[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (816)

Search Parameters:
Keywords = tropical cyclones

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 6253 KiB  
Article
WRF-ROMS-SWAN Coupled Model Simulation Study: Effect of Atmosphere–Ocean Coupling on Sea Level Predictions Under Tropical Cyclone and Northeast Monsoon Conditions in Hong Kong
by Ngo-Ching Leung, Chi-Kin Chow, Dick-Shum Lau, Ching-Chi Lam and Pak-Wai Chan
Atmosphere 2024, 15(10), 1242; https://doi.org/10.3390/atmos15101242 - 17 Oct 2024
Viewed by 192
Abstract
The Hong Kong Observatory has been using a parametric storm surge model to forecast the rise of sea level due to the passage of tropical cyclones. This model includes an offset parameter to account for the rise in sea level due to other [...] Read more.
The Hong Kong Observatory has been using a parametric storm surge model to forecast the rise of sea level due to the passage of tropical cyclones. This model includes an offset parameter to account for the rise in sea level due to other meteorological factors. By adding the sea level rise forecast to the astronomical tide prediction using the harmonic analysis method, coastal sea level prediction can be produced for the sites with tidal observations, which supports the high water level forecast operation and alert service for risk assessment of sea flooding in Hong Kong. The Coupled Ocean-Atmosphere-Wave-Sediment Transport (COAWST) Modelling System, which comprises the Weather Research and Forecasting (WRF) Model and Regional Ocean Modelling System (ROMS), which in itself is coupled with wave model WaveWatch III and nearshore wave model SWAN, was tested with tropical cyclone cases where there was significant water level rise in Hong Kong. This case study includes two super typhoons, namely Hato in 2017 and Mangkhut in 2018, three cases of the combined effect of tropical cyclone and northeast monsoon, including Typhoon Kompasu in 2021, Typhoon Nesat and Severe Tropical Storm Nalgae in 2022, as well as two cases of monsoon-induced sea level anomalies in February 2022 and February 2023. This study aims to evaluate the ability of the WRF-ROMS-SWAN model to downscale the meteorological fields and the performance of the coupled models in capturing the maximum sea levels under the influence of significant weather events. The results suggested that both configurations could reproduce the sea level variations with a high coefficient of determination (R2) of around 0.9. However, the WRF-ROMS-SWAN model gave better results with a reduced RMSE in the surface wind and sea level anomaly predictions. Except for some cases where the atmospheric model has introduced errors during the downscaling of the ERA5 dataset, bias in the peak sea levels could be reduced by the WRF-ROMS-SWAN coupled model. The study result serves as one of the bases for the implementation of the three-way coupled atmosphere–ocean–wave modelling system for producing an integrated forecast of storm surge or sea level anomalies due to meteorological factors, as well as meteorological and oceanographic parameters as an upgrade to the two-way coupled Operational Marine Forecasting System in the Hong Kong Observatory. Full article
Show Figures

Figure 1

Figure 1
<p>Model domains used in the atmospheric, ocean and wave models, in which ‘SCS_PAC’ domain covers the South China Sea and western North Pacific, ‘NSCS’ the northern part of the South China Sea, ‘SHK’ the south of Hong Kong, and ‘HKW’ the Hong Kong waters.</p>
Full article ">Figure 2
<p>A schematic diagram illustrating the coupling and data flow between different models. Two-way coupling refers to current–wave coupling, and three-way coupling is current–wave–atmosphere coupling.</p>
Full article ">Figure 3
<p>Location of tide observations in the innermost domain ‘HKW’ of the ocean model used in this study.</p>
Full article ">Figure 4
<p>Comparison of TC tracks from ERA5 input for two-way coupling runs (circle) and predicted by WRF in three-way coupling runs (triangle), with HKO TC best-track data (cross) for different TC cases in this study.</p>
Full article ">Figure 5
<p>Comparison of mean sea level pressure and 10 m wind speed from ERA5 input for two-way coupling model simulations (red) and predicted by WRF model in three-way coupling model simulations (blue) against observations (black) at weather station WGL for different weather cases.</p>
Full article ">Figure 6
<p>Scatter plots of 10 m wind speed from ERA5 input for two-way coupling model simulations (red) and predicted by WRF in three-way coupling model simulations (blue) against observations at weather stations WGL for different weather cases.</p>
Full article ">Figure 7
<p>Scatter plots of sea level predicted by two-way (red) and three-way coupling simulations (blue) against observations at various tide stations for different weather cases. Values before 24 h forecast were excluded to eliminate the effect of model spin-up.</p>
Full article ">Figure 8
<p>Scatter plots of sea level bias (in metres) during maxima at different tide stations from the forecasts by three-way coupling model versus two-way coupling model for extreme storm surge cases (square) Hato (orange) and Mangkhut (red), combined effect of TC and monsoon cases (circle) Kompasu (deep blue), Nesat (light blue) and Nalgae (green), and monsoon-induced sea level anomalies cases (triangle) in February 2022 (pink) and February 2023 (violet).</p>
Full article ">Figure 9
<p>Comparison of sea levels predicted by two-way (red) and three-way (blue) coupling simulations against observation (black) and predicted astronomical tide (green) at QUB, TBT and TPK tide stations. The times when SMS or TC warning signals were in force are shaded in light and dark grey, respectively.</p>
Full article ">Figure 10
<p>Comparison of the surface wind field from the ERA5 reanalysis dataset in the two-way coupling simulation (left) and that predicted by the atmospheric model in the three-way coupling simulation (right) when Hato (2017) was near Hong Kong.</p>
Full article ">Figure 11
<p>Comparison of the track of Mangkhut from the three-way coupling simulations using the 0.25-degree resolution ERA5 reanalysis dataset (triangle) for the initial and boundary conditions and that using the 0.125-degree resolution ECMWF forecasts (circle) against the HKO’s TC best track.</p>
Full article ">
20 pages, 5623 KiB  
Article
Tropical Cyclone Wind Direction Retrieval Based on Wind Streaks and Rain Bands in SAR Images
by Zhancai Liu, Hongwei Yang, Weihua Ai, Kaijun Ren, Shensen Hu and Li Wang
Remote Sens. 2024, 16(20), 3837; https://doi.org/10.3390/rs16203837 (registering DOI) - 15 Oct 2024
Viewed by 325
Abstract
Tropical cyclones (TCs) are associated with severe weather phenomena, making accurate wind field retrieval crucial for TC monitoring. SAR’s high-resolution imaging capability provides detailed information for TC observation, and wind speed calculations require wind direction as prior information. Therefore, utilizing SAR images to [...] Read more.
Tropical cyclones (TCs) are associated with severe weather phenomena, making accurate wind field retrieval crucial for TC monitoring. SAR’s high-resolution imaging capability provides detailed information for TC observation, and wind speed calculations require wind direction as prior information. Therefore, utilizing SAR images to retrieve TC wind fields is of significant importance. This study introduces a novel approach for retrieving wind direction from SAR images of TCs through the classification of TC sub-images. The method utilizes a transfer learning-based Inception V3 model to identify wind streaks (WSs) and rain bands in SAR images under TC conditions. For sub-images containing WSs, the Mexican-hat wavelet transform is applied, while for sub-images containing rain bands, an edge detection technique is used to locate the center of the TC eye and subsequently the tangent to the spiral rain bands is employed to determine the wind direction associated with the rain bands. Wind direction retrieval from 10 SAR TC images showed an RMSD of 19.52° and a correlation coefficient of 0.96 when compared with ECMWF and HRD observation wind directions, demonstrating satisfactory consistency and providing highly accurate TC wind directions. These results confirm the method’s potential applications in TC wind direction retrieval. Full article
Show Figures

Figure 1

Figure 1
<p>Example of geophysical phenomena in SAR images. The first row represents wind streaks (G), the second row depicts rain bands (I), and the third row illustrates other geophysical phenomena (A).</p>
Full article ">Figure 2
<p>Flowchart of retrain recognition model based on transfer learning and wind direction retrieval from TCs SAR images.</p>
Full article ">Figure 3
<p>The architecture of transfer learning.</p>
Full article ">Figure 4
<p>The wind direction of rain band locations existing in Northern Hemisphere TCs.</p>
Full article ">Figure 5
<p>Accuracy and loss of training set (blue lines) and validation set (orange lines).</p>
Full article ">Figure 6
<p>Sub-image recognition results of SAR TC images. “G” represents WSs, “I” represents rain bands and “A” denotes other geophysical phenomena.</p>
Full article ">Figure 7
<p>Wind direction retrieval from SAR TC sub-images using 2-D Mexican-hat wavelet transform. (<b>a</b>) SAR sub-image; (<b>b</b>) The result of FFT; (<b>c</b>) The result of Mexico-hat wavelet transformation; (<b>d</b>) The wind direction of the sub-image.</p>
Full article ">Figure 8
<p>The Canny edge detection results for TC Douglas. The NRCS for VV and VH polarizations are presented in (<b>a</b>,<b>d</b>), respectively; the rain band distributions for VV and VH polarizations are shown in (<b>b</b>,<b>e</b>), respectively; the TC eye positions for VV and VH polarizations are depicted in (<b>c</b>,<b>f</b>), respectively.</p>
Full article ">Figure 9
<p>The Canny edge detection results for TC Larry. The NRCS for VV and VH polarizations are presented in (<b>a</b>,<b>d</b>), respectively; the rain band distributions for VV and VH polarizations are shown in (<b>b</b>,<b>e</b>), respectively; the TC eye positions for VV and VH polarizations are depicted in (<b>c</b>,<b>f</b>), respectively.</p>
Full article ">Figure 10
<p>Schematic diagram of wind directions with 180° ambiguity and reference wind direction. For the two predicted wind directions <math display="inline"><semantics> <msub> <mi>θ</mi> <msub> <mi>p</mi> <mn>1</mn> </msub> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>θ</mi> <msub> <mi>p</mi> <mn>2</mn> </msub> </msub> </semantics></math> that are aligned but point in opposite directions, the smaller the <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>θ</mi> </mrow> </semantics></math> calculated relative to the reference wind direction <math display="inline"><semantics> <msub> <mi>θ</mi> <mi>t</mi> </msub> </semantics></math>, the closer it is to the true wind direction.</p>
Full article ">Figure 11
<p>The wind field rotational pattern of TCs. (<b>a</b>) TCs in the Northern Hemisphere. (<b>b</b>) TCs in the Southern Hemisphere.</p>
Full article ">Figure 12
<p>The wind direction retrieval results for TC Douglas, acquired on 25 July 2020. (<b>a</b>) Quick-look from the VV polarized SAR image over TC Douglas; (<b>b</b>) The wind direction retrieval results; (<b>c</b>) The ECMWF wind direction; (<b>d</b>) Comparison of the retrieved wind direction with ECMWF and HRD observation wind direction.</p>
Full article ">Figure 13
<p>The wind direction retrieval results for TC Larry, acquired on 7 September 2021. (<b>a</b>) Quick-look from the VV polarized SAR image over TC Larry; (<b>b</b>) The wind direction retrieval results; (<b>c</b>) The ECMWF wind direction; (<b>d</b>) Comparison of the retrieved wind direction with ECMWF and HRD observation wind direction.</p>
Full article ">Figure 14
<p>Comparison of wind directions retrieved from 10 SAR TCs images with ECMWF reanalysis and HRD observation wind directions.</p>
Full article ">
22 pages, 7116 KiB  
Article
Regional Mean Sea Level Variability Due to Tropical Cyclones: Insights from August Typhoons
by MyeongHee Han, SungHyun Nam and Hak-Soo Lim
J. Mar. Sci. Eng. 2024, 12(10), 1830; https://doi.org/10.3390/jmse12101830 - 14 Oct 2024
Viewed by 458
Abstract
This study investigates the interannual variations in regional mean sea levels (MSLs) of the northeast Asian marginal seas (NEAMS) during August, focusing on the role of typhoon activity from 1993 to 2019. The NEAMS are connected to the Pacific through the East China [...] Read more.
This study investigates the interannual variations in regional mean sea levels (MSLs) of the northeast Asian marginal seas (NEAMS) during August, focusing on the role of typhoon activity from 1993 to 2019. The NEAMS are connected to the Pacific through the East China Sea (ECS) and narrow, shallow straits in the east, where inflow from the southern boundary (ECS), unless balanced by eastern outflow, leads to significant convergence or divergence, as well as subsequent changes in regional MSLs. Satellite altimetry and tide-gauge data reveal that typhoon-induced Ekman transport plays a key role in MSL variability, with increased inflow raising MSLs during active typhoon seasons. In contrast, weak typhoon activity reduces inflow, resulting in lower MSLs. This study’s findings have significant implications for coastal management, as the projected changes in tropical cyclone frequency and intensity due to climate change could exacerbate sea level rise and flooding risks. Coastal communities in the NEAMS region will need to prioritize enhanced flood defenses, early warning systems, and adaptive land use strategies to mitigate these risks. This is the first study to link typhoon frequency directly to NEAMS MSL variability, highlighting the critical role of wind-driven processes in regional sea level changes. Full article
(This article belongs to the Special Issue Air-Sea Interaction and Marine Dynamics)
Show Figures

Figure 1

Figure 1
<p>Domains of the NEAMS (gray shading), including Area A, where typhoon activity was assessed; Areas B and C, where inflow and outflow occur, respectively (black arrows indicate basic flow transport); and Areas D and E, which are related to time series of zonal and meridional winds. Typhoon tracks are superimposed with symbol sizes and color scales according to maximum sustained wind speeds of 17, 33, 43, 49, 58, and 70 m s<sup>−1</sup> for the months of August from 1993 to 2019. ES, YS, BS, ECS, SCS, SO, and PO denote the East Sea (Sea of Japan), Yellow Sea, Bohai Sea, East China Sea, South China Sea, Sea of Okhotsk, and Pacific Ocean, respectively. KS, TAS, TSS, and SS represent the Korea/Tsushima Strait, Taiwan Strait, Tsugaru Strait, and Soya Straits, respectively.</p>
Full article ">Figure 2
<p>Time series of the summer (August) MSLs in the NEAMS region from 1993 to 2019, based on satellite altimetry (red open circles) and tide-gauge observations (black open squares), both adjusted to exclude the inverted barometer effect. The tide-gauge data have been referenced to a common vertical datum to match the MSL with satellite altimetry over the 27-year period. Dashed red and black dotted lines indicate the respective trends for satellite and tide-gauge measurements.</p>
Full article ">Figure 3
<p>Time series of the detrended summer NEAMS MSL anomalies (black filled diamond) derived from satellite altimetry data, SLC by Ekman transport anomaly in Area B (red filled square), and SLC by Ekman transport anomaly differences between Areas B and C (red open triangle), derived from ERA5 data from 1993 to 2019. The correlation coefficients between the NEAMS MSL and SLC by Ekman transport anomaly in Area B and between the NEAMS MSL and SLC by Ekman transport anomaly differences between Areas B and C are both 0.65 (<span class="html-italic">p</span>-value &lt; 0.01).</p>
Full article ">Figure 4
<p>(<b>a</b>) Time series of detrended summer NEAMS MSL anomalies (black solid line) derived from satellite altimetry data and typhoon occurrence in the ECS (red dashed line) from 1993 to 2019. Summers with relatively high MSL changes (<span class="html-italic">Period H</span>; &gt;2 cm, black solid diamonds) are denoted by orange and green shading, while summers with relatively low MSL changes (<span class="html-italic">Period L</span>; &lt;−2 cm) are denoted by black-filled diamonds. During each summer throughout <span class="html-italic">Period H</span>, three or more typhoons passed through Area A (red solid circles), with the exception of the summers of 2001 and 2002 (green shading,) while two or fewer typhoons occurred during each summer of <span class="html-italic">Period L</span> (red open circles) (correlation coefficient = 0.54; <span class="html-italic">p</span>-value &lt; 0.01). Error bar indicates the positive and negative standard deviations of daily detrended summer NEAMS MSL. Time series of detrended summer (<b>b</b>) ES (green triangle, correlation coefficient = 0.49; <span class="html-italic">p</span> = 0.01) and (<b>c</b>) YS (blue square, correlation coefficient = 0.55; <span class="html-italic">p</span>-value &lt; 0.01) MSL anomalies derived from satellite altimetry data from 1993 to 2019. Error bar indicates the positive and negative standard deviations of daily detrended summer ES and YS MSLs.</p>
Full article ">Figure 5
<p>Surface wind stress vectors (black arrows) and typhoon tracks (dotted lines) for (<b>a</b>) composite <span class="html-italic">Period H</span> and (<b>b</b>) composite <span class="html-italic">Period L</span>. The legend for dotted lines can be found in the upper-left corner with criteria for maximum wind speed in m s<sup>−1</sup> similar to <a href="#jmse-12-01830-f001" class="html-fig">Figure 1</a>. Area A and coastlines are denoted by red dashed box and gray lines, respectively.</p>
Full article ">Figure 6
<p>Surface wind stress vectors (black arrows) and typhoon tracks (dotted lines) in (<b>a</b>) 2001 and (<b>b</b>) 2002. Area A and coastlines are denoted by a red dashed box and gray lines, respectively.</p>
Full article ">Figure 7
<p>Time series of hourly zonal and meridional winds averaged over (<b>a</b>) Area D, summer 2001, and (<b>b</b>) Area E, summer 2002, are plotted using blue squares and red circles, respectively. Monthly mean zonal and meridional winds and zero wind speed lines are represented by the dotted blue and red horizontal lines and solid black lines, respectively. Gray shading represents typhoons Pabuk and Rusa during 19–20 August 2001 and 30–31 August 2002, respectively.</p>
Full article ">Figure 8
<p>(<b>a</b>) Correlation map of MSL atmospheric pressure changes related to NEAMS MSL changes during August 1993–2019. In (<b>a</b>), contour intervals are 0.05, and correlations with confidence levels &lt; 90% are discarded. Schematics demonstrate the ocean inflow and outflow (black filled arrows) of the NEAMS and NEAMS MSL anomalies (red and blue) in August driven by (<b>b</b>,<b>c</b>) convergence and (<b>d</b>) divergence during <span class="html-italic">Periods H</span> and <span class="html-italic">L</span>, respectively, related to inflow Ekman transport (black open arrows) induced by wind (blue shaded arrows). In (<b>b</b>), the large L represents the typhoon center. Composite sea level (ADT) anomalies for (<b>b</b>) <span class="html-italic">Period H</span>, (<b>c</b>) summers of 2001 and 2002, and (<b>d</b>) <span class="html-italic">Period L</span> in August obtained from satellite altimeters are indicated using colors.</p>
Full article ">Figure 9
<p>Time series of detrended summer NEAMS MSL anomalies derived from satellite altimetry data (black open diamonds, NEAMS) by heat transport (red open circles, HT) and salt transport (blue open squares, ST) differences in anomalies over the ESC (Area B) and the Tsugaru and Soya Straits, by net surface heat flux (magenta asterisk, HF) derived from the Ocean Reanalysis System 5 (ORAS5), and by subtracting MSL by HT, ST, and HF from altimetry MSL (green open triangle, mass transport (MT)) from 1993 to 2019.</p>
Full article ">
27 pages, 7826 KiB  
Article
Comprehensive Comparison of Seven Widely-Used Planetary Boundary Layer Parameterizations in Typhoon Mangkhut Intensification Simulation
by Lei Ye, Yubin Li, Ping Zhu, Zhiqiu Gao and Zhihua Zeng
Atmosphere 2024, 15(10), 1182; https://doi.org/10.3390/atmos15101182 - 30 Sep 2024
Viewed by 463
Abstract
Numerical experiments using the WRF model were conducted to analyze the sensitivity of Typhoon Mangkhut intensification simulations to seven widely used planetary boundary layer (PBL) parameterization schemes, including YSU, MYJ, QNSE, MYNN2, MYNN3, ACM2, and BouLac. The results showed that all simulations generally [...] Read more.
Numerical experiments using the WRF model were conducted to analyze the sensitivity of Typhoon Mangkhut intensification simulations to seven widely used planetary boundary layer (PBL) parameterization schemes, including YSU, MYJ, QNSE, MYNN2, MYNN3, ACM2, and BouLac. The results showed that all simulations generally reproduced the tropical cyclone (TC) track and intensity, with YSU, QNSE, and BouLac schemes better capturing intensification processes and closely matching observed TC intensity. In terms of surface layer parameterization, the QNSE scheme produced the highest Ck/Cd ratio, resulting in stronger TC intensity based on Emanuel’s potential intensity theory. In terms of PBL parameterization, the YSU and BouLac schemes, with the same revised MM5 surface layer scheme, simulated weaker turbulent diffusivity Km and shallower mixing height, leading to stronger TC intensity. During the intensification period, the BouLac, YSU, and QNSE PBL schemes exhibited stronger tangential wind, radial inflow within the boundary layer, and updraft around the eye wall, consistent with TC intensity results. Both PBL and surface layer parameterization significantly influenced simulated TC intensity. The QNSE scheme, with the largest Ck/Cd ratio, and the YSU and BouLac schemes, with weaker turbulent diffusivity, generated stronger radial inflow, updraft, and warm core structures, contributing to higher storm intensity. Full article
Show Figures

Figure 1

Figure 1
<p>Model domain configuration and the best track of Mangkhut from JTWC during 0000 UTC 7th to 0000 UTC 13 September 2018. Colored dots indicate the 6 hourly TC track, and colors indicate the categories of TC intensity (DB: disturbance; TD: tropical depression; TS: tropical storm; TY: typhoon; ST: super typhoon). And d01, d02, and d03 represent the domain 01, 02, and 03, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) TC track, (<b>b</b>) time series of track error (unit: km), (<b>c</b>) minimum central sea level pressure (MSLP) (unit: hPa), and (<b>d</b>) maximum sustained wind speed (VMAX) (unit: m s<sup>−1</sup>) from the JTWC best track data, ERA5 reanalysis data, and the numerical simulations from 0600 UTC 7th to 0000 UTC 13 September 2018.</p>
Full article ">Figure 3
<p>Time series of area-averaged (<b>a</b>) surface enthalpy flux (unit: W m<sup>−2</sup>) and (<b>b</b>) surface momentum flux (unit: kg m<sup>−1</sup> s<sup>−2</sup>) from TC center to 200 km radius from 0600 UTC 7th to 0000 UTC 13 September 2018.</p>
Full article ">Figure 4
<p>The variation of surface exchange coefficients for (<b>a1</b>–<b>g1</b>) momentum <span class="html-italic">C<sub>d</sub></span> and (<b>a2</b>–<b>g2</b>) heat <span class="html-italic">C<sub>k</sub></span> with 10 m wind speed at 1200 UTC 12 September 2018; numerical simulations (thick gray dots) compared with data from Large and Pond [<a href="#B69-atmosphere-15-01182" class="html-bibr">69</a>], Powell et al. [<a href="#B70-atmosphere-15-01182" class="html-bibr">70</a>], Donelan et al. [<a href="#B71-atmosphere-15-01182" class="html-bibr">71</a>], Black et al. [<a href="#B72-atmosphere-15-01182" class="html-bibr">72</a>], Zhang et al. [<a href="#B73-atmosphere-15-01182" class="html-bibr">73</a>], Haus et al. [<a href="#B74-atmosphere-15-01182" class="html-bibr">74</a>], Bell et al. [<a href="#B75-atmosphere-15-01182" class="html-bibr">75</a>], and Richter et al. [<a href="#B76-atmosphere-15-01182" class="html-bibr">76</a>].</p>
Full article ">Figure 5
<p>The variation of surface exchange coefficient ratio (ratio of <span class="html-italic">C<sub>k</sub></span>/<span class="html-italic">C<sub>d</sub></span>) with 10 m wind speed at 1200 UTC 12 September 2018 from different numerical simulations.</p>
Full article ">Figure 6
<p>Height–radius distribution of azimuthally averaged turbulent diffusivity for momentum (unit: m<sup>2</sup> s<sup>−1</sup>) from (<b>a1</b>–<b>a4</b>) YSU, (<b>b1</b>–<b>b4</b>) MYJ, (<b>c1</b>–<b>c4</b>) QNSE, (<b>d1</b>–<b>d4</b>) MYNN2, (<b>e1</b>–<b>e4</b>) MYNN3, (<b>f1</b>–<b>f4</b>) ACM2, and (<b>g1</b>–<b>g4</b>) BouLac, at (<b>a1</b>–<b>g1</b>) 0600 UTC 7th, (<b>a2</b>–<b>g2</b>) 0000 UTC 9th, (<b>a3</b>–<b>g3</b>) 0000 UTC 11th, and (<b>a4</b>–<b>g4</b>) 1200 UTC 12 September 2018.</p>
Full article ">Figure 7
<p>Height–radius distribution of azimuthally averaged turbulent diffusivity for momentum (unit: m<sup>2</sup> s<sup>−1</sup>) from (<b>a1</b>–<b>a4</b>) YSU, (<b>b1</b>–<b>b4</b>) MYJ, (<b>c1</b>–<b>c4</b>) QNSE, (<b>d1</b>–<b>d4</b>) MYNN2, (<b>e1</b>–<b>e4</b>) MYNN3, (<b>f1</b>–<b>f4</b>) ACM2, and (<b>g1</b>–<b>g4</b>) BouLac, at (<b>a1</b>–<b>g1</b>) 0300 UTC 7th, (<b>a2</b>–<b>g2</b>) 0400 UTC 7th, (<b>a3</b>–<b>g3</b>) 0500 UTC 7th, and (<b>a4</b>–<b>g4</b>) 0600 UTC 7 September 2018.</p>
Full article ">Figure 8
<p>Comparison of turbulent diffusivity for momentum (<span class="html-italic">K<sub>m</sub></span>) between numerical simulations and observations at 0600 UTC 7th (in red), 0000 UTC 9th (in green), 0000 UTC 11th (in yellow), and 1200 UTC 12th (in blue) September 2018. In (<b>a</b>–<b>g</b>), OBS (in black) represents observations from Zhang et al. [<a href="#B80-atmosphere-15-01182" class="html-bibr">80</a>], which are based on in situ data at 450 m altitude.</p>
Full article ">Figure 9
<p>Height–radius distribution of azimuthally averaged tangential wind (unit: m s<sup>−1</sup>) from (<b>a1</b>–<b>a4</b>) ERA5, (<b>b1</b>–<b>b4</b>) YSU, (<b>c1</b>–<b>c4</b>) MYJ, (<b>d1</b>–<b>d4</b>) QNSE, (<b>e1</b>–<b>e4</b>) MYNN2, (<b>f1</b>–<b>f4</b>) MYNN3, (<b>g1</b>–<b>g4</b>) ACM2, and (<b>h1</b>–<b>h4</b>) BouLac at (<b>a1</b>–<b>h1</b>) 0600 UTC 7th, (<b>a2</b>–<b>h2</b>) 0000 UTC 9th, (<b>a3</b>–<b>h3</b>) 0000 UTC 11th, and (<b>a4</b>–<b>h4</b>) 1200 UTC 12 September 2018.</p>
Full article ">Figure 10
<p>As in <a href="#atmosphere-15-01182-f009" class="html-fig">Figure 9</a> but for radial wind (unit: m s<sup>−1</sup>), with (<b>a1</b>–<b>a4</b>) ERA5, (<b>b1</b>–<b>b4</b>) YSU, (<b>c1</b>–<b>c4</b>) MYJ, (<b>d1</b>–<b>d4</b>) QNSE, (<b>e1</b>–<b>e4</b>) MYNN2, (<b>f1</b>–<b>f4</b>) MYNN3, (<b>g1</b>–<b>g4</b>) ACM2, and (<b>h1</b>–<b>h4</b>) BouLac at (<b>a1</b>–<b>h1</b>) 0600 UTC 7th, (<b>a2</b>–<b>h2</b>) 0000 UTC 9th, (<b>a3</b>–<b>h3</b>) 0000 UTC 11th, and (<b>a4</b>–<b>h4</b>) 1200 UTC 12 September 2018.</p>
Full article ">Figure 11
<p>As in <a href="#atmosphere-15-01182-f009" class="html-fig">Figure 9</a> but for upward vertical wind (unit: m s<sup>−1</sup>), with (<b>a1</b>–<b>a4</b>) ERA5, (<b>b1</b>–<b>b4</b>) YSU, (<b>c1</b>–<b>c4</b>) MYJ, (<b>d1</b>–<b>d4</b>) QNSE, (<b>e1</b>–<b>e4</b>) MYNN2, (<b>f1</b>–<b>f4</b>) MYNN3, (<b>g1</b>–<b>g4</b>) ACM2, and (<b>h1</b>–<b>h4</b>) BouLac at (<b>a1</b>–<b>h1</b>) 0600 UTC 7th, (<b>a2</b>–<b>h2</b>) 0000 UTC 9th, (<b>a3</b>–<b>h3</b>) 0000 UTC 11th, and (<b>a4</b>–<b>h4</b>) 1200 UTC 12 September 2018.</p>
Full article ">Figure 12
<p>Height–radius distribution of azimuthally averaged equivalent potential temperature (unit: K) from (<b>a1</b>–<b>a4</b>) YSU, (<b>b1</b>–<b>b4</b>) MYJ, (<b>c1</b>–<b>c4</b>) QNSE, (<b>d1</b>–<b>d4</b>) MYNN2, (<b>e1</b>–<b>e4</b>) MYNN3, (<b>f1</b>–<b>f4</b>) ACM2, and (<b>g1</b>–<b>g4</b>) BouLac, at (<b>a1</b>–<b>g1</b>) 0000 UTC 11th, (<b>a2</b>–<b>g2</b>) 1200 UTC 11th, (<b>a3</b>–<b>g3</b>) 0000 UTC 12th, and (<b>a4</b>–<b>g4</b>) 1200 UTC 12 September 2018.</p>
Full article ">Figure 13
<p>Height–radius distribution of azimuthally averaged specific humidity difference (unit: g kg<sup>−1</sup>) between YSU and (<b>a</b>) MYJ, (<b>b</b>) QNSE, (<b>c</b>) MYNN2.5, (<b>d</b>) MYNN3, (<b>e</b>) ACM2, and (<b>f</b>) BouLac at (<b>1</b>) 0600 UTC 7, (<b>2</b>) 0000 UTC 9, (<b>3</b>) 0000 UTC 11, and (<b>4</b>) 1200 UTC 12, September 2018.</p>
Full article ">
16 pages, 9032 KiB  
Article
Assessing Vulnerability to Cyclone Hazards in the World’s Largest Mangrove Forest, The Sundarbans: A Geospatial Analysis
by Mohammed, Fahmida Sultana, Ariful Khan, Sohag Ahammed, Md. Shamim Reza Saimun, Md Saifuzzaman Bhuiyan, Sanjeev K. Srivastava, Sharif A. Mukul and Mohammed A. S. Arfin-Khan
Forests 2024, 15(10), 1722; https://doi.org/10.3390/f15101722 - 29 Sep 2024
Viewed by 849
Abstract
The Sundarbans is the world’s largest contiguous mangrove forest with an area of about 10,000 square kilometers and shared between Bangladesh and India. This world-renowned mangrove forest, located on the lower Ganges floodplain and facing the Bay of Bengal, has long served as [...] Read more.
The Sundarbans is the world’s largest contiguous mangrove forest with an area of about 10,000 square kilometers and shared between Bangladesh and India. This world-renowned mangrove forest, located on the lower Ganges floodplain and facing the Bay of Bengal, has long served as a crucial barrier, shielding southern coastal Bangladesh from cyclone hazards. However, the Sundarbans mangrove ecosystem is now increasingly threatened by climate-induced hazards, particularly tropical cyclones originating from the Indian Ocean. To assess the cyclone vulnerability of this unique ecosystem, using geospatial techniques, we analyzed the damage caused by past cyclones and the subsequent recovery across three salinity zones, i.e., Oligohaline, Mesohaline, and Polyhaline. Our study also examined the relationship between cyclone intensity with the extent of damage and forest recovery. The findings of our study indicate that the Polyhaline zone, the largest in terms of area and with the lowest elevation, suffered the most significant damage from cyclones in the Sundarbans region, likely due to its proximity to the most cyclone paths. A correlation analysis revealed that cyclone damage positively correlated with wind speed and negatively correlated with the distance of landfall from the center of the Sundarbans. With the expectation of more extreme weather events in the near future, the Sundarbans mangrove forest faces a potentially devastating outlook unless both natural protection processes and human interventions are undertaken to safeguard this critical ecosystem. Full article
(This article belongs to the Special Issue Biodiversity, Health, and Ecosystem Services of Mangroves)
Show Figures

Figure 1

Figure 1
<p>Path of the studied cyclones in the Bangladesh Sundarbans and the coverage of three saline zones (i.e., Polyhaline, Mesohaline, and Oligohaline) in the area. The red dot mark indicates the center point of the Sundarbans mangrove forest.</p>
Full article ">Figure 2
<p>Flowchart of the working procedures of cyclone damage and recovery analysis.</p>
Full article ">Figure 3
<p>Transition of the four NDVI classes after the occurrence of cyclones and their transitioned amount in square kilometers (km<sup>2</sup>). C1, C2, C3, and C4 represent Class 1, Class 2, Class 3, and Class 4 of pre-cyclone and post-cyclone NDVI classes, respectively. The total amount of area (in km<sup>2</sup>) of each class of pre-cyclone and post-cyclone NDVI is given under the class identifier (C1, C2, C3, and C4). Gray-colored inset boxes represent the unchanged area of each NDVI class after cyclone occurrence. A distinct colored number at the end of each line illustrates the amount of area (in km<sup>2</sup>) shifted to other classes.</p>
Full article ">Figure 4
<p>Transitioned area under different NDVI classes according to salinity zones in Bangladesh Sundarbans after the occurrence of 12 studied cyclones, where (<b>a</b>) represents the extent of the three saline zones.</p>
Full article ">Figure 5
<p>Four NDVI classes of both, pre-cyclone NDVI and post-cyclone NDVI, for the three saline zones of Bangladesh Sundarbans, where (<b>a</b>) represents the extent of the three saline zones.</p>
Full article ">Figure 6
<p>Cyclone-damaged areas in different saline zones are sorted by the year the cyclone occurred.</p>
Full article ">Figure 7
<p>Cyclone-recovered areas in different saline zones sorted by the year of cyclone occurred.</p>
Full article ">Figure 8
<p>Mean saline zone-wise values of (<b>a</b>) the damaged area and (<b>b</b>) the recovered area. Here, similar letters were used to represent no significant difference among saline zone-wise damaged and recovered areas, respectively.</p>
Full article ">Figure 9
<p>Results of correlation analysis of damage percentage (%) with cyclone variables, distance, and wind speed shown by (<b>a</b>) correlation matrix showing correlation coefficient values among different values; (<b>b</b>) results of Pearson’s correlation among damage percentage and wind speed (<b>c</b>); results of Pearson’s correlation among damage percentage and distance.</p>
Full article ">
18 pages, 5148 KiB  
Article
Trends and Periodicities of Tropical Cyclone Frequencies and the Correlations with Ocean Drivers
by Guoyou Li, Huabin Shi and Zhiguo He
J. Mar. Sci. Eng. 2024, 12(10), 1707; https://doi.org/10.3390/jmse12101707 - 26 Sep 2024
Viewed by 515
Abstract
This study presents a comprehensive analysis on the variations in the tropical cyclone (TC) frequencies during 1980–2021, including the linear trends, periodicities, and their variabilities on both global and basin-wise scales. An increasing trend in the annual number of global TCs is identified, [...] Read more.
This study presents a comprehensive analysis on the variations in the tropical cyclone (TC) frequencies during 1980–2021, including the linear trends, periodicities, and their variabilities on both global and basin-wise scales. An increasing trend in the annual number of global TCs is identified, with a significant rising trend in the numbers of tropical storms (maximum sustained wind 35 ktsUmax<64 kts) and intense typhoons (Umax96 kts) and a deceasing trend for weak typhoons (64 ktsUmax<96 kts). There is no statistically significant trend shown in the global Accumulated Cyclone Energy (ACE). On a regional scale, the Western North Pacific (WNP) and Eastern North Pacific (ENP) are the regions of the first- and second-largest numbers of TCs, respectively, while the increased TC activity in the North Atlantic (NA) contributes the most to the global increase in TCs. It is revealed in the wavelet transformation for periodicity analysis that the variations in the annual number of TCs with different intensities mostly show an inter-annual period of 3–7 years and an inter-decadal one of 10–13 years. The inter-annual and inter-decadal periods are consistent with those in the ENSO-related ocean drivers (via the Niño 3.4 index), Southern Oscillation Index (SOI), and Inter-decadal Pacific Oscillation (IPO) index. The inter-decadal variation in 10–13 years is also observed in the North Atlantic Oscillation (NAO) index. The Tropical North Atlantic (TNA) index and Atlantic Multi-decadal Oscillation (AMO) index, on the other hand, present the same inter-annual period of 7–10 years as that in the frequencies of all the named TCs in the NA. Further, the correlations between TC frequencies and ocean drivers are also quantified using the Pearson correlation coefficient. These findings contribute to an enhanced understanding of TC activity, thereby facilitating efforts to predict particular TC activity and mitigate the inflicted damage. Full article
(This article belongs to the Section Physical Oceanography)
Show Figures

Figure 1

Figure 1
<p>Locations of defined ocean basins. NIO: North Indian Ocean; WNP: Western North Pacific Ocean; ENP: Eastern North Pacific Ocean; NA: North Atlantic Ocean; SIO: South Indian Ocean; SP: South Pacific Ocean.</p>
Full article ">Figure 2
<p>Time series (solid lines) and its linear trend (dashed lines) of (<b>a</b>) annual numbers of all named TCs; the annual numbers as well as proportions of (<b>b</b>) TSs, (<b>c</b>) WTYs, and (<b>d</b>) ITYs on the globe scale. In subfigures (<b>b</b>–<b>d</b>), the blue lines represent the series of annual numbers, while the red ones are for the proportions. The shaded areas represent the 95% confidence interval in the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-value for the statistical significance of the linear trend is included.</p>
Full article ">Figure 3
<p>Time series (solid line) and the linear trend (dashed line) of annual values of global ACE. The shaded area represents the 95% confidence interval of the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-value for the statistical significance of the linear trend is included.</p>
Full article ">Figure 4
<p>(<b>a</b>) Linear trend in the annual numbers of all named TCs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of all named TC numbers generated in hemispheres and six ocean basins in global total TCs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 5
<p>(<b>a</b>) Linear trend in the annual numbers of TSs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. Two asterisks above the bars indicate statistical significance levels of the linear trend of 5%; (<b>b</b>) the 42-year-averaged proportions (columns) of TS numbers generated in hemispheres and six ocean basins in global total TSs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 6
<p>(<b>a</b>) Linear trend in the annual numbers of WTYs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of WTY numbers generated in hemispheres and six ocean basins in global total WTYs and their linear trends (numbers above the columns, in the unit of %/year). One and two asterisks above the numbers indicate statistical significance levels of the linear trend of 10% and 5%, respectively.</p>
Full article ">Figure 7
<p>(<b>a</b>) Linear trend in the annual numbers of ITYs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of ITY numbers generated in hemispheres and six ocean basins in global total ITYs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 8
<p>(<b>a</b>) Linear trend in the annual values of ACE in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. Two asterisks above the bars indicate statistical significance levels of the linear trend of 5%; (<b>b</b>) the 42-year-averaged proportions (columns) of ACE in hemispheres and six ocean basins in global ACE and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 9
<p>Wavelet coefficient magnitude scalograms and power spectrums of annual numbers of all named TCs across the entire globe, the two hemispheres, and six ocean basins. The white dashed lines in magnitude scalograms indicate the cone of influence.</p>
Full article ">Figure 10
<p>Time series (solid lines) and its linear trend (dashed lines) of (<b>a</b>) Niño 3.4 index, (<b>b</b>) SOI, (<b>c</b>) IPO index, (<b>d</b>) IOD index, (<b>e</b>) TNA index, (<b>f</b>) NAO index, and (<b>g</b>) AMO index. The shaded areas represent the 95% confidence interval in the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-values for the statistical significance of the linear trend are included.</p>
Full article ">Figure 11
<p>Wavelet coefficient magnitude scalograms and power spectrums of annual values of (<b>a</b>) Niño 3.4 index, (<b>b</b>) SOI, (<b>c</b>) IPO index, (<b>d</b>) IOD index, (<b>e</b>) TNA index, (<b>f</b>) NAO index, and (<b>g</b>) AMO index. The white dashed lines in magnitude scalograms indicate the cone of influence.</p>
Full article ">
24 pages, 3135 KiB  
Review
Current Status of Remote Sensing for Studying the Impacts of Hurricanes on Mangrove Forests in the Coastal United States
by Abhilash Dutta Roy, Daria Agnieszka Karpowicz, Ian Hendy, Stefanie M. Rog, Michael S. Watt, Ruth Reef, Eben North Broadbent, Emma F. Asbridge, Amare Gebrie, Tarig Ali and Midhun Mohan
Remote Sens. 2024, 16(19), 3596; https://doi.org/10.3390/rs16193596 - 26 Sep 2024
Viewed by 969
Abstract
Hurricane incidents have become increasingly frequent along the coastal United States and have had a negative impact on the mangrove forests and their ecosystem services across the southeastern region. Mangroves play a key role in providing coastal protection during hurricanes by attenuating storm [...] Read more.
Hurricane incidents have become increasingly frequent along the coastal United States and have had a negative impact on the mangrove forests and their ecosystem services across the southeastern region. Mangroves play a key role in providing coastal protection during hurricanes by attenuating storm surges and reducing erosion. However, their resilience is being increasingly compromised due to climate change through sea level rises and the greater intensity of storms. This article examines the role of remote sensing tools in studying the impacts of hurricanes on mangrove forests in the coastal United States. Our results show that various remote sensing tools including satellite imagery, Light detection and ranging (LiDAR) and unmanned aerial vehicles (UAVs) have been used to detect mangrove damage, monitor their recovery and analyze their 3D structural changes. Landsat 8 OLI (14%) has been particularly useful in long-term assessments, followed by Landsat 5 TM (9%) and NASA G-LiHT LiDAR (8%). Random forest (24%) and linear regression (24%) models were the most common modeling techniques, with the former being the most frequently used method for classifying satellite images. Some studies have shown significant mangrove canopy loss after major hurricanes, and damage was seen to vary spatially based on factors such as proximity to oceans, elevation and canopy structure, with taller mangroves typically experiencing greater damage. Recovery rates after hurricane-induced damage also vary, as some areas were seen to show rapid regrowth within months while others remained impacted after many years. The current challenges include capturing fine-scale changes owing to the dearth of remote sensing data with high temporal and spatial resolution. This review provides insights into the current remote sensing applications used in hurricane-prone mangrove habitats and is intended to guide future research directions, inform coastal management strategies and support conservation efforts. Full article
Show Figures

Figure 1

Figure 1
<p>PRISMA workflow representing the systematic literature review process.</p>
Full article ">Figure 2
<p>Applications of remote sensing for studying impacts of hurricanes on mangroves.</p>
Full article ">Figure 3
<p>Coastal southeastern United States showing some locations where studies on hurricane impact on mangroves were carried out, that included (<b>A</b>) Everglades National Park, Florida, (<b>B</b>) Florida Keys, (<b>C</b>) Port Fourchon, Louisiana, (<b>D</b>) (Inset): Puerto Rico.</p>
Full article ">Figure 4
<p>The regional frequency of remote sensing based peer-reviewed articles published on studying impacts of hurricanes on mangroves in the United States from January 2010 to September 2024.</p>
Full article ">Figure 5
<p>Percentage breakdown of sensors used for studying impacts of hurricanes on mangroves in the United States from January 2010 to September 2024.</p>
Full article ">Figure 6
<p>Data analysis methods used to study the impacts of hurricanes on mangroves in the United States from January 2010 to September 2024.</p>
Full article ">
19 pages, 7507 KiB  
Article
Spatiotemporal Climatology of Georgia Tropical Cyclones and Associated Rainfall
by Reilly Corkran, Jill Trepanier and Vincent Brown
J. Mar. Sci. Eng. 2024, 12(10), 1693; https://doi.org/10.3390/jmse12101693 - 24 Sep 2024
Viewed by 862
Abstract
Tropical cyclones (TCs), often characterized by high wind speeds and heavy rainfall, cause widespread devastation, affecting millions of people and leading to economic losses worldwide. TC-specific research in Georgia is scarce, likely due to the minimal geographical extent of its coast and the [...] Read more.
Tropical cyclones (TCs), often characterized by high wind speeds and heavy rainfall, cause widespread devastation, affecting millions of people and leading to economic losses worldwide. TC-specific research in Georgia is scarce, likely due to the minimal geographical extent of its coast and the infrequency of direct landfalls. Research on Georgia TCs does not account for storms that make landfall in other southeastern states (e.g., Florida) and continue north, northeast, or northwest into Georgia. This study used the North Atlantic Basin hurricane database (HURDAT2) to quantify the spatiotemporal patterns of direct and indirect landfalling of Georgia tropical cyclones (>16 ms−1) from 1851 to 2021. TC-induced rainfall was also quantified using rainfall data (nClimGrid-Daily and nClimGrid) from 1951 to 2021 to estimate the proportion of Georgia’s total annual and monthly rainfall attributed to TCs. A multi-methodological approach, incorporating statistics and mapping, is employed to assess the trends of Georgia’s tropical cyclones and the associated rainfall. The study analyzed 113 TCs and found that, on average, less than one TC annually (x¯ = 0.66) traverses the state. September averaged the highest percentage (25%) of TC-induced rainfall, followed by October (14%), and August (13%). This pattern aligns with the TC season, with the highest frequency of TCs occurring in September (n = 35), followed by August (n = 25), and October (n = 18). We found that 10% of tropical storms make landfall on the coastline, while the remaining 91% enter Georgia by making landfall in Florida (92%), Louisiana (7%), or South Carolina (1%) first. A threat of TCs during the peak of the season emphasizes the importance of heightened awareness, increased planning practices, and resource allocation during these periods to protect Georgia’s history and natural beauty, and its residents. Full article
(This article belongs to the Special Issue Coastal Disaster Assessment and Response)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Track map of direct landfalling tropical cyclones on the coastline of Georgia from 1851 to 2021; (<b>b</b>) track map of indirect landfalling tropical cyclones in the state of Georgia from 1851 to 2021.</p>
Full article ">Figure 2
<p>Heat map of first tropical cyclone entry points into Georgia, 1851–2021.</p>
Full article ">Figure 3
<p>Heat map of tropical cyclone exit points along the Georgia border, 1851–2021.</p>
Full article ">Figure 4
<p>Georgia’s annual tropical cyclone occurrence (1851–2021) categorized by intensity.</p>
Full article ">Figure 5
<p>The seasonality of tropical cyclones within the Georgia state boundary separated by intensity.</p>
Full article ">Figure 6
<p>Distribution of tropical cyclone maximum intensity within the state of Georgia from 1851 to 2021.</p>
Full article ">Figure 7
<p>Tropical cyclone tracks used to find the annual TC precipitation in Georgia from 1951 to 2021.</p>
Full article ">Figure 8
<p>The five geographic regions of Georgia with county delineation and the Coastal Plain divided into the upper and lower regions.</p>
Full article ">Figure 9
<p>Annual TC precipitation percentage in Georgia from 1951 to 2021.</p>
Full article ">Figure 10
<p>TC Tracks that occurred from May to November in Georgia from 1951 to 2021; monthly average rainfall data attributed to tropical cyclones from May to November.</p>
Full article ">Figure 11
<p>Average TC precipitation percentage for May–November in Georgia from 1951 to 2021.</p>
Full article ">
13 pages, 1606 KiB  
Article
Phylogeography of Coccoloba uvifera (Polygonaceae) Sampled across the Caribbean Basin
by Danny J. Gustafson, Logan A. Dix, Derek P. Webster, Benjamin K. Scott, Isabella E. Gustafson, Aidan D. Farrell and Daniel M. Koenemann
Diversity 2024, 16(9), 562; https://doi.org/10.3390/d16090562 - 8 Sep 2024
Viewed by 590
Abstract
Coccoloba uvifera L. (seagrape) is a primarily dioecious neotropical tree species which often grows in the beach–forest transitional ecotone of coastal strand vegetation. We used five maternally inherited non-coding chloroplast regions to characterize the phytogeography of C. uvifera collected across the Caribbean Basin [...] Read more.
Coccoloba uvifera L. (seagrape) is a primarily dioecious neotropical tree species which often grows in the beach–forest transitional ecotone of coastal strand vegetation. We used five maternally inherited non-coding chloroplast regions to characterize the phytogeography of C. uvifera collected across the Caribbean Basin and Florida. Bayesian analysis revealed divergence between the Aruba–Trinidad–Tobago–Antigua–Jamaica island group and the continental Belize–Florida–US Virgin Islands (USVI) group at 1.78 million years before present (mybp), divergence between the Belize and Florida–USVI groups at 1.08 mybp, and a split of Antigua–Jamaica from Aruba–Trinidad–Tobago at 0.217 mybp. Haplotype network analysis supports the three clades, with the island group possessing the oldest haplotype. Based on geology and proximity, these clades correspond to South American (oldest), Central American, and North American (most recent). Coccoloba uvifera demographic expansion occurred during the Pleistocene epoch and peaked near the end of the last glacial maximum (ca. 0.026–0.019 mybp) when the global sea levels were 125 m lower than today. Our findings also reveal that tropical cyclones, which often impact coastal strand vegetation, did not affect genetic diversity. However, there was a positive association between latitude and the average number of substitutions, further enriching our understanding of the species’ phytogeography. Full article
(This article belongs to the Section Biogeography and Macroecology)
Show Figures

Figure 1

Figure 1
<p>Geographic distribution of 51 <span class="html-italic">Coccoloba uvifera</span> plants collected from nine locations across the Caribbean Basin and Florida. Twenty-six haplotypes were identified based on five concatenated non-coding chloroplast regions with a total length of approximately 4650 bps.</p>
Full article ">Figure 2
<p>TCS haplotype maximum parsimony network analysis of 26 haplotypes based on the sequence of chloroplast non-coding regions from 51 <span class="html-italic">Coccoloba uvifera</span> plants sampled across the Caribbean Basin. The size of the circle reflects the number of individuals that share this haplotype. Color represents the source population, white nodes represent hypothesized unsampled haplotypes, and hash marks indicate the number of mutational differences.</p>
Full article ">Figure 3
<p>Spherical phylogeography coalescent analysis of <span class="html-italic">Coccoloba uvifera</span>, using bModelTest evolutionary model averaging, indicated divergence between the Central American (Belize), North America (Florida), and U.S. Virgin Islands (USVI) group from the Aruba, Antigua, Trinidad, Tobago, and Jamaica island group approximately 1.78 mybp. The Belize group diverged from the Florida–USVI group around 1.08 mybp. The Antigua–Jamaica island group diverged from the Trinidad, Tobago, and Aruba island group by around 0.217 mybp. Bars indicate the 95% HPD intervals for the divergence times. The red arrow indicates the last glacial maximum (LGM).</p>
Full article ">Figure 4
<p>Coalescent Bayesian Skyline analysis of 51 <span class="html-italic">Coccoloba uvifera</span> plants sampled across the Caribbean Basin, based on five non-coding chloroplast regions. Effective population size increases from 1 million individuals (0.324 mybp) to 20.2 million individuals (0.028 mybp). Solid line represents the mean effective population size and dashed lines represent the 95% highest posterior density (HPD). The time scale is thousands of years before present (Kybp).</p>
Full article ">
17 pages, 4369 KiB  
Article
A Multi-Hazard Approach to Climate Migration: Testing the Intersection of Climate Hazards, Population Change, and Location Desirability from 2000 to 2020
by Zachary M. Hirsch, Jeremy R. Porter, Jasmina M. Buresch, Danielle N. Medgyesi, Evelyn G. Shu and Matthew E. Hauer
Climate 2024, 12(9), 140; https://doi.org/10.3390/cli12090140 - 7 Sep 2024
Viewed by 1056
Abstract
Climate change intensifies the frequency and severity of extreme weather events, profoundly altering demographic landscapes globally and within the United States. This study investigates their impact on migration patterns, using propensity score matching and LASSO techniques within a larger regression modeling framework. Here, [...] Read more.
Climate change intensifies the frequency and severity of extreme weather events, profoundly altering demographic landscapes globally and within the United States. This study investigates their impact on migration patterns, using propensity score matching and LASSO techniques within a larger regression modeling framework. Here, we analyze historical population trends in relation to climate risk and exposure metrics for various hazards. Our findings reveal nuanced patterns of climate-induced population change, including “risky growth” areas where economic opportunities mitigate climate risks, sustaining growth in the face of observed exposure; “tipping point” areas where the amenities are slowly giving way to the disamenity of escalating hazards; and “Climate abandonment” areas experiencing exacerbated out-migration from climate risks, compounded by other out-migration market factors. Even within a single county, these patterns vary significantly, underscoring the importance of localized analyses. Projecting population impacts due to climate risk to 2055, flood risks are projected to impact the largest percentage of areas (82.6%), followed by heatwaves (47.4%), drought (46.6%), wildfires (32.7%), wildfire smoke (21.7%), and tropical cyclone winds (11.1%). The results underscore the importance of understanding hyperlocal patterns of risk and change in order to better forecast future patterns. Full article
Show Figures

Figure 1

Figure 1
<p>Census block group relative population change from years 2000 to 2020 (%).</p>
Full article ">Figure 2
<p>County-level projected population change resulting from the combined climate effect over the next 30 years.</p>
Full article ">Figure 3
<p>County-level projected population change (%) resulting from the combined climate effect, socioeconomic impact under SSP2, and population redistribution due to climate migration over the next 30 years.</p>
Full article ">Figure 4
<p>Population projection trends in Miami-Dade County neighborhoods for areas of continual growth (blue), risky growth with tipping points (gray), and climate abandonment (red).</p>
Full article ">Figure 5
<p>Miami-Dade County block groups’ combined climate effect and projected population trend designation (risky growth, tipping point, or climate abandonment area).</p>
Full article ">
17 pages, 9836 KiB  
Article
An Algorithm to Retrieve Range Ocean Current Speed under Tropical Cyclone Conditions from Sentinel-1 Synthetic Aperture Radar Measurements Based on XGBoost
by Yuhang Zhou, Weizeng Shao, Ferdinando Nunziata, Weili Wang and Cheng Li
Remote Sens. 2024, 16(17), 3271; https://doi.org/10.3390/rs16173271 - 3 Sep 2024
Viewed by 444
Abstract
In this study, a novel algorithm to retrieve the current speed along the range direction under extreme sea states is developed from C-band synthetic aperture radar imagery. To this aim, a Sentinel-1 (S-1) dual-polarized synthetic aperture radar (SAR) dataset consisting of 2300 images [...] Read more.
In this study, a novel algorithm to retrieve the current speed along the range direction under extreme sea states is developed from C-band synthetic aperture radar imagery. To this aim, a Sentinel-1 (S-1) dual-polarized synthetic aperture radar (SAR) dataset consisting of 2300 images is collected during 200 tropical cyclones (TCs). The dataset is complemented with collocated wave simulations from the Wavewatch-III (WW3) model and reanalysis currents from the HYbrid Coordinate Ocean Model (HYCOM). The corresponding TC winds are officially released by IFRMER, while the Stokes drift following the wave propagation direction is estimated from the waves simulated by WW3. In this study, first the dependence of wind, Stokes drift, and range current on the Doppler centroid anomaly is investigated, and then the extreme gradient boosting (XGBoost) machine learning model is trained on 87% of the S-1 dataset for range current retrieval purposes. The rest of the dataset is used for testing the retrieval algorithm, showing a root mean square error (RMSE) and a correlation coefficient (r) of 0.11 m/s and 0.97, respectively, with the HYCOM outputs. A validation against measurements collected from two high-frequency (HF) phased-array radars is also performed, resulting in an RMSE and r of 0.12 m/s and 0.75, respectively. Those validation results are better than the 0.22 m/s RMSE and 0.28 r achieved by the empirical CDOP model. Hence, the experimental results confirm the soundness of the XGBoost, exhibiting a certain improvement over the empirical model. Full article
(This article belongs to the Special Issue SAR Monitoring of Marine and Coastal Environments)
Show Figures

Figure 1

Figure 1
<p>The frames of the 2300 S-1 SAR images overlaid on the TC tracks and maximum wind speeds, in which the red boxes indicate the training set and the black boxes indicate the test set.</p>
Full article ">Figure 2
<p>(<b>a</b>) S-1 VV-polarized NRCS image collected during the TC Infa (24 July 2021, 09:55 UTC); (<b>b</b>) S-1 VH-polarized NRCS image collected during the TC Infa (24 July 2021, 09:55 UTC); (<b>c</b>) corresponding CyclObs wind map.</p>
Full article ">Figure 3
<p>(<b>a</b>) SWH map simulated by WW3 at 10:00 UTC on 24 July 2021 relevant to the TC Infa, where the footprint of the HY-2B altimeter track is highlighted. (<b>b</b>) Validation of WW3-simulated SWH against the HY-2B products during the period of June–August 2022 in China seas.</p>
Full article ">Figure 4
<p>Current speed map simulated for 24 July 2021, 09:00 UTC, using HYCOM. The black box represents the footprint of the S-1 SAR image shown in <a href="#remotesensing-16-03271-f002" class="html-fig">Figure 2</a>a.</p>
Full article ">Figure 5
<p>Current speed maps measured by the HF phased-array radars over the TC Infa on (<b>a</b>) 24 July 2021, 09:55 UTC, and (<b>b</b>) 25 July 2021, 21:51 UTC. The black line and red dot are the typhoon track, and the red triangle is the current typhoon time position.</p>
Full article ">Figure 6
<p>DCA versus (<b>a</b>) wind speed, (<b>b</b>) Stokes drift, and (<b>c</b>) current speed, projected onto the range direction. Black lines stand for the linear regression results.</p>
Full article ">Figure 7
<p>The flow chart of the eXtreme gradient boosting (XGBoost).</p>
Full article ">Figure 8
<p>(<b>a</b>) The behavior of the XGBoost training process and (<b>b</b>) the SHAP value map.</p>
Full article ">Figure 9
<p>(<b>a</b>) Range current speed map by XGBoost method and (<b>b</b>) the empirical algorithm obtained from the S-1 SAR scene collected over TC Infa on 24 July 2021, 09:55 UTC; (<b>c</b>) the HYCOM range current speed map on 24 July 2021, 09:00.</p>
Full article ">Figure 10
<p>Latitude variation given the longitude—see the dashed red line in (<a href="#remotesensing-16-03271-f009" class="html-fig">Figure 9</a>a)—of the range current speed retrieved by XGBoost (red line), by CDOP model (green line), and simulated with HYCOM (black line).</p>
Full article ">Figure 11
<p>Validation of SAR-derived range current wind speeds by (<b>a</b>) XGBoost and (<b>b</b>) the CDOP model against HYCOM data.</p>
Full article ">Figure 12
<p>Validation of SAR-derived range current wind speeds by (<b>a</b>) XGBoost and (<b>b</b>) CDOP model against HF radar measurements.</p>
Full article ">Figure 13
<p>Taylor diagram with respect to current speed up to 3 m/s at intervals of 1 m/s, in which the red and blue symbols represent the result using XGBoost and the CDOP model.</p>
Full article ">Figure 14
<p>Variation in the bias (SAR-derived minus HYCOM data) of the range current speed with respect to (<b>a</b>) the HYCOM current speed for 0.375 m/s, (<b>b</b>) the wind speed for 10 m/s, (<b>c</b>) the Stokes drift for a 0.125 m/s bin, and (<b>d</b>) the DCA for a 12.5 Hz bin.</p>
Full article ">
28 pages, 18362 KiB  
Communication
Observations and Simulations of the Winds at a Bridge in Hong Kong during Two Tropical Cyclone Events in 2023
by Ka-Wai Lo, Pak-Wai Chan and Kai-Kwong Lai
Appl. Sci. 2024, 14(17), 7789; https://doi.org/10.3390/app14177789 - 3 Sep 2024
Viewed by 612
Abstract
The impact of tropical cyclones on the operation of bridges and railways are mostly dependent on the forecast wind speed trend (upward, downward or staying steady) at present in Hong Kong. There are requests to forecast the exceedance of wind speed thresholds for [...] Read more.
The impact of tropical cyclones on the operation of bridges and railways are mostly dependent on the forecast wind speed trend (upward, downward or staying steady) at present in Hong Kong. There are requests to forecast the exceedance of wind speed thresholds for such operations with a lead time of many hours ahead. This study considers the technical feasibility of forecasting the wind speed along a recently built bridge in Hong Kong with a coupled mesoscale numerical weather prediction (NWP)–computational fluid dynamics (CFD) model. This bridge features numerous anemometers where the coupled model can be verified. It is found that these two tropical cyclone cases are very challenging, especially in representing the wind structure of the cyclone because of its relatively small circulation. As such, the timing of the maximum wind could be 2 to 4 h earlier than the actual observation. The maximum wind speed from CFD modeling could be higher than that from the NWP model alone by 5 to 10 m/s, which shows that CFD modeling could add value in the forecasting of wind speed exceedance, although the maximum simulated value is still lower than the actual observation by as much as 10 m/s. As a result, while the general wind speed trend may be forecast, the exceedance of definite values of wind speed limits is still practically rather challenging, given the present two cases of tropical cyclones. Full article
Show Figures

Figure 1

Figure 1
<p>The surface isobaric charts at (<b>a</b>) 08H, 1 September 2023 and (<b>b</b>) 08H, 8 October 2023. Chinese translations are shown in the figure. True color satellite image of Himawari-9 satellite of Japan Meteorological Agency at (<b>c</b>) 08H, 1 September 2023 and (<b>d</b>) 08H, 8 October 2023. All times are local time (+8 UTC).</p>
Full article ">Figure 1 Cont.
<p>The surface isobaric charts at (<b>a</b>) 08H, 1 September 2023 and (<b>b</b>) 08H, 8 October 2023. Chinese translations are shown in the figure. True color satellite image of Himawari-9 satellite of Japan Meteorological Agency at (<b>c</b>) 08H, 1 September 2023 and (<b>d</b>) 08H, 8 October 2023. All times are local time (+8 UTC).</p>
Full article ">Figure 1 Cont.
<p>The surface isobaric charts at (<b>a</b>) 08H, 1 September 2023 and (<b>b</b>) 08H, 8 October 2023. Chinese translations are shown in the figure. True color satellite image of Himawari-9 satellite of Japan Meteorological Agency at (<b>c</b>) 08H, 1 September 2023 and (<b>d</b>) 08H, 8 October 2023. All times are local time (+8 UTC).</p>
Full article ">Figure 2
<p>(<b>a</b>) Locations of eight anemometers along the bridge under study. The cyan line represents a cross section perpendicular to the bridge where the simulation winds on it will be shown in Figures 11b and 18b. (<b>b</b>) Vertical sectional view of the bridge projected onto the horizontal red line in (<b>a</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) The 10 min mean wind speeds and directions for the eight anemometers at 20:10H, 1 September 2023 in local time and (<b>b</b>) at 21:30H, 1 September 2023 in local time.</p>
Full article ">Figure 4
<p>(<b>a</b>) The 10 min mean wind speeds and directions for the eight anemometers at 17:50H, 8 October 2023 in local time and (<b>b</b>) at 22:25H, 8 October 2023 in local time.</p>
Full article ">Figure 5
<p>(<b>a</b>) The 1st, 2nd, and 3rd nested domains of the RAMS simulation. (<b>b</b>) The 3rd, 4th, and 5th nested domains of the RAMS simulation. (<b>c</b>) The domain of the PALM simulation is denoted by a red rectangle, where the boundary is the 5th nested domain of the RAMS simulation.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) The 1st, 2nd, and 3rd nested domains of the RAMS simulation. (<b>b</b>) The 3rd, 4th, and 5th nested domains of the RAMS simulation. (<b>c</b>) The domain of the PALM simulation is denoted by a red rectangle, where the boundary is the 5th nested domain of the RAMS simulation.</p>
Full article ">Figure 6
<p>(<b>a</b>) Horizontal winds and (<b>b</b>) eddy dissipation rate at a height of around 60 m above the mean sea level at 12:10Z, 1 September 2023 in RAMS simulation.</p>
Full article ">Figure 7
<p>Simulated 10 min mean wind speed and direction at the locations of the eight anemometers by PALM simulation at (<b>a</b>) 20:10H, 1 September 2023 and (<b>b</b>) 21:30H, 1 September 2023. Both are in local time.</p>
Full article ">Figure 8
<p>Time series of 10 min mean wind speed for the six anemometers (<b>a</b>) CB1, (<b>b</b>) CB3, (<b>c</b>) CB5, (<b>d</b>) CB6, (<b>e</b>) CB7, and (<b>f</b>) CB8 on 1 September 2023 in local time.</p>
Full article ">Figure 9
<p>Time series of wind gusts for the six anemometers (<b>a</b>) CB1, (<b>b</b>) CB3, (<b>c</b>) CB5, (<b>d</b>) CB6, (<b>e</b>) CB7, and (<b>f</b>) CB8 on 1 September 2023 in local time.</p>
Full article ">Figure 10
<p>Time series of 10 min wind direction for CB5 on 1 September 2023 in local time.</p>
Full article ">Figure 11
<p>(<b>a</b>) Simulated 10 min mean wind speed and direction distribution by PALM at a height of about 70 m above mean sea level and (<b>b</b>) vertical cross-section of wind speed and direction across the bridge at 20:10H, 1 September 2023 in local time. The cross-section location is depicted by the cyan line in <a href="#applsci-14-07789-f002" class="html-fig">Figure 2</a>a, where a smaller value on the x-axis indicates a location further south.</p>
Full article ">Figure 12
<p>(<b>a</b>) Horizontal wind speed and (<b>b</b>) eddy dissipation rate at a height of around 60 m above the mean sea level at 09:50Z, 8 October 2023 in RAMS simulation.</p>
Full article ">Figure 12 Cont.
<p>(<b>a</b>) Horizontal wind speed and (<b>b</b>) eddy dissipation rate at a height of around 60 m above the mean sea level at 09:50Z, 8 October 2023 in RAMS simulation.</p>
Full article ">Figure 13
<p>Simulated 10 min mean wind speed and direction at the locations of the eight anemometers by PALM simulation at (<b>a</b>) 17:50H, 8 October 2023 and (<b>b</b>) 22:25H, 8 October 2023. Both are in local time.</p>
Full article ">Figure 14
<p>Time series of 10 min mean wind speed for the six anemometers (<b>a</b>) CB1, (<b>b</b>) CB3, (<b>c</b>) CB5, (<b>d</b>) CB6, (<b>e</b>) CB7, and (<b>f</b>) CB8 on 8 October 2023 in local time.</p>
Full article ">Figure 15
<p>Time series of wind gusts for the six anemometers (<b>a</b>) CB1, (<b>b</b>) CB3, (<b>c</b>) CB5, (<b>d</b>) CB6, (<b>e</b>) CB7, and (<b>f</b>) CB8 on 8 October 2023 in local time..</p>
Full article ">Figure 16
<p>Time series of 10 min wind direction for CB5 on 8 October 2023 in local time.</p>
Full article ">Figure 17
<p>Comparison of provisional best track from HKO and NCEP model forecast track for (<b>a</b>) Saola, where the NCEP model is initialized at 2023090106UTC, and (<b>b</b>) Koinu, where the NCEP model is initialized at 2023100806UTC.</p>
Full article ">Figure 18
<p>(<b>a</b>) Simulated 10 min mean wind speed and direction distribution by PALM at a height of about 70 m above mean sea level and (<b>b</b>) vertical cross-section of wind speed and direction across the bridge at 17:50H, 8 October 2023 in local time. The cross-section location is depicted by the cyan line in <a href="#applsci-14-07789-f002" class="html-fig">Figure 2</a>a, where a smaller value on the x-axis indicates a location further south.</p>
Full article ">Figure 19
<p>Time series of measured and simulated 10 min mean wind speed for the six anemometers (<b>a</b>) CB1, (<b>b</b>) CB3, (<b>c</b>) CB5, (<b>d</b>) CB6, (<b>e</b>) CB7, and (<b>f</b>) CB8 on 1 September 2023 in local time. The simulated wind speeds are driven by GRAPES-Meso.</p>
Full article ">
33 pages, 11955 KiB  
Article
The Adiabatic Evolution of 3D Annular Vortices with a Double-Eyewall Structure
by Gabriel J. Williams
Dynamics 2024, 4(3), 698-730; https://doi.org/10.3390/dynamics4030035 - 2 Sep 2024
Viewed by 398
Abstract
Tropical cyclones (TCs) can be characterized as a 3D annular structure of elevated potential vorticity (PV). However, strong mature TCs often develop a secondary eyewall, leading to a 3D annular vortex with a double-eyewall structure. Using 2D linear stability analysis, it is shown [...] Read more.
Tropical cyclones (TCs) can be characterized as a 3D annular structure of elevated potential vorticity (PV). However, strong mature TCs often develop a secondary eyewall, leading to a 3D annular vortex with a double-eyewall structure. Using 2D linear stability analysis, it is shown that three types of barotropic instability (BI) are present for annular vortices with a double-eyewall structure: Type-1 BI across the secondary eyewall, Type-2 BI across the moat of the vortex, and Type-3 BI across the primary eyewall. The overall stability of these vortices (and the type of BI that develops) depends principally upon five vortex parameters: the thickness of the primary eyewall, the thickness of the secondary eyewall, the moat width, the vorticity ratio between the eye and the primary eyewall, and the vorticity ratio between the primary and secondary eyewall. The adiabatic evolution of 3D annular vortices with a double-eyewall structure is examined using a primitive equation model in normalized isobaric coordinates. It is shown that Type-2 BI is the most common type of BI for 3D annular vortices whose vortex parameters mimic TCs with a double-eyewall structure. During the onset of Type-2 BI, eddy kinetic energy budget analysis indicates that barotropic energy conversion from the mean azimuthal flow is the dominant energy source of the eddies, which produces a radial velocity field with a quadrupole structure. Absolute angular momentum budget analysis indicates that Type-2 BI generates azimuthally averaged radial outflow across the moat, and the eddies transport absolute angular momentum radially outward towards the secondary eyewall. The combination of these processes leads to the dissipation of the primary eyewall and the maintenance of the secondary eyewall for the vortex. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the axisymmetric vertical vorticity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ζ</mi> </mrow> <mo>¯</mo> </mover> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math> (with solid black curve) for the six-region 2D annular vortex with a double-eyewall structure. The accompanying angular velocity <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ω</mi> </mrow> <mo>¯</mo> </mover> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math> is given by the solid blue curve.</p>
Full article ">Figure 2
<p>Isolines of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of primary eyewall thickness <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> </mrow> </semantics></math> and moat width <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> </mrow> </semantics></math>. The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>m</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> <mo>=</mo> <mn>2.00</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Isolines of the maximum value of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of eyewall vorticity ratio <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> </mrow> </semantics></math> and the moat strength <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> </mrow> </semantics></math> f. The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>m</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.333</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.750</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Isolines of the maximum value of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of eyewall vorticity ratio <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> </mrow> </semantics></math> and primary eyewall thickness <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> </mrow> </semantics></math> The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>m</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.750</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Isolines of the maximum value of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of eyewall vorticity ratio <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> </mrow> </semantics></math> and moat width <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> </mrow> </semantics></math>. The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.750</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Isolines of the maximum value of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of eyewall vorticity ratio <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> </mrow> </semantics></math> and moat width <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> </mrow> </semantics></math>. The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>m</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.333</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Isolines of the maximum value of the dimensionless growth rate <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> computed from Equation (9) as a function of eyewall vorticity ratio <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>13</mn> </mrow> </msub> </mrow> </semantics></math> and primary eyewall thickness <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> </mrow> </semantics></math>. The minimum isoline value is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ν</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, and the contour interval is 0.05. The color shading indicates the azimuthal wavenumber associated with the most unstable mode <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>m</mi> </mrow> <mrow> <mi>*</mi> </mrow> </msup> </mrow> </semantics></math> where the white color indicates the stable region. For each panel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>01</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>23</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mn>43</mn> </mrow> </msub> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mn>34</mn> </mrow> </msub> <mo>=</mo> <mn>0.500</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>The initial condition for the control experiment. (<b>a</b>) shows the azimuthal-mean PV (in PVU where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi>P</mi> <mi>V</mi> <mi>U</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> <mi mathvariant="normal">K</mi> <mo> </mo> <mi mathvariant="normal">k</mi> <msup> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>b</b>) shows the azimuthal-mean azimuthal velocity with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>c</b>) shows the azimuthal-mean-normalized Okubo–Weiss parameter. (<b>d</b>) shows the PV on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.85</mn> </mrow> </semantics></math>, and (<b>e</b>) shows the potential vorticity on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.35</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>The evolution of the vortex for the control experiment at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>middle panel</b>), and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>18.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>). (<b>a</b>,<b>e</b>,<b>i</b>) show the azimuthal-mean PV (in PVU where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi>P</mi> <mi>V</mi> <mi>U</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> <mi mathvariant="normal">K</mi> <mo> </mo> <mi mathvariant="normal">k</mi> <msup> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>b</b>,<b>f</b>,<b>j</b>) show the azimuthal-mean-normalized Okubo–Weiss parameter with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>c</b>,<b>g</b>,<b>k</b>) show the PV on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>. (<b>d</b>,<b>h</b>,<b>l</b>) show the radial velocity <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>R</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>The Hovmoller diagram associated with the azimuthal wind on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math> from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> up to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>18</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>M</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math>. The explanation for each term is found in the text.</p>
Full article ">Figure 12
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>K</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math>. See the text for the meaning for each term in the budget analysis.</p>
Full article ">Figure 13
<p>The initial condition for Vortex I, II, and III. (<b>a</b>–<b>c</b>) show the azimuthal-mean PV for Vortex I, II, and III, respectively. (<b>d</b>–<b>f</b>) show the azimuthal-mean azimuthal wind <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>T</mi> </mrow> </msub> </mrow> </semantics></math> with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>) for Vortex I, II, and III, respectively.</p>
Full article ">Figure 14
<p>The evolution of the vortex for Vortex I at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>2.50</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>middle panel</b>), and (<b>bottom panel</b>). (<b>a</b>,<b>e</b>,<b>i</b>) show the azimuthal-mean PV (in PVU where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi>P</mi> <mi>V</mi> <mi>U</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> <mi mathvariant="normal">K</mi> <mo> </mo> <mi mathvariant="normal">k</mi> <msup> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>b</b>,<b>f</b>,<b>j</b>) show the azimuthal-mean-normalized Okubo–Weiss parameter with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>c</b>,<b>g</b>,<b>k</b>) show the PV on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>. (<b>d</b>,<b>h</b>,<b>l</b>) show the radial velocity <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>R</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>M</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>) and from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>) for Vortex I. The top panel corresponds to the time interval in which Type-3 BI is the dominant instability, whereas the bottom panel corresponds to the time interval in which Type-2 BI is the dominant instability. See the text for the meaning of each term in the budget analysis.</p>
Full article ">Figure 16
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>K</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>) and from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>) for Vortex I. The top panel corresponds to the time interval in which Type-3 BI is the dominant instability, whereas the bottom panel corresponds to the time interval in which Type-2 BI is the dominant instability. See the text for the meaning of each term in the budget analysis.</p>
Full article ">Figure 17
<p>The evolution of the vortex for Vortex II at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>4.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>6.50</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>middle panel</b>), and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>18.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>). (<b>a</b>,<b>e</b>,<b>i</b>) show the azimuthal-mean PV (in PVU where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi>P</mi> <mi>V</mi> <mi>U</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> <mi mathvariant="normal">K</mi> <mo> </mo> <mi mathvariant="normal">k</mi> <msup> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>b</b>,<b>f</b>,<b>j</b>) show the azimuthal-mean-normalized Okubo–Weiss parameter with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>c</b>,<b>g</b>,<b>k</b>) show the PV on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>. (<b>d</b>,<b>h</b>,<b>l</b>) show the radial velocity <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>R</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 18
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>M</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis (<b>top panel</b>) and the integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>K</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis (<b>bottom panel</b>) from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>8.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> for Vortex I. See the text for the meaning of each term in the budget analysis.</p>
Full article ">Figure 19
<p>The evolution of the vortex for Vortex III at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>7.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>middle panel</b>), and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>18.0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>). (<b>a</b>,<b>e</b>,<b>i</b>) show the azimuthal-mean PV (in PVU where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo> </mo> <mi>P</mi> <mi>V</mi> <mi>U</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> <mi mathvariant="normal">K</mi> <mo> </mo> <mi mathvariant="normal">k</mi> <msup> <mrow> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>b</b>,<b>f</b>,<b>j</b>) show the azimuthal-mean-normalized Okubo–Weiss parameter with isosurfaces of absolute angular momentum (in units of <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <msup> <mrow> <mi mathvariant="normal">s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). (<b>c</b>,<b>g</b>,<b>k</b>) show the PV on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>. (<b>d</b>,<b>h</b>,<b>l</b>) show the radial velocity <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>R</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 20
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>M</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>) and from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>4.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>) for Vortex I. The top panel corresponds to the time interval in which Type-1 BI is the dominant instability, whereas the bottom panel corresponds to the time interval in which Type-2 BI is the dominant instability. See the text for the meaning of each term in the budget analysis.</p>
Full article ">Figure 21
<p>The integrated <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>K</mi> </mrow> <mo>¯</mo> </mover> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> budget analysis from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>3.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>top panel</b>) and from <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>4.00</mn> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.00</mn> <mo> </mo> <mi mathvariant="normal">h</mi> </mrow> </semantics></math> (<b>bottom panel</b>) for Vortex III. The top panel corresponds to the time interval in which Type-3 BI is the dominant instability, whereas the bottom panel corresponds to the time interval in which Type-2 BI is the dominant instability. See the text for the meaning of each term in the budget analysis.</p>
Full article ">
15 pages, 4855 KiB  
Article
An Assessment of Tropical Cyclone Frequency in the Bay of Bengal and Its Impact on Coastal Bangladesh
by Md Wahiduzzaman and Alea Yeasmin
Coasts 2024, 4(3), 594-608; https://doi.org/10.3390/coasts4030030 - 2 Sep 2024
Viewed by 421
Abstract
This study examines the frequency of tropical cyclones in the Bay of Bengal and their impact on Bangladesh. The extent of environmental harm led to the selection of two specific areas: the Panpatty Union and Galachipa Upzilla in the Patuakhali district, and the [...] Read more.
This study examines the frequency of tropical cyclones in the Bay of Bengal and their impact on Bangladesh. The extent of environmental harm led to the selection of two specific areas: the Panpatty Union and Galachipa Upzilla in the Patuakhali district, and the Sariakat Union and Swandip Upzilla in the Chittagong district. The results indicate that cyclonic storms are more common in May and November. The results also demonstrate that the studied regions are vulnerable to the effects of tropical cyclones and suffer significant consequences. The differences in influence between the two locations are statistically significant with a confidence level of 90%. The findings have significant ramifications for policymaking decisions. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the Bay of Bengal.</p>
Full article ">Figure 2
<p>Study area (<b>a</b>) Panpatty Union, Galachipa Upazilla in Patuakhali District, (<b>b</b>) Saraikat Union, Sandwip Upazilla in Chittangong District and (<b>c</b>) Bangladesh.</p>
Full article ">Figure 3
<p>Monthly frequency of depression and deep depression.</p>
Full article ">Figure 4
<p>Monthly distribution of cyclonic storm.</p>
Full article ">Figure 5
<p>Monthly distribution of severe cyclonic storm (SCS).</p>
Full article ">Figure 6
<p>Relief analysis between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union.</p>
Full article ">Figure 7
<p>Agriculture land analysis between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union. There are three types of land identified—highly rich/fertile, medium rich/fertile and lower rich/fertile land.</p>
Full article ">Figure 8
<p>Fishing zone comparison between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union.</p>
Full article ">Figure 9
<p>Hazard zone comparison between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union.</p>
Full article ">Figure 10
<p>Cyclone vulnerability comparison between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union.</p>
Full article ">Figure 11
<p>Food crisis comparison between (<b>a</b>) Panpatty and (<b>b</b>) Sariakat union.</p>
Full article ">Figure 12
<p>Home materials in Panpatty Union, Galachipa Upazilla in Patuakhali District (<b>a</b>) and Saraikat Union, Sandwip Upazilla in Chittangong District (<b>b</b>).</p>
Full article ">Figure 13
<p>Nature of impacts of cyclone on house in Panpatty Union (<b>a</b>) and Saraikat Union (<b>b</b>).</p>
Full article ">Figure 14
<p>Impacts of cyclone on agriculture in (<b>a</b>) Panpatty Union and (<b>b</b>) Saraikat Union.</p>
Full article ">Figure 15
<p>Impacts on fisheries in Panpatty Union (<b>a</b>) and Saraikat Union (<b>b</b>).</p>
Full article ">Figure 16
<p>Institutional damage in Panpatty Union (<b>a</b>) and Saraikat Union (<b>b</b>).</p>
Full article ">Figure 17
<p>Impacts of cyclone on (<b>a</b>) health and (<b>b</b>) environment.</p>
Full article ">
21 pages, 7364 KiB  
Article
Deriving Tropical Cyclone-Associated Flood Hazard Information Using Clustered GPM-IMERG Rainfall Signatures: Case Study in Dominica
by Catherine Nabukulu, Victor G. Jetten, Janneke Ettema, Bastian van den Bout and Reindert J. Haarsma
Atmosphere 2024, 15(9), 1042; https://doi.org/10.3390/atmos15091042 - 29 Aug 2024
Viewed by 868
Abstract
Various stakeholders seek effective methods to communicate the potential impacts of tropical cyclone (TC) rainfall and subsequent flood hazards. While current methods, such as Intensity–Duration–Frequency curves, offer insights, they do not fully capture TC rainfall complexity and variability. This research introduces an innovative [...] Read more.
Various stakeholders seek effective methods to communicate the potential impacts of tropical cyclone (TC) rainfall and subsequent flood hazards. While current methods, such as Intensity–Duration–Frequency curves, offer insights, they do not fully capture TC rainfall complexity and variability. This research introduces an innovative workflow utilizing GPM-IMERG satellite precipitation estimates to cluster TC rainfall spatial–temporal patterns, thereby illustrating their potential for flood hazard assessment by simulating associated flood responses. The methodology is tested using rainfall time series from a single TC as it traversed a 500 km diameter buffer zone around Dominica. Spatial partitional clustering with K-means identified the spatial clusters of rainfall time series with similar temporal patterns. The optimal value of K = 4 was most suitable for grouping the rainfall time series of the tested TC. Representative precipitation signals (RPSs) from the quantile analysis generalized the cluster temporal patterns. RPSs served as the rainfall input for the openLISEM, an event-based hydrological model simulating related flood characteristics. The tested TC exhibited three spatially distinct levels of rainfall magnitude, i.e., extreme, intermediate, and least intense, each resulting in different flood responses. Therefore, TC rainfall varies in space and time, affecting local flood hazards; flood assessments should incorporate variability to improve response and recovery. Full article
(This article belongs to the Special Issue Advances in Rainfall-Induced Hazard Research)
Show Figures

Figure 1

Figure 1
<p>Flow chart for the proof of concept showing the sequence of analyses that form the developed workflow.</p>
Full article ">Figure 2
<p>Illustration of the procedure for selecting cluster representative precipitation signals. (<b>a</b>) Visualization of a selection of pixel rainfall time series of a given cluster. (<b>b</b>) Rainfall series after applying a starting threshold. (<b>c</b>) Example for timestep quantile series for probabilities 0.5, 0.75, and 0.9 calculated after applying the starting threshold.</p>
Full article ">Figure 3
<p>Map of the region of interest defined by the buffer boundary. The cumulative total rainfall for any given pixel in the study area over the selected time window is represented with a green-red colour ramp. The clusters labelled T1, T2, T3, T4, and T5 resulted from using <span class="html-italic">K</span> = 5 when conducting the spatial clustering analysis of the pixel rainfall time series. The bottom inset is the location of the Grand Bay catchment for which the flood modelling was performed.</p>
Full article ">Figure 4
<p>Elbow graph of the optimal <span class="html-italic">K</span> values used to experiment with the developed workflow. The total within-cluster sum of squares (TWSS) is the total sum of the squared distances between the data points and the centroid of their assigned clusters.</p>
Full article ">Figure 5
<p>The outcome of the time alignment and quantile analysis when applying <span class="html-italic">K</span> = 5. Only Q<sub>0.5</sub> and Q<sub>0.75</sub> were selected for each cluster as the representative precipitation signals (RPSs). Clusters T1, T3, T4, and T5 are in graphs (<b>a</b>), (<b>b</b>), (<b>c</b>), and (<b>d</b>), respectively.</p>
Full article ">Figure 6
<p>Flood depth maps for Grand Bay catchment on Dominica as simulated using the Q<sub>0.75</sub> RPSs for clusters T1, T3, and T4 resulting from applying <span class="html-italic">K</span> = 4.</p>
Full article ">Figure 7
<p>Precipitation time series (intensity/duration) and cumulative plots of the TC-associated rainfall scenarios for TS Erika resulting from <span class="html-italic">K</span> = 4. The blue bar graphs represent the half-hourly rainfall intensities, and the black line is for the cumulative precipitation. (<b>a</b>) Extreme, (<b>b</b>) intermediate, and (<b>c</b>) least intense.</p>
Full article ">Figure 8
<p>Plots of (<b>a</b>) the average hourly rainfall on Dominica due to TS Erika and the line plot of the cumulative rain from <a href="#atmosphere-15-01042-t001" class="html-table">Table 1</a> of the report by [<a href="#B28-atmosphere-15-01042" class="html-bibr">28</a>]. Plots (<b>b</b>,<b>c</b>) compare the output cluster-based design events with the design storms derived from IDF curves available for the region in <a href="#atmosphere-15-01042-f003" class="html-fig">Figure 3</a> from [<a href="#B20-atmosphere-15-01042" class="html-bibr">20</a>] by applying the ABM. The line plots (<b>b</b>,<b>c</b>) are the corresponding cumulative rainfall, black for the output cluster-based design events and orange for the design storms.</p>
Full article ">Figure A1
<p>The maps below represent the cluster boundaries resulting from applying <span class="html-italic">K</span> = 4 (<b>left</b>) and <span class="html-italic">K</span> = 3 (<b>right</b>) when conducting the spatial partitional clustering.</p>
Full article ">
Back to TopTop