divergence along 24.835°N where the GJ was located is shown in Figure 2d. The results indicate a
divergence and convergence center at 200 and 900 hPa, respectively, forming favorable conditions for con-
vergence in the lower atmosphere and divergence in the upper atmosphere. The upward motion caused
by the strong divergence in the middle and upper levels provided dynamic conditions for the development
of a thunderstorm.
The Tlogpdiagram from Qingyuan (within 200 km of the GJ parent thunderstorm) at 20:00 BJT 13 August
2016 (Figure 3) shows that the temperature dew point difference was small below 500 hPa, indicating that
relative humidity was high while the large temperature dewpoint difference above 500 hPa was indicative
of low relative humidity. The wind rose diagram in the upper right corner of Figure 3 shows that there
was a large wind shear in the middle and upper level (12 m/s between 400 and 200 hPa). The shear between
the central charge center and upper charge or even the cloud top may be important for the GJ (Krehbiel
et al., 2008; Lazarus et al., 2015; Van der Velde et al., 2010).
Some key environmental parameters obtained from Qingyuan sounding at 20:00 BJT are listed in Table 1.
The CAPE value was high of 2,428 J/kg, and K index was 42.5 °C, which indicates that the atmosphere
is unstable. The 0 °C height was 5.3 km, and the PWAT (precipitable water) was over 60 mm. Overall, the
atmospheric environment was very moist and featured tropical characteristics.
5. Characteristics of the
GJProducing Thunderstorm
5.1. Overall Characteristic of the Thunderstorm
The radar composite reectivity images in Figure 4 show that the
storm began with several small convective cells (named S1) that
appeared at about 15:00 BJT on 13 August 2016 near Shaoguan,
which then merged and reached a maximum radar reectivity of 52
dBZ at 18:24 BJT with the coldest cloud top brightness temperature
of 66°C at 18:24 BJT (Figure 4c). Meanwhile, a new small storm cell
(the GJproducing thunderstorm, named S2) occurred northeast of
S1. Then, thunderstorm S1 gradually weakened and almost disap-
peared at 21:00 BJT. S2 moved westward and developed rapidly after
20:00 BJT with the coldest cloud top brightness temperature of 64°C
(Figure 5a) and the maximum radar reectivity of 57 dBZ at 22:42
BJT (Figure 4e). The GJ occurred at 22:43 during the maturity stage
of the thunderstorm. After the GJ occurrence, the parent thunder-
storm (S2) began to weaken and disappeared completely at 03:00
BJT on August 14. As seen in Figure 4e, the GJ occurred in the area
where the radar composite reectivity was between 40 and 45 dBZ.
5.2. Thunderstorm Structure
Figure 5a shows cloud top brightness temperature at 23:00 BJT with
lightning activity 3 min before the GJ. The results show that the cold-
est brightness temperature region corresponded very well with the
convection region in the GJproducing thunderstorm (S2)
Table 1
The Meteorological Parameters of the Background Environment of the GJ
Parent Thunderstorm
Station
LI
(°C)
CAPE
(J/kg)
PWAT
(mm)
C
height
(km)
06km
shear (m/s)
K index
(°C)
Qingyuan 5.7 2,428 66.3 5.3 4.6 42.5
Figure 3. Upper air sounding (skew Tlogp) temperature (°C, the blue line),
dewpoint (°C, the red line), parcel adiabatic lapse rate (the black line in the
right), and wind (kt) obtained from Qingyuan sounding at 20:00 BJT on 13
August 2016. The circular diagram in the upper right corner is a wind rose
diagram.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 5of14
(Figure 4e). The GJ location was marked by cyan ×in Figure 5a and was located in the area of the lowest
cloud top brightness temperature with 64°C, which was consistent with the previous results (Boggs
et al., 2019; Singh et al., 2017; Soula et al., 2011).
The cloud top temperature determined from MTSAT satellite gives an estimation of the cloud top altitude,
and the coldest temperature indicates the highest altitude (Soula et al., 2009). The previous results indicate
that the GJ location was most likely related to the coldest cloud top temperature; that is, the occurrence of GJ
may be related to the vertical development of thundercloud (He et al., 2019). Figure 5b depicts the area
Figure 4. Radar composite reectivity images at different times on 1314 August 2016. The black ×in panel (e) stands
for the location of GJ.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 6of14
evolution of cloud top brightness temperature of the GJproducing thunderstorm (S2) from 19:0024:00 BJT.
According to the Qingyuan sounding at 20:00 BJT, the temperature of 55°C was at about 13 km altitude.
The area of cloud top brightness temperature colder than 55°C (altitudes above 13 km) continued to
increase until 20:00 BJT. After 20:00 BJT, the area of these temperature intervals began to decrease.
However, the area of the whole system constantly increased until 23:00 BJT. We also noticed that the
coldest cloud top brightness temperature (temperature below 60°C) reappeared at 23:00 BJT (the closest
time to the occurrence of the GJ), indicating the strong vertical development in GJ parent thunderstorm.
Overall, the GJ appeared in the mature stage of the thunderstorm, which is different from the GJ reported
by Soula et al. (2011) and Meyer et al. (2013). The thunderstorm began to weaken and dissipated
gradually after the occurrence of GJ. The production of GJ was associated with the coldest cloud top
brightest temperature (the highest cloud top), which is consistent with previous reports (He et al., 2019;
Meyer et al., 2013; Van der Velde et al., 2010).
In order to clearly show the thunderstorm characteristics, the radar composite reectivity and vertical cross
sections along the line of sight of two amateur astronomers are shown in Figure 6. The pink dotted lines AB
and CD represent the line of sight of the observers at Jiahe and Shikengkong, and the intersection of the two
lines is the location of GJ. Cummer et al. (2009) and Van der Velde et al. (2007, 2010) estimated the position
where the radar reectivity of the azimuth crossed by the GJ was the strongest position for the occurrence of
GJ. Figure 6g shows that the GJ was located near the strongest radar reectivity, which is similar to the pre-
vious results. The location of GJ in the present paper was obtained by triangulation, which is more accurate
than the result estimated by single station observation to a certain extent. It can be seen that the GJ did not
necessarily occur at the position with the strongest radar reectivity.
Previous studies show that most GJs are associated with overshooting tops of the thundercloud (He
et al., 2019; Meyer et al., 2013; Peng et al., 2018; Van der Velde et al., 2007). We use the ECMWF reanalysis
data to calculate the tropopause height. The results show that the tropopause height in the thunderstorm
area was about 15.6 km at 20:00 BJT, and the monthly average tropopause height in this area was about
15.7 km in August 2016. The vertical cross sections shown in Figure 6 suggest that the thunderstorm region
close to the GJ in time interval of 22:3022:42 BJT had strong convection and over shooting tops, indicated
by Figure 6c in which the reectivity of 25 dBZ reached 18 km and 35 dBZ echo top (ET) also reached 16 km
that exceeded the tropopause (15.6 km, as shown by the black horizontal line in the gure). The results
above indicate that the GJ occurred in close proximity to the strong convective zone in the thunderstorm,
which is consistent with the previous reports (Cummer et al., 2009; Huang et al., 2012; Lazarus et al., 2015;
Su et al., 2003). After the occurrence of GJ (22:48 BJT), the ET fell below 15.6 km, and the thunderstorm
decreased gradually. The cloud top reported by Boggs et al. (2018) was 15 km, which is lower than the results
of the present study.
Figure 5. (a) Cloud top brightness temperature of the GJproducing thunderstorm (S2) at 23:00 BJT on 13 August. The
black .and rosered +stand for CG strokes and +CG strokes, respectively. The GJ location was marked by cyan
×. (b) Area evolution of MTSAT cloud top brightness temperature of the GJproducing thunderstorm (S2) at different
temperature intervals during 19:0024:00 BJT.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 7of14
Figure 6. The radar composite reectivity and the vertical section along the lines AB and CD at different times. The vertical section along line AB is shown in
panels (b), (e), (h), and (k); the vertical section along line CD is shown in panel (c), (f), (i), and (l). The black horizontal line in the gure represents the local
tropopause height (~15.6 km).
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 8of14
Quantitative analysis of radar data can help understanding of the thunderstorm. ETs indirectly reect the
strong vertical updraft in a thunderstorm. The area evolution of different ET of the GJproducing storm is
shown in Figure 7a. The results show that the ET area of 1014 km altitude increased rapidly before the
GJ, indicating the convection strengthened around this time. Vertically integrated liquid (VIL) indicates
atmospheric water content and precipitation that weather radar can measure. Figure 7b shows the evolution
of maximum VIL of the GJproducing thunderstorm. The maximum value of VIL decreased sharply 30 min
before the GJ, suggesting the liquid water content in the cloud decreased.
In addition, we also quantitatively analyze the radar composite reectivity of the GJproducing thunder-
storm (Figure 7c). The results show that the area of 4550 dBZ increased, while the area of 3035 and 35
40 dBZ decreased during the time window containing GJ. The above results indicate that GJ occurrence
was associated with strong reectivity.
5.3. Characteristics of the Lightning Activity
Detailed analysis of the lightning activity of the parent thunderstorm has been made, and the results are
shown in Figure 8a. The results show that CG ashes dominated from 20:0024:00 BJT. The peak value
of CG ashes occurred between 22:00 and 23:00 BJT, and the number of CG ashes was about twice that
of +CG ashes. The parent thunderstorm was in its developmentmaturation stage during that time period.
After 23:00, both CG and +CG ashes activity decreased, suggesting the parent thunderstorm was
dissipating.
Figure 8b shows the cumulative ash rates within ±0.25° centered at the GJ location. It shows CG and
+CG ash rates of on average 8 and 6.5 min
1
, respectively. Negative CG ashes almost ceased during a
Figure 7. Area evolution of (a) different echo tops (ET), (b) maximum vertically integrated liquid (VIL), and (c) different
composite reectivity in ±0.5° range of GJ during 22:1223:12 BJT.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 9of14
short period (22:4122:42 BJT), while positive CG ashes increased (16 +CG ashes). Negative CG ashes
increased, and positive CG ashes decreased 1 min before the GJ. The CG ashes continued to decrease
after the GJ. The results of Krehbiel et al. (2008) and Riousset et al. (2010) show that there is a
competitive relationship between the GJ and CG ashes, the charges needed to generate them come from
the same charge source in the cloud, and the charges transferred by the GJ were very large (Cummer
et al., 2009; Lu et al., 2011; Liu et al., 2015). Thus, the sharp decrease of CG ashes may suggest the
accumulation of charges in the cloud and provides enough charge and energy for the generation of GJs.
The change of positive and negative CG ashes rate before and after the GJ may provide some
information for the charge accumulation during the occurrence of GJ. However, since the polarity of the
GJ in this study is unknown and there is no other observation to prove the charge structure in the cloud,
so further study is needed.
Figure 8c shows the peak currents of the positive and negative CG ashes within ±0.5° centered at the GJ
location. A total of 1,227 CG ashes were detected within 30 min before and after the GJ, including 759
CG ashes (about 62%) with the maximum peak current of 174.1 kA, and 468 +CG ashes (about
38%) with the maximum peak current of +159 kA. Before and after the GJ, the average of peak current of
+CG and CG was +29.3 kA (+26.4 kA) and 11.27 kA (25.8 kA), respectively. It can be seen that the
average peak current of +CG did not change much before and after the GJ (+29.3 to +26.4 kA), while the
average peak current of CG was about twice that before the GJ. Based on the results of Krehbiel et al. (2008),
negative GJ and CG have similar charge structure layers in storms. The average peak current of CG was
small before the GJ, indicating that more negative charge accumulated in the clouds and more negative
Figure 8. (a) Evolution of CG ashes per 6 min in the parent thunderstorm (S2), (b) cumulative ash rates within ±0.25°
centered at the GJ, location, and (c) evolution of peak current of CG ashes in the GJ parent thunderstorm (S2) within
±0.5° centered at the GJ location. The vertical black lines in (a), (b), and (b) represent the time of GJ occurrence. The
horizontal blue dotted line in panel (c) represents the value of +15 kA.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 10 of 14
charge was released after the GJ. The above results are only derived from the average peak currents of CG
and +CG, and further studies are needed in the future.
6. Possible Relationship Between GJ and NBE
NBEs are one type of incloud discharge process with relatively short timescale (typically about 1020 μs)
(Smith et al., 1999). NBEs produce extremely powerful natural radio frequency radiation (Rison et al.,
2016; Smith et al., 1999; Zhu et al., 2010). By analyzing the data obtained by Guangzhou Meteorological
Administration, we fortuitously found 11 cases of NBEs during the thunderstorm lifecycle. Among the 11
cases of NBEs, three NBEs were detected within 30 s (22:43:0022:44:00 BJT) before and after the GJ.
Figure 9 shows the waveforms of three NBEs, and it indicates that all the three NBEs in this study were of
positive polarity. The characteristics of the bipolar pulse waveforms were consistent with the results reported
by Wu et al. (2012). The pulse 0 was generated when the observation station received the wave signal of the
NBE (ground wave), and the pulses a and b (reected wave) were the signals that are reected by the iono-
sphere and the ground ionosphere, respectively. The time, location, and electric eld strength of the NBEs
are given in Table 2. Lu et al. (2011) reported that NBE is likely to be the initial event of a GJ. However,
the occurrence time of the GJ cannot be determined exactly, so the correlation between NBE and the GJ can-
not be further studied in this study and need to be completed by other observations.
The result showed that positive NBEs occurred between the upper main positive charge region and the mid-
level main negative charge region in the tripole electrical structure of thunderstorm (Rison et al., 1999; Wu
et al., 2011), mainly starting at 815 km, with an average height of 12.8 km. It occurred as the initiating event
of intracloud lightning discharges. Negative NBE occurred between the upper main positive charge region
and the negative screening charge region (Smith et al., 1999; Wu et al., 2012), starting at 1519 km, with
Figure 9. The electric eld waveform of three positive NBEs detected within 30 s before and after the GJ. (a) 22:43:00
BJT; (b) 22:43:13 BJT; and (c) 22:43:54 BJT.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 11 of 14
an average height of about 16.4 km. The method for calculating the
source height of NBE in the present study is similar to that used by
Wu et al. (2012). The basic principle is to determine the source height
by using the time delay between the original signal (ground wave) of
an NBE and its ionosphere and ground reection signal (reected
wave). Suppose his the source height of NBE, His the height of iono-
spheric reection, t
a
is the time difference between pulses a and 0,
and t
b
is the time difference between pulses b and 0, using geometric
relation to get the following equation:
ctaffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2HhðÞ
2
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h2þr2
p

¼0;(1)
ctbffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2Hþh2þr2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h2þr2
p¼0;(2)
where ris the spherical distance between sensor location (lon0, lat0) and NBE source location (lon1, lat1):
r¼R· arccos sin lat1ðÞsin lat0ðÞþcos lon1ðÞcos lat0ðÞcos lon1ðÞlon0½;(3)
where Ris the Earth radius. Therefore, according to the longitude and latitude of NBE events provided by
the network and the time delay of reected and ground waves, the source height of NBE can be calculated
by combining functions (1) and (2).
According to Equations 1 and 2, the height of three positive NBEs is calculated to be 1113 km as shown in
Table 2, with an average altitude of 12.2 km. Smith et al. (1999) found that NBE tended to occur in the per-
ipheral region of high radar reectivity. Figure 10a shows overlaps of radar composite reectivity at 22:42
BJT with NBEs. Three positive NBEs were located in the periphery of the strong echo region with radar
reectivity of 4550 dBZ. In order to explore the relationship between the source height of positive NBE
and the charge structure corresponding to the thunderstorm area, overlap of the vertical section along line
AB in Figure 10a with NBE is shown in Figure 10b. The results show that positive NBEs were located in the
middle and upper regions of the thunderstorm and occurred in the region with reectivity of 3035 dBZ,
which is similar to the results of Lü et al. (2013).
7. Conclusions
One GJ event captured over a thunderstorm in south China at about 22:43:30 BJT on 13 August 2016 is
reported in this paper. The triangulation of the GJ was obtained through the synchronous observation from
Table 2
Information on the Three Positive NBEs on 13 August 2016
Time (BJT) Location Height (km) E (V/m) Location error
22:43:00 24.80°N, 113.36°E 12.8 450 <500 m
22:43:13 24.79°N, 113.40°E 11.4 135 <500 m
22:43:54 24.91°N, 113.37°E 12.5 298 <500 m
Figure 10. (a) Overlapped image of three positive NBEs generated between 22:43 and 22:44 BJT on 13 August 2016 with
the radar composite reectivity at 22:42 BJT. (b) Vertical cross section along line AB. The positive NBEs are marked by
black ×,the blue +shows the location of GJ.
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 12 of 14
two amateur astronomers in Shikengkong, Guangdong Province, and Jiahe County, Hunan Province,
respectively. We analyzed the meteorological background environment, characteristics of the parent thun-
derstorm, and lightning activity by using multiple data. In addition, three interesting NBEs have been found
in a time window containing the GJ, and the characteristics of NBEs were also analyzed. The main results
obtained from our analysis are summarized as follows:
1. The GJ parent thunderstorm was in a very humid environment (PWAT greater than 60 mm) with high
CAPE (2,428 J/kg) and weak 06 km vertical wind shear (4.6 m/s). The area of coldest cloud top bright-
ness temperature (below 60°C) appeared at the nearest to the occurrence of GJ, suggesting the GJ was
associated with strong vertical development of the thunderstorm.
2. There was an indication of the overshooting top in the parent thunderstorm near the time of GJ, which
reveals the existence of strong convection in the cloud. Quantitative analysis of radar reectivity showed
that before the GJ initiation, the ET area of 1014 km altitude increased rapidly, indicating the convec-
tion enhanced around this time.
3. Analysis of lightning activity showed that the thunderstorm was dominated by CG ashes. A total of
1,227 CG ashes were detected within 30 min before and after the GJ, including 759 CG ashes (about
62%) with the maximum peak current of 174.1 kA, and 468 CG ashes (about 38%) with the maximum
peak current of +159 kA.
4. Three positive NBEs were detected within the 30 s before and after the GJ. It indicates that the NBEs were
distributed between the positive and negative charge region (1113 km). The relationship between NBE
and thunderstorm macrometeorology has the following characteristics: three positive NBEs were located
in the periphery of the strong echo region with radar reectivity of 4550 dBZ and occurred in the upper
and middle layers of a thunderstorm with radar reectivity of 3035 dBZ.
Data Availability Statements
The optical images of the gigantic jet are available online (from the link https://zenodo.org/record/
3734288#.XoMqfLB0yUk).
References
Boggs, L. D., Liu, N., Peterson, M., Lazarus, S., Splitt, M., Lucena , F., et al. (2019). First observations of gigantic jets from geostationary orbit.
Geophysical Research Letters,46, 39994006. https://doi.org/10.1029/2019GL082278
Boggs, L. D., Liu, N. Y., Riousset, J. A., Shi, F., Lazarus, S., Splitt, M., & Rassoul, H. K. (2018). Thunderstorm charge structures produ-
cinggigantic jets. Scientic Reports,8(1), 18085. https://doi.org/10.1038/s41598-018-36309-z
Chen, A. B., Kuo, C. L., Lee, Y. J., Su, H. T., Hsu, R. R., Chern, J. L., et al. (2008). Global distributions and occurrence rates of transient
luminous events. Journal of Geophysical Research,113, A08306. https://doi.org/10.1029/2008JA013101
Cummer, S. A., Li, J., Han, F., Lu, G., Jaugey, N., Lyons, W. A., & Nelson, T. E. (2009). Quantication of the tropospheretoionosphere
charge transfer in a gigantic jet. Nature Geoscience,2(9), 617620. https://doi.org/10.1038/ngeo607
He, Q., Yang, J., Lu, G., Chen, Z., Wang, Y., Sato, M., & Qie, X. (2019). Analysis of the rst positive polarity gigantic jet recorded near the
Yellow Sea in mainland China. Journal of Atmospheric and SolarTerrestrial Physics,190,615. https://doi.org/10.1016/j.
jastp.2019.04.015
Huang, S. M., Hsu, R. R., Lee, L. J., Su, H. T., Kuo, C. L., Wu, C. C., et al. (2012). Optical and radio signatures of negative gigantic jets: Cases
from Typhoon Lionrock (2010). Journal of Geophysical Research,117, A08307. https://doi.org/10.1029/2012JA017600
Krehbiel, P. R., Riousset, J. A., Pasko, V. P., Thomas, R. J., Rison, W., Stanley, M. A., & Edens, H. E. (2008). Upward electrical discharges
from thunderstorms. Nature Geoscience,1(4), 233237. https://doi.org/10.1038/ngeo162
Kuo, C. L., Chou, J. K., Tsai, L. Y., Chen, A. B., Su, H. T., Hsu, R. R., et al. (2009). Discharge processes, electric eld, and electron energy in
ISUALrecorded gigantic jets. Journal of Geophysical Research,114, A04314. https://doi.org/10.1029/2008JA013791
Lazarus, S. M., Splitt, M. E., Brownlee, J., Spiva, N., & Liu, N. (2015). A thermodynamic, kinematic and microphysical analysis of a jet and
gigantic jetproducing Florida thunderstorm. Journal of Geophysical Research: Atmospheres,120,84698490. https://doi.org/10.1002/
2015JD023383
Liu, F., Zhu, B., Lu, G., Qin, Z., Lei, J., Peng, K. M., et al. (2018). Observations of blue discharges associated with negative narrow bipolar
events in active deep convection. Geophysical Research Letters,45, 28422851. https://doi.org/10.1002/2017GL076207
Liu, N., Spiva, N., Dwyer, J. R., Rassoul, H. K., Free, D., & Cummer, S. A. (2015). Upward electrical discharges observed above Tropical
Depression Dorian. Nature Communications,6(1), 5995. https://doi.org/10.1038/ncomms6995
Lü, F., Zhu, B. Y., Ma, M., Wei, L. X., & Ma, D. (2013). Observations of narrow bipolar events during two thunde rstorms in Northeast China.
Science China Earth Sciences,56(8), 14591470. https://doi.org/10.1007/s11430-012-4484-2
Lu, G., Cummer, S. A., Lyons, W. A., Krehbiel, P. R., Li, J., Rison, W., et al. (2011). Lightning development associated with two negative
gigantic jets. Geophysical Research Letters,38, L12801. https://doi.org/10.1029/2011GL047662
Lyons, W. A., Nelson, T. E., Armstrong, R. A., Pasko, V. P., & Stanley, M. A. (2003). Upward electrical discharges from thunderstorm tops.
Bulletin of the American Meteorological Society,84(4), 445454. https://doi.org/10.1175/BAMS-84-4-445
Meyer, T. C., Lang, T. J., Rutledge, S. A., Lyons, W. A., Cummer, S. A., Lu, G., & Lindsey, D. T. (2013). Radar and lightning analyses of
gigantic jetproducing storms. Journal of Geophysical Research: Atmospheres,118, 28722888. https://doi.org/10.1002/jgrd.50302
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 13 of 14
Acknowledgments
This work was supported jointly by the
Strategic Priority Research Program of
the Chinese Academy of Sciences
(Grant XDA17010101) and the National
Natural Science Foundation of China
(Grant 41574141). We agree with the
AGU data policy.
Pasko, V. P., Stanley, M. A., Mathews, J. D., Inan, U. S., & Wood, T. G. (2002). Electrical discharge from a thundercloud top to the lower
ionosphere. Nature,416(6877), 152154. https://doi.org/10.1038/416152a
Peng, K. M., Hsu, R. R., Chang, W. Y., Su, H. T., Chen, A. B. C., Chou, J. K., et al. (2018). Triangulation and coupling of gigantic jets near the
lower ionosphere altitudes. Journal of Geophysical Research: Space Physics,123, 69046916. https://doi.org/10.1029/201 8JA025624
Riousset, J. A., Pasko, V. P., Krehbiel, P. R., Rison, W., & Stanley, M. A. (2010). Modeling of thundercloud screening charges: Implications
for blue and gigantic jets. Journal of Geophysical Research,115, A00E10. https://doi.org/10.1029/2009JA01428 6
Rison, W., Krehbiel, P. R., Stock, M. G., Edens, H. E., Shao, X. M., Thomas, R. J., et al. (2016). Observations of narrow bipolar events reveal
how lightning is initiated in thunderstorms. Nature Communications,7(1), 10,721. https://doi.org/10.1038/ncomms10721
Singh, R., Maurya, A. K., Chanrion, O., Neubert, T., Cummer, S. A., Mlynarczyk, J., et al. (2017). Assessment of unusual gigantic jets
observed during the monsoon season: First observations from Indian subcontinent. Scientic Reports,7(1), 16,436. https://doi.org/
10.1038/541598-017-16696-s
Smith, D. A., Shao, X. M., Holden, D. N., Rhodes, C. T., Brook, M., Krehbiel, P. R., et al. (1999). A distinct class of isolated intracloud
lightning discharges and their associated radio emissions. Journal of Geophysical Research,116(D19), 41894212. https://doi.org/
10.1029/2010JD015581
Soula, S., Van Der Velde, O., Montanya, J., Huet, P., Barthe, C., & Bór, J. (2011). Gigantic jets produced by an isolated tropical thunderstorm
near Réunion Island. Journal of Geophysical Research,116, D19103. https://doi.org/10.1029/2010JD015581
Soula, S., van der Velde, O., Montanyà, J., Neubert, T., Chanrion, O., & Ganot, M. (2009). Analysis of thunderstorm and lightning activity
associated with sprites observed during the EuroSprite campaigns: Two case studies. Atmospheric Research,91(24), 514528. https://
doi.org/10.1016/j.atmosres.2008.06.017
Su, H. T., Hsu, R. R., Chen, A. B., Wang, Y. C., Hsiao, W. S., Lai, W. C., et al. (2003). Gigantic jets between a thundercloud and the iono-
sphere. Nature,423(6943), 974976. https://doi.org/10.1038/nature01759
Van der Velde, O. A., Bór, J., Li, J., Cummer, S. A., Arnone, E., Zanotti, F., et al. (2010). Multiinstrumental observations of a positive
gigantic jet produced by a winter thunderstorm in Europe. Journal of Geophysical Research,115, D24301. https://doi.org/10.1029/
2010JD014442
Van der Velde, O. A., Lyons, W. A., Nelson, T. E., Cummer, S. A., Li, J., & Bunnell, J. (2007). Analysis of the rst gigantic jet recorded over
continental North America. Journal of Geophysical Research,112, D20104. https://doi.org/10.1029/200 7JD008575
Van der Velde, O. A., Montanyà, J., López, J. A., & Cummer, S. A. (2019). Gigantic jet discharges evolve stepwise through the middle
atmosphere. Nature Communications,10(1), 4350. https://doi.org/10.1038/s41467-019-12261-y
Wescott, E. M., Sentman, D., Osborne, D., Hampton, D., & Heavner, M. (1995). Preliminary results from the Sprites94 aircraft campaign: 2.
Blue jets. Geophysical Research Letters,22(10), 12091212. https://doi.org/10.1029/95GL00582
Wescott, E. M., Sentman, D. D., Heavner, M. J., Hampton, D. L., Osborne, D. L., & Vaughan, O. H. Jr. (1996). Blue starters: Brief upward
discharges from an intense Arkansas thunderstorm. Geophysical Research Letters,23(16), 21532156. https://doi.org/10.1029/
96GL01969
Wescott, E. M., Sentman, D. D., Heavner, M. J., Hampton, D. L., & Vaughan, O. H. Jr. (1998). Blue jets: Their relationship to lightning and
very large hailfall, and their physical mechanisms for their production. Journal of Atmospheric and Solarterrestrial Physics,60(79),
713724. https://doi.org/10.1016/S1364-6826(98)00018-2
Wescott, E. M., Sentman, D. D., StenbaekNielsen, H. C., Huet, P., Heavner, M. J., & Moudry , D. R. (2001). New evidence for the brightness
and ionization of blue starters and blue jets. Journal of Geophysical Research,106(A10), 21,54921,554. https://doi.org/10.1029/
2000JA000429
Wu, T., Dong, W., Zhang, Y., Funaki, T., Yoshida, S., Morimoto, T., et al. (2012). Discharge height of lightning narrow bipolar events.
Journal of Geophysical Research,117, D05119. https://doi.org/10.1029/2011JD017054
Wu, T., Dong, W., Zhang, Y., & Wang, T. (2011). Comparison of positive and negative compact intracloud discharges. Journal of
Geophysical Research,116, D03111. https://doi.org/10.1029/2010JD015233
Yang, J., & Feng, G. (2012). A gigantic jet event observed over a thunderstormin Chinese mainland. Chinese Science Bulletin,57(36),
47914800.
Yang, J., Qie, X., Zhang, G., Yang, Z., & Tong, Z. (2008). Red sprites over thunderstorms in the coast of Shandong Province, China. Chinese
Science Bulletin, 2008,53(7), 10791086. https://doi.org/10.1007/s11434-008-0141-8
Yang, J., Sato, M., Liu, N., Lu, G., Wang, Y., & Wang, Z. (2018). A gigantic jet observed over a MCS in middle latitude region. Journal of
Geophysical Research: Atmospheres,123, 977996. https://doi.org/10.1002/2017JD026878
Zhang, Y., Yang, S., Lu, W., Zheng, D., Dong, W., Li, B., et al. (2014). Experiments of articially triggered lightning and its application in
Conghua, Guangdong, China. Atmospheric Research,135136, 330343. https://doi.org/10.1016/j.atmosres.2013.02.010
Zhu, B., Zhou, H., Ma, M., & Tao, S. (2010). Observations of narrow bipolar events in East China. Journal of Atmospheric and
Solarterrestrial Physics,72(23), 271278. https://doi.org/10.1016/j.jastp.2009.12.002
10.1029/2019JD031538
Journal of Geophysical Research: Atmospheres
YANG ET AL. 14 of 14
... Halos are caused by both positive and negative CG lightning (Williams et al., 2007(Williams et al., , 2012Yang et al., 2018). Blue jets and gigantic jets shoot upwards from cloud tops to the stratosphere, or even to the ionosphere (e.g., Mishin & Milikh, 2008;Pasko, 2008;Yang et al., 2020). They are associated rather with the dynamics and intensity of charge separation in the thundercloud, which determine the energy of intra-cloud discharges, for example, narrow bipolar events (Chou et al., 2018;F. ...
... Most research on the physical features of TLEs and their association with thunderstorms and lightning discharges has focused on individual cases (e.g., Chou et al., 2011;Sato et al., 2016;van der Velde et al., 2006;Yang et al., 2020), which always show their specificities and making it difficult to learn their climatology. The studies of TLEs and related thunderstorms over different regions from ground-based observation campaigns conducted over several years, like that over the Eastern Mediterranean in winter (Yair et al., 2015), above Europe and parts of the Mediterranean Sea (Arnone et al., 2020), revealed that the climatology of TLEs and lightning varies with geography. ...
Article
Full-text available
This study investigates the relationship between transient luminous events (TLEs) and lightning strokes, and the characteristics of TLE‐producing thunderstorms over the Tibetan Plateau (TP), and compares them to those over the Yangtze‐River Delta (YRD) and East China Sea (ECS) where at the same latitude during 2005–2015. The data were collected using the Imager of Sprites and Upper Atmospheric Lightning (ISUAL) and the World Wide Lightning Location Network (WWLLN) observations. Elves, sprites and halo (abbreviated as ESHs) were dominantly detected over the southeastern TP (∼88%), mostly in August and September (∼81%). Different from the southeastern TP, the detected sprites and sprite‐to‐lightning stroke ratios over the YRD were larger in spring than those in summer and autumn. Halos were frequently observed in August over all study regions. Blue jets were only detected over the YRD. The density of TLEs over the southeastern TP was three times smaller than those over the YRD and ECS, while the density of elves over the southeastern TP was slightly larger than that over the YRD. The average energy of TLE‐related lightning strokes based on WWLLN was found to be larger over the southeastern TP compared to the YRD and ECS. The ESH‐producing clouds over the southeastern TP had a larger scale with a lower lightning frequency than those over the YRD.
... Since the summer season of (Chen et al., 2014), the amateurs in China, including some professional photographers, have been enthusiastically contributing to the research of TLEs, either by providing the critical observations (e.g., Chou et al., 2016;Yang et al., 2020), or arousing the attention from the publicity. Yang et al. (2020) reported the analysis of a gigantic jet event captured by two amateurs concurrently from different sites in southern part of China. ...
... Since the summer season of (Chen et al., 2014), the amateurs in China, including some professional photographers, have been enthusiastically contributing to the research of TLEs, either by providing the critical observations (e.g., Chou et al., 2016;Yang et al., 2020), or arousing the attention from the publicity. Yang et al. (2020) reported the analysis of a gigantic jet event captured by two amateurs concurrently from different sites in southern part of China. ...
Article
Full-text available
The observations of transient luminous events from space-borne platform extend our exploration on the mysteries of sprite phenomenology from continental thunderstorms to oceanic thunderstorms. By combining with ground-based measurements of causative strokes for hundreds of red sprites observed by the Imager of Sprites and Upper Atmospheric Lightnings (ISUAL) during 2004–2016, there is a consensus that negative cloud-to-ground (CG) strokes spawned by oceanic thunderstorms are more readily to produce sprites. The existing ground-based observations in both Caribbean Sea and near the coast of South China, mainly due to the contributions from numerous amateurs, are generally consistent with the implications of ISUAL observations. However, the physical mechanisms that might cause the enhancement of negative CG strength in the ocean remain not completely understood. There have been analyses on several cases of oceanic thunderstorms abundant in producing negative sprites. It seems that the production of negative sprites heavily depends on the size of parent thunderstorms, and they are often generated by thunderstorm conditions that are also favorable for gigantic jets.
... NBEs are segregated into positive and negative polarities originating at different altitudes of thundercloud regions in normally electrified thunderstorms [9][10][11]. Positive NBEs can occur as the initial discharge event of normal bi-level intra-cloud lightning flashes [5,[12][13][14][15], and also serve as the initial event of negative gigantic jets [16][17][18]. The negative NBEs are far from being as frequent as their positive counterpart, whereas they are usually produced at higher altitudes (most in the altitude range of 14-18 km) between the upper positive charge layer and the negative screening charge region near the top of thunderclouds [9,19]. ...
... The question of what charge condition near cloud top produces negative NBEs is still unclear. Boggs et al. [45] examined the charge structure and mixing conditions of thunderclouds producing gigantic jets, one kind of upward electrical discharge triggered by positive NBEs [16][17][18]. They found that the large spectrum width is always associated with gigantic jets, suggesting strong mixing between the main negative charge layer and the upper positive charge layer. ...
Article
Full-text available
Lightning discharges are the electrical production in thunderclouds. They radiate the bulk of radio signals in the very low-frequency and low-frequency (VLF/LF) that can be detected by ground-based receivers. One kind of special intra-cloud lightning discharges known as narrow bipolar events (NBEs) have been shown to be rare but closely linked to the convective activity that leads to hazardous weather. However, there is still lack of understanding on the meteorological conditions for thunderstorm-producing NBEs, especially for those of negative polarity, due to their rare occurrence. In this work, we aim to investigate what meteorological and electrical conditions of thunderclouds favor the production of negative NBEs. Combining with the VLF/LF radio signal measured by Jianghuai Area Sferic Array (JASA), S-band Doppler radar observation and balloon sounding data, two mid-latitude thunderstorms with outbreaks of negative NBEs at midnight in East China were analyzed. The comparison with the vertical radar profile shows that the bursts of negative NBEs occurred near thunderclouds with overshooting tops higher than 18 km. Manifestation of negative NBEs is observed with a relatively low spectrum width near thundercloud tops. Our findings suggest that the detection of negative NBEs would provide a unique electrical means to remotely probe overshooting tops with implications for the exchange of troposphere and stratosphere.
... As for atmospheric processes, VLF measurements have been traditionally utilized to monitor lightning discharge (e.g., Bozóki et al., 2023;Qie et al., 2013) and thunderstorm activity (e.g., Kubisz et al., 2024;Qie et al., 2022). In addition, various high-energy radiation and transient luminous events in the Earth's atmosphere are associated with unique VLF signatures (Xu, Qie, et al., 2023;Yang et al., 2020;Zhang et al., 2020). VLF measurements have been utilized to understand the underlying mechanism of these atmospheric events (Berge et al., 2022;Zhang et al., 2021). ...
Article
Full-text available
Measurements of Very Low Frequency (VLF) signals from navy transmitters carry direct information about the D‐region ionosphere and have been widely utilized for detecting the electron density at D‐region altitudes, but not frequently for the atmospheric waves therein. Atmospheric waves have been extensively studied using the total ionospheric electron content, but if and how they are correlated with the D‐region ionosphere and VLF measurements still remains poorly investigated. In this study, we have conducted a comprehensive analysis using 7‐year measurements (2016–2022) of VLF signals from the JJI, NWC, and VTX transmitters as being recorded in Suizhou, China. These three transmitter‐receiver paths are representative and the corresponding observations constitute a valuable data set to investigate the periodicities of VLF data. Different from previous studies, we have utilized the ensemble empirical mode decomposition and Lomb‐Scargle methods to determine the periodicities of these data. By contrasting these paths, prominent periodicities ranging from 2 to 730 days have been found, with clear diurnal variation and suggestive latitudinal/longitudinal dependence. Moreover, we have found that the mesospheric temperature is closely related with the annual oscillation of VLF measurements, while this oscillation has a low correlation with solar Lyman‐α fluxes or geomagnetic activity. The oscillations with relatively shorter periods are likely atmospheric waves such as gravity waves, planetary waves, or harmonics of these waves. Our results suggest that, in addition to the electron density, the subionospheric VLF technique can be potentially utilized to remotely sense atmospheric waves that propagate up to or through the D‐region ionosphere.
... Strong lightning strokes, characterized by their large energy (e.g., >10 3 kJ for superbolts) or peak currents (e.g., >75 kA or >100 kA) [3][4][5], were commonly believed to be associated with discharge events initiating in the upper atmosphere, like elves, sprites, halos, and terrestrial gamma-ray flashes [6][7][8][9][10][11][12]. This highlights the potential impact of strong lightning strokes on aerospace, communication, etc., which are associated with the electromagnetic environment in the upper atmosphere. ...
Article
Full-text available
Lightning stroke strength, characterized by energy and peak currents, over the Tibetan Plateau (TP), is investigated by utilizing datasets from the World Wide Lightning Location Network and the Chinese Cloud-to-Ground Lightning Location System during 2016–2019. Focused on the south-central (SC) and southeast (SE) of the TP, it reveals that SE-TP experiences strokes with larger average energy and peak currents. Strong strokes (energy ≥ 100 kJ or peak currents ≥ |100| kA), exhibiting bimodal distribution in winter and summer, are more frequent and have larger average values over the SE-TP than the SC-TP, with diurnal distribution indicating peaks in energy and positive strokes in the middle of the night and negative strokes peaking in the morning. Utilizing the ECMWF/ERA-5 and MERRA-2 reanalysis, we find that stronger strokes correlate with thinner charge zone depths and larger CIWCFs but stable warm cloud depths and zero-degree levels over the SC-TP. Over the SE-TP, stronger strokes are associated with smaller CIWCFs and show turning points for warm cloud depths and zero-degree levels. Thicker charge zone depths correlate with stronger negative strokes but weaker positive strokes. Generating strokes of similar strength over the SC-TP requires larger CIWCFs, thinner warm cloud depths, and lower zero-degree levels than over the SE-TP.
... It was produced in a thunderstorm context consistent with a typical summer thunderstorm, and during a CG-increasing period when overshooting appears. Yang et al. (2020) analyzed a GJ in southern China in terms of its morphology, meteorology, storm evolution, lightning, and narrow bipolar events. It was found that the GJ initiated in the I N P R E S S region with the coldest cloud top brightness temperature and near the strong convection region. ...
Article
Full-text available
Atmospheric electricity is composed of a series of electric phenomena in the atmosphere. Significant advances in atmospheric electricity research conducted in China have been achieved in recent years. In this paper, the research progress on atmospheric electricity achieved in China during 2019–22 is reviewed focusing on the following aspects: (1) lightning detection and location techniques, (2) thunderstorm electricity, (3) lightning forecasting methods and techniques, (4) physical processes of lightning discharge, (5) high energy emissions and effects of thunderstorms on the upper atmosphere, and (6) the effect of aerosol on lightning.
... S1). Other observations with color footage of gigantic jets have a similar appearance (10,31,62). The bright white portion is the leader channel that climbs higher in altitude after the ionospheric connection, and the optical (777.4-nm ...
Article
Full-text available
Occasionally, lightning will exit the top of a thunderstorm and connect to the lower edge of space, forming a gigantic jet. Here, we report on observations of a negative gigantic jet that transferred an extraordinary amount of charge between the troposphere and ionosphere (∼300 C). It occurred in unusual circumstances, emerging from an area of weak convection. As the discharge ascended from the cloud top, tens of very high frequency (VHF) radio sources were detected from 22 to 45 km altitude, while simultaneous optical emissions (777.4 nm OI emitted from lightning leaders) remained near cloud top (15 to 20 km altitude). This implies that the high-altitude VHF sources were produced by streamers and the streamer discharge activity can extend all the way from near cloud top to the ionosphere. The simultaneous three-dimensional radio and optical data indicate that VHF lightning networks detect emissions from streamer corona rather than the leader channel, which has broad implications to lightning physics beyond that of gigantic jets.
... The first real color image of a cluster of sprites shown in Figure 2 was recorded in 1994 . Soon after the discovery of TLEs in North America, different studies and observations started in Europe (Neubert et al., 2001), the eastern Mediterranean (Ganot et al., 2007), China and oceans around Taiwan (Su et al., 2002;Yang et al., 2008Yang et al., , 2020, Japan (Suzuki et al., 2006) and South America (Pinto et al., 2004). ...
Article
Full-text available
Atmospheric electricity has been intensively studied during the last 30 years after the discovery in 1989 of different forms of upper atmospheric electrical discharges (so–called Transient Luminous Events) triggered by lightning in the troposphere. In spite of the significant number of investigations that led to important new results unveiling how lightning produces a zoo of transient electrical discharges from the upper troposphere to the mesosphere, there is still no clear understanding about how all sort of TLEs – including those that occur inside thunderclouds – can contribute to the chemistry of the atmosphere both at the local and global scale. This review paper aims at presenting a perspective on the TLE atmospheric chemistry research done in the past, in the present as well as to describe some of the challenges that await ahead to find the true scientific importance of the non-equilibrium atmospheric chemistry triggered by TLEs. This review comes to conclude that while the global chemical impact of elves and halos are almost negligible, the large scale chemical impact of sprites, blue jets and blue starters and that of impulsive cloud corona discharges might be non–negligible in terms of their possibly measurable contribution to important greenhouse gases such as ozone and nitrous oxide (N2O). Being the third strongest greenhouse gas (after carbon dioxide and methane) and by having the ability to deplete ozone, precise determination of atmospheric N2O sources is of increasing and pressing demand. A new era in atmospheric electricity is just emerging in which dedicated scientific space missions (ISS–LIS, ASIM) together with geostationary lightning sensors (since 2016) and new micro–scale and parameterizations of TLEs in general atmospheric chemistry circulation models will hopefully help to start clarifying the full role of TLEs in the chemistry of the atmosphere.
Article
Full-text available
To evaluate the impact of a blue jet (BJ) discharge on the chemical system in the whole stratosphere as a function of altitude, we developed a detailed ion‐neutral chemistry model. The BJ discharge is formed as a streamer discharge up to 50 km with a leader part up to 28 km associated with high‐temperature chemistry. The simulations are performed in a 2‐day duration to investigate diurnal variations of chemical perturbations at the altitudes of every 2 km from 20 to 50 km. The specific chemistry processes during the leader (considering the molecular diffusion) and streamer discharges react with the whole stratospheric chemical families (oxygen, nitrogen, chlorine, and bromine). We systematically compare the simulations with and without the BJ discharge. The results obtained during the first 100 s indicate the ozone enhanced in the middle stratosphere, while no obvious change in the lower and close to the top of the stratosphere. After 2 days, simulations show that the entire neutral chemical stratospheric system is modified with the enhancement of nitrogen oxides, chlorine, and bromine reservoirs. As a consequence, ozone depletion appears in the middle and upper stratosphere due to the catalytic cycle associated with reactive NOx (=NO + NO2). Each chemical family results in a new equilibrium, and the ozone layer appears to be “shifted” to a lower altitude with its maximum less abundance. Due to the long lifetimes of the chlorine and bromine reservoirs in the stratosphere, the chemical perturbations caused by the BJ discharge at all studied altitudes are maintained.
Article
Full-text available
Coordinated TLE (transient luminous event) optical observations in Taiwan have been heldsince 2011, with an aim to achieve triangulation. Currently, there are four observation stations withbaselines varying from 100 to 400 km between them. The system recorded eight gigantic jets (GJs) thatwere recorded by at least two stations on the night of 20 August 2014. The weather radar data indicate thatthese GJs occurred around the troposphere overshooting tops of a vigorous cumulonimbus cloud. Aleader-to-streamer transition was discerned as the appearance of these GJs changed from jet-like (leader) tofan-like (streamer) at ~40 km altitude. Most of these GJs terminated at the lower ionosphere boundary(80–90 km), but one GJ topped with a 10 km thick diffuse region extended higher than 100 km. Moreover,three sets of the GJs occurred within 0.5–100 s in the same general region. The residual plasma patchesfrom the preceding GJs appear to cause the subsequent GJs to contain more bead structures and to bebrighter. Also, three streamer columns of a subsequent GJ that occurred more than 100 s after the precedingGJ were identified to have rebrightened at 55 to 70 km altitudes. The rebrightened streamers and the beadstructure increments in the subsequent GJs suggest that there were GJ-produced long-lasting plasmapatches in the mesosphere.
Article
Full-text available
Halogenated very short lived substances (VSLS) affect the ozone budget in the atmosphere. Brominated VSLS are naturally emitted from the ocean, and current oceanic emission inventories vary dramatically. We present a new global oceanic emission inventory of Br‐VSLS (bromoform and dibromomethane), considering the physical forcing in the ocean and the atmosphere, as well as the ocean biogeochemistry control. A data‐oriented machine‐learning emulator was developed to couple the air‐sea exchange with the ocean biogeochemistry. The predicted surface seawater concentrations and the surface atmospheric mixing ratios of Br‐VSLS are evaluated with long‐term, global‐scale observations; and the predicted vertical distributions of Br‐VSLS are compared to the global airborne observations in both boreal summer and winter. The global marine emissions of bromoform and dibromomethane are estimated to be 385 and 54 Gg Br per year, respectively. The new oceanic emission inventory of Br‐VSLS is more skillful than the widely used top‐down approaches for representing the seasonal/spatial variations and the annual means of atmospheric concentrations. The new approach improves the model predictability for the coupled Earth system model and can be used as a basis for investigating the past and future ocean emissions and feedbacks under climate change. This model framework can be used to calculate the bidirectional oceanic fluxes for other compounds of interest.
Article
Full-text available
In 2002 it was discovered that a lightning discharge can rise out of the top of tropical thunderstorms and branch out spectacularly to the base of the ionosphere at 90 km altitude. Several dozens of such gigantic jets have been recorded or photographed since, but eluded capture by high-speed video cameras. Here we report on 4 gigantic jets recorded in Colombia at a temporal resolution of 200 µs to 1 ms. During the rising stage, one or more luminous steps are revealed at 32-40 km, before a continuous final jump of negative streamers to the ionosphere, starting in a bidirectional (bipolar) fashion. The subsequent trailing jet extends upward from the jump onset, with a current density well below that of lightning leaders. Magnetic field signals tracking the charge transfer and optical Geostationary Lightning Mapper data are now matched unambiguously to the precisely timed final jump process in a gigantic jet.
Article
Full-text available
Gigantic jets are the clearest manifestation of direct electrical coupling between tropospheric thunderstorms and the ionosphere. They are leaders that emerge from electrical breakdown near the top of thunderstorms and extend all the way to the lower edge of the ionosphere near 90 km altitude. By contrast, blue jets and other related events terminate at much lower altitudes. Gigantic jets have been observed from the ground and from orbit. Some seem to be consistent with an upward-propagating negative discharge of 1,000 to 2,000 C km total charge moment change, but others have not been connected to distinguishable electromagnetic signatures. Here we report simultaneous low-light video images and low-frequency magnetic field measurements of a gigantic jet that demonstrate the presence and dynamics of a substantial electric charge transfer between the troposphere and the ionosphere. The signatures presented here confirm the negative polarity of gigantic jets and constrain the lightning processes associated with them. The observed total charge transfer from the thunderstorm to the ionosphere is 144 C for the assumed channel length of 75 km, which is comparable to the charge transfer in strong cloud-to-ground lightning strokes.
Article
Full-text available
Here we report the first observations of gigantic jets (GJ) by the Geostationary Lightning Mapper (GLM) on board the Geostationary Operational Environmental Satellite − R series (GOES −R). Fourteen GJs produced by Tropical Storm Harvey on 19 August 2017 were observed by both GLM and a ground‐based low‐light‐level camera system. The majority of the GJs produced distinguishable signatures in the GLM data, which included long continuous emissions, large peak flash optical energies, and small lateral propagation distances in comparison with other flashes observed by GLM. For two GJs with the best ground‐based images, each have a single pixel that contains the largest optical energy throughout the duration of the GJ and also coincides with the azimuth of the GJ from the video images. The optical energy of the pixel increases as the gigantic jet propagates upward, reaches its peak when the GJ connects to the ionosphere, and then fades away.
Article
Full-text available
Gigantic jets are atmospheric electrical discharges that propagate from the top of thunderclouds to the lower ionosphere. They begin as lightning leaders inside the thundercloud, and the thundercloud charge structure primarily determines if the leader is able to escape upward and form a gigantic jet. No observationally verified studies have been reported on the thundercloud charge structures of the parent storms of gigantic jets. Here we present meteorological observations and lightning simulation results to identify a probable thundercloud charge structure of those storms. The charge structure features a narrow upper charge region that forms near the end of an intense convective pulse. The convective pulse produces strong storm top divergence and turbulence, as indicated by large values of storm top radial velocity differentials and spectrum width. The simulations show the charge structure produces leader trees closely matching observations. This charge structure may occur at brief intervals during a thunderstorm’s evolution due to the brief nature of convective pulses, which may explain the rarity of gigantic jets compared to other forms of atmospheric electrical discharges.
Article
Full-text available
This paper provides the first concrete evidence that preliminary breakdown (PB) pulses of either polarity in positive cloud-to-ground (+CG) lightning are produced by negative leaders. Three-dimensional location results of PB pulses in 46 +CG flashes are analyzed. The majority (40) of the +CG flashes started with positive PB pulses (+PBPs), the same polarity as the positive return stroke. Location results showed that +PBPs were produced by leaders propagating upward, which were determined to be negative leaders based on PB pulse polarities. Similarly, for the negative PB pulses (−PBPs) found in six +CG flashes, location results showed that they were produced by leaders propagating downward, and we determined that these were also negative leaders. Upward negative leaders producing +PBPs in +CG lightning are very similar to those in intracloud lightning. They usually propagate upward before turning in a horizontal direction. Downward negative leaders producing −PBPs in +CG lightning are more complicated. They usually move back upward after a period of downward propagation. Positive leaders could not be detected, but their possible propagations are analyzed along with possible charge structures for different types of PB pulses. We also demonstrate that PB pulse studies based on single-site records are potentially unreliable.
Article
Full-text available
On 19 August 2012, the Imager of Sprites and Upper Atmospheric Lightning on board the FORMOSAT-2 satellite captured a sequence of seven blue discharges within 1 min that emanated from a parent thunderstorm over Lake Taihu in East China. The analysis of lightning activity produced in the thunderstorm indicates that at least six of these events occurred in association with negative narrow bipolar events (NBEs) that were concurrent with the blue discharge by less than 1 ms, and negative cloud-to-ground occurred within 6 s before each blue discharge, which is in agreement with the modeling presented by Krehbiel et al. (2008). Therefore, the frequent occurrence of negative cloud-to-ground could provide the favorable condition for the production of blue discharges, and negative NBEs are probably the initial event of blue discharges. The detection of negative NBEs might provide a convenient approach to detect the occurrence of blue discharges as lightning bolt shooting upward from the top of energetic thunderstorms.
Article
Full-text available
Gigantic jets (GJs) are mostly observed over summer tropical or tropical-like thunderstorms. This study reports observation of a GJ over a mesoscale convective system (MCS) in the middle latitude region in eastern China. The GJ is observed over a relatively weak radar reflectivity region ahead of the leading line, and the maximum radar echo top along the GJ azimuth was lower than the tropopause in the same region, significantly different from past studies that indicate summer GJs are usually associated with convective surges or overshooting tops. Also different from most of previous observations showing GJ-producing summer thunderstorms only produced GJ type of transient luminous events during their life cycles, two sprites were also captured in a time window of 15 mins containing the GJ, indicating that the MCS provides favorable conditions not only for the GJ but also for the sprites. The balloon-borne soundings of the MCS show that there were large wind shears in the middle and upper level of the thundercloud, which may have played important roles for the GJ production.