[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

Search Results (192)

Search Parameters:
Keywords = typhoon track

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 3842 KiB  
Article
An Innovative Dynamic Model for Predicting Typhoon Track Deflections over Complex Terrain
by Hung-Cheng Chen
Atmosphere 2024, 15(11), 1372; https://doi.org/10.3390/atmos15111372 - 14 Nov 2024
Viewed by 462
Abstract
This study presents an innovative, dynamic model for predicting typhoon track deflections over complex terrain. Based on potential vorticity conservation, the model incorporates a topographic adjusting parameter (α) and a meridional adjusting velocity (MAV) to capture the vortex’s response to terrain variations. Simulations [...] Read more.
This study presents an innovative, dynamic model for predicting typhoon track deflections over complex terrain. Based on potential vorticity conservation, the model incorporates a topographic adjusting parameter (α) and a meridional adjusting velocity (MAV) to capture the vortex’s response to terrain variations. Simulations using an idealized bell-shaped mountain and Taiwan’s realistic topography reveal that steeper terrain gradients consistently deflect typhoon tracks southward. This steering effect intensifies with increasing vortex strength due to a larger α, leading to enhanced MAV. Shallower approach angles also amplify deflections due to prolonged terrain interaction. Results highlight the significant role of Taiwan’s Central Mountain Range in shaping typhoon trajectories. This model offers a refined approach for predicting typhoon behavior near complex terrain, advancing forecasting capabilities, and enhancing disaster preparedness strategies. Full article
(This article belongs to the Special Issue Typhoon/Hurricane Dynamics and Prediction (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Schematic of the dynamic model that illustrates the interaction between a cyclonic vortex and isolated topography. (<b>a</b>) The vortex approaches the terrain at an impinging angle <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math>. The vortex has a radius of maximum winds <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math> and a maximum azimuthal velocity <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>, interacting with a bell-shaped mountain of maximum height <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>h</mi> </mrow> <mrow> <mi>M</mi> </mrow> </msub> </mrow> </semantics></math> and base height <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>h</mi> </mrow> <mrow> <mi>B</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>b</b>) Cross-sectional profile of the topography along the vortex path. <math display="inline"><semantics> <mrow> <mi>D</mi> </mrow> </semantics></math> represents the unperturbed fluid depth, and <math display="inline"><semantics> <mrow> <mi>H</mi> </mrow> </semantics></math> is the total fluid depth accounting for topography and surface effects. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>v</mi> </mrow> </msub> </mrow> </semantics></math> denotes the maximum free surface depression caused by the vortex. <math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math> is the inclined angle of topography.</p>
Full article ">Figure 2
<p>(<b>a</b>) The 3D bell-shaped mountain’s topography shows the latitudinal cross sections AA’ to EE’. Solid contour lines represent 100-m intervals, whereas dashed lines mark 1000-m elevations. (<b>b</b>) Cross-sectional topographic profiles along each of the five latitudes. These profiles illustrate the steep elevation of the CMR, peaking above 3000 m, with the highest elevations near the CC’ and DD’ sections, showing strong variation in terrain across the island.</p>
Full article ">Figure 3
<p>Tracks of vortex M passing over isolated topography for different impinging angles: (<b>a</b>) 195°, (<b>b</b>) 170°, (<b>c</b>) 145°, and (<b>d</b>) 120°. The corresponding track deflections (<span class="html-italic">δ</span>) are also shown. Numbers 1–12 represent the name of cases, respectively. The dashed lines represent the tracks and deflections for the northern landfall cases (1, 4, 7, and 10). The solid lines show the tracks and deflections for the central landfall cases (2, 5, 8, and 11). The dashed-dotted lines correspond to the tracks and deflections for the southern landfall cases (3, 6, 9, and 12).</p>
Full article ">Figure 4
<p>(<b>a</b>) Topography of Taiwan: Contour lines representing elevations of 100 m (solid lines) and 1000 m (dashed lines) are shown. Five latitudinal cross-sections (AA’, BB’, CC’, DD’, and EE’) are marked. (<b>b</b>) Elevational profiles: Topographic profiles along the five cross-sections indicated in panel (<b>a</b>) are displayed.</p>
Full article ">Figure 5
<p>Simulated tracks (scanned trajectories) of vortices interacting with the realistic topography of Taiwan at varying initial latitudes but fixed impinging angles. Each panel represents a different vortex central relative vorticity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ζ</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>): (<b>a</b>) S1 (<math display="inline"><semantics> <mrow> <mn>2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>), (<b>b</b>) S2 (<math display="inline"><semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>), (<b>c</b>) S3 (<math display="inline"><semantics> <mrow> <mn>6</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>). The impinging angle (γ) for the panels (<b>a</b>–<b>c</b>) is 195°. Panels (<b>d</b>–<b>f</b>) use the same vortex intensities as (<b>a</b>–<b>c</b>), respectively, but with an impinging angle of 170°. Numbers 1–5 are scanned tracks of vortices impinging from north to south latitudes. S1, S2 and S3 are names of vortices.</p>
Full article ">Figure 6
<p>The figure is similar to <a href="#atmosphere-15-01372-f005" class="html-fig">Figure 5</a> but illustrates tracks for shallower impinging angles. Panels (<b>a</b>–<b>c</b>) display the tracks for vortices S1, S2, and S3, respectively, with a fixed impinging angle of 145°. Panels (<b>d</b>–<b>f</b>) present the corresponding tracks for the same vortices, but with an impinging angle of 120°. Numbers 1-5 are scanned tracks of vortices impinging from north to south latitudes. S1, S2 and S3 are names of vortices.</p>
Full article ">Figure 7
<p>Comparison of the dynamic model’s simulated track with the observed track of Typhoon Soudelor (2015). (<b>a</b>) Simulated track (red solid line) and observed typhoon center positions (open pink stars) overlaid on the topography of Taiwan. (<b>b</b>) Elevational profile (black solid line) along the simulated track and the corresponding track deflection (red solid line). Track deflection is defined as the southward deviation of the simulated track from a straight westward path.</p>
Full article ">
16 pages, 5064 KiB  
Article
Impacts of Forecast Time and Verification Area Setting on the Targeted Observation of Typhoon
by Jiaqi Kang, Jianxia Guo, Jia Wang and Chao Zhang
Atmosphere 2024, 15(11), 1335; https://doi.org/10.3390/atmos15111335 - 7 Nov 2024
Viewed by 265
Abstract
The results of the identification of sensitive areas are affected by the forecast time and verification area settings in targeted observations. Understanding this setting issue is important for improving the effectiveness of the identification of sensitive areas in real-time field campaigns. To determine [...] Read more.
The results of the identification of sensitive areas are affected by the forecast time and verification area settings in targeted observations. Understanding this setting issue is important for improving the effectiveness of the identification of sensitive areas in real-time field campaigns. To determine this, a series of experiments were carried out based on the Ensemble Transform Sensitivity (ETS) method, and the results are as follows: (1) First, Observation System Simulation Experiments (OSSEs) were conducted to assimilate simulated dropsondes in sensitive areas (SENS) or non-sensitive areas (OTHR). The results showed that the SENS experiment improved forecasts of typhoon intensity, track, precipitation score, and RMSE of forecast elements. However, the OTHR experiment only improved the forecast in some aspects and even had negative effects on other aspects. This indicates that the sensitive areas identified by the ETS method are effective. (2) Different forecast time experiments were carried out. There were significant differences between the sensitive areas of fixed verification times and variable targeted observation times, indicating that the sensitive areas changed greatly with time. In the field campaign, it was necessary to calculate the sensitive area for multiple times in advance and to design or adjust the observation scheme according to the time. (3) Finally, comparative experiments of position deviation and size change in the verification area were carried out. It was found that for a big deviation, too large or too small a verification area will result in significant differences in the sensitive areas. Based on the study in this article, a verification area size of about 6° × 6° is recommended; this can not only accommodate the position deviation of the verification area from the typhoon center caused by forecast errors, but also does not contain too much noise unrelated to typhoons, which may affect the accuracy of identification of sensitive areas. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of dropsonde points The vector arrow represents the 10 m wind field forecast via ensemble control at 06:00 on the 24th, unit: m/s. The shadow represents the observation sensitive signals calculated via the ETS method of the verification time at 06:00 on 25 July and the targeted observation time at 06:00 on 24 July. The blue dots represent dropsondes in the sensitive area; the red diamond dots represent dropsondes in the non-sensitive area. The red solid rectangle is the verification area form <a href="#sec3dot1-atmosphere-15-01335" class="html-sec">Section 3.1</a> and <a href="#sec3dot2dot1-atmosphere-15-01335" class="html-sec">Section 3.2.1</a>. The blue dashed rectangle is the verification area form <a href="#sec3dot2dot2-atmosphere-15-01335" class="html-sec">Section 3.2.2</a>; and the black dotted line is the actual track of Typhoon In-Fa).</p>
Full article ">Figure 2
<p>Schematic diagram of different verification areas experiments (the black rectangle indicates the verification area of control experiment A, and the black dotted line indicates the track of the typhoon 24 h before landfall; (<b>a</b>) represents the experiments of the verification area shifted forward and backward in the direction of the typhoon moving track; (<b>b</b>) represents the experiments of the verification area shifted left and right over the typhoon moving track; (<b>c</b>) represents the experiments where the center position is unchanged and the size of the verification area is enlarged or reduced; (<b>d</b>) represents experiments where the center position and size of the verification area have both changed).</p>
Full article ">Figure 2 Cont.
<p>Schematic diagram of different verification areas experiments (the black rectangle indicates the verification area of control experiment A, and the black dotted line indicates the track of the typhoon 24 h before landfall; (<b>a</b>) represents the experiments of the verification area shifted forward and backward in the direction of the typhoon moving track; (<b>b</b>) represents the experiments of the verification area shifted left and right over the typhoon moving track; (<b>c</b>) represents the experiments where the center position is unchanged and the size of the verification area is enlarged or reduced; (<b>d</b>) represents experiments where the center position and size of the verification area have both changed).</p>
Full article ">Figure 3
<p>Time variation in typhoon forecast intensity (<b>a</b>) and typhoon forecast track (<b>b</b>) for each experiment (black line represents IBTrACS data, orange line represents ERA5 data, blue line represents CTRL experiment, red line represents SENS experiment; and green line represents OTHR experiment).</p>
Full article ">Figure 4
<p>The RMSE of different element forecasts for each experiment on 25 July 2021 at 06:00. (blue line represents CTRL experiment, red line represents SENS experiment, green line represents OTHR experiment; (<b>a</b>) is u-wind, (<b>b</b>) is v-wind; (<b>c</b>) is vertical wind, (<b>d</b>) is temperature; (<b>e</b>) is geopotential height; (<b>f</b>) is specific humidity).</p>
Full article ">Figure 5
<p>Distribution of sensitive areas at 24, 18, 12, and 06 h before the verification time (from Figure (<b>a</b>–<b>d</b>)) of Typhoon In-Fa. Overlay with the sea surface wind field of ensemble control forecast.</p>
Full article ">Figure 6
<p>Distribution of sensitive areas for targeted observation time of 06, 12, 18, and 24 h forecast (from Figure (<b>a</b>–<b>d</b>)). Overlay with the sea surface wind field of ensemble control forecast.</p>
Full article ">Figure 7
<p>Distribution of sensitive areas of each verification area experiment at 12 h before verification time overlaid with the sea surface wind field of ensemble control forecast (the red rectangle represents the verification area of each experiment; the verification area of experiment D5 exceeds the drawing range in this figure, which has been shown in <a href="#atmosphere-15-01335-f002" class="html-fig">Figure 2</a>d; and the numbers in the figure are the correlation coefficient between this experiment and control experiment A).</p>
Full article ">Figure 7 Cont.
<p>Distribution of sensitive areas of each verification area experiment at 12 h before verification time overlaid with the sea surface wind field of ensemble control forecast (the red rectangle represents the verification area of each experiment; the verification area of experiment D5 exceeds the drawing range in this figure, which has been shown in <a href="#atmosphere-15-01335-f002" class="html-fig">Figure 2</a>d; and the numbers in the figure are the correlation coefficient between this experiment and control experiment A).</p>
Full article ">Figure 8
<p>Bar chart of the correlation coefficient of sensitive signals between each verification area experiment and control experiment A of Typhoon In-Fa.</p>
Full article ">
18 pages, 10136 KiB  
Article
The Combination Application of FY-4 Satellite Products on Typhoon Saola Forecast on the Sea
by Chun Yang, Bingying Shi and Jinzhong Min
Remote Sens. 2024, 16(21), 4105; https://doi.org/10.3390/rs16214105 - 2 Nov 2024
Viewed by 604
Abstract
Satellite data play an irreplaceable role in global observation data systems. Effective comprehensive application of satellite products will inevitably improve numerical weather prediction. FengYun-4 (FY-4) series satellites can provide not only radiance data but also retrieval data with high temporal and spatial resolutions. [...] Read more.
Satellite data play an irreplaceable role in global observation data systems. Effective comprehensive application of satellite products will inevitably improve numerical weather prediction. FengYun-4 (FY-4) series satellites can provide not only radiance data but also retrieval data with high temporal and spatial resolutions. To evaluate the potential benefits of the combination application of FY-4 Advanced Geostationary Radiance Imager (AGRI) products on Typhoon Saola analysis and forecast, two group of experiments are set up with the Weather Research and Forecasting model (WRF). Compared with the benchmark experiment, whose sea surface temperature (SST) is from the National Centers for Environmental Prediction (NCEP) reanalysis data, the SST replacement experiments with FY-4 A/B SST products significantly improve the track and precipitation forecast, especially with the FY-4B SST product. Based on the above results, AGRI clear-sky and all-sky assimilations with FY-4B SST are implemented with a self-constructed AGRI assimilation module. The results show that the AGRI all-sky assimilation experiment can obtain better analyses and forecasts. Furthermore, it is proven that the combination application of AGRI radiance and SST products is beneficial for typhoon prediction. Full article
Show Figures

Figure 1

Figure 1
<p>The weighting function of channels 9-14 of FY-4A AGRI with RTTOV and the U.S. standard atmospheric profile.</p>
Full article ">Figure 2
<p>(<b>a</b>) The evolution of the best track, (<b>b</b>) the central sea level pressure (units: hPa) and maximum wind (units: knot) for Typhoon Saola from 0000 UTC 22 August to 1200 UTC 3 September 2023.</p>
Full article ">Figure 3
<p>Initial SST (units: K) from (<b>a</b>) <span class="html-italic">CON</span>, (<b>b</b>) SSTA, and (<b>c</b>) SSTB.</p>
Full article ">Figure 4
<p>The predicted (<b>a</b>) track, (<b>b</b>) track errors (units: km), (<b>c</b>) CSLP errors (units: hPa), and (<b>d</b>) MW errors (units: knot) in <span class="html-italic">CON</span> (light blue lines), SSTA (red lines), and SSTB (light green lines) are compared to the JMA best track estimates (blue lines) from 0600 UTC 30 August to 1200 UTC 2 September 2023.</p>
Full article ">Figure 5
<p>Time series of the U and V components of average steering flow (units: m/s) from 0600 UTC 30 August to 1200 UTC 2 September 2023.</p>
Full article ">Figure 6
<p>The 24 h accumulated precipitation (units: mm) from 1200 UTC 1 September to 1200 UTC 2 September 2023 of (<b>a</b>) the Micaps observation; (<b>b</b>) the interpolated Micaps observation with a horizontal resolution of 0.5° × 0.5°; (<b>c</b>) <span class="html-italic">CON</span>; (<b>d</b>) SSTA; and (<b>e</b>) SSTB. The dots with different colors in (<b>a</b>) represent different accumulated precipitation, as shown in the color bar.</p>
Full article ">Figure 7
<p>Performance diagram for the 24 h accumulated precipitation for the <span class="html-italic">CON</span> (light blue), SSTA (red), and SSTB (light green) with a threshold of (<b>a</b>) 0.01 mm; (<b>b</b>) 10 mm; (<b>c</b>) 25 mm; (<b>d</b>) 50 mm; and (<b>e</b>) 75 mm from 1200 UTC 1 September to 1200 UTC 2 September 2023.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>b</b>) The AGRI observed brightness temperature (units: K) distributions at channel 9 after QC in (<b>a</b>) CLR and (<b>b</b>) ALL valid at 1500 UTC 30 August 2023. (<b>c</b>) The counts of assimilated AGRI observations at channel 9 in ALL and CLR with different cloud mask types every 3 hr from 0900 UTC 30 August to 1500 UTC 30 August 2023.</p>
Full article ">Figure 9
<p>The IPs (units: %) over <span class="html-italic">CON</span> of individual experiments every 3 h from 0900 UTC 30 August to 1500 UTC 30 August 2023 in (<b>a</b>) <span class="html-italic">CTTs</span> and (<b>b</b>) agreements on sky conditions.</p>
Full article ">Figure 10
<p>The predicted (<b>a</b>) track, (<b>b</b>) track errors (units: km), (<b>c</b>) CSLP errors (units: hPa), and (<b>d</b>) MW errors (units: knot) in <span class="html-italic">CON</span> (light blue lines), SSTB (light green lines), CLR (light yellow lines), ALL (orange lines), CLR + SSTB (light red lines), and ALL + SSTB (brown lines) are compared to the JMA best track estimates (blue lines) from 1800 UTC 30 August to 1200 UTC 2 September 2023.</p>
Full article ">Figure 11
<p>The (<b>a</b>) U and (<b>b</b>) V components of steering flows (units: m/s) from 700 to 200 hPa with an interval of 50 hPa in individual experiments at 0600 UTC 1 September 2023.</p>
Full article ">Figure 12
<p>The same as <a href="#remotesensing-16-04105-f007" class="html-fig">Figure 7</a> but for (<b>a</b>) the interpolated Micaps observation with a horizontal resolution of 0.5° × 0.5°; (<b>b</b>) <span class="html-italic">CON</span>; (<b>c</b>) CLR; (<b>d</b>) ALL; (<b>e</b>) SSTB; (<b>f</b>) CLR + SSTB; and (<b>g</b>) ALL + SSTB.</p>
Full article ">Figure 13
<p>The same as <a href="#remotesensing-16-04105-f006" class="html-fig">Figure 6</a> but for <span class="html-italic">CON</span> (light blue), SSTB (light green), CLR (light yellow), ALL (orange), CLR + SSTB (light red), and ALL + SSTB (brown) with a threshold of (<b>a</b>) 0.01 mm; (<b>b</b>) 10 mm; (<b>c</b>) 25 mm; (<b>d</b>) 50 mm; and (<b>e</b>) 75 mm.</p>
Full article ">
20 pages, 14376 KiB  
Article
Impact of Directly Assimilating Radar Reflectivity Using a Reflectivity Operator Based on a Double-Moment Microphysics Scheme on the Analysis and Forecast of Typhoon Lekima (1909)
by Jingyao Luo, Hong Li, Yijie Zhu and Lijian Zhu
Remote Sens. 2024, 16(21), 3918; https://doi.org/10.3390/rs16213918 - 22 Oct 2024
Viewed by 633
Abstract
In previous studies, radar reflectivity is often directly assimilated using reflectivity operators based on a single-moment (SM) microphysics scheme, though the forecast model uses a double-moment (DM) microphysics scheme. With the fixed number concentrations, only the mixing ratios of hydrometeors are directly updated [...] Read more.
In previous studies, radar reflectivity is often directly assimilated using reflectivity operators based on a single-moment (SM) microphysics scheme, though the forecast model uses a double-moment (DM) microphysics scheme. With the fixed number concentrations, only the mixing ratios of hydrometeors are directly updated during the assimilation, which leads to a mismatch between the analyzed microphysical state variables and the microphysics scheme of the forecast model. In this study, the radar reflectivity is directly assimilated through an observation operator consistent with the DM Thompson microphysics scheme used in numerical integrations, and the impact of reflectivity operators based on SM and DM schemes on the analysis performance of the ensemble Kalman filter for typhoon Lekima on 9 August 2019 is evaluated. Reflectivity observations from a single operational weather radar in Wenzhou City, Zhejiang Province of China are assimilated. In addition, the dual-polarization observations from the same radar are used to evaluate the quality of the analysis. The analyzed reflectivity and dual-polarization characteristics obtained by different reflectivity operators are examined in detail. Compared with the experiments applying the reflectivity operator based on the SM Lin scheme, the use of a reflectivity operator consistent with the DM Thompson scheme adopted in the forecast model results in analyzed reflectivity and polarization characteristics that are more consistent with the observed characteristics in terms of general structure, location, and intensity. Forecasted reflectivity, 3 h accumulated precipitation, and typhoon intensity and track are also evaluated. The application of the reflectivity operator based on the DM scheme makes better forecasts of typhoon intensity, precipitation, and reflectivity, which also improves the forecast skills on typhoon tracks to a certain extent. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of the cycled data assimilation and 12 h forecast experiments.</p>
Full article ">Figure 2
<p>Root mean square innovations (unit: dBZ) of radar reflectivity of the background and the analysis of LIN and TM experiments. LIN_TM indicates that the reflectivity is calculated using the Thompson operator for LIN experiment.</p>
Full article ">Figure 3
<p>Composite reflectivity (unit: dBZ) by (<b>a</b>) the observation, the background of (<b>b</b>) LIN and (<b>c</b>) TM experiments, and the analysis fields of (<b>d</b>) LIN_TM (LIN analysis with the Thompson operator), (<b>e</b>) LIN, and (<b>f</b>) TM experiments at 1100 UTC on 9 August. The black dashed line in (<b>a</b>) represents the location of the vertical cross-section in <a href="#remotesensing-16-03918-f004" class="html-fig">Figure 4</a> and <a href="#remotesensing-16-03918-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 4
<p>Vertical cross-sections of reflectivity (unit: dBZ) by (<b>a</b>) the observation, the backgrounds of (<b>b</b>) LIN, and (<b>c</b>) TM experiments, and the analysis fields of (<b>d</b>) LIN_TM, (<b>e</b>) LIN, and (<b>f</b>) TM experiments along the line shown in <a href="#remotesensing-16-03918-f003" class="html-fig">Figure 3</a>a.</p>
Full article ">Figure 5
<p>Cross-sections of reflectivity components (shaded, unit: dBZ) and mixing ratios (black contour, unit: g kg<sup>−1</sup>) of (<b>a</b>–<b>d</b>) rain, (<b>e</b>–<b>h</b>) graupel, and (<b>i</b>–<b>l</b>) snow by the background of (<b>a</b>,<b>e</b>,<b>i</b>) LIN and (<b>b</b>,<b>f</b>,<b>j</b>) TM experiments, and the analysis fields of (<b>c</b>,<b>g</b>,<b>k</b>) LIN, and (<b>d</b>,<b>h</b>,<b>l</b>) TM experiments along the line illustrated in <a href="#remotesensing-16-03918-f003" class="html-fig">Figure 3</a>a. The starting values of the contours are 0.5 g kg<sup>−1</sup>, 0.01 g kg<sup>−1</sup>, and 0.1 g kg<sup>−1</sup> and the intervals are 0.8 g kg<sup>−1</sup>, 0.3 g kg<sup>−1</sup>, and 1.0 g kg<sup>−1</sup> for the mixing ratios of rain, graupel, and snow, respectively. The hot-pink contours in (<b>a</b>–<b>d</b>) represent the number concentration of rainwater with the starting value of 1.25 m<sup>3</sup> and an interval of 1.0 m<sup>3</sup>.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) Horizontal reflectivity (ZH, unit: dBZ), (<b>d</b>–<b>f</b>) differential phase (KDP, unit: ° km<sup>−1</sup>), and differential reflectivity (ZDR, unit: dB) by (<b>a</b>,<b>d</b>,<b>g</b>) the observation and the analysis fields of (<b>b</b>,<b>e</b>,<b>h</b>) LIN and (<b>c</b>,<b>f</b>,<b>i</b>) TM experiments at the vertical height of 5.5 km after first assimilation of reflectivity at 1100 UTC.</p>
Full article ">Figure 7
<p>Analysis increments of potential temperature (shaded, unit: °C) and horizontal wind speed increments (brown contour, starting from −2 m s<sup>−1</sup> and with an interval of 3 m s<sup>−1</sup>) obtained by the (<b>a</b>) LIN and (<b>b</b>) TM experiments. The analysis time is 1100 UTC on 9 August 2019.</p>
Full article ">Figure 8
<p>Horizontal reflectivity (unit: dBZ) by (<b>a</b>,<b>d</b>) the observation and the analysis fields of (<b>b</b>,<b>e</b>) LIN and (<b>c</b>,<b>f</b>) TM experiments at the vertical heights of (<b>a</b>–<b>c</b>) 3 km and (<b>d</b>–<b>f</b>) 5.5 km at 1200 UTC.</p>
Full article ">Figure 9
<p>Mean-mass diameters (shaded, unit: mm) and mixing ratios (contour) of (<b>a</b>,<b>b</b>) raindrops at the vertical height of 3 km and (<b>c</b>,<b>d</b>) graupels at the vertical height of 5.5 km by the analysis fields of (<b>a</b>,<b>c</b>) LIN and (<b>b</b>,<b>d</b>) TM experiments at 1200 UTC. The starting values and intervals of the mixing ratios of rainwater and graupel are all 1.0 g kg<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Zonal cross-sections of horizontal wind speed (shaded, unit: m s<sup>−1</sup>) and potential temperature (contour, unit: °C) through vortex center by (<b>a</b>) LIN and (<b>b</b>) TM experiments at 1200 UTC.</p>
Full article ">Figure 11
<p>Root mean squares of the hourly reflectivity forecasted by the LIN (orange) and TM (purple) experiments with a threshold of 15 dBZ.</p>
Full article ">Figure 12
<p>Twelve-hour forecast of (<b>a</b>) maximum wind speed and (<b>b</b>) track for the 12 h forecasts from 1200 UTC August to 0000 UTC 10 August 2019 by the LIN (orange lines) and TM (purple lines) experiments. The black lines denote the results of the Tropical Cyclone Best Track Dataset from the China Meteorological Administration.</p>
Full article ">Figure 13
<p>Three-hour accumulated precipitation from 1200 UTC to 1500 UTC on 9 August by (<b>a</b>) observations, (<b>b</b>) LIN experiment, and (<b>c</b>) TM experiment. The precipitation observations are the multi-source merged hourly analysis product with the resolution of 0.1° × 0.1° from CMA.</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 617
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 ">
20 pages, 6954 KiB  
Article
Typhoon Trajectory Prediction by Three CNN+ Deep-Learning Approaches
by Gang Lin, Yanchun Liang, Adriano Tavares, Carlos Lima and Dong Xia
Electronics 2024, 13(19), 3851; https://doi.org/10.3390/electronics13193851 - 28 Sep 2024
Viewed by 879
Abstract
The accuracy in predicting the typhoon track can be key to minimizing their frequent disastrous effects. This article aims to study the accuracy of typhoon trajectory prediction obtained by combining several algorithms based on current deep-learning techniques. The combination of a Convolutional Neural [...] Read more.
The accuracy in predicting the typhoon track can be key to minimizing their frequent disastrous effects. This article aims to study the accuracy of typhoon trajectory prediction obtained by combining several algorithms based on current deep-learning techniques. The combination of a Convolutional Neural Network with Long Short-Term Memory (CNN+LSTM), Patch Time-Series Transformer (CNN+PatchTST) and Transformer (CNN+Transformer) were the models chosen for this work. These algorithms were tested on the best typhoon track data from the China Meteorological Administration (CMA), ERA5 data from the European Centre for Medium-Range Weather Forecasts (ECMWF), and structured meteorological data from the Zhuhai Meteorological Bureau (ZMB) as an extension of existing studies that were based only on public data sources. The experimental results were obtained by testing two complete years of data (2021 and 2022), as an alternative to the frequent selection of a small number of typhoons in several years. Using the R-squared metric, results were obtained as significant as CNN+LSTM (0.991), CNN+PatchTST (0.989) and CNN+Transformer (0.969). CNN+LSTM without ZMB data can only obtain 0.987, i.e., 0.004 less than 0.991. Overall, our findings indicate that appropriately augmenting data near land and ocean boundaries around the coast improves typhoon track prediction. Full article
Show Figures

Figure 1

Figure 1
<p>Components of typhoon trajectory prediction system.</p>
Full article ">Figure 2
<p>Format of CMA best track data.</p>
Full article ">Figure 3
<p>The data relationship among CMA BST, ECMWF ERA5 in addition to a pseudo example of a ZMB configuration business data file.</p>
Full article ">Figure 4
<p>The whole process of reading business data.</p>
Full article ">Figure 5
<p>Typhoon trajectory segment samples with a length of nine track points.</p>
Full article ">Figure 6
<p>Key model classes and dataset classes in the typhoon prediction system.</p>
Full article ">Figure 7
<p>The overall process of trajectory prediction.</p>
Full article ">Figure 8
<p>The entire data flow of the application.</p>
Full article ">Figure 9
<p>A sample of a CNN layer.</p>
Full article ">Figure 10
<p>Forecast charts of some typhoon tracks in 2021 and 2022 on the scenario 12 by CNN+LSTM.</p>
Full article ">Figure 11
<p>Latitude and longitude prediction errors between 2021 and 2022 by CNN+LSTM.</p>
Full article ">
27 pages, 13088 KiB  
Article
Effects of Surface Layer Physics Schemes on the Simulated Intensity and Structure of Typhoon Rai (2021)
by Thi-Huyen Hoang, Ching-Yuang Huang and Thi-Chinh Nguyen
Atmosphere 2024, 15(9), 1140; https://doi.org/10.3390/atmos15091140 - 20 Sep 2024
Viewed by 652
Abstract
The influences of surface layer (SL) physics schemes on the simulated intensity and structure of Typhoon Rai (2021) are investigated using the WRF model. Numerical experiments using different SL physics schemes—revised MM5 scheme (MM5), Eta similarity scheme (CTL), and Mellor–Yamada–Nakanishi–Niino scheme (MYNN)—are conducted. [...] Read more.
The influences of surface layer (SL) physics schemes on the simulated intensity and structure of Typhoon Rai (2021) are investigated using the WRF model. Numerical experiments using different SL physics schemes—revised MM5 scheme (MM5), Eta similarity scheme (CTL), and Mellor–Yamada–Nakanishi–Niino scheme (MYNN)—are conducted. The results show that the intensity forecast of Typhoon Rai is largely influenced by SL physics schemes, while its track forecast is not significantly affected. All three experiments can successfully capture the movement of Rai, while CTL provides better intensity simulation compared to the other two experiments. The higher ratio of enthalpy exchange coefficient to drag coefficient (CK/CD) in CTL than MM5 and MYNN leads to significantly increased surface enthalpy fluxes, which are crucial for the typhoon intensification of the former. To explore the influence of SL physics on the structural evolution of the typhoon, the azimuthal-mean angular momentum (AM) budget is utilized. The results indicate that asymmetric eddy terms may also largely contribute to the AM tendencies, which are relatively more comparable in the weaker TC for MM5, compared to the stronger TC with the dominant symmetric mean terms for CTL. Furthermore, the extended Sawyer–Eliassen (SE) equation is solved to quantify the transverse circulations of the typhoon induced by different forcing sources for CTL and MM5. The SE solution indicates that the transverse circulation above and within the boundary layer is predominantly induced by diabatic heating and turbulent friction, respectively, for both CTL and MM5, while all other physical forcing terms are relatively insignificant for the induced transverse circulation for CTL, except for the large contribution from the eddy forcing in the upper-tropospheric outflow for MM5. With the stronger connective heating in the eyewall and boundary-layer radial inflow, the linear SE analysis agrees much better with the nonlinear simulation for CTL than MM5. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>The WRF model domains for Rai at the initial time. The outermost box (d01) denotes the outermost domain, while the red and blue boxes (d02 and d03, respectively) denote the two inner moving domains. The dashed black line (JMA) with cycles at intervals of 24 h indicates the best track from JMA from 0000 UTC 14 December to 0000 UTC 18 December 2021.</p>
Full article ">Figure 2
<p>(<b>a</b>) Tracks of Typhoon Rai, including the best track data from JTWC (dashed black line) and JMA (solid black line), as well as simulated tracks for CTL (red line), MM5 (blue line), and MYNN (green line), during the period from 0000 UTC 14 December to 0000 UTC 18 December 2021. Circle symbols in (<b>a</b>) indicate the time every 24 h. (<b>b</b>) as in (<b>a</b>), but for the 10-m maximum wind speed (V<sub>max</sub>, m s<sup>−1</sup>).</p>
Full article ">Figure 3
<p>12-h accumulated precipitation (mm) during 48–60 h from (<b>a</b>) multi-satellite precipitation product GSMaP, (<b>b</b>) CTL, (<b>c</b>) MM5, and (<b>d</b>) MYNN. (<b>e</b>), (<b>f</b>), (<b>g</b>), and (<b>h</b>) as in (<b>a</b>), (<b>b</b>), (<b>c</b>), and (<b>d</b>), respectively, but during 60–72 h. (<b>i</b>), (<b>j</b>), (<b>k</b>), and (<b>l</b>) as in (<b>a</b>), (<b>b</b>), (<b>c</b>), and (<b>d</b>), respectively, but during 72–84 h.</p>
Full article ">Figure 4
<p>Simulated ratio of enthalpy exchange coefficient to drag coefficient (C<sub>K</sub>/C<sub>D</sub>) as a function of 10-m wind speed for CTL (red line), MM5 (blue line), and MYNN (green line) at 54 h.</p>
Full article ">Figure 5
<p>Time evolutions of (<b>a</b>) friction velocity (m s<sup>−1</sup>), (<b>b</b>) surface sensible heat flux (W m<sup>−2</sup>), and (<b>c</b>) surface latent heat flux (W m<sup>−2</sup>) for CTL (red line), MM5 (blue line), and MYNN (green line), averaged within the area of 300 × 300 km around the typhoon center from 24 h to 72 h.</p>
Full article ">Figure 6
<p>Horizontal distribution of 10-m wind speed (shaded color, m s<sup>−1</sup>) for (<b>a</b>) CTL, (<b>b</b>) MM5, and (<b>c</b>) MYNN at 54 h. (<b>d</b>), (<b>e</b>), and (<b>f</b>) as in (<b>a</b>), (<b>b</b>), and (<b>c</b>), respectively, but for friction velocity (m s<sup>−1</sup>). (<b>g</b>), (<b>h</b>), and (<b>i</b>) as in (<b>a</b>), (<b>b</b>), and (<b>c</b>), respectively, but for surface sensible heat flux (W m<sup>−2</sup>). (<b>j</b>), (<b>k</b>), and (<b>l</b>) as in (<b>a</b>), (<b>b</b>), and (<b>c</b>), respectively, but for surface latent heat flux (W m<sup>−2</sup>). The black vectors in (<b>a</b>–<b>c</b>) denote the 10-m horizontal wind speed.</p>
Full article ">Figure 7
<p>Azimuthal-mean tangential velocity (m s<sup>−1</sup>) in the radius–height cross-section at 54 h for (<b>a</b>) CTL and (<b>b</b>) MM5. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for the radial velocity (m s<sup>−1</sup>). The black line in (<b>a</b>,<b>b</b>) represents the height of the maximum tangential wind speed (h<sub>vt</sub>). The black line in (<b>c</b>,<b>d</b>) represents the inflow layer depth (h<sub>vr</sub>).</p>
Full article ">Figure 8
<p>Time–radius Hovmöller diagrams of azimuthal-mean tangential wind (m s<sup>−1</sup>) at 2-km height for (<b>a</b>) CTL and (<b>b</b>) MM5 from 24 h to 72 h. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for radial wind (m s<sup>−1</sup>) at 0.25-km height. The black line in (<b>a</b>,<b>b</b>) represents the radius of maximum tangential wind speed (RMW) at 2 km height. The green line in (<b>c</b>,<b>d</b>) represents the maximum inflow at 0.25-km height.</p>
Full article ">Figure 9
<p>Azimuthal-mean potential temperature (K) in the radius–height cross-section for (<b>a</b>) CTL and (<b>b</b>) MM5 at 54 h. The thick black line represents the RMW.</p>
Full article ">Figure 10
<p>Time–height Hovmöller diagrams of azimuthal-mean potential temperature (K) inside RMW for (<b>a</b>) CTL and (<b>b</b>) MM5. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for warm core (K) averaged inside a radius of 1.5 degrees from the typhoon center.</p>
Full article ">Figure 11
<p>Azimuthal-mean flow (shaded color) in the radius–height cross-section of radial velocity (m s<sup>−1</sup>) for (<b>a</b>) CTL and (<b>b</b>) MM5 at 54 h. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for tangential velocity (m s<sup>−1</sup>). (<b>e</b>) and (<b>f</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for vertical velocity (m s<sup>−1</sup>). (<b>g</b>) and (<b>h</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but for latent heating rate (K h<sup>−1</sup>). The green line represents the RMW. The reference vector in the lower right corner represents the radial and vertical wind components.</p>
Full article ">Figure 12
<p>Radius–height cross-sections of azimuthal-mean angular momentum (AAM) (shaded color, 10<sup>6</sup> m<sup>2</sup> s<sup>−1</sup>) for (<b>a</b>) CTL and (<b>b</b>) MM5 at 54 h. The green line represents the RMW. The reference vector in the lower right corner represents the radial and vertical wind components.</p>
Full article ">Figure 13
<p>Radius–height cross-section of azimuthal-mean AM budget terms (shaded color, m<sup>2</sup> s<sup>−2</sup>), including (<b>a</b>) radial advection of mean AM, (<b>b</b>) radial advection of eddy AM, (<b>c</b>) vertical advection of mean AM, (<b>d</b>) vertical advection of eddy AM, (<b>e</b>) mean Coriolis force term, and (<b>f</b>) sum of all AM budget terms for CTL at 54 h. The green line represents the RMW. The reference vector in the lower right corner represents the radial and vertical wind components.</p>
Full article ">Figure 14
<p>As in <a href="#atmosphere-15-01140-f013" class="html-fig">Figure 13</a>, but for MM5 including (<b>a</b>) radial advection of mean AM, (<b>b</b>) radial advection of eddy AM, (<b>c</b>) vertical advection of mean AM, (<b>d</b>) vertical advection of eddy AM, (<b>e</b>) mean Coriolis force term, and (<b>f</b>) sum of all AM budget terms for CTL at 54 h. The green line represents the RMW. The reference vector in the lower right corner represents the radial and vertical wind components.</p>
Full article ">Figure 15
<p>Azimuthal-mean radial velocity (shaded colors, m s<sup>−1</sup>) at 54 h from the Sawyer–Eliassen (SE) solution with total forcing sources for (<b>a</b>) CTL and (<b>b</b>) MM5. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with symmetric diabatic heating only. (<b>e</b>) and (<b>f</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with asymmetric eddy momentum and heating only. (<b>g</b>) and (<b>h</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with turbulent momentum diffusion only. (<b>i</b>) and (<b>j</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with residual terms only. The wind vectors (m s<sup>−1</sup>) induced by the total forcing sources overlapped in each panel indicate the radial and vertical wind components (m s<sup>−1</sup>) with their reference vectors given at the lower right corner.</p>
Full article ">Figure 16
<p>Azimuthal-mean vertical velocity (shaded colors, m s<sup>−1</sup>) at 54 h from the Sawyer–Eliassen (SE) solution with total forcing sources for (<b>a</b>) CTL and (<b>b</b>) MM5. (<b>c</b>) and (<b>d</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with symmetric diabatic heating only. (<b>e</b>) and (<b>f</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with asymmetric eddy momentum and heating only. (<b>g</b>) and (<b>h</b>) as in (<b>a</b>) and (<b>b</b>), respectively, but with residual terms only. The wind vectors (m s<sup>−1</sup>) induced by the total forcing sources overlapped in each panel indicate the radial and vertical wind components (m s<sup>−1</sup>) with their reference vectors given at the lower right corner.</p>
Full article ">
27 pages, 18384 KiB  
Article
Calibration of Typhoon Track Forecasts Based on Deep Learning Methods
by Chengchen Tao, Zhizu Wang, Yilun Tian, Yaoyao Han, Keke Wang, Qiang Li and Juncheng Zuo
Atmosphere 2024, 15(9), 1125; https://doi.org/10.3390/atmos15091125 - 17 Sep 2024
Viewed by 781
Abstract
An accurate forecast of typhoon tracks is crucial for disaster warning and mitigation. However, existing numerical weather prediction models, such as the Weather Research and Forecasting (WRF) model, still exhibit significant errors in track forecasts. This study aims to improve forecast accuracy by [...] Read more.
An accurate forecast of typhoon tracks is crucial for disaster warning and mitigation. However, existing numerical weather prediction models, such as the Weather Research and Forecasting (WRF) model, still exhibit significant errors in track forecasts. This study aims to improve forecast accuracy by correcting WRF-forecasted tracks using deep learning models, including Bidirectional Long Short-Term Memory (BiLSTM) + Convolutional Long Short-Term Memory (ConvLSTM) + Wide and Deep Learning (WDL), BiLSTM + Convolutional Gated Recurrent Unit (ConvGRU) + WDL, and BiLSTM + ConvLSTM + Extreme Deep Factorization Machine (xDeepFM), with a comparison to the Kalman Filter. The results demonstrate that the BiLSTM + ConvLSTM + WDL model reduces the 72 h track prediction error (TPE) from 255.18 km to 159.23 km, representing a 37.6% improvement over the original WRF model, and exhibits significant advantages across all evaluation metrics, particularly in key indicators such as Bias2, Mean Squared Error (MSE), and Sequence. The decomposition of MSE further validates the importance of the BiLSTM, ConvLSTM, WDL, and Temporal Normalization (TN) layers in enhancing the model’s spatio-temporal feature-capturing ability. Full article
(This article belongs to the Special Issue Applications of Artificial Intelligence in Atmospheric Sciences)
Show Figures

Figure 1

Figure 1
<p>A visualization of all forecasted typhoons. Adapted from the figure by Xu [<a href="#B34-atmosphere-15-01125" class="html-bibr">34</a>] et al.</p>
Full article ">Figure 2
<p>The three-dimensional time-series structure of a typhoon.</p>
Full article ">Figure 3
<p>Methodological Workflow Diagram.</p>
Full article ">Figure 4
<p>Typhoon Track Prediction Model Based on the Wide and Deep Framework.</p>
Full article ">Figure 5
<p>Typhoon Track Prediction Model Based on the Neural Factorization Machine Framework.</p>
Full article ">Figure 6
<p>Typhoon Track Prediction Model Based on the Extreme Deep Factorization Machine Framework.</p>
Full article ">Figure 7
<p>Comparative Analysis of Typhoon TPE and Latitude–Longitude RMSE at Different Integration Times in WRF Forecast Results.</p>
Full article ">Figure 8
<p>Temporal Trends of WRF-Forecasted Typhoon TPE from 2000 to 2022.</p>
Full article ">Figure 9
<p>Scatter Plot Comparing Latitude and Longitude of WRF Forecasts with Best Track at Different Integration Times; the red line indicates the regression line, reflecting the linear relationship between observed and predicted values. (<b>a</b>–<b>c</b>) represent scatter plots for 72 h, 48 h, and 24 h Integration Times along the longitude, and (<b>a1</b>–<b>c1</b>) represent the same along the latitude.</p>
Full article ">Figure 10
<p>Comparative Analysis of Overall MSE ((<b>a</b>,<b>a1</b>), Unit: °<sup>2</sup>), Bias<sup>2</sup> ((<b>b</b>,<b>b1</b>), Unit: °<sup>2</sup>), Distribution ((<b>c</b>,<b>c1</b>), Unit: °<sup>2</sup>), and Sequence ((<b>d</b>,<b>d1</b>), Unit: °<sup>2</sup>) for Latitude and Longitude Directions Across Different Models on the 2018–2022 Test Set. The top panels show metrics for the latitude direction, while the bottom panels show metrics for the longitude direction. The models labeled in the figure are BiLSTM + ConvGRU + WDL (A), BiLSTM + ConvLSTM + WDL (B), WRF (C), BiLSTM + ConvLSTM + xDeepFM (D), and Kalman Filter (E). In each subplot, the curves of different colors correspond to 72-h, 48-h, and 24-h integration times. (<b>a</b>–<b>d</b>) represent latitudinal variables, and (<b>a1</b>–<b>d1</b>) represent longitudinal variables.</p>
Full article ">Figure 11
<p>Spatial Distribution Comparison of MSE (Unit: °<sup>2</sup>) for Latitude and Longitude Directions Across BiLSTM + ConvGRU + WDL, BiLSTM + ConvLSTM + WDL, BiLSTM + ConvLSTM + DeepFM, WRF, and Kalman Filter on the 2018–2022 Typhoon Test Set. The (<b>top panels</b>) show MSE distribution for latitude, while the (<b>bottom panels</b>) show MSE distribution for longitude.</p>
Full article ">Figure 12
<p>Spatial Distribution Comparison of Bias<sup>2</sup> (Unit: °<sup>2</sup>) for Latitude and Longitude Directions Across BiLSTM + ConvGRU + WDL, BiLSTM + ConvLSTM + WDL, BiLSTM + ConvLSTM + DeepFM, WRF, and Kalman Filter on the 2018–2022 Typhoon Test Set. The (<b>top panels</b>) show Bias<sup>2</sup> distribution for latitude, while the (<b>bottom panels</b>) show Bias<sup>2</sup> distribution for longitude.</p>
Full article ">Figure 13
<p>Spatial Distribution Comparison of Distribution (Unit: °<sup>2</sup>) for Latitude and Longitude Directions Across BiLSTM + ConvGRU + WDL, BiLSTM + ConvLSTM + WDL, BiLSTM + ConvLSTM + DeepFM, WRF, and Kalman Filter on the 2018–2022 Typhoon Test Set. The (<b>top panels</b>) show the distribution metric for latitude, while the (<b>bottom panels</b>) show the distribution metric for longitude.</p>
Full article ">Figure 14
<p>Spatial Distribution Comparison of Sequence (Unit: °<sup>2</sup>) for Latitude and Longitude Directions Across BiLSTM + ConvGRU + WDL, BiLSTM + ConvLSTM + WDL, BiLSTM + ConvLSTM + DeepFM, WRF, and Kalman Filter on the 2018–2022 Typhoon Test Set. The (<b>top panels</b>) show the sequence metric for latitude, while the (<b>bottom panels</b>) show the sequence metric for longitude.</p>
Full article ">Figure 15
<p>Comparison of Track Prediction Performance in Latitude and Longitude Directions Between the WRF Model and Correction Models (BiLSTM + ConvLSTM + WDL and BiLSTM + ConvGRU + WDL) on the 2018–2022 Typhoon Test Set. The red line indicates the regression line, reflecting the linear relationship between observed and predicted values; the red shaded area represents the 95% confidence interval of the regression line.</p>
Full article ">Figure 16
<p>Path Comparison of Typhoon In-fa (2021), Showing the Differences Between the WRF, BiLSTM + ConvLSTM + WDL, and BiLSTM + ConvGRU + WDL Methods and Historical Typhoon Tracks. Adapted from the figure by Xu [<a href="#B34-atmosphere-15-01125" class="html-bibr">34</a>] et al.</p>
Full article ">Figure 17
<p>Path Comparison of Typhoon Chanthu (2021), Showing the Differences Between the WRF, BiLSTM + ConvLSTM + WDL, and BiLSTM + ConvGRU + WDL Methods and Historical Typhoon Tracks. Adapted from the figure by Xu [<a href="#B34-atmosphere-15-01125" class="html-bibr">34</a>] et al.</p>
Full article ">
27 pages, 14463 KiB  
Article
Numerical Investigation of Track and Intensity Evolution of Typhoon Doksuri (2023)
by Dieu-Hong Vu, Ching-Yuang Huang and Thi-Chinh Nguyen
Atmosphere 2024, 15(9), 1105; https://doi.org/10.3390/atmos15091105 - 11 Sep 2024
Viewed by 630
Abstract
This study utilized the WRF model to investigate the track evolution and rapid intensification (RI) of Typhoon Doksuri (2023) as it moved across the Luzon Strait and through the South China Sea (SCS). The simulation results indicate that Doksuri has a smaller track [...] Read more.
This study utilized the WRF model to investigate the track evolution and rapid intensification (RI) of Typhoon Doksuri (2023) as it moved across the Luzon Strait and through the South China Sea (SCS). The simulation results indicate that Doksuri has a smaller track sensitivity to the use of different physics schemes, while having a greater intensity sensitivity. Sensitivity numerical experiments with different physics schemes can well capture its northwestward movement in the first two days, but they predict less westward track deflection as the typhoon moves across the Luzon Strait and through the SCS. Moreover, all the experiments successfully simulated Doksuri’s RI, albeit with quite different rates and a time lag of 12 h. Among different combinations of physics schemes, there exists an optimal set of cumulus parameterization and cloud microphysics schemes for track and intensity predictions. Doksuri’s track changes as the typhoon moved across the Luzon Strait and through the SCS were influenced by the topographic effects of the terrain of the Philippines and Taiwan, to different extents. The track changes of Doksuri are explained by the wavenumber-one potential vorticity (PV) tendency budget from different physical processes, highlighting that the horizontal PV advection dominates the PV tendency throughout most of the simulation time due to the offset of vertical PV advection and differential diabatic heating. In addition, this study applies the extended Sawyer–Eliassen (SE) equation to compare the transverse circulations of the typhoon induced by various forcing sources. The SE solution indicates that radial inflow was largely driven in the lower-tropospheric vortex by strong diabatic heating, while being significantly enhanced in the lower boundary layer due to turbulent friction. All other physical forcing terms were relatively insignificant for the induced transverse circulation. The coordinated radial inflow at low levels may have led to the eyewall development in unbalanced dynamics. Intense diabatic heating thus was vital to the severe RI of Doksuri under a weak vertical wind shear. Full article
(This article belongs to the Special Issue Typhoon/Hurricane Dynamics and Prediction (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>The two nested domains of the WRF model for Doksuri during the simulation time. The outermost box and white box denote the outer and the inner domains, respectively. The black line with color dots at intervals of 12 h indicates the best track from the IBTrACS from 1200 UTC 23 July to 1200 UTC 28 July 2023. The color of the dots represents different typhoon intensity categories according to the Saffir–Simpson scale for Typhoon Doksuri (2023).</p>
Full article ">Figure 2
<p>Tracks of Typhoon Doksuri including the best track data from IBTrACS (solid black line) and JMA (dashed black line) as well as simulated tracks from sensitivity experiments that combined the CPSs (<b>a</b>) KF, (<b>b</b>) GF, (<b>c</b>) GD, and (<b>d</b>) New Tiedtke with different MPSs (solid colored lines) during the period from 1200 UTC 23 July to 1200 UTC 28 July 2023 (0 to 120 forecast hours). The circle symbols indicate the time every 24 h. (<b>e</b>–<b>h</b>) as in (<b>a</b>–<b>d</b>), respectively, but for the 10 m maximum wind speed (V<sub>max</sub>, m s<sup>−1</sup>). The blue, green, magenta, yellow, red, and cyan lines denote the cloud microphysics schemes Lin, WSM6, Goddard, Thompson, NSSL2, and P3, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Track of Typhoon Doksuri including the best track data from IBTrACS (solid black line) and JMA (dashed black line) as well as simulated tracks from the sensitivity experiments using the GF cumulus schemes combined with nine different MPSs (solid colored lines) during the period from 1200 UTC 23 July to 1200 UTC 28 July 2023 (0 to 120 forecast hours). The circle symbols indicate the time every 24 h. The 9 different MPSs are listed in <a href="#atmosphere-15-01105-t001" class="html-table">Table 1</a>. (<b>b</b>) as in (<b>a</b>) but for the 10 m maximum wind speed (V<sub>max</sub>, m s<sup>−1</sup>). (<b>c</b>) as in (<b>a</b>) but for the track error (km) for all the sensitivity experiments combining the GF cumulus scheme with the fifteen different MPSs, as noted in <a href="#atmosphere-15-01105-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Accumulated hourly precipitation (mm) from GSMaP in (<b>a</b>) 12 h (from 0000 UTC to 0059 UTC 24 July, (<b>b</b>) 24 h (from 1200 UTC to 1259 UTC 24 July), (<b>c</b>) 36 h (from 0000 UTC to 0059 UTC 25 July), and (<b>d</b>) 48 h (from 1200 UTC to 1259 UTC 25 July). (<b>e</b>–<b>h</b>) as in (<b>a</b>–<b>d</b>), respectively, but for simulated rainfall (mm) for CTL.</p>
Full article ">Figure 5
<p>The simulated horizontal wind speed (shaded colors, m s<sup>−1</sup>) averaged in 1–8 km height for CTL at (<b>a</b>) 36 h, (<b>b</b>) 48 h, (<b>c</b>) 60 h, and (<b>d</b>) 72 h. (<b>e</b>–<b>h</b>) as in (<b>a</b>–<b>d</b>) respectively, but for vertical velocity (colors shaded, m s<sup>−1</sup>). The vector shows the simulated horizontal wind averaged in 1–8 km height.</p>
Full article ">Figure 6
<p>Hovmöller plots of (<b>a</b>) horizontal wind speed (m s<sup>−1</sup>), (<b>b</b>) vertical velocity (m s<sup>−1</sup>), and (<b>c</b>) diabatic heating (K h<sup>−1</sup>) averaged in 1–8 km height and radii of 0.5°–1.5° concerning azimuth for Doksuri.</p>
Full article ">Figure 7
<p>Time evolution of the translation velocity (vectors, m s<sup>−1</sup>) induced by different PV budget terms, including differential diabatic heating (HDIA), vertical PV advection (VADV), horizontal PV advection (HADV), and the sum of the former three terms (SUM), during the forecast time of 120 h for CTL.</p>
Full article ">Figure 8
<p>Simulated tracks for CTL (red line), NoTW (blue line), NoPhi (green line), NoPhi-TW (magenta line), and the best track from IBTrACS (black line). The simulated time was from 1200 UTC 23 July to 1200 UTC 28 July 2023. The circle symbols indicate the time every 24 h.</p>
Full article ">Figure 9
<p>As in <a href="#atmosphere-15-01105-f007" class="html-fig">Figure 7</a> but for (<b>a</b>) NoTW and (<b>b</b>) NoPhi.</p>
Full article ">Figure 10
<p>LHF (shaded colors, W m<sup>−2</sup>) for CTL at (<b>a</b>) 3 h, (<b>b</b>) 12 h, (<b>c</b>) 18 h, (<b>d</b>) 24 h, (<b>e</b>) 36 h, and (<b>f</b>) 48 h. Solid blue circles mark the 100 and 200 km radii from the typhoon center. The vector at the typhoon center denotes the VWS averaged within the radius of 0–800 km. The number in the top-right inset of each panel denotes the magnitude of VWS.</p>
Full article ">Figure 11
<p>Azimuthal-mean diabatic heating (shaded colors, K h<sup>−1</sup>) and tangential wind (contours, at intervals of 5 m s<sup>−1</sup>) for CTL at (<b>a</b>) 3 h, (<b>b</b>) 12 h, (<b>c</b>) 18 h, (<b>d</b>) 24 h, (<b>e</b>) 36 h, and (<b>f</b>) 48 h.</p>
Full article ">Figure 12
<p>Azimuthal-mean radial velocity (shaded colors, m s<sup>−1</sup>) at 24 h from (<b>a</b>) the nonlinear simulation, (<b>b</b>) the SE solution with the total sources, (<b>c</b>) as in (<b>b</b>) but with symmetric diabatic heating only, (<b>d</b>) as in (<b>b</b>) but with turbulent momentum diffusion only, (<b>e</b>) as in (<b>b</b>) but with asymmetric eddy heating only, (<b>f</b>) as in (<b>b</b>) but with asymmetric eddy momentum only, (<b>g</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>U</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>W</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only, (<b>h</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only. The wind vectors (m s<sup>−1</sup>) induced by the total forcing sources overlapped in each panel indicate the radial and vertical wind components (m s<sup>−1</sup>) with their reference vectors given at the lower right corner.</p>
Full article ">Figure 13
<p>As in <a href="#atmosphere-15-01105-f012" class="html-fig">Figure 12</a> but for vertical velocity (m s<sup>−1</sup>) at 24 h for (<b>a</b>) the nonlinear simulation, (<b>b</b>) the SE solution with the total sources, (<b>c</b>) as in (<b>b</b>) but with symmetric diabatic heating only, (<b>d</b>) as in (<b>b</b>) but with turbulent momentum diffusion only, (<b>e</b>) as in (<b>b</b>) but with asymmetric eddy heating only, (<b>f</b>) as in (<b>b</b>) but with asymmetric eddy momentum only, (<b>g</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>U</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>W</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only, (<b>h</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only.</p>
Full article ">Figure 14
<p>As in <a href="#atmosphere-15-01105-f012" class="html-fig">Figure 12</a> but at 48 h for (<b>a</b>) the nonlinear simulation, (<b>b</b>) the SE solution with the total sources, (<b>c</b>) as in (<b>b</b>) but with symmetric diabatic heating only, (<b>d</b>) as in (<b>b</b>) but with turbulent momentum diffusion only, (<b>e</b>) as in (<b>b</b>) but with asymmetric eddy heating only, (<b>f</b>) as in (<b>b</b>) but with asymmetric eddy momentum only, (<b>g</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>U</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>W</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only, (<b>h</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only.</p>
Full article ">Figure 15
<p>As in <a href="#atmosphere-15-01105-f013" class="html-fig">Figure 13</a> but for vertical velocity (m s<sup>−1</sup>) at 48 h for (<b>a</b>) the nonlinear simulation, (<b>b</b>) the SE solution with the total sources, (<b>c</b>) as in (<b>b</b>) but with symmetric diabatic heating only, (<b>d</b>) as in (<b>b</b>) but with turbulent momentum diffusion only, (<b>e</b>) as in (<b>b</b>) but with asymmetric eddy heating only, (<b>f</b>) as in (<b>b</b>) but with asymmetric eddy momentum only, (<b>g</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>U</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>W</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only, (<b>h</b>) as in (<b>b</b>) but with <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>V</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> only.</p>
Full article ">Figure 16
<p>Contoured frequency by altitude diagrams (CFAD, %) of vertical velocity within a radius of 150 km from the typhoon center at 3 h for (<b>a</b>) WDM6 and (<b>b</b>) NSSL2. (<b>c</b>,<b>d</b>) as in (<b>a</b>,<b>b</b>), respectively, but at 24 h.</p>
Full article ">Figure 17
<p>Vertical profiles of hydrometeors (10<sup>−3</sup> kg kg<sup>−1</sup>) averaged in a 200 km radius of the typhoon center for (<b>a</b>) cloud water mixing ratio; (<b>b</b>) rainwater mixing ratio; (<b>c</b>) total of ice mixing ratio, snow mixing ratio, and graupel mixing ratio at 3 h. (<b>d</b>–<b>f</b>) as in (<b>a</b>–<b>c</b>), respectively, but at 24 h.</p>
Full article ">
19 pages, 10160 KiB  
Article
Performance Evaluation of TGFS Typhoon Track Forecasts over the Western North Pacific with Sensitivity Tests on Cumulus Parameterization
by Yu-Han Chen, Sheng-Hao Sha, Chang-Hung Lin, Ling-Feng Hsiao, Ching-Yuang Huang and Hung-Chi Kuo
Atmosphere 2024, 15(9), 1075; https://doi.org/10.3390/atmos15091075 - 5 Sep 2024
Viewed by 830
Abstract
This study employed the new generation Taiwan global forecast system (TGFS) to focus on its performance in forecasting the tracks of western North Pacific typhoons during 2022–2023. TGFS demonstrated better forecasting performance in typhoon track compared to central weather administration (CWA) GFS. For [...] Read more.
This study employed the new generation Taiwan global forecast system (TGFS) to focus on its performance in forecasting the tracks of western North Pacific typhoons during 2022–2023. TGFS demonstrated better forecasting performance in typhoon track compared to central weather administration (CWA) GFS. For forecasts with large track errors by TGFS at the 120th h, it was found that most of them originated during the early stages of typhoon development when the typhoons were of mild intensity. The tracks deviated predominantly towards the northeast and occasionally towards the southwest, which were speculated to be due to inadequate environmental steering guidance resulting from the failure to capture synoptic environmental features. The tracks could be corrected by replacing the original new simplified Arakawa–Schubert (NSAS) scheme with the new Tiedtke (NTDK) scheme to change the synoptic environmental field, not only for Typhoon Khanun, which occurred in the typhoon season of 2023, but also for Typhoon Bolaven, which occurred after the typhoon season, in October 2023, under atypical circulation characteristics over the western Pacific. The diagnosis of vorticity budget primarily analyzed the periods where divergence in typhoon tracks between control (CTRL) and NTDK experiments occurred. The different synoptic environmental fields in the NTDK experiment affected the wavenumber-1 vorticity distribution in the horizontal advection term, thereby enhancing the accuracy of typhoon translation velocity forecasts. This preliminary study suggests that utilizing the NTDK scheme might improve the forecasting skill of TGFS for typhoon tracks. To gain a more comprehensive understanding of the impact of NTDK on typhoon tracks, further examination for more typhoons is still in need. Full article
(This article belongs to the Special Issue Typhoon/Hurricane Dynamics and Prediction (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Track errors for the western North Pacific typhoons during 2022–2023 in TGFS (red), NCEP GFS (yellow), IFS (blue), and CWAGFS (green). The number of cases at each forecast hour was marked in brackets on the horizontal axis.</p>
Full article ">Figure 2
<p>Differences in longitude and latitude of TGFS typhoon centers relative to the best track (blue) at the 120th forecast hour. Red (black) represents the track error was greater (less) than the mean error. Solid dot represents forecasts initiated under mild typhoon category conditions (17.2 to 32.6 m s<sup>−1</sup>), while hollow diamond represents forecasts with mature typhoon phase at the forecast initiation.</p>
Full article ">Figure 3
<p>The 500 hPa geopotential height (contour) and wind (arrow, m s<sup>−1</sup>) fields averaged from June to October in 2022 and 2023 by (<b>a</b>) ERA5 reanalysis data, (<b>b</b>) TGFS, (<b>c</b>) NCEP GFS, and (<b>d</b>) IFS. The thick black line represents 5880 gpm. The thick red line in (<b>b</b>–<b>d</b>) represents 5880 gpm by ERA5 as (<b>a</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) Track errors for the selected DTGs of Typhoon Doksuri (2023), Khanun (2023), Lan (2023), and Bolaven (2023) in CTRL (red), NCEP GFS (yellow), IFS (blue), and NTDK experiment (green). The number of cases at each forecast hour was marked in brackets on the horizontal axis. (<b>b</b>) Mean typhoon track error differences between CTRL and NTDK experiments (bars). Error bars denote the 95% confidence interval of the mean difference.</p>
Full article ">Figure 5
<p>Model tracks of total 120 h for the selected DTGs (<a href="#atmosphere-15-01075-t002" class="html-table">Table 2</a>) in CTRL (red) and NTDK (green) experiments and the best track of CWA (black) of Typhoon Doksuri (2023), Khanun (2023), Lan (2023), and Bolaven (2023). The black line starts from the initial time of the first DTG and ends at the last time of the last DTG of each typhoon.</p>
Full article ">Figure 6
<p>Model tracks of total 120 h in CTRL (red) and NTDK (green) experiments and the best track of CWA (black: during the 120 h forecasts; grey: out of the forecast period) from 1800 UTC, 28 July 2023 of Typhoon Khanun (2023). Dots were marked every six hours.</p>
Full article ">Figure 7
<p>The 500 hPa geopotential height (red), streamline (black), and 12 h accumulated precipitation (shaded, mm) fields in (<b>a</b>) CTRL and (<b>b</b>) NTDK experiments at 1800 UTC 30 July 2023 (the 48th forecast hour). The thick red contour represents 5880 gpm. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for 0600 UTC, 31 July 2023 (the 60th forecast hour).</p>
Full article ">Figure 8
<p>The translation velocity (m s<sup>−1</sup>) for Typhoon Khanun (2023) regressed by wavenumber-1 decomposition of vorticity budget terms averaged from 1–8 km height in CTRL (red) and NTDK (green) experiments. SUM, TILT, STRT, VAVD, and HAVD represent summation, tilting, stretching, vertical advection, and horizontal advection, respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>) Wavenumber-1 HAVD of vorticity (s<sup>−2</sup>, ×10<sup>−10</sup>) and horizontal wind (m s<sup>−1</sup>) averaged over 1–8 km height at 1800 UTC, 30 July 2023 (the 48th forecast hour) for CTRL. The light blue arrow represents the translation velocity (m s<sup>−1</sup>) of Typhoon Khanun (2023) due to the HAVD term as in <a href="#atmosphere-15-01075-f008" class="html-fig">Figure 8</a>. (<b>b</b>) As in (<b>a</b>), but for NTDK. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for 0600 UTC, 31 July 2023 (the 60th forecast hour).</p>
Full article ">Figure 10
<p>Model tracks of total 120 h in CTRL (red) and NTDK (green) experiments and the best track of CWA (black: during the 120 h forecasts; grey: out of the forecast period) from 0000 UTC, 9 October 2023 of Typhoon Bolaven (2023). Dots were marked every six hours.</p>
Full article ">Figure 11
<p>The 500 hPa geopotential height (red) and streamline (black) fields in (<b>a</b>) CTRL and (<b>b</b>) NTDK experiments at 000 UTC, 11 October 2023 (the 48th forecast hour). The thick red contour represents 5880 gpm. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for 0600 UTC, 11 October 2023 (the 54th forecast hour).</p>
Full article ">Figure 12
<p>The translation velocity (m s<sup>−1</sup>) for Typhoon Bolaven (2023) regressed by wavenumber-1 decomposition of vorticity budget terms averaged from 1–8 km height in CTRL (red) and NTDK (green) experiments. SUM, TILT, STRT, VAVD, and HAVD represent summation, tilting, stretching, vertical advection, and horizontal advection, respectively.</p>
Full article ">Figure 13
<p>(<b>a</b>) Wavenumber-1 HAVD of vorticity (s<sup>−2</sup>, ×10<sup>−10</sup>) and horizontal wind (m s<sup>−1</sup>) averaged over 1–8-km height at 0000 UTC, 11 October 2023 (the 48th forecast hour) for CTRL. The light blue arrow represents the translation velocity (m s<sup>−1</sup>) of Typhoon Bolaven (2023) due to the HAVD term as in <a href="#atmosphere-15-01075-f012" class="html-fig">Figure 12</a>. (<b>b</b>) As in (<b>a</b>), but for NTDK. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for 0600 UTC, 11 October 2023 (the 54th forecast hour).</p>
Full article ">
23 pages, 9292 KiB  
Article
Potential Impacts of Future Climate Change on Super-Typhoons in the Western North Pacific: Cloud-Resolving Case Studies Using Pseudo-Global Warming Experiments
by Chung-Chieh Wang, Min-Ru Hsieh, Yi Ting Thean, Zhe-Wen Zheng, Shin-Yi Huang and Kazuhisa Tsuboki
Atmosphere 2024, 15(9), 1029; https://doi.org/10.3390/atmos15091029 - 25 Aug 2024
Viewed by 1065
Abstract
Potential impacts of projected long-term climate change toward the end of the 21st century on rainfall and peak intensity of six super-typhoons in the western North Pacific (WNP) are assessed using a cloud-resolving model (CRM) and the pseudo-global warming (PGW) method, under two [...] Read more.
Potential impacts of projected long-term climate change toward the end of the 21st century on rainfall and peak intensity of six super-typhoons in the western North Pacific (WNP) are assessed using a cloud-resolving model (CRM) and the pseudo-global warming (PGW) method, under two representative concentration pathway (RCP) emission scenarios of RCP4.5 and RCP8.5. Linear long-term trends in June–October are calculated from 38 Coupled Model Intercomparison Project phase 5 (CMIP5) models from 1981–2000 to 2081–2100, with warmings of about 3 °C in sea surface temperature, 4 °C in air temperature in the lower troposphere, and increases of 20% in moisture in RCP8.5. The changes in RCP4.5 are about half the amounts. For each typhoon, three experiments are carried out: a control run (CTL) using analysis data as initial and boundary conditions (IC/BCs), and two future runs with the trend added to the IC/BCs, one for RCP4.5 and the other for RCP8.5, respectively. Their results are compared for potential impacts of climate change. In future scenarios, all six typhoons produce more rain rather consistently, by around 10% in RCP4.5 and 20% in RCP8.5 inside 200–250 km from the center, with increased variability toward larger radii. Such increases are tested to be highly significant and can be largely explained by the increased moisture and water vapor convergence in future scenarios. However, using this method, the results on peak intensity are mixed and inconsistent, with the majority of cases becoming somewhat weaker in future runs. It is believed that in the procedure to determine the best initial time for CTL, which yielded the strongest TC, often within a few hPa in minimum central sea-level pressure to the best track data, an advantage was introduced to the CTL unintentionally. Once the long-term trends were added in future runs, the environment of the storm was altered and became not as favorable for subsequent intensification. Thus, the PGW approach may have some bias in assessing the peak intensity of such super-typhoon cases, and caution should be practiced. Full article
(This article belongs to the Section Climatology)
Show Figures

Figure 1

Figure 1
<p>The averaged long-term trend (Δ values) of (<b>a</b>) height (gpm, blue contours, every 1 gpm) and horizontal wind (m s<sup>−1</sup>, vector and color, reference vector length and scale at bottom) at 1000 hPa and (<b>b</b>) SST (K) in the WNP, between Jun and Oct of 1981–2000 and 2081–2100 from 38 CMIP5 models for the RCP8.5 scenario. Initial positions of the six typhoons in CTL are marked (typhoon symbols). (<b>c</b>–<b>h</b>) Vertical profiles of areal-mean Δ values over the domain of 6°–16° N, 135°–155° E, i.e., dashed box in (<b>b</b>), for the RCP4.5 (blue) and RCP8.5 (scarlet) scenarios for the changes in (<b>c</b>) temperature (Δ<span class="html-italic">T</span>, K), (<b>d</b>) specific humidity (Δ<span class="html-italic">q<sub>v</sub></span>, g kg<sup>−1</sup>), (<b>e</b>) <span class="html-italic">u</span>- and (<b>f</b>) <span class="html-italic">v</span>-components of wind (Δ<span class="html-italic">u</span> and Δ<span class="html-italic">v</span>, m s<sup>−1</sup>), (<b>g</b>) saturation specific humidity (Δ<span class="html-italic">q<sub>s</sub></span>, g kg<sup>−1</sup>), and (<b>h</b>) deficit in specific humidity to saturation (Δ<span class="html-italic">q<sub>d</sub></span>, Δ<span class="html-italic">q<sub>d</sub></span> = Δ<span class="html-italic">q<sub>v</sub></span> − Δ<span class="html-italic">q<sub>s</sub></span>, g kg<sup>−1</sup>), respectively.</p>
Full article ">Figure 2
<p>Comparison between JTWC (red) and JMA (green) best tracks and the CReSS-simulated track in CTL (blue) for each of the six super-typhoons: (<b>a</b>) Megi (2010), (<b>b</b>) Haiyan (2013), (<b>c</b>) Vongfong (2014), (<b>d</b>) Soudelor (2015), (<b>e</b>) Meranti (2016), and (<b>f</b>) Yutu, respectively. Typhoon center locations are given every 6 h in UTC (circles) during the simulation period, with solid dots at 0000 UTC with the date labeled. The topography (m) is also plotted (scale at lower right).</p>
Full article ">Figure 3
<p>Similar to <a href="#atmosphere-15-01029-f002" class="html-fig">Figure 2</a>, but for comparison of TC intensity in minimum central (sea-level) pressure <span class="html-italic">p<sub>min</sub></span> (hPa, thick curves and left axis) and maximum wind speed <span class="html-italic">V<sub>max</sub></span> (m s<sup>−1</sup>, thin curves and right axis) between JTWC (red) and JMA (green) best tracks and the CTL simulation (blue) for four typhoons: (<b>a</b>) Megi (2010) in October, (<b>b</b>) Haiyan (2013) in November, (<b>c</b>) Soudelor (2015) in August, and (<b>d</b>) Meranti (2016) in September, respectively. Data points are 6 h apart, and the time is in UTC.</p>
Full article ">Figure 4
<p>(<b>a</b>) TRMM satellite brightness temperature observation (<span class="html-italic">T<sub>B</sub></span>, K) at 2242 UTC and (<b>b</b>) column-maximum mixing ratio of precipitation (g kg<sup>−1</sup>, rain + snow + graupel) in CTL at 2100 UTC, both on 17 Oct for STY Megi (2010). (<b>c</b>–<b>f</b>) As in (<b>a</b>,<b>b</b>), except (<b>c</b>) at 1101 UTC for TRMM <span class="html-italic">T<sub>B</sub></span> and (<b>d</b>) at 1200 UTC for model mixing ratio in CTL on 7 Nov for STY Haiyan (2013), and (<b>e</b>) at 1650 UTC for TRMM <span class="html-italic">T<sub>B</sub></span> and (<b>f</b>) at 1800 UTC for model mixing ratio in CTL on 13 Sep for STY Meranti (2016), respectively. The domain of the upper panels is approximately 15° × 15° and that of lower panels is 750 km × 750 km, with the model simulation time (h) also labeled inside. (Source of TRMM images: NRL).</p>
Full article ">Figure 5
<p>As in <a href="#atmosphere-15-01029-f002" class="html-fig">Figure 2</a>, except showing (<b>a</b>) model-simulated tracks of the six super-typhoons in CTL (color), and (<b>b</b>–<b>d</b>) for comparison between tracks in CTL (blue), R4.5 (green), and R8.5 (red) for STYs (<b>b</b>) Megi (2010), (<b>c</b>) Haiyan (2013), and (<b>d</b>) Meranti (2016), respectively. Typhoon locations are given every 3 h in UTC (circles), with solid dots at 0000 UTC (date labeled). The scale for topography (m) is at the bottom of panel (<b>a</b>).</p>
Full article ">Figure 6
<p>As in <a href="#atmosphere-15-01029-f003" class="html-fig">Figure 3</a>, but for comparison of intensity in <span class="html-italic">p<sub>min</sub></span> (hPa, thick curves) and <span class="html-italic">V<sub>max</sub></span> (m s<sup>−1</sup>, thin curves) between CTL (blue), R4.5 (green), and R8.5 (red) for the six typhoons: (<b>a</b>) Megi (2010) in October, (<b>b</b>) Haiyan (2013) in November, (<b>c</b>) Vongfong (2014) in October, (<b>d</b>) Soudelor (2015) in August, (<b>e</b>) Meranti (2016) in September, and (<b>f</b>) Yutu (2018) in October, respectively. Data points are 3 h apart.</p>
Full article ">Figure 7
<p>(<b>a</b>) Radial profile of azimuthally-averaged rainfall from 0 to 500 km and (<b>b</b>) time series of areal-mean rainfall (both in mm per 3 h) inside the radius of 350 km over the full simulation period for TY Megi (2010) in CTL (blue), R4.5 (green), and R8.5 (red), respectively. The observation from GPM IMERG is also plotted (black) in (<b>b</b>). (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), except for TY Haiyan (2013) and inside 250 km. (<b>e</b>,<b>f</b>) As in (<b>a</b>,<b>b</b>), except for TY Meranti (2016) and inside 300 km.</p>
Full article ">Figure 8
<p>The mean radius–height profiles of (<b>a</b>) tangential wind (m s<sup>−1</sup>, isotachs every 4 m s<sup>−1</sup>) and radial wind and vertical velocity (<span class="html-italic">w</span>, m s<sup>−1</sup>, vectors, reference length at lower right of panel), with <span class="html-italic">w</span> colored (scale at bottom), and mixing ratio of (<b>b</b>) graupel and (<b>c</b>) rain (both in g kg<sup>−1</sup>, scale at bottom), respectively, from 0 to 400 km and averaged azimuthally and over the full simulation period for Haiyan in CTL. (<b>d</b>–<b>f</b>) As in (<b>a</b>–<b>c</b>), except for their differences of R85 minus CTL.</p>
Full article ">
25 pages, 6747 KiB  
Article
Spatiotemporal Patterns of Typhoon-Induced Extreme Precipitation in Hainan Island, China, 2000–2020, Using Satellite-Derived Precipitation Data
by Mengyu Xu, Yunxiang Tan, Chenxiao Shi, Yihang Xing, Ming Shang, Jing Wu, Yue Yang, Jianhua Du and Lei Bai
Atmosphere 2024, 15(8), 891; https://doi.org/10.3390/atmos15080891 - 25 Jul 2024
Viewed by 1177
Abstract
Extreme precipitation events induced by tropical cyclones have increased frequency and intensity, significantly impacting human socioeconomic activities and ecological environments. This study systematically examines the spatiotemporal characteristics of these events across Hainan Island and their influencing factors using GsMAP satellite precipitation data and [...] Read more.
Extreme precipitation events induced by tropical cyclones have increased frequency and intensity, significantly impacting human socioeconomic activities and ecological environments. This study systematically examines the spatiotemporal characteristics of these events across Hainan Island and their influencing factors using GsMAP satellite precipitation data and tropical cyclone track data. The results indicate that while the frequency of typhoon events in Hainan decreased by 0.3 events decade−1 from 1949 to 2020, extreme precipitation events have increased significantly since 2000, especially in the eastern and central regions. Different typhoon tracks have distinct impacts on the island, with Track 1 (Northeastern track) and Track 2 (Central track) primarily affecting the western and central regions and Track 3 (Southern track) impacting the western region. The impact of typhoon precipitation on extreme events increased over time, being the greatest in the eastern region, followed by the central and western regions. Incorporating typhoon precipitation data shortened the recurrence interval of extreme precipitation in the central and eastern regions. Diurnal peaks occur in the early morning and late evening, primarily affecting coastal areas. Typhoon duration (CC_max = 0.850) and wind speed (CC_max = 0.369) positively correlated with extreme precipitation, while the pressure was negatively correlated. High sea surface temperature areas were closely associated with extreme precipitation events. The atmospheric circulation indices showed a significant negative correlation with extreme precipitation, particularly in the western and central regions. ENSO events, especially sea surface temperature changes in the Niño 1 + 2 region (−0.340 to −0.406), have significantly influenced typhoon precipitation characteristics. These findings can inform region-specific disaster prevention and mitigation strategies for Hainan Island. Full article
(This article belongs to the Special Issue Extreme Weather Events in a Warming Climate)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Topographic map of Hainan Island and spatial distribution of meteorological stations; (<b>b</b>) multi-year average TIP.</p>
Full article ">Figure 2
<p>(<b>a</b>) Multi-year average typhoon precipitation ratio; (<b>b</b>) Spatial distribution zones of typhoon-induced extreme precipitation (TIEP).</p>
Full article ">Figure 3
<p>Spatial patterns of clustered typhoon tracks over Hainan Island.</p>
Full article ">Figure 4
<p>Number of typhoons affecting Hainan Island from 1950 to 2020. Notes: The dashed trend line is derived from Theil–Sen regression after three-point average smoothing. Dotted lines represent segmented regression after breakpoint detection, which is used to analyze the trend changes in different periods.</p>
Full article ">Figure 5
<p>Frequency of extreme precipitation events distinguished with the 95th (<b>a</b>) and 99th (<b>b</b>) percentile threshold, respectively.</p>
Full article ">Figure 6
<p>Typhoon-induced extreme precipitation in three sub-regions of Hainan Island: Analysis using 1-hour precipitation thresholds at the 95th (<b>a</b>) and 99th (<b>b</b>) percentiles.</p>
Full article ">Figure 7
<p>Temporal evolution of the ratio between typhoon-induced extreme precipitation (TIEP) and total typhoon rainfall, based on 1-h 95th (<b>a</b>) and 99th (<b>b</b>) percentile thresholds.</p>
Full article ">Figure 8
<p>Intra-day temporal pattern of extreme precipitation based on 1-h data with 95th percentile (<b>a</b>) and 99th percentile (<b>b</b>) thresholds.</p>
Full article ">Figure 9
<p>The maximum hourly precipitation and maximum daily precipitation in different typhoon tracks. Notes: The subplots of (<b>a</b>–<b>c</b>) are the maximum hourly precipitation in the typhoon Tracks 1–3, respectively, while the subplots of (<b>d</b>–<b>f</b>) show the maximum daily precipitation in the typhoon Tracks 1–3, respectively.</p>
Full article ">Figure 10
<p>Spatial distribution of typhoon-induced extreme precipitation (TIEP) at a threshold of the 95th percentile.</p>
Full article ">Figure 11
<p>Spatial distribution of TIEP at a threshold of the 99th percentile.</p>
Full article ">Figure 12
<p>Temporal pattern of typhoon contribution to extreme precipitation for different tracks.</p>
Full article ">Figure 13
<p>TIEP in different return periods and different temporal scales.</p>
Full article ">Figure 14
<p>Spatial distribution of correlations between extreme precipitation and SST in the western, central, and eastern regions, respectively. Notes: (1) Subplots of (<b>a</b>–<b>c</b>) show the correlation coefficients between extreme precipitation and SST in the western, central and eastern regions, respectively. (2) The black lines in the left diagram represent the representative paths of the three typhoon paths.</p>
Full article ">Figure 15
<p>Comparisons between the TIEP and Northern Hemisphere Polar Vortex Area Index (<b>a</b>), Pacific Subtropical High Ridge Position Index (<b>b</b>), and Niño 1 + 2 SST Index (<b>c</b>).</p>
Full article ">
23 pages, 8912 KiB  
Article
A New Post-Processing Method for Improving Track and Rainfall Ensemble Forecasts for Typhoons over Eastern China
by Chun Liu, Hanqing Deng, Xuexing Qiu, Yanyu Lu and Jiayun Li
Atmosphere 2024, 15(8), 874; https://doi.org/10.3390/atmos15080874 - 23 Jul 2024
Viewed by 611
Abstract
This paper proposes a new post-processing method for model data in order to improve typhoon track and rainfall forecasts. The model data used in the article include low-resolution ensemble forecasts and high-resolution forecasts. The entire improvement method contains the following three steps. The [...] Read more.
This paper proposes a new post-processing method for model data in order to improve typhoon track and rainfall forecasts. The model data used in the article include low-resolution ensemble forecasts and high-resolution forecasts. The entire improvement method contains the following three steps. The first step is to correct the typhoon track forecast: three ensemble member optimization methods are applied to the low-resolution ensemble forecasts, and then the best optimization method is selected with the principle of the smallest average distance error. The results of rainfall forecasts show that the corrected rainfall forecast performs better than the original forecasts. The second step is to derive the high-resolution probability rainfall forecast: the neighborhood method is applied to the deterministic high-resolution rainfall forecast. The last step is to correct the typhoon rainfall forecast: the low- and high-resolution forecasts are blended using the probability-matching method with two different schemes. The results show that the forecasts of the two schemes perform better than the original forecast under all rainfall thresholds and all forecast lead times. In terms of bias score, a rain forecast from one scheme corrects the rainfall deviation from observation better for light and moderate rainfall, whereas a rain forecast from another scheme corrects the rainfall deviation better for heavy and torrential rainfall. The better performance of corrected rain forecasts in the case of Typhoon Lekima and Rumbia over eastern China is demonstrated. Full article
Show Figures

Figure 1

Figure 1
<p>Accumulated observed rainfall (in mm) of Typhoon Lekima from 08:00 CST 9 August to 08:00 CST 12 August.</p>
Full article ">Figure 2
<p>Schematics of the EMOM. The radius of the dashed circle represents the N selected members with minimum average track distance error. In this figure, as an example, the four selected members (m<sub>1</sub>, m<sub>2</sub>, m<sub>3</sub>, and m<sub>4</sub>) have the smallest average distance error (N = 4).</p>
Full article ">Figure 3
<p>Schematic example of neighborhood determination and fractional value computation for a model forecast. Rainfall exceeds the threshold in the shaded boxes (Figure given in Theis [<a href="#B27-atmosphere-15-00874" class="html-bibr">27</a>]).</p>
Full article ">Figure 4
<p>Average distance error with different selected ensemble members for EMOM_OP (red), EMOM_SF (yellow), and EMOM_OPSF (blue). (<b>a</b>–<b>c</b>) Distance error of EC_EPS for lead times of 24, 48, and 72 h; (<b>d</b>–<b>f</b>) distance error of NCEP_EPS for lead times of 24, 48, and 72 h; (<b>g</b>–<b>i</b>) distance error of multi-model EPS (EC_EPS and NCEP_EPS) for lead times of 24, 48, and 72 h.</p>
Full article ">Figure 5
<p>Rainfall Brier score comparison between the new ensemble forecast from EMOM_OPSF and the original ensemble forecast (EC_EPS) with a 6 h interval for different rainfall thresholds. (<b>a</b>) 0.1 mm per 6 h; (<b>b</b>) 5 mm per 6 h; (<b>c</b>) 10 mm per 6 h; (<b>d</b>) 25 mm per 6 h.</p>
Full article ">Figure 6
<p>Spatial distribution of Brier score difference between the new ensemble forecast from EMOM_OPSF and the original ensemble forecast (EC_EPS). (<b>a</b>,<b>e</b>,<b>i</b>) 0.1 mm rainfall; (<b>b</b>,<b>f</b>,<b>g</b>) 5 mm rainfall; (<b>c</b>,<b>g</b>,<b>k</b>) 10 mm rainfall; (<b>d</b>,<b>h</b>,<b>l</b>) 25 mm rainfall. (<b>a</b>–<b>d</b>) 24 h lead time; (<b>e</b>–<b>h</b>) 48 h lead time; (<b>i</b>–<b>l</b>) 72 h lead time.</p>
Full article ">Figure 7
<p>ROC area for light, moderate, and heavy rainfall during the previous 6 h. (<b>a</b>) Range for the forecast with a starting time of 20:00 CST on 9 August; (<b>b</b>) range for the forecast with a starting time of 08:00 CST on 10 August.</p>
Full article ">Figure 8
<p>Comparison of ETS of S1, S2, EC_EPS ensemble mean, and WARMS for different rainfall thresholds. (<b>a</b>) 0.1 mm per 6 h; (<b>b</b>) 5 mm per 6 h; (<b>c</b>) 10 mm per 6 h; (<b>d</b>) 25 mm per 6 h.</p>
Full article ">Figure 9
<p>Track of Typhoon Lekima from 14:00 CST 4 August to 08:00 CST 13 August with an interval of 3 h.</p>
Full article ">Figure 10
<p>Comparison of track distance errors between the new ensemble forecast based on EMOM_OPSF and the original ensemble for Typhoon Lekima.</p>
Full article ">Figure 11
<p>Comparison of rainfall Brier score between the new ensemble forecast from EMOM_OPSF and the original ensemble forecast (EC_EPS) with a 6 h interval for different rainfall thresholds for Typhoon Lekima. (<b>a</b>) 0.1 mm per 6 h; (<b>b</b>) 5 mm per 6 h; (<b>c</b>) 10 mm per 6 h; (<b>d</b>) 25 mm per 6 h.</p>
Full article ">Figure 12
<p>Comparison of the ETS values of S1, S2, EC_EPS mean, and WARMS for different rainfall thresholds for Typhoon Lekima. (<b>a</b>) 0.1 mm per 6 h; (<b>b</b>) 5 mm per 6 h; (<b>c</b>) 10 mm per 6 h; (<b>d</b>) 25 mm per 6 h.</p>
Full article ">Figure 13
<p>Rainfall patterns for Typhoon Lekima (for the three periods from 08:00 CST on 9 August to 08:00 CST on 10 August; from 08:00 CST on 10 August to 08:00 CST on 11 August; and from 08:00 CST on 11 August to 08:00 CST on 12 August), showing the comparison between observed rainfall and original rainfall forecasts. (<b>a</b>–<b>c</b>) Observed rainfall; (<b>d</b>–<b>f</b>) EC_EPS mean; (<b>g</b>–<b>i</b>) WARMS. The province locations are shown in <a href="#atmosphere-15-00874-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 14
<p>Rainfall patterns for Typhoon Lekima (for the three periods from 08:00 CST on 9 August to 08:00 CST on 10 August; from 08:00 CST on 10 August to 08:00 CST on 11 August; and from 08:00 CST on 11 August to 08:00 CST on 12 August), showing the performance of corrected rainfall forecasts. (<b>a</b>–<b>c</b>) S1; (<b>d</b>–<b>f</b>) S2. The province locations are shown in <a href="#atmosphere-15-00874-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 15
<p>(<b>a</b>) Track of Typhoon Rumbia from 15th August; (<b>b</b>) accumulated observed rainfall (mm) from 08:00 CST 19 August to 08:00 CST 21 August.</p>
Full article ">Figure 16
<p>(<b>a</b>) Mean rainfall of EC_EPS forecast for 19 August to 21 August; (<b>b</b>) WARMS rainfall forecast for 19 August to 21 August.</p>
Full article ">Figure 17
<p>Comparison of the ETS values of S1, S2, EC_EPS mean, and WARMS for different rainfall thresholds for Typhoon Rumbia. (<b>a</b>) 0.1 mm per 6 h; (<b>b</b>) 5 mm per 6 h; (<b>c)</b> 10 mm per 6 h; (<b>d</b>) 25 mm per 6 h.</p>
Full article ">
27 pages, 17193 KiB  
Article
Response of Cyclonic Eddies to Typhoon Surigae and Their Weakening Effect on the Kuroshio Current in the Western North Pacific Ocean
by Yanzeng Zhang and Shuzong Han
J. Mar. Sci. Eng. 2024, 12(7), 1202; https://doi.org/10.3390/jmse12071202 - 17 Jul 2024
Viewed by 757
Abstract
This study investigated the dynamic and thermal responses of cyclonic eddies (CEs) to Typhoon Surigae in the western North Pacific Ocean using satellite data and a coupled ocean–atmosphere model. Observations and simulations revealed that the typhoon enhanced the two preexisting CEs (C1 and [...] Read more.
This study investigated the dynamic and thermal responses of cyclonic eddies (CEs) to Typhoon Surigae in the western North Pacific Ocean using satellite data and a coupled ocean–atmosphere model. Observations and simulations revealed that the typhoon enhanced the two preexisting CEs (C1 and C2). After the typhoon passed the two eddies, the sea surface height (SSH) lowered and the eddy velocity increased above 200 m. C1 was stretched with elliptical deformation accompanied by an SSH trough and jets on the sides of the typhoon track at the eddy edge. The comparative experiments indicated that the typhoon caused the SSH of C1 and C2 to lower by 53.52% and 25.14% compared to conditions without the typhoon, respectively, and the kinetic energy of C1 and C2 to increase by 12 times and 65.76%, respectively. The positive vorticity anomaly input from the typhoon to the CEs was the main mechanism for the enhancement of the CEs. The enhanced CEs modulated the typhoon-induced sea surface temperature (SST) cooling, causing the temperature within the eddies to decrease by upwelling and mixing, and the SST cooling became significant at the center of the CEs and propagated westward with the eddies. This study also revealed that typhoons can significantly perturb eddy dynamic structures by enhancing or generating cyclonic cold eddies and eradicating anticyclonic eddies, thereby weakening the Kuroshio Current transport via eddy–Kuroshio interactions. Full article
(This article belongs to the Special Issue Air-Sea Interaction and Marine Dynamics)
Show Figures

Figure 1

Figure 1
<p>Tracks of Typhoon Surigae (displayed at 6 h intervals from IBTrACS) and the mean sea level anomaly (SLA) from 18 April to 23 April. The gray solid line indicates the track of Surigae, the colored dots represent the typhoon intensity category according to the Saffir–Simpson Hurricane scale (SSHS), and the date labels (month/day) indicate 00:00 UTC of the day. C1 and C2 are the two cyclonic cold eddies. The arrows indicate the geostrophic velocity vector, whereas the blue arrows denote the Kuroshio Current. The green dots are Argo floats within 200 km of the typhoon track from 16 April to 30 April, the purple dots are Argo 5904698, and the blue pentagram indicates mooring station M1.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>k</b>) Distributions of SLA before, during, and after Surigae, and (<b>l</b>) the typhoon-induced SLA reduction. The solid gray lines indicate the typhoon tracks, the black dashed lines indicate the edges of the identified cyclonic eddies, and the colored dots indicate the position of Surigae at 00:00 UTC on that day and the typhoon intensity. (<b>m</b>,<b>n</b>) Meridional average SLA between 10 April and 10 May for C1 and C2.</p>
Full article ">Figure 3
<p>Distributions of (<b>a</b>–<b>d</b>) eddy kinetic energy (EKE), (<b>e</b>–<b>h</b>) Ro, and (<b>i</b>–<b>l</b>) <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> </semantics></math> before, during, and after Surigae. Arrows indicate geostrophic velocity vectors and black dashed lines indicate identified cyclonic eddy edges.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>k</b>) Distributions of the sea surface temperature (SST) before, during, and after Surigae and (<b>l</b>) the SST cooling. The solid gray lines indicate the typhoon track, the black dashed lines are the edges of the identified cyclonic eddies, and the colored dots indicate the position of Surigae at 00:00 UTC on that day and the typhoon intensity.</p>
Full article ">Figure 5
<p>Comparisons of the (<b>a</b>) tracks, (<b>b</b>) maximum wind speeds at 10 m, and (<b>c</b>) central air pressure of Surigae between the observations and the EXP-TC results.</p>
Full article ">Figure 6
<p>Distributions of the mean sea surface height (SSH) and surface velocity from 17 April to 30 April based on (<b>a</b>) satellite observations and (<b>b</b>) the EXP-TC simulation and variations in (<b>c</b>) the mean EKE and (<b>d</b>) the mean SSH in C1 and C2 from 17 April to 30 April.</p>
Full article ">Figure 7
<p>Comparisons of (<b>a</b>) temperature from Argo, (<b>b</b>) salinity from Argo, (<b>c</b>) temperature from M1, and (<b>d</b>) velocity from M1 with the EXP-TC simulation. The black solid lines indicate the regression line, and the magenta dashed lines are the 95% confidence intervals. In the regression equation, y is the model result, and x is the corresponding measured result. R<sup>2</sup> is the correlation coefficient between the model result and the measured result, and RMSE is the root mean square error.</p>
Full article ">Figure 8
<p>Distributions of the SSH and surface velocity of C1 for the EXP-TC simulation (daily averaged over 6 h time-resolved output results) on (<b>a</b>) 18 April, (<b>b</b>) 22 April, (<b>c</b>) 25 April, and (<b>d</b>) 30 April; (<b>e</b>–<b>h</b>) evolution of the 3D potential temperature and velocity; (<b>i</b>–<b>l</b>) distributions of the EKE above 500 m for the sections shown by the blue dashed lines, and the black solid lines are the contours of meridional velocity (m·s<sup>−1</sup>) at the same times as (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 9
<p>SSH and surface velocity (daily averaged) simulated by (<b>a</b>–<b>d</b>) EXP-TC and (<b>e</b>–<b>h</b>) EXP-NoTC.</p>
Full article ">Figure 10
<p>Distributions of the SSH and surface velocity of C2 for the EXP-TC simulation (daily averaged over 6 h time-resolved output results) on (<b>a</b>) 20 April, (<b>b</b>) 22 April, (<b>c</b>) 25 April, and (<b>d</b>) 30 April; (<b>e</b>–<b>h</b>) evolution of the 3D potential temperature and velocity; (<b>i</b>–<b>l</b>) distributions of the EKE above 500 m for the sections shown by the blue dashed lines, and the black solid lines are the contours of meridional velocity (m·s<sup>−1</sup>) at the same times as (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 11
<p>Distributions of the SSH and surface velocity of C2 for the EXP-NoTC simulation (daily averaged over 6 h time-resolved output results) on (<b>a</b>) 20 April, (<b>b</b>) 22 April, (<b>c</b>) 25 April, and (<b>d</b>) 30 April; (<b>e</b>–<b>h</b>) evolution of the 3D potential temperature and velocity; (<b>i</b>–<b>l</b>) distributions of the EKE above 500 m for the sections shown by the blue dashed lines, and the black solid lines are the contours of meridional velocity (m·s<sup>−1</sup>) at the same times as (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 12
<p>Wind speeds and Ekman pumping velocity (EPV) as Typhoon Surigae passed (<b>a</b>–<b>c</b>) C1 and (<b>d</b>–<b>f</b>) C2 simulated by EXP-TC. The gray arrows indicate 10 m wind vectors, the gray solid lines indicate typhoon tracks, and the dots are the typhoon positions at the corresponding times.</p>
Full article ">Figure 13
<p>Vertical distributions of the potential vorticity (PV) anomalies in (<b>a</b>–<b>d</b>) C1 and (<b>e</b>–<b>h</b>) C2 before and after the typhoon. The locations of the sections are consistent with those in <a href="#jmse-12-01202-f008" class="html-fig">Figure 8</a>a and <a href="#jmse-12-01202-f010" class="html-fig">Figure 10</a>a. The black solid lines indicate the contours of potential density (kg·m<sup>−3</sup>).</p>
Full article ">Figure 14
<p>Temperature variations in the centers of (<b>a</b>) C1 and (<b>d</b>) C2 and the temporal evolution of the temperature tendency (black), total advection (red), horizontal advection (blue), vertical advection (cyan) and vertical diffusion (green) in the mixed layer (ML) temperature budget at the top of the ML (2 m) of (<b>b</b>) C1 and (<b>e</b>) C2 and at the bottom of the ML (50 m) of (<b>c</b>) C1 and (<b>f</b>) C2. The terms in the temperature budget equation were low-pass filtered for 48 h using a Butterworth filter of third order. Note that in (<b>b</b>,<b>c</b>) and (<b>e</b>,<b>f</b>), the values of the terms represented by the solid lines are shown by the blue axis on the left, the terms represented by the dashed lines are shown by the red axis on the right, and the gray dashed lines represent the time the typhoon passed.</p>
Full article ">Figure 15
<p>(<b>a</b>) Forcing time of Typhoon Surigae on the ocean; (<b>b</b>) energy input of Surigae to the ocean currents. In (<b>a</b>,<b>b</b>), the gray solid line indicates the track of Surigae, the colored dots represent the typhoon intensity. Distribution of the buoyancy frequency N<sup>2</sup> in the cross-eddy sections (as in <a href="#jmse-12-01202-f008" class="html-fig">Figure 8</a>a and <a href="#jmse-12-01202-f010" class="html-fig">Figure 10</a>a) on (<b>c</b>) 17 April before the typhoon’s passage at C1 and (<b>d</b>) 20 April before the typhoon’s passage at C2. The red solid lines indicate the mixed layer depth and the black solid lines denote the potential density contours (kg·m<sup>−3</sup>).</p>
Full article ">Figure 16
<p>Spatial and temporal evolution of (<b>a</b>–<b>d</b>) C1, (<b>e</b>–<b>h</b>) new eddy, and (<b>i</b>–<b>l</b>) C2 based on SSH and geostrophic velocity. The magenta dots represent the eddy centers.</p>
Full article ">Figure 17
<p>Distribution of (<b>a</b>) the surface geostrophic velocity based on satellite data and (<b>b</b>) the surface velocity based on the CMEMS in the pre-TC stage (averaged between 1 March and 18 April) and (<b>c</b>,<b>d</b>) as in (<b>a</b>,<b>b</b>) but for the post-TC stage (averaged between 19 April and 25 May). Meridional velocity in sections (<b>e</b>) 15° N, (<b>f</b>) 18° N, and (<b>g</b>) 21° N (shown by solid magenta lines in (<b>b</b>,<b>d</b>)) in the pre-TC stage and (<b>h</b>–<b>j</b>) as in (<b>e</b>–<b>g</b>) but for the post-TC stage.</p>
Full article ">Figure 18
<p>(<b>a</b>) Kuroshio transport in pre-TC and post-TC stages in three sections; (<b>b</b>) Kuroshio transport in spring (from March to May) from 1993 to 2022 in three sections. Sv: Sverdrup, 1Sv = 10<sup>6</sup> m<sup>3</sup>/s. In (<b>b</b>), the vertical red dashed line represents 2021 in this case.</p>
Full article ">
25 pages, 11697 KiB  
Article
Improving Typhoon Muifa (2022) Forecasts with FY-3D and FY-3E MWHS-2 Satellite Data Assimilation under Clear Sky Conditions
by Feifei Shen, Xiaolin Yuan, Hong Li, Dongmei Xu, Jingyao Luo, Aiqing Shu and Lizhen Huang
Remote Sens. 2024, 16(14), 2614; https://doi.org/10.3390/rs16142614 - 17 Jul 2024
Viewed by 864
Abstract
This study investigates the impacts of assimilating the Microwave Humidity Sounder II (MWHS-2) radiance data carried on the FY-3D and FY-3E satellites on the analyses and forecasts of Typhoon Muifa in 2022 under clear-sky conditions. Data assimilation experiments are conducted using the Weather [...] Read more.
This study investigates the impacts of assimilating the Microwave Humidity Sounder II (MWHS-2) radiance data carried on the FY-3D and FY-3E satellites on the analyses and forecasts of Typhoon Muifa in 2022 under clear-sky conditions. Data assimilation experiments are conducted using the Weather Research and Forecasting (WRF) model coupled with the Three-Dimensional Variational (3D-Var) Data Assimilation method to compare the different behaviors of FY-3D and FY-3E radiances. Additionally, the data assimilation strategies are assessed in terms of the sequence of applying the conventional and MWHS-2 radiance data. The results show that assimilating MWHS-2 data is able to enhance the dynamic and thermal structures of the typhoon system. The experiment with FY-3E MWHS-2 assimilated demonstrated superior performance in terms of simulating the typhoon’s structure and providing a prediction of the typhoon’s intensity and track than the experiment with FY-3D MWHS-2 did. The two-step assimilation strategy that assimilates conventional observations before the radiance data has improved the track and intensity forecasts at certain times, particularly with the FY-3E MWHS-2 radiance. It appears that large-scale atmospheric conditions are more refined by initially assimilating the Global Telecommunication System (GTS) data, with subsequent satellite data assimilation further adjusting the model state. This strategy has also confirmed improvements in precipitation prediction as it enhances the dynamic and thermal structures of the typhoon system. Full article
Show Figures

Figure 1

Figure 1
<p>The evolution of Typhoon Muifa’s intensity levels throughout its track, recorded from 0600 on 6 September to 2100 on 16 September UTC. Time (day and hour) and central pressure of typhoon are noted at nodes where the typhoon intensity shifts.</p>
Full article ">Figure 2
<p>Terrain height (filled colors, unit: m) and the best track of Typhoon Muifa 24 h interval (black dot), from 0000 UTC on 8 September to 0000 UTC 16 September within the model domain. Blue points represent FY-3D MWHS-2 data, while red points represent FY-3E MWHS-2 data.</p>
Full article ">Figure 3
<p>The flow chart for the data assimilation experiments. One-step experiments including GTS_DA (yellow solid line), 3D_DA (blue solid line), and 3E_DA (red solid line) are depicted in (<b>a</b>), two-step experiments including 3D_R_DA (blue dashed line) and 3E_R_DA (red dashed line) are depicted in (<b>b</b>).</p>
Full article ">Figure 4
<p>OMB of (<b>a</b>) 3D_DA, (<b>c</b>) 3D_R_DA, (<b>e</b>) 3E_DA, and (<b>g</b>) 3E_R_DA; OMA of (<b>b</b>) 3D_DA, (<b>d</b>) 3D_R_DA, (<b>f</b>) 3E_DA, and (<b>h</b>) 3E_R_DA for the brightness temperature (units: K) of channel 11. The blue dot and red dot represent the location of the typhoon center at 0600 UTC on 14 September 2022 and 0900 UTC on 14 September 2022, respectively.</p>
Full article ">Figure 5
<p>Scatter distribution of the brightness temperature (unit: K) of channel 11 simulated from (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) the background before the bias correction, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) the background after bias correction, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) the analysis after the bias correction in the y-axis versus the observed radiances in the x-axis. (<b>a</b>–<b>c</b>) 3D_DA experiment, (<b>d</b>–<b>f</b>) 3D_R_DA experiment, (<b>g</b>–<b>i</b>) 3E_DA experiment, and (<b>j</b>–<b>l</b>) 3E_R_DA experiment.</p>
Full article ">Figure 6
<p>(<b>a</b>) The number of assimilated satellite data for the experiment; (<b>b</b>,<b>c</b>) Mean for the experiment and (<b>d</b>,<b>e</b>) stdv for the experiment.</p>
Full article ">Figure 7
<p>The 500 hPa water vapor mixing ratio differences (analysis minus background, shading, unit: g/kg) and wind speed (vector, unit: m/s) for (<b>a</b>) GTS_DA, (<b>b</b>) 3D_DA, and (<b>c</b>) 3E_DA. The differences in the (<b>d</b>) first step, (<b>e</b>) second step, and (<b>f</b>) two-steps in 3D_R_DA at 0600 UTC 14 September. The differences in the (<b>g</b>) first step, (<b>h</b>) second step, and (<b>i</b>) two-steps in 3E_R_DA at 0900 UTC 14 September. The red symbol represents the position of the typhoon sourced from the CMA at 0600 UTC 14 September (<b>a</b>,<b>b</b>,<b>d</b>–<b>f</b>) and 0900 UTC 14 September (<b>c</b>,<b>g</b>–<b>i</b>).</p>
Full article ">Figure 8
<p>Same as <a href="#remotesensing-16-02614-f007" class="html-fig">Figure 7</a>, but for the temperature (unit: K).</p>
Full article ">Figure 9
<p>Same as <a href="#remotesensing-16-02614-f007" class="html-fig">Figure 7</a>, but for geopotential height (contours, units: m<sup>2</sup> s<sup>−2</sup>; contours every 2 m<sup>2</sup> s<sup>−2</sup>) and geopotential height differences (analysis minus background, shading, unit: m).</p>
Full article ">Figure 10
<p>RMSE vertical profiles of (<b>a</b>) u-wind (m/s), (<b>b</b>) v-wind (m/s), (<b>c</b>) temperature (K), and (<b>d</b>) specific humidity (g/kg) forecasts versus sounding and surface synoptic (SYNOP) observations at 1200 UTC on 15 September 2022. Error dots indicate statistical significance at the 95% confidence level.</p>
Full article ">Figure 11
<p>(<b>a</b>) Typhoon Muifa’s track from 0900 UTC on 14 September to 0000 UTC on 16 September 2022 and (<b>b</b>) the corresponding track error, the numbers on the x-axis represent day and hour, respectively.</p>
Full article ">Figure 12
<p>The 39 h forecast errors of (<b>a</b>) MSLP (unit: hPa) and (<b>b</b>) MSWS (unit: m/s), initialized at 0900 UTC on 14 September 2022.</p>
Full article ">Figure 13
<p>The vertical cross section of the wind speed (shading, unit: m/s) and potential temperature (contours, unit: K, intervals of 4 K) for (<b>a</b>) CTRL, (<b>b</b>) GTS_DA, (<b>c</b>) 3D_DA, and (<b>d</b>) 3D_R_DA at 0600 UTC 14 September, (<b>e</b>) 3E_DA, and (<b>f</b>) 3E_R_DA at 0900 UTC 14 September. (<b>g</b>) is the cross section with the wind speed (blue wind barbs, unit: m/s) and potential temperature (shading, unit: K). The red symbol is the position of the typhoon at 0600 UTC 14 September.</p>
Full article ">Figure 14
<p>The 24 h accumulated precipitation distribution (unit: mm) from 0900 UTC on 14 September to 0900 UTC on 15 September in the experiments (<b>a</b>) CTRL, (<b>b</b>) GTS_DA, (<b>c</b>) 3D_DA, (<b>d</b>) 3D_R_DA, (<b>e</b>) 3E_DA, (<b>f</b>) 3E_R_DA, and in (<b>g</b>) the observation.</p>
Full article ">Figure 15
<p>ETS of 24 h accumulated precipitation at different thresholds.</p>
Full article ">
Back to TopTop