Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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 (133)

Search Parameters:
Keywords = southern South China Sea

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 20585 KiB  
Article
Screening and Analysis of Potential Aquaculture Spaces for Larimichthys crocea in China’s Surrounding Waters Based on Environmental Temperature Suitability
by Ling Yang, Weifeng Zhou, Xuesen Cui, Yanan Lu and Qin Liu
Biology 2025, 14(2), 205; https://doi.org/10.3390/biology14020205 (registering DOI) - 15 Feb 2025
Abstract
This research evaluates the potential spaces of deep offshore waters for cultivating the Larimichthys crocea, analyzing ocean profile temperature data from 2000 to 2022 according to the species’ environmental temperature suitability. There are significant seasonal variations and differences in habitat distributions of [...] Read more.
This research evaluates the potential spaces of deep offshore waters for cultivating the Larimichthys crocea, analyzing ocean profile temperature data from 2000 to 2022 according to the species’ environmental temperature suitability. There are significant seasonal variations and differences in habitat distributions of different temperature ranges in China’s surrounding waters. The range of maximum living space obtained according to the tolerance temperature shows a trend of being larger in summer and smaller in winter; and the range of viable habitat space obtained based on the suitable and optimal temperature shows a trend of being smaller in summer and larger in winter. Broad areas meeting tolerance temperatures offer broad, yet impractical, site selection options. In contrast, areas with optimal temperatures are limited, which means the availability of ideal site locations is very restricted. Regions consistently within the 20–28 °C range are best for practical site selection. Year-round suitable areas are primarily found at depths of 30 to 90 m in the southern East China Sea and the South China Sea, particularly within the 40 to 50 m depth range. Water mass like the South China Sea Surface Water and the Kuroshio Surface Water consistently maintain suitable temperatures, making them ideal for aquaculture. Full article
(This article belongs to the Section Ecology)
Show Figures

Figure 1

Figure 1
<p>The study area of China’s surrounding waters.</p>
Full article ">Figure 2
<p>Technical flow chart of this research.</p>
Full article ">Figure 3
<p>Inter-annual and seasonal changes in grid points’ numbers under the condition of the biological tolerance temperature from 2000 to 2022.</p>
Full article ">Figure 4
<p>In 2022 Geographic Map of Tolerance Temperature by Depth. This figure illustrates the monthly spatial distribution of temperature tolerance across various ocean depths throughout 2022: (<b>a</b>) 0–24 m depth; (<b>b</b>) 25–50 m depth; (<b>c</b>) 51–90 m depth; (<b>d</b>) 91–130 m depth; (<b>e</b>) 131–199 m depth; (<b>f</b>) 200–2000 m depth.</p>
Full article ">Figure 5
<p>Inter-annual and seasonal changes in grid points’ numbers under the condition of the biological suitable temperature from 2000 to 2022.</p>
Full article ">Figure 6
<p>In 2022 Geographic Map of Suitable Temperature by Depth. This figure illustrates the monthly spatial distribution of temperature suitable across various ocean depths throughout 2022: (<b>a</b>) 0–24 m depth; (<b>b</b>) 25–50 m depth; (<b>c</b>) 51–90 m depth; (<b>d</b>) 91–130 m depth; (<b>e</b>) 131–199 m depth; (<b>f</b>) 200–2000 m depth.</p>
Full article ">Figure 7
<p>Inter-annual and seasonal changes in grid points’ numbers under the condition of the biological optimal temperature range from 2000 to 2022.</p>
Full article ">Figure 8
<p>In 2022 Geographic Map of Optimal Temperature by Depth. This figure illustrates the monthly spatial distribution of temperature optimal across various ocean depths throughout 2022: (<b>a</b>) 0–24 m depth; (<b>b</b>) 25–50 m depth; (<b>c</b>) 51–90 m depth; (<b>d</b>) 91–2000 m depth.</p>
Full article ">Figure 9
<p>Common map of areas with suitable temperatures by depth from 2000 to 2022.</p>
Full article ">Figure 10
<p>Comparative diagram of suitable and optimal spatial conditions by month for 2022: (<b>a</b>) suitable conditions; (<b>b</b>) optimal conditions.</p>
Full article ">Figure 11
<p>Seasonal average temperature clustering and profile analysis in China’s adjacent seas. This figure displays the spatial distribution of six K-Means clustering categories for average temperatures in different seasons from 2000 to 2022 in the seas adjacent to China: (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 11 Cont.
<p>Seasonal average temperature clustering and profile analysis in China’s adjacent seas. This figure displays the spatial distribution of six K-Means clustering categories for average temperatures in different seasons from 2000 to 2022 in the seas adjacent to China: (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">
14 pages, 19711 KiB  
Article
Shallow Gas Distribution Influenced by the Interface of Sedimentary Facies in the Southwest of the Qiongdongnan Basin
by Taotao Yang, Xiaohan Li, Jiapeng Jin, Jianwei Chen, Zhi Gong, Li Zhao, Wenlong Wang, Bo Liu, Jinzi Hu, Wenlu Wang and Xiujuan Wang
J. Mar. Sci. Eng. 2025, 13(2), 301; https://doi.org/10.3390/jmse13020301 - 6 Feb 2025
Viewed by 413
Abstract
Shallow gas, with huge resources, has been confirmed using three dimensional (3D) seismic data and more than 20 drilling sites in the deep water of the LS36 gas field, the Qiongdongnan Basin, the South China Sea. The interface of sedimentary facies in the [...] Read more.
Shallow gas, with huge resources, has been confirmed using three dimensional (3D) seismic data and more than 20 drilling sites in the deep water of the LS36 gas field, the Qiongdongnan Basin, the South China Sea. The interface of sedimentary facies in the southern boundary of the basin controls the distribution within the basin of clastic sediments coming from the north and west of the land uplifted. In this study, seismic data and geophysical attributes were used to investigate the controlling effect of the interface of sedimentary facies on the distribution of shallow gas within the basin. Our study shows that the shallow gas is mainly distributed in the Quaternary Ledong Formation in the southwest of the Qiongdongnan Basin, which was observed from acoustic impedance, amplitude versus offset (AVO), and seismic interpretations. The channelized submarine fans that onlap the interface of the sedimentary facies are distributed in a vertically stacked manner and are the main reservoirs for the shallow gas. Therefore, these sedimentary studies show that the sand-rich sediments are distributed along the interface of the sedimentary facies from the southwest to the northeast and are limited to the shallow gas within the basin. The Central Canyon provides an important deep gas source, while the flank of the canyon, gas chimney, and normal faults related to basement uplift provide pathways for vertical and lateral gas migration to form the shallow gas. This study shows that shallow gas may be widely distributed in other marginal sea basins, and sedimentary systems should be further studied in the future. Full article
(This article belongs to the Special Issue Advances in Marine Gas Hydrate Exploration and Discovery)
Show Figures

Figure 1

Figure 1
<p>The current topographic map of the northwestern continental margin of the South China Sea showing the Qiongdongnan Basin and its internal tectonic regions. The Vietnam and Hainan Uplifts supply terrigenous clastics for sedimentation, and submarine fans, slope fans, and MTDs are distributed in the Qiongdongnan Basin. The channels are revised from [<a href="#B24-jmse-13-00301" class="html-bibr">24</a>] and the sand distribution in the northern slope is revised from [<a href="#B7-jmse-13-00301" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>Seismic profiles (<b>a</b>) and interpretation profiles (<b>b</b>) across the Qiongdongnan Basin and the Zhongjian Basin showing that a large amount of clastic sediment is deposited in the northern sag of the interface of sedimentary facies transition within the Qiongdongnan Basin. (<b>c</b>) is the enlarged seismic profile, which shows a large number of submarine fan deposits in the north of the interface of sedimentary facies transition. The location of the seismic profile is shown in <a href="#jmse-13-00301-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Seismic profiles showing the sedimentary facies’ transition interface and internal shallow gas, submarine fans, channels, and MTDs. The location of the seismic profile is shown in <a href="#jmse-13-00301-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 4
<p>Shallow gas distributions in three layers and a sand-rich reservoir identified from 2D/3D seismic data showing the submarine fan-like distribution. The shallow gas in the LS 36-1 is revised from [<a href="#B7-jmse-13-00301" class="html-bibr">7</a>,<a href="#B8-jmse-13-00301" class="html-bibr">8</a>]. The thick white line represents the location of the whole seismic profiles, such as Figures 3, 5, and 8, while the green line indicates the scope of the seismic profiles used in Figure 7.</p>
Full article ">Figure 5
<p>(<b>a</b>) Seismic profiles showing enhanced reflections related to submarine fan systems and shallow gas. (<b>b</b>) An enlarged seismic profile showing positive and negative amplitude anomalies and pull-down reflections. (<b>c</b>) The instantaneous frequency profile shows low frequency for the gas-bearing layers and high frequency for the gas hydrate-bearing layers. (<b>d</b>) The inverted P-wave velocity from constrained sparse spike inversion (CSSI) showing the anomalies. The location of the seismic profile is shown in <a href="#jmse-13-00301-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>AVO attribute profiles show intercept profile (P), gradient profile (G), and product of intercept and gradient profile (P × G) for seismic profile of <a href="#jmse-13-00301-f005" class="html-fig">Figure 5</a>b.</p>
Full article ">Figure 7
<p>Continuous seismic profiles showing the spatial distribution of the sedimentary facies’ transition interface, which separates the continuous fine-grained sediments in the south from the coarse-grained sediments in the north of the Qiongdongnan Basin. The seismic profiles from (<b>a</b>–<b>g</b>) are distributed from west to east, and the location of the seismic profiles are shown in <a href="#jmse-13-00301-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>A shallow gas accumulation model in the southwest of the Qiongdonganan Basin showing submarine fan-shaped distributions under the control of the sedimentary facies’ transition interface. The location of the seismic profile is shown in <a href="#jmse-13-00301-f004" class="html-fig">Figure 4</a>.</p>
Full article ">
19 pages, 11765 KiB  
Article
Descriptive Methodology for Risk Situation of Disastrous Sea Waves in the China Sea
by Juanjuan Wang and Mengmeng Wu
J. Mar. Sci. Eng. 2025, 13(2), 188; https://doi.org/10.3390/jmse13020188 - 21 Jan 2025
Viewed by 418
Abstract
To meet the needs of marine disaster prevention and mitigation, this paper proposes a systematic methodological framework to describe the annual risk situation of Disastrous Sea Waves (DSWs) from four perspectives. Its application is demonstrated for the China Sea in 2023 as a [...] Read more.
To meet the needs of marine disaster prevention and mitigation, this paper proposes a systematic methodological framework to describe the annual risk situation of Disastrous Sea Waves (DSWs) from four perspectives. Its application is demonstrated for the China Sea in 2023 as a case study. The systematic approach is reflected in the following: (1) a comprehensive description of DSW risks based on three dimensions: occurrence frequency, maximum intensity, and hazard index; (2) an overview of the DSW risk characteristics for the year through spatial and monthly distributions; (3) a comparative analysis of the year’s DSWs, with historical data based on anomalies and return periods used to assess the risk characteristics and extremities; and (4) an analysis of the causes of the year’s characteristics based on monthly anomalies and weather systems. Through its application to the China Sea in 2023, the analysis process is introduced as follows. (1) High-Frequency and Intensity Areas: DSWs frequently occurred in the northeastern South China Sea (SCS) and Taiwan Strait, exceeding 450 h. The maximum significant wave height (Hs), reaching 11.3 m, was recorded in the southern East China Sea (ECS) in August. (2) Extremity in Frequency and Attribution: The occurrence frequency was extremely high, with the cumulative hours exceeding the historical average by 159 h (9.1%). The southwestern SCS showed the most significant excess, up to 168 h (>120%). The reason for this was that DSWs in January caused by prolonged cold air lasted 236 h longer (121%). (3) Extremity in Intensity and Attribution: The maximum Hs in the southern ECS and Taiwan Strait was 2 m (30%) higher than the historical average. The intensified cold air waves caused the higher intensities. (4) Hazard Levels: Higher risk occurred in the southwestern SCS, southern ECS, and Taiwan Strait, while the highest extremity occurred in the Bohai Sea. Full article
(This article belongs to the Section Coastal Engineering)
Show Figures

Figure 1

Figure 1
<p>Map of six sub-sea areas in the China Sea. The land is gray and the ocean is blue.</p>
Full article ">Figure 2
<p>Spatial distribution of DSW <span class="html-italic">T</span><sub>4m</sub> (<b>a</b>) and maximum <span class="html-italic">H</span><sub>s</sub> (<b>b</b>) in 2023.</p>
Full article ">Figure 3
<p>Spatial distribution of Hazard Index in 2023.</p>
Full article ">Figure 4
<p>Cumulated statistics of DSW <span class="html-italic">T</span><sub>4m</sub> and maximum <span class="html-italic">H<sub>s</sub></span> in the China Sea from January to December in 2023.</p>
Full article ">Figure 5
<p>Comparison of cumulative <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> in the China Sea in 2023 with historical values.</p>
Full article ">Figure 6
<p>Comparison of cumulative <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> in the six sub-sea areas in 2023 with historical values.</p>
Full article ">Figure 7
<p>Comparison of maximum <span class="html-italic">H<sub>s</sub></span> in the China Sea in 2023 with historical values.</p>
Full article ">Figure 8
<p>Comparisons of maximum <span class="html-italic">H<sub>s</sub></span> in the six sea areas in 2023 with historical values.</p>
Full article ">Figure 9
<p>Spatial distribution of anomalies in DSW <span class="html-italic">T</span><sub>4m</sub> (<b>a</b>) and maximum <span class="html-italic">H</span><sub>s</sub> (<b>b</b>) in 2023.</p>
Full article ">Figure 10
<p>Spatial distribution of anomaly percentages in DSW <span class="html-italic">T</span><sub>4m</sub> (<b>a</b>) and maximum <span class="html-italic">H</span><sub>s</sub> (<b>b</b>) in 2023.</p>
Full article ">Figure 11
<p>Spatial distribution of anomalies in hazard index (<span class="html-italic">HI</span>) in 2023.</p>
Full article ">Figure 12
<p>Corresponding historical return periods for <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> (<b>a</b>) and maximum <span class="html-italic">H<sub>s</sub></span> (<b>b</b>) in 2023, unit: year.</p>
Full article ">Figure 13
<p>Monthly distribution of anomaly percentages of <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> (<b>a</b>) and maximum <span class="html-italic">H<sub>s</sub></span> (<b>b</b>) in 2023. The text in red with a plus sign represents an increase. The blue text with a minus sign represents a decrease.</p>
Full article ">Figure 13 Cont.
<p>Monthly distribution of anomaly percentages of <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> (<b>a</b>) and maximum <span class="html-italic">H<sub>s</sub></span> (<b>b</b>) in 2023. The text in red with a plus sign represents an increase. The blue text with a minus sign represents a decrease.</p>
Full article ">Figure 14
<p>Spatial distribution of anomalies in <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> (<b>a</b>) and maximum <span class="html-italic">H<sub>s</sub></span> (<b>b</b>) in January.</p>
Full article ">Figure 15
<p>Spatial distribution of anomalies in <span class="html-italic">T</span><sub>4<span class="html-italic">m</span></sub> (<b>a</b>) and maximum <span class="html-italic">H<sub>s</sub></span> (<b>b</b>) in September.</p>
Full article ">
21 pages, 9471 KiB  
Article
The Seasonal Correlation Between El Niño and Southern Oscillation Events and Sea Surface Temperature Anomalies in the South China Sea from 1958 to 2024
by Jun Song, Lingxiang Yao, Junru Guo, Yanzhao Fu, Yu Cai and Meng Wang
J. Mar. Sci. Eng. 2025, 13(1), 153; https://doi.org/10.3390/jmse13010153 - 16 Jan 2025
Viewed by 521
Abstract
This study utilizes high-resolution sea surface temperature (SST) reanalysis data (0.25° × 0.25°) to investigate the relationship between SST anomalies in the South China Sea and ENSO events. The main findings are as follows: First, there is a delayed correlation between ENSO and [...] Read more.
This study utilizes high-resolution sea surface temperature (SST) reanalysis data (0.25° × 0.25°) to investigate the relationship between SST anomalies in the South China Sea and ENSO events. The main findings are as follows: First, there is a delayed correlation between ENSO and SST anomalies in the South China Sea, with the correlation being stronger during El Niño years than during La Niña years. Second, the correlation with the peak values of the Oceanic Niño Index (ONI) is strongest for El Niño events with a 9-month lead, while for La Niña events, it is strongest with a 2-month lead. Seasonally, during El Niño events, the strongest correlations are observed in summer with a 3-month lead and in winter with a 1-month lag. For La Niña events, the strongest correlations are seen in summer with an 8-month lag and in winter with a 9-month lag. Finally, an analysis of atmospheric anomalies and shear kinetic energy anomalies relative to SST anomalies reveals a significant seasonal SST response, particularly during the summer of El Niño years and the winter of La Niña years. Overall, these results enhance our understanding of ENSO’s influence on the South China Sea and provide valuable insights for climate prediction and ecosystem protection in the region. Full article
(This article belongs to the Section Physical Oceanography)
Show Figures

Figure 1

Figure 1
<p>Research area.</p>
Full article ">Figure 2
<p>Oceanic El Niño Index (ONI) (according to the Climate Prediction Center of NOAA, El Niño and La Niña events are classified when five consecutive ONI values exceed +0.5 °C (El Niño) or fall below −0.5 °C (La Niña); <a href="https://www.cpc.ncep.noaa.gov/" target="_blank">https://www.cpc.ncep.noaa.gov/</a> accessed on 1 December 2024).</p>
Full article ">Figure 3
<p>Spatial distribution of mean SST in four seasons from 1958 to 2021: (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 4
<p>The seasonal mean SSTAs in the South China Sea during 1958–2021 in (<b>a</b>) spring, (<b>b</b>) summer, (<b>c</b>) autumn, and (<b>d</b>) winter (the upward orange arrow represents a strong El Niño event, and the downward blue arrow represents a strong La Niña event).</p>
Full article ">Figure 5
<p>SST of the “−1 year” to “+1 year” periods of El Niño events and La Niña events during 1958–2021 (where (<b>a</b>) is El Niño events and (<b>b</b>) is La Niña events; among them, J., M., M., J., S., and N. represent January, March, May, July, September, and November, respectively, and then the cycle repeats).</p>
Full article ">Figure 6
<p>The correlation lag of SSTAs in the South China Sea with El Niño events and La Niña events.</p>
Full article ">Figure 7
<p>Four-season SSTAs for El Niño “−1 year” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 8
<p>Four-season SSTAs for El Niño “year 0” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 9
<p>Four-season sea SSTAs for El Niño “+1 year” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 10
<p>Four-season SSTAs for La Niña “−1 year” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 11
<p>Four-season SSTAs for La Niña “0 year” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 12
<p>Four-season SSTAs for La Niña “+1 year” from 1958 to 2021, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 13
<p>Seasonal SSTA lag correlation of El Niño “−1” years and La Niña “−1” years events during 1958–2021 (where (<b>a</b>) is El Niño and (<b>b</b>) is La Niña).</p>
Full article ">Figure 14
<p>Seasonal SSTA lag correlation of El Niño “0” years and La Niña “0” years during 1958–2021 (where (<b>a</b>) is El Niño and (<b>b</b>) is La Niña).</p>
Full article ">Figure 15
<p>Seasonal SSTA lag correlation of El Niño “+1” years and La Niña “+1” years during 1958–2021 (where (<b>a</b>) is El Niño and (<b>b</b>) is La Niña).</p>
Full article ">Figure 16
<p>Verification of SSTA and ONI lead and lag in the four seasons of El Niño events, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 17
<p>Verification of SSTA and ONI lead and lag in the four seasons of La Niña events, where (<b>a</b>) is spring, (<b>b</b>) is summer, (<b>c</b>) is autumn, and (<b>d</b>) is winter.</p>
Full article ">Figure 18
<p>Correlation between shear kinetic energy anomalies and SSTAs in summer (left) and winter (right) for El Niño “−1 year” to El Niño “+1 year”, where (<b>a</b>,<b>b</b>) is the year of El Niño “−1”, (<b>c</b>,<b>d</b>) is the year of El Niño “0”, and (<b>e</b>,<b>f</b>) is the year of El Niño “1”.</p>
Full article ">Figure 19
<p>Correlation between shear kinetic energy anomalies and SSTAs in summer (<b>left</b>) and winter (<b>right</b>) from La Niña “−1 year” to La Niña “1 year”, where (<b>a</b>,<b>b</b>) is the year of La Niña “−1”, (<b>c</b>,<b>d</b>) is the year of La Niña “0”, and (<b>e</b>,<b>f</b>) is the year of La Niña “1”.</p>
Full article ">
21 pages, 4929 KiB  
Article
Climatic Background and Prediction of Boreal Winter PM2.5 Concentrations in Hubei Province, China
by Yuanyue Huang, Zijun Tang, Zhengxuan Yuan and Qianqian Zhang
Atmosphere 2025, 16(1), 52; https://doi.org/10.3390/atmos16010052 - 7 Jan 2025
Viewed by 451
Abstract
This study investigates the climatic background of winter PM2.5 (particulate matter with a diameter of 2.5 micrometers or smaller) concentrations in Hubei Province (DJF-HBPMC) and evaluates its predictability. The key findings are as follows: (1) Elevated DJF-HBPMC levels are associated with an upper-tropospheric [...] Read more.
This study investigates the climatic background of winter PM2.5 (particulate matter with a diameter of 2.5 micrometers or smaller) concentrations in Hubei Province (DJF-HBPMC) and evaluates its predictability. The key findings are as follows: (1) Elevated DJF-HBPMC levels are associated with an upper-tropospheric northerly anomaly, a deepened southern branch trough (SBT) that facilitates southwesterly flow into central and eastern China, and a weakened East Asian winter monsoon (EAWM), which reduces the frequency and intensity of cold air intrusions. Near-surface easterlies and an anomalous anticyclonic circulation over Hubei contribute to reduced precipitation, thereby decreasing the dispersion of pollutants and leading to higher PM2.5 concentrations. (2) Significant correlations are observed between DJF-HBPMC and sea surface temperature (SST) anomalies in specific oceanic regions, as well as sea-ice concentration (SIC) anomalies near the Antarctic. For the atmospheric pattern anomalies over Hubei Province, the North Atlantic SST mode (NA) promotes the southward intrusion of northerlies, while the Northwest Pacific (NWP) and South Pacific (SPC) SST modes enhance wet deposition through increased precipitation, showing a negative correlation with DJF-HBPMC. Conversely, the South Atlantic–Southwest Indian Ocean SST mode (SAIO) and the Ross Sea sea-ice mode (ROSIC) contribute to more stable local atmospheric conditions, which reduce pollutant dispersion and increase PM2.5 accumulation, thus exhibiting a positive correlation with DJF-HBPMC. (3) A multiple linear regression (MLR) model, using selected seasonal SST and SIC indices, effectively predicts DJF-HBPMC, showing high correlation coefficients (CORR) and anomaly sign consistency rates (AS) compared to real-time values. (4) In daily HBPMC forecasting, both the Reversed Unrestricted Mixed-Frequency Data Sampling (RU-MIDAS) and Reversed Restricted-MIDAS (RR-MIDAS) models exhibit superior skill using only monthly precipitation, and the RR-MIDAS offers the best balance in prediction accuracy and trend consistency when incorporating monthly precipitation along with monthly SST and SIC indices. Full article
Show Figures

Figure 1

Figure 1
<p>Plots of DJF-HBPMCI from 2000 to 2021 (Solid line denotes DJF-HBPMCI while dotted line denotes linear trend of DJF-HBPMCI).</p>
Full article ">Figure 2
<p>Correlation fields between DJF-HBPMCI and UVH. ((<b>a</b>–<b>d</b>) for global H200 to H1000, (<b>e</b>–<b>h</b>) for UVH200 to UVH1000. Color denotes geopotential height, vector denotes wind speed. Red denotes positive correlation while blue denotes negative correlation. Plotted-crossed regions and black vectors are significant at the 90% confidence level. Hubei Province is marked with light yellow).</p>
Full article ">Figure 3
<p>Correlation fields between DJF-HBPMCI and meridional–vertical speed and climatic omega over the longitudes 110~115° E (Color denotes climatic omega of 100 times larger, vector denotes the correlation coefficient of meridional–vertical speed. Red denotes downward motion while blue denotes upward motion. Black vectors are significant at the 90% confidence level).</p>
Full article ">Figure 4
<p>Correlation fields between DJF-HBPMCI and global SST and SIC ((<b>a</b>) for SST, (<b>b</b>) for north polar SIC, and (<b>c</b>) for south polar SIC. Red denotes positive correlation while blue denotes negative correlation. Dotted regions are significant at the 90% confidence level).</p>
Full article ">Figure 5
<p>Correlation fields between DJF-NAI and atmospheric circulation. Color denotes geopotential height, vector denotes wind speed. Red denotes positive correlation while blue denotes negative correlation. ((<b>a</b>–<b>c</b>) for global H200 to H1000, (<b>d</b>–<b>f</b>) for UVH200 to UVH1000 Plotted-crossed regions and black vectors are significant at the 90% confidence level. Hubei Province is marked with light yellow).</p>
Full article ">Figure 6
<p>Correlation fields between DJF-NWPI and atmospheric circulation Color denotes geopotential height, vector denotes wind speed. Red denotes positive correlation while blue denotes negative correlation. ((<b>a</b>–<b>c</b>) for global H200 to H1000, (<b>d</b>–<b>f</b>) for UVH200 to UVH1000 Plotted-crossed regions and black vectors are significant at the 90% confidence level. Hubei Province is marked with light yellow).</p>
Full article ">Figure 7
<p>Correlation fields between DJF-SPCI and atmospheric circulation Color denotes geopotential height, vector denotes wind speed. Red denotes positive correlation while blue denotes negative correlation. ((<b>a</b>–<b>c</b>) for global H200 to H1000, (<b>d</b>–<b>f</b>) for UVH200 to UVH1000 Plotted-crossed regions and black vectors are significant at the 90% confidence level. Hubei Province is marked with light yellow).</p>
Full article ">Figure 8
<p>Correlation fields between DJF-ROSICI and atmospheric circulation. Color denotes geopotential height, vector denotes wind speed. Red denotes positive correlation while blue denotes negative correlation. ((<b>a</b>–<b>c</b>) for global H200 to H1000, (<b>d</b>–<b>f</b>) for UVH200 to UVH1000 Plotted-crossed regions and black vectors are significant at the 90% confidence level. Hubei Province is marked with light yellow).</p>
Full article ">Figure 9
<p>Multiple linear regression (MLR) for DJF-HBPMCI prediction. (<b>a</b>,<b>d</b>) for correlation between indices and DJF-HBPMCI reaching the 90% confidence level. (<b>b</b>,<b>e</b>) for 95%. (<b>c</b>,<b>f</b>) for 99%. (<span class="html-italic">x</span>-axis denotes DJF-HBPMCI in (<b>a</b>–<b>c</b>) and denotes DJF-HBPMCI anomaly in (<b>d</b>–<b>f</b>); <span class="html-italic">y</span>-axis denotes years. Green bars denote real-time DJF-HBPMCI, blue solid-dotted line denotes hindcast DJF-HBPMCI, red dot denotes prediction of DJF-HBPMCI in (<b>a</b>–<b>c</b>), while red solid-dotted line denotes prediction of DJF-HBPMCI anomaly in (<b>d</b>–<b>f</b>)).</p>
Full article ">Figure 10
<p>Predictive results for DJF-HBPMC by monthly precipitation. (<b>a</b>,<b>c</b>) Monthly precipitation of Hubei Province and chosen SST-SIC indices (<b>b</b>,<b>d</b>) based on MIDAS. ((<b>a</b>,<b>b</b>) for RR-MIDAS while (<b>c</b>,<b>d</b>) for RU-MIDAS, blue lines denote real-time daily HBPMCI while red lines denote predictions).</p>
Full article ">
19 pages, 9780 KiB  
Article
Sedimentary Signatures of Super Typhoon Haiyan: Insight from Core Record in South China Sea
by Yu-Huang Chen, Chih-Chieh Su, Pai-Sen Yu, Ta-Wei Hsu, Sheng-Ting Hsu, Hsing-Chien Juan, Yuan-Pin Chang, Yu-Fang Ma and Shye-Donq Chiu
J. Mar. Sci. Eng. 2025, 13(1), 10; https://doi.org/10.3390/jmse13010010 - 25 Dec 2024
Viewed by 652
Abstract
Sedimentary records of event deposits are crucial for regional natural disaster risk assessments and hazard history reconstructions. After Super Typhoon Haiyan passed through the South China Sea in 2013, five gravity cores were collected along the typhoon path in the southern South China [...] Read more.
Sedimentary records of event deposits are crucial for regional natural disaster risk assessments and hazard history reconstructions. After Super Typhoon Haiyan passed through the South China Sea in 2013, five gravity cores were collected along the typhoon path in the southern South China Sea basin (>3800 mbsl). The results showed that Super Typhoon Haiyan deposits with clear graded bedding are preserved at the top of all cores. The thickness of the typhoon layers ranges from 20 to 240 cm and is related to changes in typhoon intensity. The lack of river-connected submarine canyon systems limited the transportation of terrestrial sediments from land to sea. Super Typhoon Haiyan-induced large surface waves played an important role in carrying suspended sediment from the Philippines. The Mn-rich layers at the bottom of the typhoon layers may be related to the soil and rock composition of the Palawan region, which experienced tsunami-like storm surges caused by Super Typhoon Haiyan. These Mn-rich layers may serve as a proxy for sediment export from large-scale extreme terrigenous events. This study provides the first sedimentary record of extreme typhoon events in the deep ocean, which may shed light on reconstructing regional hazard history. Full article
(This article belongs to the Section Geological Oceanography)
Show Figures

Figure 1

Figure 1
<p>The sampling location of gravity cores used in this study and the path of Super Typhoon Haiyan. The gray dots indicate the locations and times of Super Typhoon Haiyan. The red dots represent the sample sites.</p>
Full article ">Figure 2
<p>X-radiographs, surface images, bulk density, grain size median, sorting, and Itrax Mn/Fe ratio and Itrax Ca/Ti ratio profiles of OR1-1133-A5 and OR1-1068-8. Sedimentary units correspond to properties of sediments. Eight sediment units could be identified based on these parameters, and surface layers of OR1-1068-8 and OR1-1133-A5 contained same sediment units.</p>
Full article ">Figure 3
<p>Grain size distribution curves of the sediment units in OR1-1068-8 and OR1-1133-A5. The average grain size distribution for each unit is represented by colored solid lines, while the gray dashed lines are the grain size distribution for each centimeter within the units. Units A, B, C, F, and H exhibit similar grain size distributions. The grain size composition for each centimeter within Unit C and F shows greater variation. Unit D displays a relatively unimodal distribution. Unit E and Unit G have the broadest distribution (poor sorting). The <span class="html-italic">X</span>-axis is the grain diameter (ϕ) and the <span class="html-italic">Y</span>-axis is the volume percentage.</p>
Full article ">Figure 4
<p>Ternary diagram showing clay mineral compositions (illite + chlorite, kaolinite, and smectite). Shaded regions represent clay mineral provinces from different source areas [<a href="#B28-jmse-13-00010" class="html-bibr">28</a>]. Symbols denote different units from cores OR1-1068-8 (dots) and OR1-1133-A5 (crosses). Unit E (black symbols) clusters near North Borneo and North Palawan compositions. Unit G (yellow symbols) shows stronger sedimentary influence from Luzon sources. Other units show wider distribution.</p>
Full article ">Figure 5
<p>XRF mapping of core OR1-1068-8. Left: Itrax Mn/Fe ratio profile showing sampling depths for XRF mapping (horizontal lines). Center: Resin-embedded samples with mapped areas (red boxes) and analyzed particles (red circles). Right: Spatial distribution of major elements (Si, Al, Fe, Mn) in mapped areas. Pie charts show elemental compositions of specific particles, with Fe being dominant, followed by Si, Mn, Al, and K. Element distributions correlate with grain patterns, where Si, Al, and K indicate feldspar presence, while Fe and Mn reflect detrital input under oxidizing conditions [<a href="#B40-jmse-13-00010" class="html-bibr">40</a>,<a href="#B41-jmse-13-00010" class="html-bibr">41</a>].</p>
Full article ">
28 pages, 19078 KiB  
Article
Analysis of PM2.5 Pollution Transport Characteristics and Potential Sources in Four Chinese Megacities During 2022: Seasonal Variations
by Kun Mao, Yuan Yao, Kun Wang, Chen Liu, Guangmin Tang, Shumin Feng, Yue Shen, Anhua Ju, Hao Zhou and Zhiyu Li
Atmosphere 2024, 15(12), 1482; https://doi.org/10.3390/atmos15121482 - 12 Dec 2024
Viewed by 841
Abstract
Atmospheric particulate pollution in China’s megacities has heightened public concern over air quality, highlighting the need for precise identification of urban pollution characteristics and pollutant transport mechanisms to enable effective control and mitigation. In this study, a new method combing the High Accuracy [...] Read more.
Atmospheric particulate pollution in China’s megacities has heightened public concern over air quality, highlighting the need for precise identification of urban pollution characteristics and pollutant transport mechanisms to enable effective control and mitigation. In this study, a new method combing the High Accuracy Surface Modeling (HASM) and Multiscale Geographically Weighted Regression (MGWR) was proposed to derive seasonal high spatial resolution PM2.5 concentrations. The Hybrid Single-Particle Lagrangian Integrated Trajectory (HYSPLIT) was applied to analyze the seasonal spatial variations, transport pathways, and potential sources of PM2.5 concentrations across China’s four megacities: Beijing, Shanghai, Xi’an, and Chengdu. The result indicates that: (1) the proposed method outperformed Kriging, inverse distance weighting (IDW), and HASM, with coefficient of determination values ranging from 0.91 to 0.94, and root mean square error values ranging from 1.98 to 2.43 µg/m3, respectively; (2) all cities show a similar seasonal pattern, with PM2.5 concentrations highest in winter, followed by spring, autumn, and summer; Beijing has higher concentrations in the south, Shanghai and Xi’an in the west, and Chengdu in central urban areas, decreasing toward the rural area; (3) potential source contribution function and concentration weighted trajectory analysis indicate that Beijing’s main potential PM2.5 sources are in Hebei Province (during winter, spring, and autumn), Shanghai’s are in the Yellow Sea and the East China Sea, Xi’an’s are in Southern Shaanxi Province, and Chengdu’s are in Northeastern and Southern Sichuan Province, with all cities experiencing higher impacts in winter; (4) there is a negative correlation between precipitation, air temperature, and seasonal PM2.5 levels, with anthropogenic emissions sources such as industry combustion, power plants, residential combustion, and transportation significantly impact on seasonal PM2.5 pollution. Full article
(This article belongs to the Section Air Quality)
Show Figures

Figure 1

Figure 1
<p>Location of the Beijing, Shanghai, Xi’an, and Chengdu in China.</p>
Full article ">Figure 2
<p>Flowchart of the proposed method downscaling process.</p>
Full article ">Figure 3
<p>Interpolation accuracy validation of PM<sub>2.5</sub> concentrations for the four megacities. (<b>a</b>) Beijing. (<b>b</b>) Shanghai. (<b>c</b>) Xi’an. (<b>d</b>) Chengdu.</p>
Full article ">Figure 4
<p>Seasonal PM<sub>2.5</sub> concentration surfaces for the four megacities predicted by the proposed method.</p>
Full article ">Figure 5
<p>Seasonal cluster–mean backward trajectories arriving in Beijing during 2022, showing the main transport pathways of air masses (II: Inner Mongolia Autonomous Region, XVIII: Hebei Province). (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Autumn. (<b>d</b>) Winter.</p>
Full article ">Figure 6
<p>Seasonal cluster–mean backward trajectories arriving in Shanghai in 2022, showing the main transport pathways of air masses (II: Inner Mongolia Autonomous Region, III: Liaoning Province, IV: Shandong Province, VII: Jiangsu Province, XIII: Zhejiang Province, XIV: Fujian Province, XV: Anhui Province, XVIII: Hebei Province). (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Autumn. (<b>d</b>) Winter.</p>
Full article ">Figure 6 Cont.
<p>Seasonal cluster–mean backward trajectories arriving in Shanghai in 2022, showing the main transport pathways of air masses (II: Inner Mongolia Autonomous Region, III: Liaoning Province, IV: Shandong Province, VII: Jiangsu Province, XIII: Zhejiang Province, XIV: Fujian Province, XV: Anhui Province, XVIII: Hebei Province). (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Autumn. (<b>d</b>) Winter.</p>
Full article ">Figure 7
<p>Seasonal cluster–mean backward trajectories arriving in Xi’an in 2022, showing the main transport pathways of air masses (II: Inner Mongolia Autonomous Region, V: Shaanxi Province, VI: Henan Province, VIII: Gansu Province, X: Hubei Province, XI: Chongqing City, XVIII: Xinjiang Uygur Autonomous Region). (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Autumn. (<b>d</b>) Winter.</p>
Full article ">Figure 8
<p>Seasonal cluster–mean backward trajectories arriving in Chengdu in 2022, showing the main transport pathways of air masses (V: Shaanxi Province, VIII: Gansu Province, IX: Sichuan Province, XI: Chongqing City, XXI: Tibet Autonomous Region, XXII: Guizhou Province). (<b>a</b>) Spring. (<b>b</b>) Summer. (<b>c</b>) Autumn. (<b>d</b>) Winter.</p>
Full article ">Figure 9
<p>Source identification for PM<sub>2.5</sub> over Beijing, Shanghai, Xi’an, and Chengdu using PSCF analysis during four seasons in 2022.</p>
Full article ">Figure 10
<p>Source identification for PM<sub>2.5</sub> over Beijing, Shanghai, Xi’an, and Chengdu using CWT analysis during four seasons in 2022.</p>
Full article ">Figure 10 Cont.
<p>Source identification for PM<sub>2.5</sub> over Beijing, Shanghai, Xi’an, and Chengdu using CWT analysis during four seasons in 2022.</p>
Full article ">Figure 11
<p>Air temperature distribution in the primary pollution transport pathway areas for four study areas during three seasons in 2022.</p>
Full article ">Figure 12
<p>Precipitation distribution in the primary pollution transport pathway areas for four study areas during three seasons in 2022.</p>
Full article ">Figure 12 Cont.
<p>Precipitation distribution in the primary pollution transport pathway areas for four study areas during three seasons in 2022.</p>
Full article ">
18 pages, 10124 KiB  
Article
Origin, Migration, and Accumulation of Crude Oils in the Chaoyang Step-Fault Zone, Fushan Depression, Beibuwan Basin: Insight from Geochemical Evidence and Basin Modeling
by Yang Shi, Hao Guo, Xiaohan Li, Huiqi Li, Meijun Li, Xin Wang, Surui Dong and Xi He
Energies 2024, 17(23), 5842; https://doi.org/10.3390/en17235842 - 21 Nov 2024
Viewed by 561
Abstract
The Fushan Depression is a hydrocarbon-rich depression in the Beibuwan Basin, South China Sea. In this study, 14 source rocks and 19 crude oils from the Chaoyang Step-Fault Zone and Southern Slope Zone were geochemically analyzed to determine their origins. The hydrocarbon generation, [...] Read more.
The Fushan Depression is a hydrocarbon-rich depression in the Beibuwan Basin, South China Sea. In this study, 14 source rocks and 19 crude oils from the Chaoyang Step-Fault Zone and Southern Slope Zone were geochemically analyzed to determine their origins. The hydrocarbon generation, migration, and accumulation processes were also determined using two-dimensional basin modeling. Crude oils from the low-step area show a close relationship with the source rocks of the first and second members of the Eocene Liushagang Formation (Els1 and Els2). The oils from the middle-step area and the Southern Slope Zone are derived from the local source rocks in those areas, in the third member of the Eocene Liushagang Formation (Els3). Hydrocarbons generated from the Els3 source rocks of the Southern Slope Zone migrated along sand bodies to the Els3 reservoir. The fault system of the Chaoyang Step-Fault Zone controls hydrocarbon migration and accumulation in the low-step and middle-step areas. The resource potential of the middle-step area is limited by its shallow burial depth. The low-step area is a more favorable exploration area due to its proximity to the source kitchen. Full article
Show Figures

Figure 1

Figure 1
<p>Geological maps showing (<b>a</b>) the location of the Fushan Depression and (<b>b</b>) the tectonic distribution units and sampling wells in the Fushan Depression.</p>
Full article ">Figure 2
<p>Stratigraphic column and sequence division of the Fushan Depression (After Ma et al., 2012 [<a href="#B12-energies-17-05842" class="html-bibr">12</a>]).</p>
Full article ">Figure 3
<p>Two-dimensional lithofacies models for the representative profiles in the Chaoyang Step-Fault Zone.</p>
Full article ">Figure 4
<p>Variation of (<b>a</b>) (S<sub>1</sub>+S<sub>2</sub>) with total organic carbon (TOC) content and (<b>b</b>) hydrogen index with Tmax for source rocks from the Fushan Depression showing the hydrocarbon-generating potential and organic matter type, respectively. (modified after Robison et al., 1999 [<a href="#B29-energies-17-05842" class="html-bibr">29</a>] and Mukhopadhyay et al., 1995 [<a href="#B30-energies-17-05842" class="html-bibr">30</a>]).</p>
Full article ">Figure 5
<p>Total ion chromatogram (TIC) of the oil samples and source rocks from the Fushan Depression. (<b>a</b>) E<span class="html-italic">ls</span><sub>1</sub> oil from low-step area, well Ch8, 2710 m; (<b>b</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from middle-step area, well Ch22, 2145 m; (<b>c</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from Southern Slope Zone, well M15-3, 3131; (<b>d</b>) E<span class="html-italic">ls</span><sub>1</sub> source rock from low-step area, well Ch12, 2545 m; (<b>e</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from middle-step area, well Ch23, 2386–2390 m; (<b>f</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from Southern Slope Zone, well M17, 3803 m.</p>
Full article ">Figure 6
<p>Mass chromatograms (<span class="html-italic">m</span>/<span class="html-italic">z</span> 191) showing the distribution of tricyclic and tetracyclic terpanes in the crude oils and source rocks from the Fushan Depression. (<b>a</b>) E<span class="html-italic">ls</span><sub>1</sub> oil from low-step area, well Ch8, 2710 m; (<b>b</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from middle-step area, well Ch22, 2145 m; (<b>c</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from Southern Slope Zone, well M15-3, 3131; (<b>d</b>) E<span class="html-italic">ls</span><sub>1</sub> source rock from low-step area, well Ch12, 2545 m; (<b>e</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from middle-step area, well Ch23, 2386–2390 m; (<b>f</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from Southern Slope Zone, well M17, 3803 m.</p>
Full article ">Figure 7
<p>Mass chromatograms (<span class="html-italic">m</span>/<span class="html-italic">z</span> 191) showing the distribution of hopanes in crude oils and source rocks from the Fushan Depression. (<b>a</b>) E<span class="html-italic">ls</span><sub>1</sub> oil from low-step area, well Ch8, 2710 m; (<b>b</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from middle-step area, well Ch22, 2145 m; (<b>c</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from Southern Slope Zone, well M15-3, 3131; (<b>d</b>) E<span class="html-italic">ls</span><sub>1</sub> source rock from low-step area, well Ch12, 2545 m; (<b>e</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from middle-step area, well Ch23, 2386–2390 m; (<b>f</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from Southern Slope Zone, well M17, 3803 m.</p>
Full article ">Figure 8
<p>Correlation between the relative abundance of tri- and tetracyclic terpanes in the studied oils and source rocks. (<b>a</b>) OL/C<sub>30</sub>H vs. C<sub>19+20</sub>/C<sub>23</sub>TT; (<b>b</b>) Z/(Z+C<sub>24</sub>TT) vs. C<sub>19+20</sub>/C<sub>21</sub>TT; (<b>c</b>) X<sub>1</sub>/(X<sub>1</sub>+C<sub>24</sub>TT) vs. Y<sub>1</sub>/(Y<sub>1</sub>+C<sub>24</sub>TT); (<b>d</b>) Y<sub>1</sub>/(Y<sub>1</sub>+C<sub>24</sub>TT) vs. OL/C<sub>30</sub>H. Note: OL = oleanane; TT = tricyclic terpane; Y<sub>1</sub> = C<sub>24</sub>-des-A-oleanane, X<sub>1</sub> = C<sub>24</sub>-des-A-lupane, Z = C<sub>24</sub>-des-A-ursane.</p>
Full article ">Figure 9
<p>Mass chromatograms (<span class="html-italic">m</span>/<span class="html-italic">z</span> 217) showing the distributions of steranes in crude oils and source rocks from the Fushan Depression. (<b>a</b>) E<span class="html-italic">ls</span><sub>1</sub> oil from low-step area, well Ch8, 2710 m; (<b>b</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from middle-step area, well Ch22, 2145 m; (<b>c</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from Southern Slope Zone, well M15-3, 3131; (<b>d</b>) E<span class="html-italic">ls</span><sub>1</sub> source rock from low-step area, well Ch12, 2545 m; (<b>e</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from middle-step area, well Ch23, 2386–2390 m; (<b>f</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from Southern Slope Zone, well M17, 3803 m.</p>
Full article ">Figure 10
<p>Mass chromatograms (<span class="html-italic">m</span>/<span class="html-italic">z</span> 245) of the aromatic fraction showing the distribution of methyl-triaromatic steroids in crude oils and source rocks from the Fushan Depression. (<b>a</b>) E<span class="html-italic">ls</span><sub>1</sub> oil from low-step area, well Ch8, 2710 m; (<b>b</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from middle-step area, well Ch22, 2145 m; (<b>c</b>) E<span class="html-italic">ls</span><sub>3</sub> oil from Southern Slope Zone, well M15-3, 3131; (<b>d</b>) E<span class="html-italic">ls</span><sub>1</sub> source rock from low-step area, well Ch12, 2545 m; (<b>e</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from middle-step area, well Ch23, 2386–2390 m; (<b>f</b>) E<span class="html-italic">ls</span><sub>3</sub> source rock from Southern Slope Zone, well M17, 3803 m. Notes: 1 = C<sub>27</sub> 3-methyltriaromatic steroids; 2 = C<sub>27</sub> 4-methyltriaromatic steroids; 3 = C<sub>27</sub> 3-methyltriaromatic steroids + C<sub>28</sub> 3,24-dimethyltriaromatic steroids; 4 = C<sub>29</sub> 4,23,24-trimethyltriaromatic steroids; 5 = C<sub>27</sub> 4-methyltriaromatic steroids + C<sub>29</sub> 4-methyl-24-ethyltriaromatic steroids; 6 = C<sub>29</sub> 4,23,24-trimethyltriaromatic steroids; 7 = C<sub>29</sub> 3-methyl-24-ethyltriaromatic steroids; 8 = C<sub>29</sub> 4,23,24-trimethyltriaromatic steroids; 9 = C<sub>29</sub> 4-methyl-24-ethyltriaromatic steroids; 10 = C<sub>28</sub> 3,24-dimethyltriaromatic steroids; 11 = C<sub>28</sub> 3,24-dimethyltriaromatic steroids; 12 = C<sub>29</sub> 4α,23,24-trimethyltriaromatic steroids; 13 = C<sub>29</sub> 4α,23,24-trimethyltriaromatic steroids; 14 = C<sub>29</sub> 3-methyl-24-ethyltriaromatic steroids; 15 = C<sub>29</sub> 4α,23,24-trimethyltriaromatic steroids; 16 = C<sub>29</sub> 4α- methyl-24-ethyltriaromatic steroids; 17 = C<sub>29</sub> 4α,23,24-trimethyltriaromatic steroids.</p>
Full article ">Figure 11
<p>Correlation between the relative abundances of methyl-triaromatic steroids relative abundance in the studied oils and source rocks.</p>
Full article ">Figure 12
<p>Cross plot of the Methyl-Phenanthrene Distribution Fraction (MPDF) parameters F1 vs. F2 showing the maturity of crude oil samples.</p>
Full article ">Figure 13
<p>Evolution of the maturity and the transformation ratio with age in Profile AA′. (<b>a</b>) Maturity history at 30 Ma; (<b>b</b>) maturity history at 23.5 Ma; (<b>c</b>) maturity history at 10 Ma; (<b>d</b>) maturity history at 0 Ma; (<b>e</b>) transformation ratio at 30 Ma; (<b>f</b>) transformation ratio at 23.5 Ma; (<b>g</b>) transformation ratio at 10 Ma; (<b>h</b>) transformation ratio at 0 Ma [<a href="#B24-energies-17-05842" class="html-bibr">24</a>].</p>
Full article ">Figure 14
<p>Hydrocarbon generation in different source rocks over geological time.</p>
Full article ">Figure 15
<p>Simulation results of hydrocarbon migration and accumulation in Profile AA’. (<b>a</b>) Hydrocarbon migration and accumulation at 25 Ma; (<b>b</b>) hydrocarbon migration and accumulation at 10 Ma; (<b>c</b>) hydrocarbon migration and accumulation at 0 Ma.</p>
Full article ">Figure 16
<p>Timing and episodes of oil charging based on fluid inclusion observation and burial history—thermal history reconstruction of the Well Ch23.</p>
Full article ">Figure 17
<p>A conceptual model showing oil migration and accumulation in the Chaoyang Step-Fault Zone.</p>
Full article ">
17 pages, 7484 KiB  
Article
Prediction of the Potentially Suitable Areas of Sesame in China Under Climate Change Scenarios Using MaxEnt Model
by Guoqiang Li, Xue Wang, Jie Zhang, Feng Hu, Hecang Zang, Tongmei Gao, Youjun Li and Ming Huang
Agriculture 2024, 14(11), 2090; https://doi.org/10.3390/agriculture14112090 - 20 Nov 2024
Viewed by 855
Abstract
Sesame (Sesamum indicum L, flora of China) is an essential oil crop in China, but its growth and development are affected by climate change. To cope with the impacts of climate change on sesame cultivation, we used the Maximum Entropy (MaxEnt) model [...] Read more.
Sesame (Sesamum indicum L, flora of China) is an essential oil crop in China, but its growth and development are affected by climate change. To cope with the impacts of climate change on sesame cultivation, we used the Maximum Entropy (MaxEnt) model to analyze the bioclimatic variables of climate suitability of sesame in China and predicted the suitable area and trend of sesame in China under current and future climate scenarios. The results showed that the MaxEnt model prediction was excellent. The most crucial bioclimatic variable influencing the distribution of sesame was max temperature in the warmest month, followed by annual mean temperature, annual precipitation, mean diurnal range, and precipitation of the driest month. Under the current climate scenario, the suitable areas of sesame were widely distributed in China, from south (Hainan) to north (Heilongjiang) and from east (Yellow Sea) to west (Tibet). The area of highly suitable areas was 64.51 × 104 km2, accounting for 6.69% of the total land area in China, and was primarily located in mainly located in southern central Henan, eastern central Hubei, northern central Anhui, northern central Jiangxi, and eastern central Hunan. The area of moderately suitable areas and lowly suitable areas accounted for 17.45% and 25.82%, respectively. Compared with the current climate scenario, the area of highly and lowly suitable areas under future climate scenarios increased by 0.10%–11.48% and 0.08%–8.67%, while the area of moderately suitable areas decreased by 0.31%–23.03%. In addition, the increased highly suitable areas were mainly distributed in northern Henan. The decreased moderately suitable areas were mainly distributed in Heilongjiang, Jilin, and Liaoning. This work is practically significant for optimizing the regional layout of sesame cultivation in response to future climate conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution points of sesame in China.</p>
Full article ">Figure 2
<p>Results of jackknife test of bioclimatic variables for the suitability of sesame.</p>
Full article ">Figure 3
<p>Responses of the five major bioclimatic variables to sesame: (<b>a</b>) max temperature of warmest month, (<b>b</b>) annual mean temperature, (<b>c</b>) annual precipitation, (<b>d</b>) mean diurnal range, and (<b>e</b>) precipitation of driest month.</p>
Full article ">Figure 4
<p>Receiver operating curve with the corresponding area under the curve.</p>
Full article ">Figure 5
<p>Potentially suitable areas for sesame under the current climate scenario in China.</p>
Full article ">Figure 6
<p>Potentially suitable areas for sesame under future climate scenarios in China.</p>
Full article ">Figure 7
<p>Changes in suitable areas for sesame in different periods.</p>
Full article ">
20 pages, 9002 KiB  
Article
Projecting the Potential Global Distribution of Sweetgum Inscriber, Acanthotomicus suncei (Coleoptera: Curculionidae: Scolytinae) Concerning the Host Liquidambar styraciflua Under Climate Change Scenarios
by Kaitong Xiao, Lei Ling, Ruixiong Deng, Beibei Huang, Yu Cao, Qiang Wu, Hang Ning and Hui Chen
Insects 2024, 15(11), 897; https://doi.org/10.3390/insects15110897 - 18 Nov 2024
Viewed by 1231
Abstract
Acanthotomicus suncei is a newly discovered bark beetle in China that significantly threatens the American sweetgum Liquidambar styraciflua. In recent years, this pest has spread from its original habitat to many surrounding cities, causing substantial economic and ecological losses. Considering the wide [...] Read more.
Acanthotomicus suncei is a newly discovered bark beetle in China that significantly threatens the American sweetgum Liquidambar styraciflua. In recent years, this pest has spread from its original habitat to many surrounding cities, causing substantial economic and ecological losses. Considering the wide global distribution of its host, Liquidambar styraciflua, this pest is likely to continue to spread and expand. Once the pest colonizes a new climatically suitable area, the consequences could be severe. Therefore, we employed the CLIMEX and Random Forests model to predict the potential suitable distribution of A. suncei globally. The results showed that A. suncei was mainly distributed in Southern China, in South Hokkaido in Japan, Southern USA, the La Plata Plain in South America, southeastern Australia, and the northern Mediterranean; these areas are located in subtropical monsoon, monsoonal humid climates, or Mediterranean climate zones. Seasonal rainfall, especially in winter, is a key environmental factor that affects the suitable distribution of A. suncei. Under future climates, the total suitable area of A. suncei is projected to decrease to a certain extent. However, changes in its original habitat require serious attention. We found that A. suncei exhibited a spreading trend in Southwest, Central, and Northeast China. Suitable areas in some countries in Southeast and South Asia bordering China are also expected to show an increased distribution. The outward spread of this pest via sea transportation cannot be ignored. Hence, quarantine efforts should be concentrated in high-suitability regions determined in this study to protect against the occurrence of hosts that may contain A. suncei, thereby avoiding its long-distance spread. Long-term sentinel surveillance and control measures should be carried out as soon as A. suncei is detected, especially in regions with high suitability. Thus, our findings establish a theoretical foundation for quarantine and control measures targeting A. suncei. Full article
(This article belongs to the Section Insect Ecology, Diversity and Conservation)
Show Figures

Figure 1

Figure 1
<p>Occurrence coordinates of <span class="html-italic">A. suncei</span> in China.</p>
Full article ">Figure 2
<p>Global presence coordinates of <span class="html-italic">L. styraciflua</span>.</p>
Full article ">Figure 3
<p>Potential global distribution of <span class="html-italic">L. styraciflua</span> predicted by CLEMEX; (<b>a</b>) distribution under current climate scenarios; (<b>b</b>) distribution under A1B climate scenarios in 2050.</p>
Full article ">Figure 4
<p>Habitat suitability difference from current to future climate scenarios: (<b>a</b>) habitat suitability difference in <span class="html-italic">L. styraciflua</span>; (<b>b</b>) habitat suitability difference in <span class="html-italic">A. suncei</span> (the value in each raster is calculated by the future suitability value minus the current suitability value).</p>
Full article ">Figure 5
<p>Potential global distribution of <span class="html-italic">A. suncei</span> predicted by Random Forests: (<b>a</b>) distribution under current climate scenario; (<b>b</b>) distribution under an RCP6.0 climate scenario in 2050.</p>
Full article ">Figure 6
<p>Contribution rates of important variables affecting the potential distribution of <span class="html-italic">A. suncei</span>.</p>
Full article ">Figure 7
<p>Potential global distribution of <span class="html-italic">A. suncei</span> concerning the host <span class="html-italic">L. styraciflua</span>: (<b>a</b>) distribution under current climate scenarios; (<b>b</b>) distribution under future climate scenarios in 2050.</p>
Full article ">Figure 8
<p>The area of <span class="html-italic">A. suncei</span> habitats of each suitability level in each continent under current climate scenarios (<b>left</b> column) and under the RCP6.0 climate scenario in 2050 (<b>right</b> column).</p>
Full article ">
16 pages, 4247 KiB  
Article
Multivariate Environmental Factors and Seasonal Spatial Dynamics Affecting the Phytoplankton Community in Yazhou Bay, South China Sea
by Zhida Yu, Zhiyuan Ouyang, Chuanyang Zheng, Zhen Wang, Xiaoming Fu, Haiping Cui, Yadong Huang, Zitao Zhang and Chenhui Xiang
Water 2024, 16(22), 3181; https://doi.org/10.3390/w16223181 - 6 Nov 2024
Viewed by 858
Abstract
This study investigated phytoplankton and water environmental factors in Yazhou Bay, South China Sea, during the winter, spring, and summer of 2023. It examined phytoplankton community structure, subgroup heterogeneity, and key environmental drivers. Phytoplankton abundance ranged from 0.08 to 14.30 × 10⁴ cells·L [...] Read more.
This study investigated phytoplankton and water environmental factors in Yazhou Bay, South China Sea, during the winter, spring, and summer of 2023. It examined phytoplankton community structure, subgroup heterogeneity, and key environmental drivers. Phytoplankton abundance ranged from 0.08 to 14.30 × 10⁴ cells·L−1, with high concentrations in estuary and nearshore zones. In summer, currents carry phytoplankton offshore, with stratification leading to high sedimentation in southern offshore waters. RDA results indicated that in winter and spring, inorganic nitrogen mainly influences phytoplankton distribution, while silicate is the primary factor in summer. Although seasonal differences in total phytoplankton abundance are minimal, significant horizontal and vertical distribution variations exist. Diverse preferences of different phytoplankton species for temperature, salinity, nitrogen, and phosphorus result in high species diversity. The Shannon–Wiener diversity index (H′) averages 3.96 ± 0.09, and the Pielou evenness index (J) averages 0.82 ± 0.01. Dominant species include Pseudo-nitzschia pungens, Skeletonema costatum, and Rhizosolenia sinica. Influenced by external oceanic water masses, estuary input, and islands, phytoplankton subgroups show regional and seasonal variations. Despite recorded harmful algal blooms (HABs) in adjacent waters, Yazhou Bay’s high biodiversity and low cell density suggest a low HAB risk, though future risks due to climate change and human activities remain. Full article
(This article belongs to the Section Oceans and Coastal Zones)
Show Figures

Figure 1

Figure 1
<p>Sampling sites in Yazhou Bay.</p>
Full article ">Figure 2
<p>Seasonal variation in Shannon–Wiener diversity index (<span class="html-italic">H</span>′) at surface and bottom water layers in Yazhou Bay. The color bars indicate the level of the Shannon–Wiener diversity index.</p>
Full article ">Figure 3
<p>Seasonal abundance distribution at surface and bottom water layers in Yazhou Bay. The color bars indicate the level of phytoplankton abundance (cells·L<sup>−1</sup>).</p>
Full article ">Figure 4
<p>Hierarchical clustering of phytoplankton communities in surface and bottom water layers of Yazhou Bay. Subgroups include an Offshore Zone (OZ), Mixed Zone (MZ), Estuary and Nearshore Zone (ENZ), and Island Zone (IZ). In winter, the estuarine and nearshore area is further divided into an Estuary Zone (EZ) and Nearshore Zone (NZ) based on subcommunity differences.</p>
Full article ">Figure 5
<p>Regional variation in environmental and biological parameters in the surface and bottom water layers of Yazhou Bay during winter: salinity, temperature, phosphate (PO<sub>4</sub><sup>3−</sup>), silicate (SiO<sub>3</sub><sup>2−</sup>), ammonium (NH<sub>4</sub><sup>+</sup>), oxynitride (NO<sub>X</sub>), dissolved inorganic nitrogen (DIN), nitrogen phosphorus ratio (N:P), nitrogen silicon ratio (N:Si), chlorophyll a (Chl a), percentage of micro-phytoplankton (Micro_chl a (%)), percentage of nano-phytoplankton (Nano_chl a (%)), percentage of pico-phytoplankton (Pico_chl a (%)). One-way ANOVA was conducted for the environmental factors of the zones covering 2 or more stations (NZ, MZ, and OZ). * represents <span class="html-italic">p</span>-value ≤ 0.05, ** represents <span class="html-italic">p</span>-value ≤ 0.01, *** represents <span class="html-italic">p</span>-value ≤ 0.001.</p>
Full article ">Figure 6
<p>Regional variation in environmental and biological parameters in the surface and bottom water layers of Yazhou Bay during spring salinity, temperature, phosphate (PO<sub>4</sub><sup>3−</sup>), silicate (SiO<sub>3</sub><sup>2−</sup>), ammonium (NH<sub>4</sub><sup>+</sup>), oxynitride (NO<sub>X</sub>), dissolved inorganic nitrogen (DIN), nitrogen phosphorus ratio (N:P), nitrogen silicon ratio (N:Si), chlorophyll a (Chl a), percentage of micro-phytoplankton (Micro_chl a (%)), percentage of nano-phytoplankton (Nano_chl a (%)), percentage of pico-phytoplankton (Pico_chl a (%)). One-way ANOVA was conducted for the environmental factors of the zones covering 2 or more stations (NZ, MZ, and OZ). * represents <span class="html-italic">p</span>-value ≤ 0.05, ** represents <span class="html-italic">p</span>-value ≤ 0.01, *** represents <span class="html-italic">p</span>-value ≤ 0.001.</p>
Full article ">Figure 7
<p>Regional variation in environmental and biological parameters in the surface and bottom water layers of Yazhou Bay during summer: salinity, temperature, phosphate (PO<sub>4</sub><sup>3−</sup>), silicate (SiO<sub>3</sub><sup>2−</sup>), ammonium (NH<sub>4</sub><sup>+</sup>), oxynitride (NO<sub>X</sub>), dissolved inorganic nitrogen (DIN), nitrogen phosphorus ratio (N:P), nitrogen silicon ratio (N:Si), chlorophyll a (Chl a), percentage of micro-phytoplankton (Micro_chl a (%)), percentage of nano-phytoplankton (Nano_chl a (%)), percentage of pico-phytoplankton (Pico_chl a (%)). One-way ANOVA was conducted for the environmental factors of the zones covering 2 or more stations (NZ, MZ, and OZ). * represents <span class="html-italic">p</span>-value ≤ 0.05, ** represents <span class="html-italic">p</span>-value ≤ 0.01, *** represents <span class="html-italic">p</span>-value ≤ 0.001.</p>
Full article ">Figure 8
<p>Results of redundancy analysis (RDA) of dominant species abundance and environmental factors (Temp represents temperature, Sal represents salinity, DO represents dissolved oxygen, NO3 represents nitrate, NH4 represents ammonium, DIN represents dissolved inorganic nitrogen, PO4 represents phosphate, SiO3 represents silicate) in surface and bottom water layers of Yazhou Bay across three seasons.</p>
Full article ">
15 pages, 3025 KiB  
Article
Integrated Genetic and Statolith Shape Analysis Reveals the Population Structure of Loliolus (Nipponololigo) uyii (Cephalopoda: Loliginidae) in the Coastal Waters of China
by Xiaorong Wang, Chi Zhang and Xiaodong Zheng
Diversity 2024, 16(11), 674; https://doi.org/10.3390/d16110674 - 2 Nov 2024
Viewed by 1092
Abstract
Understanding population structure is a priority for evaluating population dynamics of commercially fished cephalopods under fishing pressure and environmental changes. This study employed a multidisciplinary approach to clarify the population structure of Loliolus (Nipponololigo) uyii, a common squid in inshore [...] Read more.
Understanding population structure is a priority for evaluating population dynamics of commercially fished cephalopods under fishing pressure and environmental changes. This study employed a multidisciplinary approach to clarify the population structure of Loliolus (Nipponololigo) uyii, a common squid in inshore fisheries. Sampling was conducted multiple times to cover the distribution range across the East China Sea and South China Sea. High haplotype diversity was revealed by three gene markers (COI, 16S and ODH). Two geographical clades with significant genetic differentiation were divided through phylogenetic trees and haplotype networks. The boundary between the two clades is delineated by the Dongshan population in the southern East China Sea. Furthermore, the neutrality tests and mismatch analysis suggested that L. (N.) uyii populations may have undergone population expansion. Correspondingly, statolith differences in lateral dome and posterior indentation, along with high classification success, further supported the genetic division. The overall difference in statolith shape also efficiently identified seasonal groups in the Beibu Gulf lacking genetic differentiation. This result offers new insights into the influence of genetic and environmental factors on statolith shape. The integrated results provide a comprehensive understanding of the population structure of L. (N.) uyii, laying the foundation for resource development and the conservation of the species. Full article
Show Figures

Figure 1

Figure 1
<p>The sampling map for <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> from different locations. Zhoushan (ZS), Ningbo (NB), Ningde (ND), Dongshan (DS), Maoming (MM) and Beihai (BH). The green color represents the northern groups (ZS, NB and ND), and the orange color represents the southern groups (DS, MM and BH). The black arrows indicate ocean currents.</p>
Full article ">Figure 2
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">COI</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 3
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">16S</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 4
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">ODH</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 5
<p>Phylogenetic trees constructed using the concatenation of mitochondrial and nuclear genes. The numbers in each node represent bootstraps of maximum likelihood (ML) and posterior probabilities of Bayesian inference (BI) analyses, respectively. The purple color represents the Clade A and the green color represents the Clade B.</p>
Full article ">Figure 6
<p>MDS plot of statolith morphology for different regions (<b>a</b>) and different seasons in BBG (<b>b</b>).</p>
Full article ">Figure 7
<p>Reconstructed statolith outlines for different regions (<b>a</b>) and different seasons in BBG (<b>b</b>).</p>
Full article ">
14 pages, 3809 KiB  
Article
Distribution Characteristics of Trichiurus japonicus and Their Relationships with Environmental Factors in the East China Sea and South-Central Yellow Sea
by Xinyu Shi, Zhanhui Lu, Zhongming Wang, Jianxiong Li, Xin Gao, Zhuang Kong and Wenbin Zhu
Fishes 2024, 9(11), 439; https://doi.org/10.3390/fishes9110439 - 29 Oct 2024
Viewed by 893
Abstract
The largehead hairtail (Trichiurus japonicus) is the most productive fish caught in China. In order to understand the seasonal distribution of T. japonicus in the East China Sea and the central and southern parts of the Yellow Sea, three species distribution [...] Read more.
The largehead hairtail (Trichiurus japonicus) is the most productive fish caught in China. In order to understand the seasonal distribution of T. japonicus in the East China Sea and the central and southern parts of the Yellow Sea, three species distribution models were used in this study, namely the random-forest model, K-nearest-neighbor algorithm, and gradient-ascending decision-tree model, based on the data of trawling surveys in the East China Sea and central and southern parts of the Yellow Sea from 2008 to 2009. Combined with a variance inflation factor and cross-check, a distribution model of T. japonicus was screened and constructed to analyze the influence of environmental factors on the distribution of T. japonicus in the East China Sea and central and southern parts of the Yellow Sea. The results showed that the random-forest model had the advantages of fitting effect and prediction ability among the three models. The analysis of this model showed that the water depth, bottom water temperature, and surface salinity had a great influence on the habitat distribution of T. japonicus. The relative resources of T. japonicus increased with the increase of bottom water temperature, reached the maximum at 23.8 °C, and first increased and then decreased with the increase of water depth and surface salinity, reaching the maximum when water depth is 72 m and surface salinity is 31.2%. This study also used the random-forest model to predict the spatial distribution of T. japonicus in the central and southern waters of the East China Sea and south-central Yellow Sea from 2008 to 2009, and the results showed that the predicted results were close to the actual situation. The research results can provide a reference for the exploitation and protection of T. japonicus resources in the East China Sea and the south-central Yellow Sea. Full article
(This article belongs to the Special Issue Biodiversity and Spatial Distribution of Fishes)
Show Figures

Figure 1

Figure 1
<p>Survey stations.</p>
Full article ">Figure 2
<p>Performance comparison of the three machine learning methods.</p>
Full article ">Figure 3
<p>Importance ranking of factors affecting the density distribution of <span class="html-italic">T. japonicus</span> in the study area.</p>
Full article ">Figure 4
<p>Impact of environmental factors on the relative resource of <span class="html-italic">T. japonicus.</span> From left to right and top to bottom are SBT SST, SSS, SBS, and SWD.</p>
Full article ">Figure 4 Cont.
<p>Impact of environmental factors on the relative resource of <span class="html-italic">T. japonicus.</span> From left to right and top to bottom are SBT SST, SSS, SBS, and SWD.</p>
Full article ">Figure 5
<p>Simulated habitat and actual survey site of <span class="html-italic">T. japonicus</span> in different seasons.</p>
Full article ">
28 pages, 13000 KiB  
Article
Dropsonde Data Impact on Rain Forecasts in Taiwan Under Southwesterly Flow Conditions with Observing System Simulation Experiments
by Fang-Ching Chien and Yen-Chao Chiu
Atmosphere 2024, 15(11), 1272; https://doi.org/10.3390/atmos15111272 - 24 Oct 2024
Viewed by 649
Abstract
This paper conducts an observing system simulation experiment (OSSE) to assess the impact of assimilating traditional sounding and surface data, along with dropsonde observations over the northern South China Sea (SCS) on heavy rain forecasts in Taiwan. Utilizing the hybrid ensemble transform Kalman [...] Read more.
This paper conducts an observing system simulation experiment (OSSE) to assess the impact of assimilating traditional sounding and surface data, along with dropsonde observations over the northern South China Sea (SCS) on heavy rain forecasts in Taiwan. Utilizing the hybrid ensemble transform Kalman filter (ETKF) and the three-dimensional variational (3DVAR) data assimilation (DA) system, this study focuses on an extreme precipitation event near Taiwan on 22 May 2020. The event was mainly influenced by strong southwesterly flow associated with an eastward-moving southwest vortex (SWV) from South China to the north of Taiwan. A nature run (NR) serves as the basis, generating virtual observations for radiosonde, surface, and dropsonde data. Three experiments—NODA (no DA), CTL (traditional observation DA), and T5D24 (additional dropsonde DA)—are configured for comparative analyses. The NODA experiment shows premature and weaker precipitation events across all regions compared with NR. The CTL experiment improved upon NODA’s forecasting capabilities, albeit with delayed onset but prolonged precipitation duration, particularly noticeable in southern Taiwan. The inclusion of dropsonde DA in the T5D24 experiment further enhanced precipitation forecasting, aligning more closely with NR, particularly in southern Taiwan. Investigations of DA impact reveal that assimilating traditional observations significantly enhances the SWV structure and wind fields, as well as the location of frontal systems, with improvements persisting for 40 to 65 h. However, low-level moisture field enhancements are moderate, leading to insufficient precipitation forecasts in southern Taiwan. Additional dropsonde DA over the northern SCS further refines low-level moisture and wind fields over the northern SCS, as well as the occurrence of frontal systems, extending positive impacts beyond 35 h and thus improving the rain forecast. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The surface map issued by the Central Weather Administration (CWA), (<b>b</b>) the Himawari infrared cloud imagery (°C), and (<b>c</b>) the 850 hPa wind vector (m s<sup>−1</sup>), wind speed (color shading, m s<sup>−1</sup>), and geopotential height (contour, interval: 10 gpm) from the ERA5 reanalysis at 0000 UTC 22 May 2020.</p>
Full article ">Figure 2
<p>(<b>a</b>) The 12 h accumulated rainfall (mm) from rain gauge stations in Taiwan ending at 0000 UTC 21 May to 0000 UTC 24 May 2020, with a 12 h interval from left to right. The black number at the lower right corner denotes the maximum rainfall in each panel. (<b>b</b>) As in (<b>a</b>), but for the NR. The green and red numbers in each panel of (<b>b</b>) denote RMSE (mm) and SCC, respectively, between the NR and observed rainfall computed over the entire domain (TW) of each individual panel. The red dot in the leftmost column of (<b>a</b>) denotes the location of Pingtung Airport. The green and red boxes in the leftmost column of (<b>a</b>) indicate the southern rain area (sRA) and northern rain area (nRA), respectively, used for areal mean rainfall calculations in subsequent analyses. Black dots in the rightmost column of (<b>a</b>) denotes the locations of rain gauge stations.</p>
Full article ">Figure 3
<p>(<b>a</b>) The flowchart of the NR and the experimental runs, including NODA, CTL, and T5D24. (<b>b</b>) Time frames of the NR (green) and the experimental runs (cyan). The ETKF DA was performed 9 times (8 cycles) from 0000 UTC 19 May to 0000 UTC 21 May 2020.</p>
Full article ">Figure 4
<p>Domain settings for the NR (green boxes) and the experimental runs (black boxes). Locations of the synthetic sounding and surface observations used in both CTL and T5D24 are denoted by red and purple dots, respectively. Green crosses indicate the locations of synthetic dropsonde observations used exclusively in the T5D24.</p>
Full article ">Figure 5
<p>Soundings from (<b>a</b>) observations and (<b>b</b>) NR at Pingtung Airport (120.47° E, 22.69° N; location shown in <a href="#atmosphere-15-01272-f002" class="html-fig">Figure 2</a>a at 0000 UTC 21 May 2020. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but showing time series of winds (full: 10 knots; half: 5 knots) and relative humidity (%) from 0000 UTC 21 May to 0000 UTC 23 May 2020. There were no sounding observations at 1200 UTC 21 May 2020.</p>
Full article ">Figure 6
<p>Time series of areal mean rain intensity (mm h<sup>−1</sup>) averaged over (<b>a</b>) TW, (<b>b</b>) sRA, and (<b>c</b>) nRA from NR. The TW, sRA and nRA regions correspond to the areas shown in the leftmost column of <a href="#atmosphere-15-01272-f002" class="html-fig">Figure 2</a>a. The abscissa shows times from 0000 UTC 21 May to 0000 UTC 24 May 2020. (<b>d</b>–<b>f</b>) As in (<b>a</b>–<b>c</b>), but for NODA. The solid line, shaded area, and dots denote the mean, the one standard deviation range, and the extreme value of the ensemble, respectively. (<b>g</b>–<b>i</b>) As in (<b>a</b>–<b>c</b>), but for CTL. (<b>j</b>–<b>l</b>) As in (<b>a</b>–<b>c</b>), but for T5D24.</p>
Full article ">Figure 7
<p>(<b>a</b>) Accumulated rainfall (mm) during P1 (0000 UTC to 1500 UTC 21 May 2020), P2 (1500 UTC 21 May to 1500 UTC 22 May 2020), and P3 (1500 UTC 22 May to 0000 UTC 24 May 2020), from left to right. The black number in the lower right corner denotes the maximum rainfall in each panel. Green/red box in the leftmost column denotes the sRA/nRA that is the same as in the leftmost column of <a href="#atmosphere-15-01272-f002" class="html-fig">Figure 2</a>a. (<b>b</b>) As in (<b>a</b>), but for NODA. (<b>c</b>) As in (<b>a</b>), but for CTL. (<b>d</b>) As in (<b>a</b>), but for T5D24.</p>
Full article ">Figure 8
<p>Threat score (TS) of the ensemble mean rain forecasts in TW (the entire domain of the leftmost column in <a href="#atmosphere-15-01272-f007" class="html-fig">Figure 7</a>a) for NODA, CTL, and T5D24 verified against NR during (<b>a</b>) P1 (0000 UTC to 1500 UTC 21 May 2020) and (<b>b</b>) P2 (1500 UTC 21 May to 1500 UTC 22 May 2020). (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for sRA (green box in <a href="#atmosphere-15-01272-f007" class="html-fig">Figure 7</a>a). (<b>e</b>,<b>f</b>) As in (<b>a</b>,<b>b</b>), but for nRA (red box in <a href="#atmosphere-15-01272-f007" class="html-fig">Figure 7</a>a).</p>
Full article ">Figure 9
<p>Performance diagrams summarizing the success ratio (SR, x-axis), probability of detection (POD, y-axis), bias, and TS for rain thresholds of (<b>a</b>) 5 mm, (<b>b</b>) 20 mm, and (<b>c</b>) 50 mm during P1 (0000 UTC to 1500 UTC 21 May 2020). Dashed lines represent bias scores (BS) with labels on the outward extension of the line, while labeled long-dashed contours are TS. Green, red, and blue crosses represent the interquartile ranges of SR and POD—with centers denoting the median—for the scores of the 32 ensemble members of NODA, CTL, and T5D24, respectively. The letters next to the crosses indicate the domains of rain verification, including TW (tw), sRA (s), and nRA (n), shown in the leftmost column of <a href="#atmosphere-15-01272-f007" class="html-fig">Figure 7</a>a. (<b>d</b>–<b>f</b>) As in (<b>a</b>–<b>c</b>), but for P2 (1500 UTC 21 May to 1500 UTC 22 May 2020) with rain thresholds of 90 mm, 180 mm, and 300 mm.</p>
Full article ">Figure 10
<p>The 850 hPa mixing ratio (color, g kg<sup>−1</sup>), 850 hPa geopotential height (contour, interval: 15 gpm), and 850 hPa winds (vector, m s<sup>−1</sup>) at 0000 UTC 21 May 2020 from (<b>a</b>) NR, and the ensemble mean of (<b>b</b>) NODA, (<b>c</b>) CTL, and (<b>d</b>) T5D24. Gray denotes the area below terrain height. Red and green boxes in (<b>a</b>) denote the South China (SC) and northern SCS (nSCS) regions for areal mean calculation, respectively. Red boxes in (<b>c</b>,<b>d</b>) show the locations of a time–height section and the Taiwan Strait (TWS) regions, respectively.</p>
Full article ">Figure 11
<p>Box plots of RMSEs of (<b>a</b>) mixing ratio (Q, g kg<sup>−1</sup>), (<b>b</b>) the east–west wind component (U, m s<sup>−1</sup>), and (<b>c</b>) the north–south wind component (V, m s<sup>−1</sup>) averaged over South China (SC; red box in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>a) at 925, 850, 700, 500, and 200 hPa for the 32 ensemble members of NODA (gray), CTL (red), and T5D24 (blue) at 0000 UTC 21 May 2020, using NR as the truth. The box extends from the first quartile to the third quartile (interquartile range), with a line denoting the median value. The right/left error bars show the data value that is 1.5 × interquartile range above/below the third/first quartile. Dots are outliers. (<b>d</b>–<b>f</b>) As in (<b>a</b>–<b>c</b>), but for RMSEs averaged over the northern SCS (nSCS; green box in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>a).</p>
Full article ">Figure 12
<p>The low-level-averaged (surface–700 hPa) mixing ratio (color shading, g kg<sup>−1</sup>), low-level-averaged (surface–700 hPa) wind (vector, m s<sup>−1</sup>), and 850 hPa geopotential height (contour, interval: 15 gpm) at 1500 UTC 21 May 2020 from (<b>a</b>) NR, and the ensemble mean of (<b>b</b>) NODA, (<b>c</b>) CTL, and (<b>d</b>) T5D24.</p>
Full article ">Figure 13
<p>Time series of the zonally averaged (119.8–120.2° E) 850 hPa geopotential height (color, gpm) and horizontal wind (vector, m s<sup>−1</sup>) along the 120° E longitude from (<b>a</b>) NR, and the ensemble mean of (<b>b</b>) NODA, (<b>c</b>) CTL, and (<b>d</b>) T5D24. (<b>e</b>–<b>g</b>) As in (<b>b</b>–<b>d</b>), but for the difference between the ensemble mean of the corresponding experiment and NR. The abscissa is time from 0000 UTC 21 May to 0000 UTC 23 May 2020, and the ordinate is latitude from 21.5 to 26.5° N (location shown in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>c). Scales are shown on the right of each panel.</p>
Full article ">Figure 14
<p>As in <a href="#atmosphere-15-01272-f013" class="html-fig">Figure 13</a> but for the 850 hPa mixing ratio (color, g kg<sup>−1</sup>) and moisture flux (vector, m s<sup>−1</sup> g kg<sup>−1</sup>).</p>
Full article ">Figure 15
<p>Skill scores (SS<sub>C</sub>; color shading) based on the member mean RMSEs of low-level-averaged (surface to 700 hPa) (<b>a</b>) x-component moisture flux (qu, g kg<sup>−1</sup> m s<sup>−1</sup>) and (<b>b</b>) y-component moisture flux (qv, g kg<sup>−1</sup> m s<sup>−1</sup>) averaged from 0800 UTC 21 May to 2200 UTC 21 May 2020 (centered at 1500 UTC ± 7 h). The green hatched regions indicate that the scores are statistically significant at a 95% confidence level. (<b>c</b>,<b>d</b>) As in (<b>a</b>,<b>b</b>), but for SS<sub>T</sub>. Red/green box in (<b>c</b>) denotes the SC/nSCS that is the same as in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>a.</p>
Full article ">Figure 16
<p>Time series of the median SS (solid line) at each full hour of the simulation for (<b>a</b>,<b>b</b>) SC region (red box in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>a), (<b>c</b>,<b>d</b>) nSCS region (green box in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>a), and (<b>e</b>,<b>f</b>) TWS region (red box in <a href="#atmosphere-15-01272-f010" class="html-fig">Figure 10</a>d). Left column (<b>a</b>,<b>c</b>,<b>e</b>) shows x-component moisture flux (qu, g kg<sup>−1</sup> m s<sup>−1</sup>), and right column (<b>b</b>,<b>d</b>,<b>f</b>) shows y-component moisture flux (qv, g kg<sup>−1</sup> m s<sup>−1</sup>). Dots denote the scores exceeding the 95% confidence level. The abscissa shows simulation time from 0 to 72 h.</p>
Full article ">Figure 17
<p>The average frontal latitude (<span class="html-italic">Lat<sub>F</sub></span>, ordinate) at each hour from 0000 UTC to 1500 UTC 22 May 2020 (abscissa) in NR (thick black line). The gray, red, and blue boxplots show the <span class="html-italic">Lat<sub>F</sub></span> of the ensemble for NODA, CTL, and T5D24, respectively. The box extends from the first quartile to the third quartile (interquartile range), with a horizontal line denoting the median value. The upper and lower error bars show the data value that is 1.5 × interquartile range above and below the third and first quartile, respectively. Dots are outliers.</p>
Full article ">
18 pages, 11141 KiB  
Article
Inter-Model Spread in Representing the Impacts of ENSO on the South China Spring Rainfall in CMIP6 Models
by Xin Yin, Xiaofei Wu, Hailin Niu, Kaiqing Yang and Linglong Yu
Atmosphere 2024, 15(10), 1199; https://doi.org/10.3390/atmos15101199 - 8 Oct 2024
Viewed by 871
Abstract
A major challenge for climate system models in simulating the impacts of El Niño–Southern Oscillation (ENSO) on the interannual variations of East Asian rainfall anomalies is the wide inter-model spread of outputs, which causes considerable uncertainty in physical mechanism understanding and short-term climate [...] Read more.
A major challenge for climate system models in simulating the impacts of El Niño–Southern Oscillation (ENSO) on the interannual variations of East Asian rainfall anomalies is the wide inter-model spread of outputs, which causes considerable uncertainty in physical mechanism understanding and short-term climate prediction. This study investigates the fidelity of 40 models from Phase 6 of the Coupled Model Intercomparison Project (CMIP6) in representing the impacts of ENSO on South China Spring Rainfall (SCSR) during the ENSO decaying spring. The response of SCSR to ENSO, as well as the sea surface temperature anomalies (SSTAs) over the tropical Indian Ocean (TIO), is quite different among the models; some models even simulate opposite SCSR anomalies compared to the observations. However, the models capturing the ENSO-related warm SSTAs over TIO tend to simulate a better SCSR-ENSO relationship, which is much closer to observation. Therefore, models are grouped based on the simulated TIO SSTAs to explore the modulating processes of the TIO SSTAs in ENSO affecting SCSR anomalies. Comparing analysis suggests that the warm TIO SSTA can force the equatorial north–south antisymmetric circulation in the lower troposphere, which is conducive to the westward extension and maintenance of the western North Pacific anticyclone (WNPAC). In addition, the TIO SSTA enhances the upper tropospheric East Asian subtropical westerly jet, leading to anomalous divergence over South China. Thus, the westward extension and strengthening of WNPAC can transport sufficient water vapor for South China, which is associated with the ascending motion caused by the upper tropospheric divergence, leading to the abnormal SCSR. Full article
(This article belongs to the Special Issue Precipitation Observations and Prediction (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Climatological distribution of the MAM (March–May) precipitation (shaded, mm day<sup>−1</sup>) and water vapor flux (vector, kg m<sup>−1</sup> s<sup>−1</sup>) over Eastern China from 1979 to 2014 for observations for the MME and individual models.</p>
Full article ">Figure 2
<p>Regression map of the MAM precipitation anomalies (shading, mm day<sup>−1</sup>) onto the standardized preceding DJF Niño3.4 index for observations, the MME, and individual models. The stippling denotes statistical significance at the 95% confidence level. The red box indicates the region used to define the SCSR.</p>
Full article ">Figure 3
<p>Scatter diagrams of the ENSO-SCSR correlation coefficients (Y−axis) and the interannual standard deviations of the DJF Niño3.4 index (X−axis, °C). Each dot represents the corresponding value for the model identified by the number (<a href="#atmosphere-15-01199-t001" class="html-table">Table 1</a>); “O” and “M” represent observation and MME.</p>
Full article ">Figure 4
<p>Regression map of MAM SSTAs (shading, °C) onto the standardized preceding DJF Niño3.4 index in observations, the MME, and individual models. The stippling denotes statistical significance at the 95% confidence level.</p>
Full article ">Figure 5
<p>As in <a href="#atmosphere-15-01199-f004" class="html-fig">Figure 4</a>, but for the MAM SSTAs regressed onto the standardized SCSR index. The stippling denotes statistical significance at the 95% confidence level.</p>
Full article ">Figure 6
<p>Scatter diagrams of the TIOI standard variations (X−axis) and (<b>a</b>) DJF Niño3.4 standard variations (Y−axis), (<b>b</b>) SCSR standard variations (Y−axis), and (<b>c</b>) ENSO-SCSR correlations (Y−axis) in the CMIP6 models. Each dot represents the corresponding value for the model identified by the number (<a href="#atmosphere-15-01199-t001" class="html-table">Table 1</a>); “O” and “M” represent observation and MME.</p>
Full article ">Figure 7
<p>Regression map of MAM precipitation (shading, mm day<sup>−1</sup>) onto the standardized TIOI in observations, the MME, and individual models. The stippling denotes statistical significance at the 95% confidence level. The red box indicates the region used to define the SCSR.</p>
Full article ">Figure 8
<p>Scatter diagrams of the TIO-SCSR correlation coefficients (Y−axis) and ENSO-SCSR correlation coefficients (X−axis). The color of each point represents the TIOI-ENSO correlations. “O” and “M” represent the observation and MME.</p>
Full article ">Figure 9
<p>Regression map of MAM precipitation anomalies (shading, unit: mm day<sup>−1</sup>) onto the standardized DJF Niño3.4 index for (<b>a</b>) observation, (<b>b</b>) “ENSO-TIO” group, (<b>c</b>) “ENSO-only” group, and (<b>d</b>) “TIO-only” group. The stippling denotes statistical significance at the 95% confidence level.</p>
Full article ">Figure 10
<p>Regression map of MAM precipitation anomalies (shading, unit: °C) onto the SSTAs for (<b>a</b>) observation, (<b>b</b>) “ENSO-TIO” group, (<b>c</b>) “ENSO-only” group, and (<b>d</b>) “TIO-only” group. The stippling denotes statistical significance at the 95% confidence level.</p>
Full article ">Figure 11
<p>Regression map of MAM 850 hPa (left column) and 200 hPa (right column) wind anomalies (vectors, unit: mm s<sup>−1</sup>) onto the standardized DJF Niño3.4 index for (<b>a</b>,<b>e</b>) observation, (<b>b</b>,<b>f</b>) “ENSO-TIO” group, (<b>c</b>,<b>g</b>) “ENSO-only” group, and (<b>d</b>,<b>h</b>) “TIO-only” group. The red arrow indicates that at least one component of the wind vector passes the 95% significance test. The black arrow indicates that no wind vector passes the 95% significance test.</p>
Full article ">Figure 12
<p>Regression map of vertical integral moisture flux (vector, kg m<sup>−1</sup> s<sup>−1</sup>) and moisture flux divergence (shading, 10<sup>−5</sup> kg m<sup>−2</sup> s<sup>−1</sup>) onto the standardized DJF Niño3.4 index for (<b>a</b>) observation, (<b>b</b>) “ENSO-TIO” group, (<b>c</b>) “ENSO-only” group, and (<b>d</b>) “TIO-only” group. The vectors indicate that at least one component of the regressed water vapor flux passes the 95% significance test. The stippling denotes statistical significance at the 95% confidence level.</p>
Full article ">
Back to TopTop