Next Article in Journal
Formation, Emission and Control Technology of Particulate Matters during Solid Fuel Combustion
Next Article in Special Issue
Disturbances of Doppler Frequency Shift of Ionospheric Signal and of Telluric Current Caused by Atmospheric Waves from Explosive Eruption of Hunga Tonga Volcano on January 15, 2022
Previous Article in Journal
Anaerobic Digestion and Alternative Manure Management Technologies for Methane Emissions Mitigation on Californian Dairies
Previous Article in Special Issue
Investigations on the Global Spread of the Hunga Tonga-Hunga Ha’apai Volcanic Eruption Using Space-Based Observations and Lagrangian Transport Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extreme Heights of 15 January 2022 Tonga Volcanic Plume and Its Initial Evolution Inferred from COSMIC-2 RO Measurements

by
Saginela Ravindra Babu
1,* and
Neng-Huei Lin
1,2
1
Department of Atmospheric Sciences, National Central University, Taoyuan 32001, Taiwan
2
Center for Environmental Monitoring and Technology, National Central University, Taoyuan 32001, Taiwan
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(1), 121; https://doi.org/10.3390/atmos14010121
Submission received: 9 December 2022 / Revised: 29 December 2022 / Accepted: 2 January 2023 / Published: 5 January 2023

Abstract

:
The Hunga Tonga–Hunga Ha’apai underwater volcano (20.57° S, 175.38° W) violently erupted on 15 January 2022. The volcanic cloud’s top height and initial evolution are delineated by using the Constellation Observing System for Meteorology, Ionosphere, and Climate (COSMIC)-2 radio occultation (RO) measurements. The bending angle (BA) anomaly over the Tonga volcanic plume (within 200 km of the eruption center) at 5:17 UTC on 15 January showed a prominent peak at higher stratospheric heights. The top of the BA anomaly revealed that negative to positive change occurred at ~38 km, indicating the first height where the RO line-of-sight encountered the volcanic plume. The BA anomaly further revealed an increase of ~50% at ~36.1 km, and confirmed that the volcanic plume reached above ~36 km. Furthermore, the evolution of BA perturbations within 24 h after the initial explosion is also discussed herein. From collocated RO profiles with the volcanic plume, we can find a clear descent of the peak altitude of the BA perturbation from ~36.1 km to ~29 km within 24 h after the initial eruption. The results from this study will provide some insights into advancing our understanding of volcanic cloud dynamics and their implementation in volcanic plume modeling.

1. Introduction

On 15 January 2022, the underwater volcano on the Hunga Tonga–Hunga Ha’apai (hereafter, the Tonga eruption) island (20.57° S, 175.38° W) erupted violently at 17:30 local time (4:30 UTC), and released enormous amounts of energy, ash, gases, and steam into the atmosphere [1]. The eruption also triggered tsunami waves around the Pacific Ocean [2,3], atmospheric gravity waves (GWs), and Lamb waves into the atmosphere [4,5,6]. [7] reported strong GWs with amplitudes greater than 30 K (twice the usual GWs) in the mesosphere, associated with the Tonga volcano eruption. Several weather stations worldwide detected atmospheric pressure changes that were among the most extraordinary ever recorded. The geostationary GOES-17 and Himawari-8 satellites captured this explosive eruption and showed volcanic plume dispersion. According to preliminary reports from [1], the volcanic plume was lofted as high as ~40 km into the stratosphere, a record in the modern satellite era which has not been seen in the previous eruptions that occurred in the 21st century [8]. Some cloud portions even reached lower mesospheric altitudes [9]. These preliminary reports are further supported by a recent study by [10]. Based on GOES-17 and Himawari-8 stereo observations, [10] reported that the main umbrella cloud of Tonga reached about 35–40 km, while parts of the center plume reached as high as ~55 km. It is unclear why this eruption was so violent and why the volcanic substances were lofted to extreme heights compared to the other explosions in the 21st century. It is suggested that ‘the involvement of water in the Tonga eruption may have increased the explosivity of the Tonga eruption on 15 January’ [11].
Global Navigation Satellite System (GNSS) radio occultation (RO) measurements, particularly the Constellation Observing System for Meteorology, Ionosphere, and Climate (COSMIC), have been widely used for understanding vertical thermal structure changes due to volcanic eruptions [12,13,14,15]. Refs. [12,13] used GNSS RO for the first time to study the effect of volcanic eruptions on atmospheric temperature by comparing RO profiles before and after the explosion. Ref. [14] extensively used COSMIC RO data to detect the volcanic cloud’s top altitude from the bending angle anomaly. Similarly, [15] reported day-to-day atmospheric temperature variability in response to the Taal volcanic eruption in 2020 by using recently launched COSMIC-2 RO data. All of the studies mentioned above reported the atmospheric thermal structure and volcanic cloud after the eruption, while mainly focusing on the upper troposphere and lower stratosphere (UTLS) region. In addition, it is noted that none of the previous eruptions in the 21st century reached such heights as the Tonga eruption. The high-density measurements from the COSMIC-2 mission provide a unique opportunity to describe the Tonga volcanic cloud and its evolution. By taking this advantage, in the present study, we attempt to demonstrate the extreme heights of the Tonga eruption and its initial evolution within 24 h by utilizing COSMIC-2 RO measurements.

2. Data and Methodology

We utilized Global Navigation Satellite System (GNSS) radio occultation (RO) atmPrf bending angle (BA) vertical profiles acquired onboard COSMIC-2 satellites. The COSMIC-2 mission, launched on 25 June 2019, collects more than 5000 RO soundings per day over the tropics and subtropics [16]. The basic advantage of the COSMIC-2 mission is that it is currently taking frequent measurements over the tropics and providing higher numbers of RO profiles in a single day compared to the previous missions, due to its low inclination of ~24°. The data were downloaded from the COSMIC Data Analysis and Archive Centre (CDAAC) website (https://data.cosmic.ucar.edu/gnss-ro/cosmic2/nrt/level2/ (accessed on 28 December 2022). These data were validated and compared with independent data, including radiosondes, model forecasts, and re-analysis [17]. The details of the temperature retrieval from the bending angle and refractivity profile obtained from the GNSS RO sounding are presented elsewhere [18,19,20,21].
Apart from COSMIC-2 RO, we also used version 5.0 water vapor measurements from the Microwave Limb Sounder (MLS) instrument, operating onboard the NASA Aura satellite. The data was downloaded from the following website: https://acdisc.gesdisc.eosdis.nasa.gov/data/Aura_MLS_Level2/ML2H2O.005/ (accessed on 28 December 2022). More details about version 5.0 WV data can be found in [22], respectively. The Ozone Monitoring Instrument (OMI) sulfur dioxide (SO2) upper tropospheric and stratospheric SO2 column (corresponding to the center of mass altitude of 18 km) data were also used. Details of the retrieval technique were documented by [23].
To distinguish different aerosol subtypes in the atmosphere during the Tonga volcanic eruption, imageries from the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP), aboard the Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observations (CALIPSO) satellite, were also utilized in this study [24,25].

3. Results and Discussion

3.1. Extreme Heights of Tonga Volcanic Plume

The Tonga volcano (20.536° S, 175.382° W) is an underwater caldera volcano located approximately 70 km north-northwest of Tonga’s capital, Nukualofa. On 15 January 2022, at approximately 4 UTC (5 pm local time), the Tonga volcano violently erupted, producing a large volcanic cloud, as shown in Figure 1. A second, smaller eruption occurred at 8 UTC, with no further large eruptions occurring thereafter. Figure 1 shows the GOES-West satellite images at each 10-min time step, starting from 4:10 UTC to 6:00 UTC. A clear expansion of the umbrella cloud around the volcano in a near-circular pattern was evident in Figure 1 (for instance, at 04:50 UTC). The upper umbrella cloud moves westward, presumably due to advection by background stratospheric easterly winds [26,27].
To understand the dispersion and transport of volcanic plumes, a better estimation of the volcanic cloud top (VCT) altitude is crucial. Previous studies have demonstrated the GNSS RO data for supporting the detection and monitoring of volcanic clouds, as well as their effects on weather and climate. The GNSS RO bending angle anomaly method (the maximum peak in the bending angle anomaly) was used to detect the VCT altitude [14,28,29]. We also applied a similar method to the Tonga volcanic cloud and detected the volcanic clouds’ heights. First, we checked the availability of COSMIC-2 RO profiles during the active eruption period. Interestingly, we found two RO profiles around the Tonga volcano during its initial eruptive period. Among the available profiles, only one RO profile was found very near the area of the eruption (hereafter eruptive profile) at 05:17 UTC (red colored dot in Figure 2c), and another, at 7:11 UTC, was found away from the volcano’s center. Quite interestingly, the time of the nearest RO profile (5:17 UTC) was close to the active phase of the Tonga volcanic eruption. We looked further at the Geostationary Operational Environmental Satellites (GOES)-17 satellite images and identified the locations of these profiles, which were shown as a red-colored dot in Figure 2a,b (top panel). We also estimated the approximate distance of the RO profile from the center of the Tonga volcano and displayed it in Figure 2c, respectively. Based on Figure 2a,c, it is very clear that the RO profile at 5:17 UTC may not be located exactly over the volcano center; instead, it looks as if it fell in between the center and the umbrella cloud of the volcano. It was reported that the Tonga eruption produced a large umbrella cloud with a diameter of around 500 km [10]. Interestingly, the available RO profile during the active Tonga volcano eruptive period at 5:17 UTC fell exactly within 200 km of the volcano center (Figure 2c). Hence, the available RO profile at 5:17 UTC can represent the umbrella cloud of an active eruptive plume. This provides us with a more detailed analysis of the available RO profiles.
In order to see the BA anomaly for the eruptive profile at 5:17 UTC (as well as 7:11 UTC) on 15 January, we must first derive a reference profile by using all the available BA profiles from 7–13 January 2022 (one week before the eruption) within the 5° latitude and longitude around the Tonga volcano. Then, we take the difference between the eruptive profile and the reference profile to find deviations near the Tonga eruption region. The obtained BA anomaly was expressed in percentage change, and a prominent peak in the BA anomaly (perturbation) defined the altitude of the volcanic cloud’s top. The obtained BA anomaly of two RO profiles is shown in Figure 2d. As shown in Figure 2d, a pronounced peak in the BA anomaly (~50% deviation from the reference profile) in the 5:17 UTC profile was observed at an altitude of about 36.1 km, suggesting the height of the volcanic plume on 15 January. The extensive peak in the BA anomaly might be due to the higher water vapor and the more likely presence of volcanic substances at the time of the active eruption. The top of the BA perturbation, where the negative to positive change occurred (~38 km), indicates the first height where the RO line-of-sight encountered the volcanic plume (Figure 2d). The BA anomaly analysis from the present study confirms that the height of the umbrella cloud of the Tonga eruption was ~36.1 km, which is in line with the recent reports [6,10,30]. Further, the 7:11 UTC profile also showed a clear change in the BA between 32–35 km, with a peak at ~34 km, respectively.
After the primary eruption, we further checked the available COSMIC-2 RO profiles over the volcanic plume. Interestingly, we discovered six collocated RO profiles with the volcanic cloud after the initial eruption on 15 January. The details of the RO profiles are presented in Table 1 (profiles 3–8), and the locations of these profiles are plotted along with the OMI SO2 data shown in Figure 3a, respectively. A similar method to that used in Figure 2 was applied for the available RO profiles, and obtained a BA anomaly for each RO profile. The obtained BA anomaly profiles are plotted in Figure 3b, respectively. The magenta-colored line in Figure 3a shows the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP) onboard the CALIPSO overpass at 15:20 UTC on January 15. Similarly, the vertical magenta-colored line shows the mean BA anomaly profile which was obtained from the average of all the available six RO profiles. A clear change in the BA anomaly can be noticed in Figure 3b. As observed in Figure 2, a significant enhancement of the BA anomaly was noticed between the 30 to 35 km region, with a maximum peak at ~32 km, respectively. It is noted that the available RO profiles are located just near the CALIPSO overpass at 15:20 UTC. One can expect that, due to CALIPSO’s low signal-to-noise ratio in the stratosphere [31], it may not detect the presence of a volcanic cloud above 31 km, as COSMIC-2 RO detected the peak BA anomaly at 32.5 km (Figure 3b). However, in a recent study reported by [32], a clear presence of a volcanic cloud above ~32 km was observed at 15:20 UTC on 15 January based on CALIPSO version 3.41 data. Our BA anomaly peak from Figure 3b is also in line with the results of [32].
To verify the BA anomaly peak estimated from COSMIC-2 RO data, we further checked the CALIPSO overpass on 16 January 2022. We found an overpass between ~2:59–3:12 UTC over the volcanic cloud region (Figure 4b). Interestingly, there was a collocated COSMIC-2 RO profile (3:07 UTC; atmPrf_C2E2.2022.016.03.10.G02_0001.0001_nc) with the volcanic cloud at the same time. The obtained BA anomaly was plotted and shown in Figure 4c, respectively. We found close agreement between the RO-estimated peak change in the BA anomaly height and the CALIPSO aerosol plume height. CALIOP observations of aerosol subtypes (Figure 4b) confirmed the presence of sulfate aerosols near ~29 km, strongly supporting the notion that the volcanic cloud height reached the upper stratosphere. COSMIC-2 RO estimated the height of the BA anomaly peak change at 3:07 UTC to be ~29 km, which very well matches the presence of a stratospheric aerosol layer at ~29 km. This further provided evidence that the estimated peak in height of the BA anomaly at 5:17 UTC was correct, and confirmed that the Tonga volcanic plume rose above an altitude of ~36 km, respectively.
The observed extreme height of the Tonga volcanic plume can be explained partially by the instantaneous warming due to the pronounced enhancement of the WV. As the Tonga volcano was a submarine volcano and erupted highly explosively, it was expected that it emitted tremendous amounts of WV into the stratosphere, as well as volcanic ash and SO2. The MLS measurements of WV observed on 16 January 2022, shortly after the initial eruption, further support this (Figure 5). Unfortunately, we do not have an MLS satellite overpass over the Tonga volcano for 15 January, during its active eruption. However, we noticed extremely high values of stratospheric WV on 16 January in both day- and nighttime overpass periods. Figure 5 shows that the stratospheric WV mixing ratios exceeded much more than 30 ppm shortly after the eruption downwind of the injection location. The enhancement in the WV was observed at 10 hPa, indicating that volcanic WV was directly injected into the upper stratospheric regions. This direct injection of WV might be due to the extreme explosion of the Tonga eruption via the phreatomagmatic interaction of magma and seawater.
Few studies have reported satellite evidence for the direct injection of WV during the volcanic eruption [33,34]. A few studies reported the injection of volcanic WV during the Pinatubo eruption [35,36]. Recent studies also highlighted the impact of the Tonga eruption on the unprecedented increase in stratospheric WV. The Tonga eruption on 15 January injected extreme amounts of WV into the stratosphere, more than previously seen in modern satellite records [27]. MLS measurements recorded (far exceeding any previous values in the 17-year MLS record) exceptionally high WV values in the stratosphere after the Tong eruption [27]. A recent preliminary study by [4] also highlighted the role of latent heat within the Tonga volcanic plume in generating the gravity waves during the 15 January Tonga eruption. They concluded that the latent heat release from the plume remained the most significant individual gravity wave source worldwide for >12 h, producing circular wavefronts visible across the Pacific basin in satellite observations.
As the Tonga volcanic vent was only tens to hundreds of meters below water, the seawater did not suppress the blast, but was flash-boiled and propelled into the stratosphere [37], rapidly launching a plume of super-heated ash and WV upwards into the atmosphere. There, it condensed, releasing latent heat near-instantaneously across a depth of tens of kilometers. This strong and short-lived force produced instantaneous warming in the volcanic plume. The instantaneous latent heat release continued to add to the positive buoyancy of the volcanic plume, allowing it to reach even greater heights, as observed during the Tonga eruption. Overall, it is concluded that the thermal energy emitted by the Tonga eruption, as well as the instantaneous release of latent heat induced by the condensation of WV within the eruptive plume, are potential sources of the extreme heights of the umbrella cloud produced by the Tonga eruption on 15 January.

3.2. Initial Evolution of Bending Angle Pertubations

Furthermore, we examined the BA perturbations for the collocated RO profiles with the volcanic cloud shortly after the initial eruption. There are a total of 16 RO profiles collocated with the volcanic cloud between 5 UTC on 15 January and 7 UTC on 16 January (approximately 24 h). The available collocated RO profiles and the OMI-observed upper tropospheric and stratospheric SO2 column on 15–16 January 2022 are shown in Figure 6a. The black dots represent the available RO profiles on 15 January and the black star dots for the available RO profiles on 16 January, respectively. The details of the available RO profiles shown in Figure 6a are presented in Table 1. To study the evolution in the BA after the eruption, we generated the reference profile for each RO profile shown in Figure 6a. We considered all the available RO profiles during the period of 7–13 January 2022 (one week before the eruption) within the 5° latitude and longitude region around each RO profile. Then, we removed the reference profile from each RO profile and estimated the change in the BA. The obtained perturbations of each profile were arranged according the time needed to show the evolution. The evolution of the percentage changes in the BA perturbation after the eruption is detailed in Figure 6b. The estimated BA peak altitudes are also overplotted in the respective plots. From the collocated RO profiles with the eruptive cloud, we find a clear descent of the peak altitude of the BA anomaly change, from ~36 km to ~29 km within 24 h after the initial eruption (Figure 6b). The evolution of the BA anomaly peak (volcanic cloud top) observed from the present study is well matched with the recent studies reported which used other satellite measurements [38,39].

4. Summary and Conclusions

The recent eruption of Hunga Tonga–Hunga Ha’apai on 15 January 2022 was the largest explosive volcanic event in the satellite era, and received immense attention from the scientific community around the world. The explosive blast sent a plume of volcanic ash well into the stratosphere, and a portion of the plume even reached lower mesospheric altitudes. Even though the Mt. Pinatubo eruption was the largest volcanic eruption since 1991, information about the initial stage of the eruption remains limited [40]. It is reported that the Stratospheric Aerosol and Gas Experiment (SAGE) missed the initial vertical evolution of the volcanic cloud and its injection height [40]. The unprecedented explosive Tonga eruption provided this rare opportunity to describe the volcanic cloud’s structure, particularly in the stratosphere, during the active eruptive phase. By taking this opportunity, the extreme heights and evolution of the Tonga volcanic cloud altitude were reported by using radio occultation measurements from the COSMIC-2 mission. We have one RO bending angle profile (at 5:17 UTC) during the active eruptive phase over the Tonga volcano region (within 200 km of the volcano). Compared to the reference profile (7–13 January 2022 (one week before the eruption)), we noticed a significant increase of ~50% at ~36.1 km in the BA profile. The height of the maximum BA anomaly in the present study is similar to that in the recent study by [10]. The BA anomaly analysis from the present study confirms that the height of the umbrella cloud for the Tonga eruption was ~36.1 km, which is in line with the recent reports [6,10].
The schematic diagram explains the plausible mechanisms for the extreme heights of the Tonga eruption on 15 January (Figure 7). As shown in Figure 5, the Tonga eruption emitted tremendous amounts of water vapor into the stratosphere, along with volcanic ash and sulfur dioxide. MLS satellite measurements from the present study clearly show a tremendous increase in water vapor in the stratosphere after the initial eruption. Recent studies have also revealed the large enhancement of stratosphere water vapor due to the Tonga eruption. In the presence of volcanic ash, the WV condensed inside the plume and released latent heat into the plume column. This latent heating can provide additional thermal energy to the plume. The continuing release of latent heat during the Tonga eruption further added positive buoyancy to the plume, allowing it to reach even greater heights. It is reported earlier that the release of latent heat adds 13% to the thermal energy released by the volcano, which leads to a further plume rise of 1.5 km [41]. Our results are also supported by the recent study by [4]. They reported that the latent heat within the plume is the most significant source that generated the gravity waves during the Tonga eruption. It is observed that the large enhancement of stratospheric WV from the MLS measurements, along with volcanic material, during the Tonga eruption clearly confirms that an instantaneous release of heat due to the condensation of WV, along with the explosive volcanic heat, might be one of the additional plausible driving forces that lofted the Tonga volcanic plume to greater heights on 15 January. Overall, based on the present results, it can be concluded that the thermal energy emitted by the Tonga eruption, as well as the instantaneous release of latent heat induced by the condensation of WV within the eruptive plume, are potential sources of higher volcanic cloud heights. The observed results from the present study provide some more insights, and lead to a better understanding of the volcanic plume dynamics. Furthermore, results suggest that the all-weather capability, high accuracy, and high vertical resolution measurements from COSMIC-2 RO play an important role in advancing our understanding of the volcanic cloud structure and its implementation in plume modeling. The results from the MLS satellite also highlight the tremendous amount of WV added to the stratosphere shortly after the initial eruption. This WV may affect a variety of stratospheric chemistry processes, particularly ozone depletion. The detailed evolution and transport of the Tonga plume and its impact on the stratospheric WV changes and thermal structure variability can be assessed in future studies.

Author Contributions

Conceptualization, S.R.B.; methodology, S.R.B.; software, S.R.B.; validation, S.R.B.; formal analysis, S.R.B.; investigation, S.R.B.; resources, N.-H.L.; data curation, S.R.B.; writing—original draft preparation, S.R.B.; writing—review and editing, S.R.B. and N.-H.L.; visualization, S.R.B.; supervision, N.-H.L.; project administration, N.-H.L.; funding acquisition, N.-H.L. All authors have read and agreed to the published version of the manuscript.

Funding

The National Science and Technology Council (earlier Ministry of Science and Technology), Taiwan primarily supports the work, under the grants of MOST 110-2811-M-008-562 and MOST 109-2811-M-008-553.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Publicly available satellite datasets were analyzed in the present study. These datasets can be found on their respective websites.

Acknowledgments

The authors thank to the National Science and Technology Council (earlier Ministry of Science and Technology), Taiwan. The authors thank the COSMIC Data Analysis and Archive Centre (CDAAC) for providing COSMIC-2 data used in the present study through its FTP site (https://data.cosmic.ucar.edu/gnss-ro/cosmic2/nrt/level2/ (accessed on 28 December 2022)). Aura MLS observations were obtained from the GES DISC through their FTP site (https://acdisc.gesdisc.eosdis.nasa.gov/data/Aura_MLS_Level2/ML2H2O.005 (accessed on 28 December 2022)). We also thank NASA for providing the Aura Ozone Monitoring Instrument (OMI) data (https://aura.gesdisc.eosdis.nasa.gov/data/Aura_OMI_Level2/OMSO2.003/2020 (accessed on 28 December 2022)) and the CALIPSO standard images (https://www-calipso.larc.nasa.gov/ (accessed on 28 December 2022)).

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Global Volcanism Program. Report on hunga tonga-hunga ha’apai (tonga). In Bulletin of the Global Volcanism Network; Smithsonian Institution: Washington, DC, USA, 2022; Volume 40, p. 1. [Google Scholar] [CrossRef]
  2. Adam, D. Tonga volcano eruption created puzzling ripples in earth’s atmosphere. Nature 2022, 601, 497. [Google Scholar] [CrossRef] [PubMed]
  3. Carvajal, M.; Sepulveda, I.; Gubler, A.; Garreaud, R. Worldwide signature of the 2022 Tonga volcanic tsunami. Geophys. Res. Lett. 2022, 49, e2022GL098153. [Google Scholar] [CrossRef]
  4. Wright, C.; Hindley, N.; Alexander, M.J.; Barlow, M.; Hoffmann, L.; Mitchell, C.; Prata, F.; Bouillon, M.; Carstens, J.; Clerbaux, C. Surface-to-space atmospheric waves from Hunga Tonga-Hunga Ha’apai eruption. Nature 2022, 609, 741–746. [Google Scholar] [CrossRef] [PubMed]
  5. Matoza, R.S.; Fee, D.; Assink, J.D.; Iezzi, A.M.; Green, D.N.; Kim, K.; Toney, L.; Lecocq, T.; Krishnamoorthy, S.; Lalande, J.M.; et al. Atmospheric waves and global seismoacoustic observations of the January 2022 Hunga eruption, Tonga. Science 2022, 377, 95–100. [Google Scholar] [CrossRef]
  6. Proud, S.R.; Prata, A.T.; Schmauß, S. The January 2022 Eruption of Hunga Tonga-Hunga Ha’apai Volcano Reached the Mesosphere. Science 2022, 378, 554–557. [Google Scholar] [CrossRef]
  7. Liu, X.; Xu, J.; Yue, J.; Kogure, M. Strong gravity waves associated with Tonga volcano eruption revealed by SABER observations. Geophys. Res. Lett. 2022, 49, e2022GL098339. [Google Scholar] [CrossRef]
  8. Tournigand, P.-Y.; Cigala, V.; Lasota, E.; Hammouti, M.; Clarisse, L.; Brenot, H.; Prata, F.; Kirchengast, G.; Steiner, A.K. A multi-sensor satellite-based archive of the largest SO2 volcanic eruptions since 2006. Earth Syst. Sci. Data 2020, 12, 3139–3159. [Google Scholar] [CrossRef]
  9. NASA Earth Observatory. Tonga Volcano Plume Reached the Mesosphere. 2022. Available online: https://earthobservatory.nasa.gov/images/149474/tonga-volcano-plume-reached-the-mesosphere (accessed on 15 January 2022).
  10. Carr, J.L.; Horváth, Á.; Wu, D.L.; Friberg, M.D. Stereo Plume Height and Motion Retrievals for the Record-Setting Hunga Tonga-Hunga Ha’apai Eruption of 15 January 2022. Geophys. Res. Lett. 2022, 49, e2022GL098131. [Google Scholar] [CrossRef]
  11. NASA Earth Observatory. Hunga Tonga-Hunga Ha‘apai Erupts. 2022. Available online: https://earthobservatory.nasa.gov/images/149347/hunga-tonga-hunga-haapai-erupts (accessed on 19 January 2022).
  12. Wang, K.Y.; Lin, S.C.; Lee, L.C. Immediate impact of the Mt Chaiten eruption on atmosphere from FORMOSAT-3/COSMIC constellation. Geophys. Res. Lett. 2009, 36, L03808. [Google Scholar] [CrossRef]
  13. Okazaki, I.; Heki, K. Atmospheric temperature changes by vol-canic eruptions: GPS radio occultation observations in the 2010 Icelandic and 2011 Chilean cases. J. Volcanol. Geoth. Res. 2012, 245, 123–127. [Google Scholar] [CrossRef] [Green Version]
  14. Biondi, R.; Steiner, A.K.; Kirchengast, G.; Brenot, H.; Rieckh, T. Supporting the detection and monitoring of volcanic clouds: A promising new application of Global Navigation Satellite System radio occultation. Adv. Sp. Res. 2017, 60, 2707–2722. [Google Scholar] [CrossRef]
  15. Ravindra Babu, S.; Liou, Y.A. Day-to-day variability of upper troposphere and lower stratosphere temperature in response to Taal volcanic eruption inferred from COSMIC-2 RO measurements. J. Volcanol. Geotherm. Res. 2022, 421, 107445. [Google Scholar] [CrossRef]
  16. Ho, S.-P.; Anthes, R.A.; Ao, C.O.; Healy, S.; Horanyi, A.; Hunt, D.; Mannucci, A.J.; Pedatella, N.; Randel, W.; Simmons, A. The COSMIC/FORMOSAT-3 radio occultation mission after 12 years: Accomplishments, remaining challenges, and potential impacts of COSMIC-2. Am. Meteorol. Soc. 2019, 101, E1107–E1136. [Google Scholar] [CrossRef] [Green Version]
  17. Schreiner, W.S.; Weiss, J.P.; Anthes, R.A.; Braun, J.; Chu, V.; Fong, J.; Zeng, Z. COSMIC-2 radio occultation constellation: First results. Geophys. Res. Lett. 2020, 47, 1–7. [Google Scholar] [CrossRef]
  18. Kursinski, E.R.; Hajj, G.A.; Schofield, J.T.; Linfield, R.P.; Hardy, K.R. Observing Earth’s atmosphere with radio occultation measurements using the Global Positioning System. J. Geophys. Res. 1997, 102, 23429–23465. [Google Scholar] [CrossRef]
  19. Kuo, Y.-H.; Wee, T.-K.; Sokolovskiy, S.; Rocken, W.; Schreiner, W.; Hunt, H.; Anthes, R.A. Inversion and Error Estimation of GPS Radio Occultation Data. J. Meteorol. Soc. Jpn. 2004, 82, 507–531. [Google Scholar] [CrossRef] [Green Version]
  20. Anthes, R.A.; Bernhardt, P.A.; Chen, Y.; Cucurull, L.; Dymond, K.F.; Ector, D.; Healy, S.B.; Ho, S.-H.; Hunt, D.C.; Kuo, Y.-H.; et al. The COSMIC/Formosat 3 mission: Early results. Am. Meteorol. Soc. 2008, 89, 313–333. [Google Scholar] [CrossRef]
  21. Schreiner, W.; Rocken, C.; Sokolovskiy, S.; Syndergaard, S.; Hunt, D. Estimates of the precision of GPS radio occultations from the COSMIC/FORMOSAT-3 mission. Geophys. Res. Lett. 2007, 34, L04808. [Google Scholar] [CrossRef] [Green Version]
  22. Livesey, N.J.; Read, W.G.; Froidevaux, L.; Lambert, A.; Santee, M.L.; Schwartz, M.J.; Millán, L.F.; Jarnot, R.F.; Wagner, P.A.; Hurst, D.F.; et al. Investigation and amelioration of long-term instrumental drifts in water vapor and nitrous oxide measurements from the Aura Microwave Limb Sounder (MLS) and their implications for studies of variability and trends. Atmos. Chem. Phys. 2021, 21, 15409–15430. [Google Scholar] [CrossRef]
  23. Li, C.; Krotkov, N.A.; Carn, S.; Zhang, Y.; Spurr, R.J.D.; Joiner, J. New-generation NASA Aura Ozone Monitoring In-strument (OMI) volcanic SO2dataset: Algorithm description, ini-tial results, and continuation with the Suomi-NPP Ozone Map-ping and Profiler Suite (OMPS). Atmos. Meas. Tech. 2017, 10, 445–458. [Google Scholar] [CrossRef] [Green Version]
  24. Anselmo, T. Cloud-Aerosol LIDAR Infrared Pathfinder Satellite Observations: Data Management System, Data Products Catalog; Document No: PC-SCI-503; NASA: Hampton, VA, USA, 2006.
  25. Young, S.A.; Vaughan, M.A. The retrieval of profiles of particulate extinction from Cloud Aerosol Lidar Infrared Pathfinder Satellite Observations (CALIPSO) data: Algorithm description. J. Atmos. Ocean. Technol. 2009, 26, 1105–1119. [Google Scholar] [CrossRef]
  26. Sellitto, P.; Podglajen, A.; Belhadji, R.; Boichu, M.; Carboni, E.; Cuesta, J.; Duchamp, C.; Kloss, C.; Siddans, R.; Bègue, N.; et al. The unexpected radiative impact of the Hunga Tonga eruption of 15th January 2022. Commun. Earth Environ. 2022, 3, 288. [Google Scholar] [CrossRef]
  27. Millán, L.; Santee, M.L.; Lambert, A.; Livesey, N.J.; Werner, F.; Schwartz, M.J.; Pumphrey, H.C.; Manney, G.L.; Wang, Y.; Su, H.; et al. The Hunga Tonga-Hunga Ha’apai Hydration of the Stratosphere. Geophys. Res. Lett. 2022, 49, e2002GL099381. [Google Scholar] [CrossRef] [PubMed]
  28. Cigala, V.; Biondi, R.; Prata, A.J.; Steiner, A.K.; Kirchengast, G.; Brenot, H. GNSS radio occultation advances the monitoring of volcanic clouds: The case of the 2008 Kasatochi eruption. Remote Sens. 2019, 11, 2199. [Google Scholar] [CrossRef] [Green Version]
  29. Prata, A.T.; Folch, A.; Prata, A.J.; Biondi, R.; Brenot, H.; Cimarelli, C.; Corradini, S.; Lapierre, J.; Costa, A. Anak Krakatau triggers volcanic freezer in the upper troposphere. Sci. Rep. 2020, 10, 3584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. NASA Earth Observatory. Dramatic Changes at Hunga Tonga-Hunga Ha‘apai. 2022. Available online: https://earthobservatory.nasa.gov/images/149367/dramatic-changes-at-hunga-tonga-hunga-haapai (accessed on 21 January 2022).
  31. Kar, J.; Lee, K.-P.; Vaughan, M.A.; Tackett, J.L.; Trepte, C.R.; Winker, D.M.; Lucker, P.L.; Getzewich, B.J. CALIPSO level 3 stratospheric aerosol profile product: Version 1.00 algorithm description and initial assessment. Atmos. Meas. Tech. 2019, 12, 6173–6191. [Google Scholar] [CrossRef] [Green Version]
  32. Taha, G.; Loughman, R.; Colarco, P.R.; Zhu, T.; Thomason, L.W.; Jaross, G. Tracking the 2022 Hunga Tonga-Hunga Ha’apai Aerosol Cloud in the Upper and Middle Stratosphere Using Space-Based Observations. Geophys. Res. Lett. 2022, 49, e2022GL100091. [Google Scholar] [CrossRef]
  33. Sioris, C.E.; Malo, A.; McLinden, C.A.; D’Amours, R. Direct injection of water vapor into the stratosphere by volcanic eruptions. Geophys. Res. Lett. 2016, 43, 7694–7700. [Google Scholar] [CrossRef]
  34. Sioris, C.E.; Zou, J.; McElroy, C.T.; Boone, C.D.; Sheese, P.E.; Bernath, P.F. Water vapour variability in the high-latitude upper troposphere—Part 2: Impact of volcanic eruptions. Atmos. Chem. Phys. 2016, 16, 2207–2219. [Google Scholar] [CrossRef]
  35. Joshi, M.M.; Jones, G.S. The climatic effects of the direct injection of water vapour into the stratosphere by large volcanic eruptions. Atmos. Chem. Phys. 2009, 9, 6109–6118. [Google Scholar] [CrossRef] [Green Version]
  36. Nedoluha, G.E.; Bevilacqua, R.M.; Gomez, R.M.; Siskind, D.E.; Hicks, B.C.; Russell, J.M.; Connor, B.J. Increases in middle atmospheric water vapor as observed by the Halogen Occultation Experiment and the ground-Based Water Vapor Millimeter-Wave Spectrometer from 1991 to 1997. J. Geophys. Res. Atmos. 1998, 103, 3531–3543. [Google Scholar] [CrossRef]
  37. Witze, A. Why the Tongan eruption will go down in the history of volcanology. Nature 2022, 602, 376–378. [Google Scholar] [CrossRef] [PubMed]
  38. Khaykin, S.; Podglajen, A.; Ploeger, F.; Grooß, J.U.; Tencé, F.; Bekki, S.; Khlopenkov, K.; Bedka, K.; Rieger, L.; Baron, A.; et al. Global perturbation of stratospheric water and aerosol burden by Hunga eruption. Commun Earth Environ. 2022, 3, 316. [Google Scholar] [CrossRef]
  39. Gupta, A.K.; Bennartz, R.; Fauria, K.E.; Mittal, T. Eruption chronology of the December 2021 to January 2022 Hunga Tonga-Hunga Ha’apai eruption sequence. Commun. Earth Environ. 2022, 3, 314. [Google Scholar] [CrossRef]
  40. Thomason, L.W. Observations of a new SAGE II aerosol extinction mode following the eruption of Mt. Pinatubo. Geophys. Res. Lett. 1992, 19, 2179–2182. [Google Scholar] [CrossRef]
  41. Herzog, M.; Graf, H.-F.; Oberhuber, J.M.; Textor, C. The effect of phase changes of water on the development of volcanic plumes. J. Volcanol. Geotherm. Res. 1998, 87, 55–74. [Google Scholar] [CrossRef]
Figure 1. Geostationary Operational Environmental Satellites (GOES)-West satellite (currently, GOES-17) images observed on 15 January 2022.
Figure 1. Geostationary Operational Environmental Satellites (GOES)-West satellite (currently, GOES-17) images observed on 15 January 2022.
Atmosphere 14 00121 g001
Figure 2. Top panel shows (a) GOES–17 Geo Color, multispectral blended infrared image, observed at 5:20 UTC, and (b) at 7:10 UTC on 15 January 2022. The red-colored (blue) dot represents the available COSMIC–2 RO profile at 5:17 (7:11) UTC. (c) The location of the aforementioned RO profile concerning Tonga volcano center. The bottom panel shows perturbations of (d) the bending angle for the aforementioned RO profile concerning the reference profile (mean of 7–13 January 2022).
Figure 2. Top panel shows (a) GOES–17 Geo Color, multispectral blended infrared image, observed at 5:20 UTC, and (b) at 7:10 UTC on 15 January 2022. The red-colored (blue) dot represents the available COSMIC–2 RO profile at 5:17 (7:11) UTC. (c) The location of the aforementioned RO profile concerning Tonga volcano center. The bottom panel shows perturbations of (d) the bending angle for the aforementioned RO profile concerning the reference profile (mean of 7–13 January 2022).
Atmosphere 14 00121 g002
Figure 3. (a) OMI observed UTLS SO2 on 15–16 January 2022 along with the available COSMIC–2 RO profiles (black dots) on 15 January. The magenta-colored line shows the CALIPSO overpass at 15:20 UTC on 15 January. (b) Bending angle (BA) percentage change for the aforementioned RO profiles. The mean profile of the BA change is shown in the magenta-colored profile.
Figure 3. (a) OMI observed UTLS SO2 on 15–16 January 2022 along with the available COSMIC–2 RO profiles (black dots) on 15 January. The magenta-colored line shows the CALIPSO overpass at 15:20 UTC on 15 January. (b) Bending angle (BA) percentage change for the aforementioned RO profiles. The mean profile of the BA change is shown in the magenta-colored profile.
Atmosphere 14 00121 g003
Figure 4. (a) CALIPSO 532 nm total attenuated backscatter coefficient; (b) the CALIPSO aerosol subtypes at @3 UTC on 16 January 2022; (c) the available COSMIC–2 RO bending angle anomaly at 3:07 UTC on 16 January 2022.
Figure 4. (a) CALIPSO 532 nm total attenuated backscatter coefficient; (b) the CALIPSO aerosol subtypes at @3 UTC on 16 January 2022; (c) the available COSMIC–2 RO bending angle anomaly at 3:07 UTC on 16 January 2022.
Atmosphere 14 00121 g004
Figure 5. Map of water vapor mixing ratio at different levels from the Microwave Limb Sounder (MLS) measurements, observed at two local times, 12 h apart, on 16 January 2022. Mid–afternoon sampling runs southeast to northwest, while nighttime observations run northeast to southwest. The red–colored pentagram represents the location of the Tonga volcano.
Figure 5. Map of water vapor mixing ratio at different levels from the Microwave Limb Sounder (MLS) measurements, observed at two local times, 12 h apart, on 16 January 2022. Mid–afternoon sampling runs southeast to northwest, while nighttime observations run northeast to southwest. The red–colored pentagram represents the location of the Tonga volcano.
Atmosphere 14 00121 g005
Figure 6. (a) OMI observed UTLS SO2 on 15–16 January 2022 along with the available COSMIC–2 RO profiles on 15–16 January 2022, a time series of the distribution of (b) bending angle (BA) percentage change and the maximum BA peak altitude (black circles), for the aforementioned RO profiles. The black dots shown in subplot (a) represent the available RO profiles on 15 January and the black star dots for the available RO profiles on 16 January. The magenta–colored pentagram shown in subplot (a) represents the location of the Tonga volcano.
Figure 6. (a) OMI observed UTLS SO2 on 15–16 January 2022 along with the available COSMIC–2 RO profiles on 15–16 January 2022, a time series of the distribution of (b) bending angle (BA) percentage change and the maximum BA peak altitude (black circles), for the aforementioned RO profiles. The black dots shown in subplot (a) represent the available RO profiles on 15 January and the black star dots for the available RO profiles on 16 January. The magenta–colored pentagram shown in subplot (a) represents the location of the Tonga volcano.
Atmosphere 14 00121 g006
Figure 7. (a) observed COSMIC–2 RO bending angle anomaly at 5:17 UTC on 15 January 2022; (b) schematic diagram summarizing the observed extreme heights of the Tonga volcanic plume on 15 January 2022.
Figure 7. (a) observed COSMIC–2 RO bending angle anomaly at 5:17 UTC on 15 January 2022; (b) schematic diagram summarizing the observed extreme heights of the Tonga volcanic plume on 15 January 2022.
Atmosphere 14 00121 g007
Table 1. Details of collocated COSMIC-2 radio occultation profiles with the volcanic cloud (OMI SO2) on 15 and 16 January. The highlighted file names are related to the RO profiles that were available during the eruptive stage, as shown in Figure 2.
Table 1. Details of collocated COSMIC-2 radio occultation profiles with the volcanic cloud (OMI SO2) on 15 and 16 January. The highlighted file names are related to the RO profiles that were available during the eruptive stage, as shown in Figure 2.
File NameLongitudeLatitudeStarting Time for RO
HourMinutes
atmPrf_C2E3.2022.015.05.20.G05_0001.0001_nc183.9595−19.299517
atmPrf_C2E3.2022.015.07.11.G08_0001.0001_nc181.5302−23.3197711
atmPrf_C2E6.2022.015.17.47.G22_0001.0001_nc171.1261−21.02821745
atmPrf_C2E6.2022.015.17.58.G15_0001.0001_nc172.4965−20.07581758
atmPrf_C2E1.2022.015.18.15.R05_0001.0001_nc172.8534−23.79791815
atmPrf_C2E4.2022.015.18.31.R05_0001.0001_nc169.7485−18.44421831
atmPrf_C2E4.2022.015.20.13.G15_0001.0001_nc168.2508−18.09952013
atmPrf_C2E4.2022.015.23.27.G17_0001.0001_nc171.9524−17.08922324
atmPrf_C2E3.2022.016.01.45.R04_0001.0001_nc165.9369−22.7931141
atmPrf_C2E4.2022.016.02.47.G02_0001.0001_nc162.7582−17.231245
atmPrf_C2E4.2022.016.02.58.G16_0001.0001_nc165.553−19.6844258
atmPrf_C2E2.2022.016.03.10.G02_0001.0001_nc167.2092−17.46337
atmPrf_C2E2.2022.016.03.23.G26_0001.0001_nc176.6833−18.2854323
atmPrf_C2E5.2022.016.03.54.G16_0001.0001_nc168.662−17.5328354
atmPrf_C2E5.2022.016.05.27.G15_0001.0001_nc165.2896−17.7996524
atmPrf_C2E2.2022.016.06.34.G12_0001.0001_nc171.0468−19.1492631
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ravindra Babu, S.; Lin, N.-H. Extreme Heights of 15 January 2022 Tonga Volcanic Plume and Its Initial Evolution Inferred from COSMIC-2 RO Measurements. Atmosphere 2023, 14, 121. https://doi.org/10.3390/atmos14010121

AMA Style

Ravindra Babu S, Lin N-H. Extreme Heights of 15 January 2022 Tonga Volcanic Plume and Its Initial Evolution Inferred from COSMIC-2 RO Measurements. Atmosphere. 2023; 14(1):121. https://doi.org/10.3390/atmos14010121

Chicago/Turabian Style

Ravindra Babu, Saginela, and Neng-Huei Lin. 2023. "Extreme Heights of 15 January 2022 Tonga Volcanic Plume and Its Initial Evolution Inferred from COSMIC-2 RO Measurements" Atmosphere 14, no. 1: 121. https://doi.org/10.3390/atmos14010121

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop