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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (293)

Search Parameters:
Keywords = tropical pacific ocean

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 6823 KiB  
Technical Note
Investigating Tropical Cyclone Warm Core and Boundary Layer Structures with Constellation Observing System for Meteorology, Ionosphere, and Climate 2 Radio Occultation Data
by Xiaoxu Qi, Shengpeng Yang and Li He
Remote Sens. 2024, 16(22), 4257; https://doi.org/10.3390/rs16224257 - 15 Nov 2024
Viewed by 319
Abstract
The Constellation Observing System for Meteorology, Ionosphere, and Climate 2 (COSMIC-2) collects data covering latitudes primarily between 40 degrees north and south, providing abundant data for tropical cyclone (TC) research. The radio occultation data provide valuable information on the boundary layer. However, quality [...] Read more.
The Constellation Observing System for Meteorology, Ionosphere, and Climate 2 (COSMIC-2) collects data covering latitudes primarily between 40 degrees north and south, providing abundant data for tropical cyclone (TC) research. The radio occultation data provide valuable information on the boundary layer. However, quality control of the data within the boundary layer remains a challenging issue. The aim of this study is to obtain a more accurate COSMIC-2 radio occultation (RO) dataset through quality control (QC) and use this dataset to validate warm core structures and explore the planetary boundary layer (PBL) structures of TCs. In this study, COSMIC-2 data are used to analyze the distribution of the relative local spectral width (LSW) and the confidence parameter characterizing the random error of the bending angle. An LSW less than 20% is set as a data QC threshold, and the warm core and PBL composite structures of TCs at three intensities in the Northwest Pacific Ocean are investigated. We reproduce the warm core intensity and warm core height characteristics of TCs. In the radial direction of the typhoon eyewall, the impact height of the PBL increases from 3.45 km to 4 km, with the tropopause ranging from 160 hPa to 100 hPa. At the bottom of the troposphere, the variations in the positive and negative bias between the RO-detected and background field bending angles correspond well to the PBL heights, and the variations in the positive bias between the RO-detected and background field refractivity reach 14%. This research provides an effective QC method and reveals that the bending angle is sensitive to the PBL height. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure diagram of this study. This flowchart outlines the process from initial data processing to final outcome.</p>
Full article ">Figure 2
<p>Spatial distribution of the COSMIC-2 (<b>a</b>) local spectral width (10<sup>−3</sup> rad) and (<b>b</b>) confidence parameter (%) over the western Pacific Ocean from 2019 to 2022. The data include the TC intensities of TDs, TSs and TYs. The horizontal coordinate represents the distance from the RO profile to the center of the TC, and the vertical coordinate represents the impact height. The black dashed line in (<b>a</b>) shows the RO-observed lowest impact height.</p>
Full article ">Figure 3
<p>Spatial distribution of the COSMIC-2 relative local spectral width (%) over the western Pacific Ocean from 2019 to 2022. The data include the TC intensities of TDs, TSs, and TYs.</p>
Full article ">Figure 4
<p>Variation in the bending angle between the RO and ECMWF analyses when the LSW is &gt;35% (green curve), between 20% and 35% (blue curve), and &lt;20% (black curve).</p>
Full article ">Figure 5
<p>Distribution of LSWs after quality control. (<b>a</b>) Spatial distribution of LSWs (%) after quality control. (<b>b</b>) RO profile counts before quality control (shaded) and the proportion of counts after quality control (black curves).</p>
Full article ">Figure 6
<p>Mean (<b>a</b>) and standard deviation (<b>b</b>) of the variation in the bending angle between the RO values and ECMWF analyses before and after quality control.</p>
Full article ">Figure 7
<p>COSMIC-2 RO profile counts for the TD (red), TS (green), and TY (blue) categories derived from collocated RO data.</p>
Full article ">Figure 8
<p>Composite temperature (<math display="inline"><semantics> <mrow> <mi>T</mi> <mo>−</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math>) cross-sections (shaded, °C), tropopause heights (black dashed lines), and boundary layer heights (black solid lines) for the TD (<b>a</b>), TS (<b>b</b>), and TY (<b>c</b>) categories.</p>
Full article ">Figure 9
<p>Composite spatial distribution of the bending angle between RO values and ECMWF analyses before (<b>a</b>–<b>c</b>) and after (<b>d</b>–<b>f</b>) quality control for the TD (<b>a</b>,<b>d</b>), TS (<b>b</b>,<b>e</b>), and TY (<b>c</b>,<b>f</b>) categories. The black lines represent the boundary layer heights.</p>
Full article ">Figure 10
<p>Composite spatial distributions of refractivity between RO values and ECMWF analyses before (<b>a</b>–<b>c</b>) and after (<b>d</b>–<b>f</b>) quality control for the TD (<b>a</b>,<b>d</b>), TS (<b>b</b>,<b>e</b>), and TY (<b>c</b>,<b>f</b>) categories. The black lines represent the boundary layer heights.</p>
Full article ">
27 pages, 12462 KiB  
Article
Long-Term Teleconnections Between Global Circulation Patterns and Interannual Variability of Surface Air Temperature over Kingdom of Saudi Arabia
by Abdullkarim K. Almaashi, Hosny M. Hasanean and Abdulhaleem H. Labban
Atmosphere 2024, 15(11), 1310; https://doi.org/10.3390/atmos15111310 - 30 Oct 2024
Viewed by 388
Abstract
Surface air temperature (SAT) variability is investigated for advancing our understanding of the climate patterns over the Kingdom of Saudi Arabia (KSA). SAT variability reveals significant warming trends, particularly from 1994 onward, as demonstrated by nonlinear and linear trend analysis. This warming is [...] Read more.
Surface air temperature (SAT) variability is investigated for advancing our understanding of the climate patterns over the Kingdom of Saudi Arabia (KSA). SAT variability reveals significant warming trends, particularly from 1994 onward, as demonstrated by nonlinear and linear trend analysis. This warming is linked to global climate patterns, which serve as significant indicators for studying the effects of climate change on surface air temperature patterns across the KSA. The empirical orthogonal function (EOF) method is employed for analyzing SAT due to its effectiveness in extracting dominant patterns of variability during the winter (DJF) and summer (JJA) seasons. The first mode (EOF1) for both seasons shows positive variability across the KSA, explaining more than 45% of the variance. The second mode (EOF2) indicates negative variability in central and northern regions. The third mode (EOF3) describes positive variability but with lower variance over time. PC1 is used to describe the physical mechanism of SAT variability and correlations with global sea surface temperature (SST). The physical mechanism shows that the variability in Mediterranean troughs during the winter season and high pressure over the Indian Ocean and central Asia controls SAT variability over the KSA. The correlation coefficients (CCs) were calculated during the winter and summer season between the SAT of the KSA and six teleconnection indices, El Niño Southern Oscillation (ENSO), Indian Ocean Dipole (IOD), Atlantic Meridional Mode (AMM), Pacific Warm Pool (PWP), North Atlantic Oscillation (NAO), and Tropical North Atlantic (TNA) SST for the period from 1994 to 2022. ENSO shifts from positive to negative correlations with SAT from winter to summer. IOD shows a diminished correlation with SAT due to the absence of upper air dynamics. PWP consistently enhances surface warming in both seasons through upper air convergence during both seasons. AMM and NAO have a non-significant impact on SAT; however, TNA contributes warming over central and northern parts during winter and summer seasons. The seasonal SAT variations emphasize the significant role of ENSO, PWP, and TNA across the seasons. The findings of this study can be helpful for seasonal predictability in the KSA. Full article
(This article belongs to the Section Climatology)
Show Figures

Figure 1

Figure 1
<p>The elevation in meters, along with solid circles representing the observation station names in Saudi Arabia. (Data source: USGS satellite topographic dataset).</p>
Full article ">Figure 2
<p>Analysis of the changes in ERA5 SAT over the KSA for (<b>a</b>) the summer season (June to August), (<b>b</b>) the winter season (December–January and February). The red lines represent the mean SAT scores for the times 1952–1993 and 1994–2022. Global analysis of the changes in ERA5 SAT of (<b>c</b>) the summer season (December–January and February) and (<b>d</b>) winter season (June to August).</p>
Full article ">Figure 3
<p>(<b>a</b>) Summer season trend magnitude (Sen’s slope °C/year) during 1952–1993, (<b>b</b>) same as (<b>a</b>) but for the period 1994–2022. (<b>c</b>) Winter season trend magnitude (Sen’s slope °C/year) during 1952–1993, (<b>d</b>) same as (<b>c</b>) but for the period 1994–2022. Dotted areas show the significance above 99% confidence level.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>): The Empirical Orthogonal Function (EOF) analysis of SAT temperature for Jun–Aug season during 1952–1993. (<b>d</b>–<b>f</b>) show the corresponding PCs of EOF analyses.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) The Empirical Orthogonal Function (EOF) analysis of SAT temperature for Jun–Aug season during 1994–2022. (<b>d</b>–<b>f</b>) show the corresponding PCs of EOFs.</p>
Full article ">Figure 6
<p>(<b>a</b>) The leading EOF mode represents the primary pattern, capturing 64.70% of the total variance. (<b>b</b>) The second EOF mode follows, explaining 14.51% of the variance. (<b>c</b>) The third EOF mode contributes to 5.75% of the total variance. (<b>d</b>–<b>f</b>) The associated principal component time series (PC1, PC2, and PC3) correspond to these leading EOF modes.</p>
Full article ">Figure 7
<p>(<b>a</b>) The leading EOF mode represents the primary pattern, capturing 54.26% of the total variance. (<b>b</b>) The second EOF mode follows, explaining 19.07% of the variance. (<b>c</b>) The third EOF mode contributes to 7.22% of the total variance. (<b>d</b>–<b>f</b>) The associated principal component time series (PC1, PC2, and PC3) correspond to these leading EOF modes.</p>
Full article ">Figure 8
<p>Correlation between PC1 of SAT and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during winter (<b>a</b>) for 1952–1993, (<b>b</b>) for 1994–2022. (<b>c,d</b>) Same as (<b>a</b>,<b>b</b>) but for summer season. The dotted areas show the regions with significance level above 95% by using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 9
<p>(<b>a</b>) Correlation between SAT and ENSO, computed from standardized anomalies of winter seasonal time series spanning 1952–2022. Dotted regions indicate significance levels exceeding 95% confidence, determined using Student’s <span class="html-italic">t</span>-test. (<b>b</b>–<b>f</b>) Same as (<b>a</b>) but for the IOD, PWP, AMM, NAO, and TNA.</p>
Full article ">Figure 10
<p>(<b>a</b>) Correlation between SAT and ENSO, computed from standardized anomalies of summer seasonal time series spanning 1952–2022. Dotted regions indicate significance levels exceeding 95% confidence, determined using Student’s <span class="html-italic">t</span>-test. (<b>b</b>–<b>f</b>) same as (<b>a</b>) but for the IOD, PWP, AMM, NAO, and TNA.</p>
Full article ">Figure 11
<p>Correlation between climate indices and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during winter from 1994 to 2022. (<b>a</b>) Nino3.4, (<b>b</b>) IOD, (<b>c</b>) PWP, (<b>d</b>) AMM, (<b>e</b>) NAO, and (<b>f</b>) TNA. The dotted areas show the regions with significance level above 95% by using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 12
<p>Correlation between climate indices and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during summer season from 1994 to 2022. (<b>a</b>) Nino3.4, (<b>b</b>) IOD, (<b>c</b>) PWP, (<b>d</b>) AMM, (<b>e</b>) NAO, and (<b>f</b>) TNA. The dotted areas show the regions with significance level above 95% using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 13
<p>The correlation coefficient map between PC1 and global SST for DJF season during (<b>a</b>) 1952–1993, (<b>b</b>) during the period 1994–2022. Stippling denotes regions where the relationship is statistically significant at 95% confidence level based on Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 14
<p>The correlation coefficient maps between PC1 and global SST for JJA season during (<b>a</b>) 1952–1993, (<b>b</b>) during the period 1994–2022. Stippling denotes regions where the relationship is statistically significant at 95% confidence level based on Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
21 pages, 19883 KiB  
Article
Larval Transport Pathways for Lutjanus peru and Lutjanus argentiventris in the Northwestern Mexico and Tropical Eastern Pacific
by Nicole Reguera-Rouzaud, Guillermo Martínez-Flores, Noé Díaz-Viloria and Adrián Munguía-Vega
Water 2024, 16(21), 3084; https://doi.org/10.3390/w16213084 - 28 Oct 2024
Viewed by 548
Abstract
Understanding how ocean currents influence larval dispersal and measuring its magnitude is critical for conservation and sustainable exploitation, especially in the Tropical Eastern Pacific (TEP), where the larval transport of rocky reef fish remains untested. For this reason, a lagrangian simulation model was [...] Read more.
Understanding how ocean currents influence larval dispersal and measuring its magnitude is critical for conservation and sustainable exploitation, especially in the Tropical Eastern Pacific (TEP), where the larval transport of rocky reef fish remains untested. For this reason, a lagrangian simulation model was implemented to estimate larval transport pathways in Northwestern Mexico and TEP. Particle trajectories were simulated with data from the Hybrid Ocean Coordinate Model, focusing on three simulation scenarios: (1) using the occurrence records of Lutjanus peru and L. argentiventris as release sites; (2) considering a continuous distribution along the study area, and (3) taking the reproduction seasonality into account in both species. It was found that the continuous distribution scenario largely explained the genetic structure previously found in both species (genetic brakes between central and southern Mexico and Central America), confirming that the ocean currents play a significant role as predictors of genetic differentiation and gene flow in Northwestern Mexico and the TEP. Due to the oceanography of the area, the southern localities supply larvae from the northern localities; therefore, disturbances in any southern localities could affect the surrounding areas and have impacts that spread beyond their political boundaries. Full article
(This article belongs to the Special Issue Aquatic Environment and Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Study area (<b>a</b>). Northwestern Mexico (<b>b</b>), Mexican Tropical Pacific (<b>c</b>), and Central America and Colombia (<b>d</b>). The blue polygons are the counting areas for the connectivity networks (the numbers represent the polygons’ order). The continuous and dashed red lines are the sites where the genetic brakes were found for <span class="html-italic">L. peru</span> and <span class="html-italic">L. argentiventris</span>, respectively [<a href="#B36-water-16-03084" class="html-bibr">36</a>]. Baja California Sur (BCS), Nayarit (NAY), Colima (CMA), Oaxaca (OAX), Panama (PAN), and Colombia (COL).</p>
Full article ">Figure 2
<p>Larval seeding points for <span class="html-italic">Lutjanus peru</span> (<b>a</b>), <span class="html-italic">Lutjanus argentiventris</span> (<b>b</b>), and centroids (<b>c</b>).</p>
Full article ">Figure 3
<p>Direction and distance traveled by the simulated trajectories obtained with the HYCOM for the zone A (<b>a</b>–<b>d</b>), zone B (<b>e</b>–<b>h</b>), and zone C (<b>i</b>–<b>l</b>). Spring (March (M), April (A), May (M)); Summer (June (J), July (J), August (A)); Autumn (September (S), October (O), November (N)); Winter (December (D), January (J), February (F)).</p>
Full article ">Figure 4
<p>Mean seasonal circulation pattern obtained with the HYCOM for zone A during (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn from 2017. The color bar represents the velocity of the mean seasonal circulation in m/s, and the arrows indicate the direction.</p>
Full article ">Figure 5
<p>Mean seasonal circulation pattern obtained with the HYCOM for zone B during (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn from 2017. The color bar represents the velocity of the mean seasonal circulation in m/s, and the arrows indicate the direction.</p>
Full article ">Figure 6
<p>Mean seasonal circulation pattern obtained with the HYCOM for zone C during (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn from 2017. The color bar represents the velocity of the mean seasonal circulation in m/s, and the arrows indicate the direction.</p>
Full article ">Figure 7
<p>Connectivity networks for the continuous distribution scenario (for both species) during winter (<b>a</b>–<b>c</b>) and spring (<b>d</b>–<b>f</b>) for 15 and 30 days of larval dispersal (PLD), respectively. The red dots represent the centroids, the colored lines represent the connectivity networks, and the tick represents the percentage of connectivity between polygons. Baja California (BC), Baja California Sur (BCS), Sonora (SON), Sinaloa (SIN), Nayarit (NAY), Colima (CMA), Guerrero (GUE), Oaxaca (OAX), Panama (PAN), and Colombia (COL).</p>
Full article ">Figure 8
<p>Connectivity networks for the occurrence records of <span class="html-italic">Lutjanus peru</span> during winter for 15 (<b>a</b>–<b>c</b>) and 30 (<b>d</b>–<b>f</b>) days of larval dispersal (PLD), respectively. The red dots represent the centroids, the colored lines represent the connectivity networks, and the tick represents the percentage of connectivity between polygons. Baja California (BC), Baja California Sur (BCS), Sonora (SON), Sinaloa (SIN), Nayarit (NAY), Colima (CMA), Guerrero (GUE), Oaxaca (OAX), Panama (PAN), and Colombia (COL).</p>
Full article ">Figure 9
<p>Connectivity networks for the occurrence records of <span class="html-italic">Lutjanus argentiventris</span> during summer (<b>a</b>–<b>c</b>) and autumn (<b>d</b>–<b>f</b>) for 15 and 30 days of larval dispersal (PLD), respectively. The red dots represent the centroids, the colored lines represent the connectivity networks, and the tick represents the percentage of connectivity between polygons. Baja California (BC), Baja California Sur (BCS), Sonora (SON), Sinaloa (SIN), Nayarit (NAY), Colima (CMA), Guerrero (GUE), Oaxaca (OAX), Panama (PAN), and Colombia (COL).</p>
Full article ">Figure 10
<p>Reproductive season scenario at 15 days of pelagic larval duration (PLD) for <span class="html-italic">Lutjanus peru</span> (<b>a</b>–<b>c</b>) and <span class="html-italic">L. argentiventris</span> (<b>d</b>–<b>f</b>). The red dots represent the centroids, the colored lines represent the connectivity networks, and the tick represents the percentage of connectivity between polygons. Baja California (BC), Baja California Sur (BCS), Sonora (SON), Sinaloa (SIN), Nayarit (NAY), Colima (CMA), Guerrero (GUE), Oaxaca (OAX), Panama (PAN), and Colombia (COL).</p>
Full article ">
5 pages, 2215 KiB  
Interesting Images
The Box Crab Calappa hepatica as a Nuclear Species for the Opportunistic Foraging Behaviour of the Flowery Flounder, Bothus mancus, in the Indo-Pacific
by Federico Betti and Bert W. Hoeksema
Diversity 2024, 16(11), 662; https://doi.org/10.3390/d16110662 - 28 Oct 2024
Viewed by 415
Abstract
Some predatory fishes may exhibit opportunistic feeding behaviour by exploiting potential prey that is distracted, displaced, or exposed by the activities of a third party that acts as a ‘nuclear’ species. Other fishes mostly perform the role of ‘nuclear’ species, but benthic invertebrates, [...] Read more.
Some predatory fishes may exhibit opportunistic feeding behaviour by exploiting potential prey that is distracted, displaced, or exposed by the activities of a third party that acts as a ‘nuclear’ species. Other fishes mostly perform the role of ‘nuclear’ species, but benthic invertebrates, such as octopuses, have also been reported. Crabs are rarely observed in this role, with only a few records from the tropical Atlantic Ocean. Here, we report the temporary association between two specimens of the flowery flounder, Bothus mancus (family Bothidae), and a box crab, Calappa hepatica (family Calappidae), from the Philippines, representing the first record of a crab–fish feeding association in the Indo-Pacific region. Full article
(This article belongs to the Collection Interesting Images from the Sea)
Show Figures

Figure 1

Figure 1
<p>Frontal view of the crab <span class="html-italic">Calappa hepatica</span> partially buried in the sand, with two <span class="html-italic">Bothus mancus</span> individuals waiting at short distances to opportunistically exploit potential prey disturbed by the crab’s movements and burying activities. Photo credit: F.B.</p>
Full article ">Figure 2
<p>(<b>a</b>) Map of the Philippines; (<b>b</b>) detail of the investigated area (white square and arrow indicate the ‘Coconut’ dive spot). Maps from Google Earth Pro 7.3.6.9796 (2024).</p>
Full article ">
23 pages, 23516 KiB  
Article
Distribution and Seasonality of the Omura’s Whale (Balaenoptera omurai) in Australia Based on Passive Acoustic Recordings
by Ciara Edan Browne, Christine Erbe and Robert D. McCauley
Animals 2024, 14(20), 2944; https://doi.org/10.3390/ani14202944 - 12 Oct 2024
Viewed by 1346
Abstract
The Omura’s whale (Balaenoptera omurai) is one of the most recently described species of baleen whale. Initially known only from stranding and whaling specimens, it has now been identified in all ocean basins excluding the central and eastern Pacific. Unlike most [...] Read more.
The Omura’s whale (Balaenoptera omurai) is one of the most recently described species of baleen whale. Initially known only from stranding and whaling specimens, it has now been identified in all ocean basins excluding the central and eastern Pacific. Unlike most baleen whales that migrate between the poles and the equator seasonally, the Omura’s whale is known to inhabit tropical to sub-tropical waters year-round. In Australian waters, there remain fewer than 30 confirmed visual sightings over the past decade. However, based on acoustic records, the Omura’s whale has been detected off areas of the northwest coast of Australia year-round. This study utilises passive acoustic recordings from 41 locations around Australia from 2005 to 2023 to assess the distribution and seasonality of the Omura’s whale. The seasonal presence of Omura’s whale vocalisations varied by location, with higher presence at lower latitudes. Vocalisations were detected year-round in the Joseph Bonaparte Gulf in the Timor Sea, and near Browse Island and Scott Reef, in the Kimberley region. In the Pilbara region, acoustic presence mostly peaked from February to April and no acoustic presence was consistently observed from July to September across all sites. The most southerly occurrence of Omura’s whale vocalisations was recorded off the North West Cape in the Gascoyne region. Vocalisations similar but not identical to those of the Omura’s whale were detected in the Great Barrier Reef. The identified seasonal distribution provides valuable information to assess environmental and anthropogenic pressures on the Omura’s whale and to aid in creating management and conservation policies for the species in Australia. Full article
Show Figures

Figure 1

Figure 1
<p>Deployment locations of USRs.</p>
Full article ">Figure 2
<p>Example spectrogram of five Omura’s whale vocalisations detected off northwest Australia. Spectrogram produced using a 512-point Hanning window with 1.172 Hz and 0.85 s frequency and time resolution, respectively; sampling frequency 600 Hz.</p>
Full article ">Figure 3
<p>(<b>a</b>) Spectrogram of a section of a recording sample with an Omura’s whale vocalisation present at 15–30 s, and (<b>b</b>) normalised amplitude envelope of the sample (blue curve) compared to the Omura’s whale vocalisation template (red curve).</p>
Full article ">Figure 4
<p>Bar plot of percentage of monthly vocal presence per recorded hours by site and year. Faded columns represent months with no data available. Sites with vocalisations in less than 15% of recordings were not included in this figure. JBG is the combined Bonaparte Gulf NE and SW sites. Scott Reef is the combined Scott Reef N and SE sites. CANPASS is the combined 1–7 sites in chronological order. Montebello is the combined Montebello, Montebello NW and NE sites. NWS W is the combined NWS W1 and W2 sites. NWS E is the combined NWS E1 and E2 sites. Ex. Plat. is the combined Exmouth Plateau and Exmouth Plateau W sites.</p>
Full article ">Figure 5
<p>Map of seasonal distribution of Omura’s whale vocal presence in the Kimberley and Timor Sea regions. Polar histograms represent the percentage of recorded hours with Omura’s whale vocal presence per month whereas the yellow segment represents January, and months follow in a clockwise direction to December in bright green. The segments in black represent months with no acoustic data available. White segments correspond to recordings without Omura’s detections. The red slashed circles represent sites with no Omura’s whale vocal presence found in available data.</p>
Full article ">Figure 6
<p>Map of seasonal distribution of Omura’s whale vocal presence in the Pilbara and Gascoyne regions. Polar histograms represent the percentage of recorded hours with Omura’s whale vocal presence per month, while the yellow segment represents January and months following in a clockwise direction to December in bright green. The segments in black represent months with no acoustic data available. White segments correspond to recordings without Omura’s detections. The red slashed circles represent sites with no Omura’s whale vocal presence found in available data.</p>
Full article ">Figure 7
<p>Sound spectrogram and waveform of potential Omura’s whale vocalisation detected off Lizard Island, Great Barrier Reef.</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 572
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 ">
18 pages, 5148 KiB  
Article
Trends and Periodicities of Tropical Cyclone Frequencies and the Correlations with Ocean Drivers
by Guoyou Li, Huabin Shi and Zhiguo He
J. Mar. Sci. Eng. 2024, 12(10), 1707; https://doi.org/10.3390/jmse12101707 - 26 Sep 2024
Viewed by 686
Abstract
This study presents a comprehensive analysis on the variations in the tropical cyclone (TC) frequencies during 1980–2021, including the linear trends, periodicities, and their variabilities on both global and basin-wise scales. An increasing trend in the annual number of global TCs is identified, [...] Read more.
This study presents a comprehensive analysis on the variations in the tropical cyclone (TC) frequencies during 1980–2021, including the linear trends, periodicities, and their variabilities on both global and basin-wise scales. An increasing trend in the annual number of global TCs is identified, with a significant rising trend in the numbers of tropical storms (maximum sustained wind 35 ktsUmax<64 kts) and intense typhoons (Umax96 kts) and a deceasing trend for weak typhoons (64 ktsUmax<96 kts). There is no statistically significant trend shown in the global Accumulated Cyclone Energy (ACE). On a regional scale, the Western North Pacific (WNP) and Eastern North Pacific (ENP) are the regions of the first- and second-largest numbers of TCs, respectively, while the increased TC activity in the North Atlantic (NA) contributes the most to the global increase in TCs. It is revealed in the wavelet transformation for periodicity analysis that the variations in the annual number of TCs with different intensities mostly show an inter-annual period of 3–7 years and an inter-decadal one of 10–13 years. The inter-annual and inter-decadal periods are consistent with those in the ENSO-related ocean drivers (via the Niño 3.4 index), Southern Oscillation Index (SOI), and Inter-decadal Pacific Oscillation (IPO) index. The inter-decadal variation in 10–13 years is also observed in the North Atlantic Oscillation (NAO) index. The Tropical North Atlantic (TNA) index and Atlantic Multi-decadal Oscillation (AMO) index, on the other hand, present the same inter-annual period of 7–10 years as that in the frequencies of all the named TCs in the NA. Further, the correlations between TC frequencies and ocean drivers are also quantified using the Pearson correlation coefficient. These findings contribute to an enhanced understanding of TC activity, thereby facilitating efforts to predict particular TC activity and mitigate the inflicted damage. Full article
(This article belongs to the Section Physical Oceanography)
Show Figures

Figure 1

Figure 1
<p>Locations of defined ocean basins. NIO: North Indian Ocean; WNP: Western North Pacific Ocean; ENP: Eastern North Pacific Ocean; NA: North Atlantic Ocean; SIO: South Indian Ocean; SP: South Pacific Ocean.</p>
Full article ">Figure 2
<p>Time series (solid lines) and its linear trend (dashed lines) of (<b>a</b>) annual numbers of all named TCs; the annual numbers as well as proportions of (<b>b</b>) TSs, (<b>c</b>) WTYs, and (<b>d</b>) ITYs on the globe scale. In subfigures (<b>b</b>–<b>d</b>), the blue lines represent the series of annual numbers, while the red ones are for the proportions. The shaded areas represent the 95% confidence interval in the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-value for the statistical significance of the linear trend is included.</p>
Full article ">Figure 3
<p>Time series (solid line) and the linear trend (dashed line) of annual values of global ACE. The shaded area represents the 95% confidence interval of the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-value for the statistical significance of the linear trend is included.</p>
Full article ">Figure 4
<p>(<b>a</b>) Linear trend in the annual numbers of all named TCs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of all named TC numbers generated in hemispheres and six ocean basins in global total TCs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 5
<p>(<b>a</b>) Linear trend in the annual numbers of TSs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. Two asterisks above the bars indicate statistical significance levels of the linear trend of 5%; (<b>b</b>) the 42-year-averaged proportions (columns) of TS numbers generated in hemispheres and six ocean basins in global total TSs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 6
<p>(<b>a</b>) Linear trend in the annual numbers of WTYs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of WTY numbers generated in hemispheres and six ocean basins in global total WTYs and their linear trends (numbers above the columns, in the unit of %/year). One and two asterisks above the numbers indicate statistical significance levels of the linear trend of 10% and 5%, respectively.</p>
Full article ">Figure 7
<p>(<b>a</b>) Linear trend in the annual numbers of ITYs in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. One and two asterisks above the bars indicate statistical significance levels of the linear trend of 10% and 5%, respectively; (<b>b</b>) the 42-year-averaged proportions (columns) of ITY numbers generated in hemispheres and six ocean basins in global total ITYs and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 8
<p>(<b>a</b>) Linear trend in the annual values of ACE in hemispheres and six ocean basins. Red columns represent increasing trends while blue ones decreasing. Two asterisks above the bars indicate statistical significance levels of the linear trend of 5%; (<b>b</b>) the 42-year-averaged proportions (columns) of ACE in hemispheres and six ocean basins in global ACE and their linear trends (numbers above the columns, in the unit of %/year). Two asterisks above the numbers indicate statistical significance levels of the linear trend of 5%.</p>
Full article ">Figure 9
<p>Wavelet coefficient magnitude scalograms and power spectrums of annual numbers of all named TCs across the entire globe, the two hemispheres, and six ocean basins. The white dashed lines in magnitude scalograms indicate the cone of influence.</p>
Full article ">Figure 10
<p>Time series (solid lines) and its linear trend (dashed lines) of (<b>a</b>) Niño 3.4 index, (<b>b</b>) SOI, (<b>c</b>) IPO index, (<b>d</b>) IOD index, (<b>e</b>) TNA index, (<b>f</b>) NAO index, and (<b>g</b>) AMO index. The shaded areas represent the 95% confidence interval in the linear regression, and the <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math>-values for the statistical significance of the linear trend are included.</p>
Full article ">Figure 11
<p>Wavelet coefficient magnitude scalograms and power spectrums of annual values of (<b>a</b>) Niño 3.4 index, (<b>b</b>) SOI, (<b>c</b>) IPO index, (<b>d</b>) IOD index, (<b>e</b>) TNA index, (<b>f</b>) NAO index, and (<b>g</b>) AMO index. The white dashed lines in magnitude scalograms indicate the cone of influence.</p>
Full article ">
23 pages, 7991 KiB  
Article
Estimating Subsurface Thermohaline Structure in the Tropical Western Pacific Using DO-ResNet Model
by Xianmei Zhou, Shanliang Zhu, Wentao Jia and Hengkai Yao
Atmosphere 2024, 15(9), 1043; https://doi.org/10.3390/atmos15091043 - 29 Aug 2024
Viewed by 495
Abstract
Estimating the ocean’s subsurface thermohaline information from satellite measurements is essential for understanding ocean dynamics and the El Niño phenomenon. This paper proposes an improved double-output residual neural network (DO-ResNet) model to concurrently estimate the subsurface temperature (ST) and subsurface salinity (SS) in [...] Read more.
Estimating the ocean’s subsurface thermohaline information from satellite measurements is essential for understanding ocean dynamics and the El Niño phenomenon. This paper proposes an improved double-output residual neural network (DO-ResNet) model to concurrently estimate the subsurface temperature (ST) and subsurface salinity (SS) in the tropical Western Pacific using multi-source remote sensing data, including sea surface temperature (SST), sea surface salinity (SSS), sea surface height anomaly (SSHA), sea surface wind (SSW), and geographical information (including longitude and latitude). In the model experiment, Argo data were used to train and validate the model, and the root mean square error (RMSE), normalized root mean square error (NRMSE), and coefficient of determination (R2) were employed to evaluate the model’s performance. The results showed that the sea surface parameters selected in this study have a positive effect on the estimation process, and the average RMSE and R2 values for estimating ST (SS) by the proposed model are 0.34 °C (0.05 psu) and 0.91 (0.95), respectively. Under the data conditions considered in this study, DO-ResNet demonstrates superior performance relative to the extreme gradient boosting model, random forest model, and artificial neural network model. Additionally, this study evaluates the model’s accuracy by comparing its estimations of ST and SS across different depths with Argo data, demonstrating the model’s ability to effectively capture the most spatial features, and by comparing NRMSE across different depths and seasons, the model demonstrates strong adaptability to seasonal variations. In conclusion, this research introduces a novel artificial intelligence technique for estimating ST and SS in the tropical Western Pacific Ocean. Full article
Show Figures

Figure 1

Figure 1
<p>Bathymetry (unit: m) in the tropical Western Pacific, as well as locations of zonal section A (red line) and meridional section B (orange line) used in this study. The black box highlights the Niño 4 region.</p>
Full article ">Figure 2
<p>The structure of DO-ResNet model, with the solid and dashed lines representing two different types of shortcut connections (<b>a</b>) and schematic diagram of a residual block (<b>b</b>), including two convolutional layers, two activation functions, and a shortcut connection.</p>
Full article ">Figure 3
<p>Flowchart of the DO-ResNet model for estimating ocean subsurface thermohaline structure in the tropical Western Pacific.</p>
Full article ">Figure 4
<p>The estimation accuracy of the annual mean ST (<b>a</b>,<b>b</b>) and SS (<b>c</b>,<b>d</b>) at different depths by the DO-ResNet model based on RMSE and R<sup>2</sup> in different cases (Cases 1, 2, and 3 shown in red, blue, and black lines, respectively) in 2020.</p>
Full article ">Figure 5
<p>Comparison of the DO-ResNet model (black lines) with three machine learning models (XGBoost: red lines, RF: blue lines, ANN: magenta lines) for the annual mean ST and SS estimation at different depths based on RMSE (<b>a</b>,<b>c</b>) and R<sup>2</sup> (<b>b</b>,<b>d</b>) in 2020.</p>
Full article ">Figure 6
<p>The 2020 annual mean Argo data (<b>a</b>,<b>d</b>), DO-ResNet model estimated (<b>b</b>,<b>e</b>), and their differences (Argo observations minus DO-ResNet estimates; (<b>c</b>,<b>f</b>)) of ST (unit: °C) along zonal and meridional sections.</p>
Full article ">Figure 7
<p>The 2020 annual mean Argo data (<b>a</b>,<b>d</b>), DO-ResNet model estimated (<b>b</b>,<b>e</b>), and their differences (Argo observations minus DO-ResNet estimates; (<b>c</b>,<b>f</b>)) for SS (unit: psu) along zonal and meridional sections.</p>
Full article ">Figure 8
<p>Annual averaged ST (unit: °C) from Argo observation (<b>a</b>–<b>f</b>) and DO-ResNet estimation (<b>g</b>–<b>l</b>) and their differences (Argo observations minus DO-ResNet estimates, (<b>m</b>–<b>r</b>)) at different depths (50 m, 200 m, 500 m, 1000 m, 1500 m, and 1900 m) in 2020.</p>
Full article ">Figure 9
<p>Annual averaged SS (unit: psu) from Argo observation (<b>a</b>–<b>f</b>) and DO-ResNet estimation (<b>g</b>–<b>l</b>) and their differences (Argo observations minus DO-ResNet estimates, (<b>m</b>–<b>r</b>)) at different depths (50 m, 200 m, 500 m, 1000 m, 1500 m, and 1900 m) in 2020.</p>
Full article ">Figure 10
<p>T−S diagrams of the Western Pacific region for winter (February 2020) and summer (August 2020), based on Argo data (<b>a</b>,<b>c</b>) and the DO-ResNet model results (<b>b</b>,<b>d</b>). Acronyms: NPCW, North Pacific Central Water; SPCW, South Pacific Central Water; NPSTUW, North Pacific Subtropical Underwater; NPSTMW, North Pacific Subtropical Mode Water; SPSTMW, South Pacific Subtropical Mode Water; NPIW, North Pacific Intermediate Water; AAIW, Antarctic Intermediate Water.</p>
Full article ">Figure 11
<p>Seasonal performance of the DO-ResNet model for ST (<b>a</b>) and SS (<b>b</b>) estimation at different depths in the tropical Western Pacific in 2020. The data are presented with error bar representing 95% confidence interval. Blue indicates February (winter), orange indicates May (spring), green indicates August (summer), and red indicates November (autumn). The histograms display the NRMSE, while the lines display R<sup>2</sup>.</p>
Full article ">Figure 12
<p>Pearson correlation coefficients between the DO-ResNet model estimated ST (<b>a</b>) and SS (<b>b</b>) at different depths and the observational sea surface parameters.</p>
Full article ">
18 pages, 15556 KiB  
Article
Spatio-Temporal Variations of Indonesian Rainfall and Their Links to Indo-Pacific Modes
by Melly Ariska, Suhadi, Supari, Muhammad Irfan and Iskhaq Iskandar
Atmosphere 2024, 15(9), 1036; https://doi.org/10.3390/atmos15091036 - 28 Aug 2024
Cited by 1 | Viewed by 800
Abstract
The analysis of rainfall patterns in the Indonesian region utilized the Empirical Orthogonal Function (EOF) method to identify spatial and temporal variations. The study evaluated the dynamic influence of the Tropical Indian Ocean (TIO) and the Tropical Pacific Ocean (TPO) on Indonesian rainfall [...] Read more.
The analysis of rainfall patterns in the Indonesian region utilized the Empirical Orthogonal Function (EOF) method to identify spatial and temporal variations. The study evaluated the dynamic influence of the Tropical Indian Ocean (TIO) and the Tropical Pacific Ocean (TPO) on Indonesian rainfall using monthly data from the Southeast Asian Climate Assessment and Dataset (SACA&D) spanning from January 1981 to December 2016 and encompassing three extreme El Niño events in 1982/1983, 1997/1998 and 2015/2016. Using combined reanalysis and gridded-observation data, this study evaluates the potential impact of the two primary modes in the tropical Indo-Pacific region, namely the Indian Ocean Dipole (IOD) and the El Niño-Southern Oscillation (ENSO) on Indonesian rainfall. The analysis using the EOF method revealed two main modes with variances of 35.23% and 13.07%, respectively. Moreover, the results indicated that rainfall in Indonesia is highly sensitive to sea surface temperatures (SST) in the southeastern tropical Indian Ocean and the central Pacific Ocean (Niño3.4 and Niño3 areas), suggesting that changes in SST could significantly alter rainfall patterns in the region. This research is useful for informing government policies related to anticipating changes in rainfall variability as part of Indonesia’s preparedness for hydrometeorological disasters. Full article
Show Figures

Figure 1

Figure 1
<p>Topography of Indonesian region. Source: <a href="https://tanahair.indonesia.go.id" target="_blank">https://tanahair.indonesia.go.id</a> (accessed on 5 January 2023).</p>
Full article ">Figure 2
<p>Theoretical–Methodological Flowchart.</p>
Full article ">Figure 3
<p>Time series of (<b>a</b>) Niño3.4 index and (<b>b</b>) Dipole Mode Index from 1940–2020.</p>
Full article ">Figure 4
<p>Spatial patterns of eigenfunctions (<b>upper panel</b>) and time series of principal component (<b>lower panel</b>) of (<b>a</b>) EOF mode 1; (<b>b</b>) EOF mode 2.</p>
Full article ">Figure 5
<p>Climatological time series of principal components of the first mode (blue) and the second mode (red).</p>
Full article ">Figure 6
<p>Spectra of the time series of principal component of (<b>a</b>) the first mode and (<b>b</b>) the second mode.</p>
Full article ">Figure 7
<p>Spatial mean (<b>left panel</b>) and spatial standard deviation (<b>right panel</b>) of reconstructed time series of the eigen-function of (<b>a</b>,<b>c</b>) the first mode, and (<b>b</b>,<b>d</b>) the second mode.</p>
Full article ">Figure 8
<p>Climatological mean of reconstructed time series of the eigenfunction of (<b>a</b>) the first mode and (<b>b</b>) the second mode.</p>
Full article ">Figure 9
<p>Monthly average of reconstructed time series of the eigenfunction of (<b>a</b>) the first mode and (<b>b</b>) the second mode. Red bars indicate decreasing in rainfall, blue bars indicate increasing in rainfall.</p>
Full article ">Figure 10
<p>Correlation between PC1 and SST for the (<b>a</b>) ASO season and (<b>b</b>) NDJ season. Red color shows positive correlation, blue color shows negative correlation. Only significant level values that are at or above the 95% level are colored. The orientation of the arrow indicates the direction of the wind, and their length is proportional to the magnitude of the correlation.</p>
Full article ">Figure 11
<p>Correlation between PC2 and SST for the (<b>a</b>) ASO season and (<b>b</b>) NDJ season. Only significant level values that are at or above the 95% level are colored. The orientation of the arrow indicates the direction of the wind, and their length is proportional to the magnitude of the correlation.</p>
Full article ">Figure 12
<p>Normalized anomalies at ASO season for (<b>a</b>) PC1, (<b>b</b>) PC2, (<b>c</b>) DMI, (<b>d</b>) Niño3.4 time series. Blue bars indicate positive anomalies, red bars indicate negative anomalies.</p>
Full article ">Figure 13
<p>Normalized anomalies at NDJ season for (<b>a</b>) PC1, (<b>b</b>) PC2, (<b>c</b>) DMI, (<b>d</b>) Niño3.4 time series. Blue bars indicate positive anomalies, red bars indicate negative anomalies.</p>
Full article ">Figure 14
<p>Composite maps of REOF1 and REOF2 anomalies during ASO (<b>left</b>) and NDJ (<b>right</b>) seasons in neutral years. REOF anomalies that are significant at the 95% level of the two-tailed <span class="html-italic">t</span>-test are marked with shaded areas.</p>
Full article ">Figure 15
<p>Composite maps of REOF1 and REOF2 anomalies during (<b>a</b>) El Niño and (<b>b</b>) La Niña. REOF anomalies that are significant at the 95% level of the two-tailed <span class="html-italic">t</span>-test are marked with shaded areas.</p>
Full article ">Figure 16
<p>Composite maps of REOF1 and REOF2 anomalies during (<b>a</b>) positive IOD and (<b>b</b>) negative IOD. REOF anomalies that are significant at the 95% level from a two-tailed <span class="html-italic">t</span>-test are marked with shaded areas.</p>
Full article ">
21 pages, 14252 KiB  
Article
Analysis of Change in Summer Extreme Precipitation in Southwest China and Human Adaptation
by Junyao Luo and Aihua Yang
Sustainability 2024, 16(17), 7329; https://doi.org/10.3390/su16177329 - 26 Aug 2024
Cited by 1 | Viewed by 868
Abstract
This study analyzed the change in and mechanisms of summer extreme precipitation in Southwest China (SWC) during 1979–2021. The trend in summer extreme precipitation showed an evident interdecadal mutation in the late 1990s; it decreased during 1979–1996 (P1) and increased during 1997–2021 (P2). [...] Read more.
This study analyzed the change in and mechanisms of summer extreme precipitation in Southwest China (SWC) during 1979–2021. The trend in summer extreme precipitation showed an evident interdecadal mutation in the late 1990s; it decreased during 1979–1996 (P1) and increased during 1997–2021 (P2). It is observed that the moisture flux in SWC is more abundant in P2 than in P1. The South Asian high (SAH) and western Pacific subtropical high (WPSH) contributed to the change in extreme precipitation in SWC. Both the SAH and WPSH weakened in 1979–1996 and enhanced in 1997–2021. The enhanced SAH and WPSH are conducive to forming updrafts in SWC and transporting moisture from the Bay of Bengal (BOB) and South China Sea (SCS) into SWC. Further research found that the causes for the interdecadal variation of the SAH and WPSH are the anomalies of sensible heat flux (SSH) over the Tibetan Plateau (TP) and sea surface temperature (SST) in the tropical western Pacific–Indian Oceans. The SSH is the main energy source of troposphere air and an essential component of the surface heat balance because it can maintain the intensity and influence range of the SAH. The increasing SST stimulated strong upward motion and thus maintained the strength of the WPSH, which also made the WPSH extend westward into mainland China. This study also summarized local human adaptation to climate change. The use of advanced science and technology to improve monitoring and forecasting ability is an important measure for human society to adapt to climate change. At the same time, increasing the participation of individuals and social organizations is also an indispensable way to increase human resilience to climate change. Full article
Show Figures

Figure 1

Figure 1
<p>Moving <span class="html-italic">t</span> test (<b>a</b>) and Yamamoto test (<b>b</b>) for extreme precipitation amount (red lines) and its frequency (blue lines). The black and gray dashed lines indicate the 99% and 95% significance level, respectively. The orange line represents the year (1997) in which the trend mutation occurs.</p>
Full article ">Figure 2
<p>Summer extreme precipitation amount ((<b>a</b>); red lines; unit: mm) and its frequency ((<b>a</b>); blue lines; unit: day) in SWC. Summer total precipitation amount ((<b>b</b>); red lines; unit: mm) and the number of rainy days ((<b>b</b>); blue lines; unit: day) in SWC. Percentage of extreme precipitation to summer total precipitation ((<b>c</b>); red lines; unit: %) and ratio of extreme precipitation frequency to rainy days ((<b>c</b>); blue lines; unit: %). Dotted lines indicate linear trends during P1 and P2. Regression function and variance contribution during P1 and P2 are shown in the left or right corners of each panel.</p>
Full article ">Figure 3
<p>Climatology of summer extreme precipitation amount ((<b>a</b>); unit: mm) and frequency ((<b>b</b>); unit: day) during 1979–2021. Distribution of the linear trend in the percentage ratio of extreme precipitation to summer total precipitation (unit: %·yr<sup>−1</sup>) during 1979–1996 (<b>c</b>) and 1997–2021 (<b>d</b>). + denote the linear trend significant at the 95% confidence level.</p>
Full article ">Figure 4
<p>Distribution of the extreme precipitation amount linear trend (shaded; unit: mm·yr<sup>−1</sup>) during 1979–1996 (<b>a</b>) and 1997–2021 (<b>c</b>). Distribution of the extreme precipitation frequency linear trend (shaded; unit: day·yr<sup>−1</sup>) during 1979–1996 (<b>b</b>) and 1997–2021 (<b>d</b>). + denote the linear trend significant at the 95% confidence level.</p>
Full article ">Figure 5
<p>Distribution of the linear trend in geopotential height (unit: gpm·yr<sup>−1</sup>) and horizontal wind velocity (unit: m·s<sup>−1</sup>·yr<sup>−1</sup>) during 1979–1996 at 200 hPa (<b>a</b>), 500 hPa (<b>b</b>), and 850 hPa (<b>c</b>). Distribution of the linear trend in geopotential height (unit: gpm·yr<sup>−1</sup>) and horizontal wind velocity (unit: m·s<sup>−1</sup>·yr<sup>−1</sup>) during 1997–2021 at 200 hPa (<b>d</b>), 500 hPa (<b>e</b>), and 850 hPa (<b>f</b>). Wind vector arrows exceeding the 95% confidence level are displayed. Dots denote the linear trends significant at the 95% confidence level. Areas bounded by pink rectangles denote SWC.</p>
Full article ">Figure 6
<p>Distribution of the moisture flux (unit: kg·m<sup>−1</sup>·s<sup>−1</sup>·yr<sup>−1</sup>) and divergence (unit: s<sup>−1</sup>·yr<sup>−1</sup>) linear trend during 1979–1996 (<b>a</b>) and 1997–2021 (<b>b</b>). Distribution of the 500 hPa omega linear trend (unit: pa·s<sup>−1</sup>·yr<sup>−1</sup>) during 1979–1997 (<b>c</b>) and 1997–2021 (<b>d</b>). Moisture flux arrows exceeding the 95% confidence level are displayed. Dots and + denote linear trends in divergence and omega significance at the 95% confidence level, respectively. Areas bounded by pink rectangles denote SWC.</p>
Full article ">Figure 7
<p>Standardized area index of SAH ((<b>a</b>); red line) and WPSH ((<b>a</b>); blue line). Standardized intensity index of SAH ((<b>b</b>); red line) and WPSH ((<b>b</b>); blue line). Regional average spring standardized SSH over TP (<b>c</b>). The dotted lines indicate linear trends during P1 and P2. Regression functions and variance contributions during P1 and P2 are marked in the left or right bottom corners of each panel.</p>
Full article ">Figure 8
<p>Distribution of SSH linear trend (unit: W·m<sup>−2</sup>·yr<sup>−1</sup>) during 1979–1996 (<b>a</b>) and 1997–2021 (<b>b</b>). Areas bounded by dashed gray rectangles denote TP. Dots denote the linear trend of SSH significance at the 95% confidence level.</p>
Full article ">Figure 9
<p>Composite geopotential height anomalies (shaded; unit: gpm) and horizontal wind anomalies (vectors; unit: m·s<sup>−1</sup>) at 200 hPa (<b>a</b>,<b>d</b>), 500 hPa (<b>b</b>,<b>e</b>), and 850 hPa (<b>c</b>,<b>f</b>) on day 0 of extreme precipitation events in SWC during Sub-P1 (left column) and Sub-P2 (right column). The region bounded by pink rectangles denotes SWC. Wind vector arrows exceeding the 95% confidence level are displayed. Dots denote the anomalies significant at the 95% confidence level. The blue and red lines represent 588 gpm (12,500 gpm) when the climate average state and extreme precipitation events occur, respectively.</p>
Full article ">Figure 10
<p>Latitude pressure (left vertical coordinate) cross-section of composite meridional wind (unit: m·s<sup>−1</sup>) and vertical motion (unit: 10<sup>−2</sup> Pa·s<sup>−1</sup>) vectors anomalies averaged over 97.5°–110° E on day 0 of REPEs during Sub-P1 (<b>a</b>) and Sub-P2 (<b>b</b>). The area within the green dashed lines denotes SWC. Dots denote anomalies significant at the 95% confidence level.</p>
Full article ">Figure 11
<p>Composite moisture flux anomalies (vectors; unit: kg·m<sup>−1</sup>·s<sup>−1</sup>) and divergence anomalies (shaded; unit: s<sup>−1</sup>) at 850–500 hPa averaged on day 0 of REPEs in SWC during sub-P1 (<b>a</b>) and sub-P2 (<b>c</b>). Composite SSH anomalies (shaded; unit: W·m<sup>−2</sup>) on day 0 of REPEs over SWC during sub-P1 (<b>b</b>) and sub-P2 (<b>d</b>). The regions bounded by pink rectangles denote SWC. The areas bounded by gray rectangles denote TP. Moisture flux arrows exceeding the 95% confidence level are displayed. Dots denote anomalies significant at the 95% confidence level.</p>
Full article ">Figure 12
<p>Correlation coefficients between SST in winter ((<b>a</b>); December–February), spring ((<b>b</b>); March–May), and summer ((<b>c</b>); June–August) and extreme precipitation in SWC. Areas bounded by pink rectangles denote SWC. Areas bounded by green rectangles denote the key regions used to define the significant impact on extreme precipitation. Distribution of the SST (DJF, MAM, and JJA) linear trend (unit: °C·decade<sup>−1</sup>) during 1979–1996 (<b>d</b>–<b>f</b>) and 1997–2021 (<b>g</b>–<b>i</b>). Dots denote the linear trend significant at the 95% confidence level.</p>
Full article ">Figure 13
<p>(<b>a</b>) Decadal difference in Hadley circulation (P2 minus P1) averaged over 97.5°–110° E. (<b>b</b>) Decadal difference in Walker circulation (P2 minus P1) averaged over 0°–20° N. The area within green dashed lines denotes SWC. The arrows represent the direction of air flow. The regions bounded by green rectangles denote SWC.</p>
Full article ">
12 pages, 3255 KiB  
Article
A New Perspective of the Spring Predictability Barrier Based on the Zonal Sea Level Pressure Gradient
by Jing Tan, Fei Zheng, Tingwei Cao, Yongyong Huang and Haiyan Wang
J. Mar. Sci. Eng. 2024, 12(9), 1463; https://doi.org/10.3390/jmse12091463 - 23 Aug 2024
Viewed by 520
Abstract
Currently, the “spring predictability barrier” (SPB) is still a controversial problem in many atmosphere–ocean coupled models and has significant impacts on degrading the El Niño–Southern Oscillation (ENSO) predictions across the boreal spring. In this study, unlike previous studies that viewed the SPB issue [...] Read more.
Currently, the “spring predictability barrier” (SPB) is still a controversial problem in many atmosphere–ocean coupled models and has significant impacts on degrading the El Niño–Southern Oscillation (ENSO) predictions across the boreal spring. In this study, unlike previous studies that viewed the SPB issue from the perspective of sea surface temperature (SST), based on the Bjerknes feedback theory and the decadal variations in Walker circulation over the tropical Pacific, a new perspective of the SPB is revealed by the seasonal variations in the observed zonal sea level pressure (SLP) gradient, which can reflect the stability and variability of the atmosphere–ocean interactions during the ENSO’s evolution. More importantly, a significant decadal variation of SPB strength (SPBS) is exhibited in the last 3 decades, from 1991 to 2020, which is strongly controlled by the dominant patterns of sea surface temperature (SST) and Walker circulation, and associated with the background mean atmosphere–ocean states. That is to say, the atmosphere–ocean interaction pattern over the tropical Pacific has undergone decadal variations over the past 3 decades which determine the decadal variations in SPBS. International Research Institute for Climate and Society/Climate Prediction Center (IRI/CPC) multi-models show stronger SPBS during 2001–2010 than during 2011–2020, indicating that the decadal variations in SPBS from statistical analysis also exist in actual model predictions, which further confirms the rationality of this perspective of SPB based on the zonal SLP gradient. Full article
(This article belongs to the Section Ocean and Global Climate)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The seasonal correlation coefficients between the SLP averaged over Indonesia and the EEP during 1991–2020 (blue line) and the seasonal standard deviation of the Niño3.4 SST (green line), and (<b>b</b>) the seasonal correlation coefficients between the SLP averaged over Indonesia and the EEP during 1991–2000 (blue line) and 2001–2010 (cyan line). Shading indicates the range of maximum and minimum correlation coefficients from the leave-one-out cross-validation, and the dashed line shows that the correlation coefficients are significant ant the 95% confidence level.</p>
Full article ">Figure 2
<p>MAM–mean (<b>a</b>) IPO index, (<b>b</b>) correlation coefficients between the SLP averaged over Indonesia and the EEP, and (<b>c</b>) standard deviation of the Niño3.4 index with 10-year moving windows for 1986–2022.</p>
Full article ">Figure 3
<p>Seasonal correlation coefficients between the SLP averaged over Indonesia and Equatorial Eastern Pacific for periods 1991–2000 (blue line), 2001–2010 (cyan line), and 2011–2020 (red line). Shading indicates the range of maximum and minimum correlation coefficients from the leave-one-out cross-validation for the period 2011–2020, and the dashed line shows that the correlation coefficients are significant at the 95% confidence level.</p>
Full article ">Figure 4
<p>Spatial patterns of the leading MEOF mode of atmospheric circulation fields (U-wind and omega; vertical) and the SST (85°E–60°W, 30°S–30°N; horizontal) during boreal spring. (<b>a</b>) The analysis period is 1991–2000, with the leading MEOF mode explaining 37.2% of the total variance. (<b>b</b>) The analysis period is 2001–2010, with the leading MEOF mode explaining 29.5% of the total variance. (<b>c</b>) The analysis period is 2011–2020, with the leading MEOF mode explaining 34.8% of the total variance.</p>
Full article ">Figure 5
<p>The east–west differences of the SST (red line) and SLP (black line) overlaid the IPO index during boreal spring of 1991–2021.</p>
Full article ">Figure 6
<p>The variation of spring predictability/persistence barrier (SPB) intensity between pre-2010 and post-2010 in IRI/CPC multi-models and observation.</p>
Full article ">
15 pages, 7765 KiB  
Article
Impact of May–June Antarctic Oscillation on July–August Heat-Drought Weather in Yangtze River Basin
by Zhengxuan Yuan, Jun Zhang, Liangmin Du, Ying Xiao and Sijing Huang
Atmosphere 2024, 15(8), 998; https://doi.org/10.3390/atmos15080998 - 20 Aug 2024
Viewed by 627
Abstract
Investigating the physical mechanism behind the formation of summer heat-drought weather (HDW) in the Yangtze River Basin (YRB) holds significant importance for predicting summer precipitation and temperature patterns in the region as well as disaster mitigation and prevention. This study focuses on spatiotemporal [...] Read more.
Investigating the physical mechanism behind the formation of summer heat-drought weather (HDW) in the Yangtze River Basin (YRB) holds significant importance for predicting summer precipitation and temperature patterns in the region as well as disaster mitigation and prevention. This study focuses on spatiotemporal patterns of July–August (JA) HDW in the YRB from 1979 to 2022, which is linked partially to the preceding May–June (MJ) Antarctic Oscillation (AAO). Key findings are summarized as follows: (1) The MJ AAO displays a marked positive correlation with the JA HDW index (HDWI) in the southern part of upper YRB (UYRB), while showing a negative correlation in the area extending from the Han River to the western lower reaches of the YRB (LYRB); (2) The signal of MJ AAO persists into late JA through a specific pattern of Sea Surface Temperature anomalies in the Southern Ocean (SOSST). This, in turn, modulates the atmospheric circulation over East Asia; (3) The SST anomalies in the South Atlantic initiate Rossby waves that cross the equator, splitting into two branches. One branch propagates from the Somali-Tropical Indian Ocean, maintaining a negative-phased East Asia–Pacific (EAP) teleconnection pattern. This enhances the moisture flow from the Pacific towards the middle and lower reaches of the Yangtze River Basin (MYRB-LYRB). The other branch propagates northward, crossing the Somali region, and induces a positive geopotential height anomaly over Urals-West Asia. This reduces the southwesterlies towards the UYRB, thereby contributing to HDW variabilities in the region. (4) Partial Least Squares Regression (PLSR) demonstrated predictive capability for JA HDW in the YRB for 2022, based on Southern Ocean SST. Full article
Show Figures

Figure 1

Figure 1
<p>Correlation coefficients between JA SPI90 and JA HDWI in YRB based on observational stations (marker ‘square’ denotes statistically significant at 95% confidence level, markers ‘plus’ are insignificance stations).</p>
Full article ">Figure 2
<p>Spatiotemporal modes for EOF results of Jul-Aug HDWI in Yangtze River Basin from 1979 to 2022 ((<b>a</b>). EOF1, (<b>b</b>). EOF2, (<b>c</b>). EOF3. The left columns are spatial patterns as correlation coefficients while the right columns are principal components of each spatial pattern. VCR denotes Variance Contribution Rate).</p>
Full article ">Figure 3
<p>Correlation coefficients July–August PC3 in Yangtze River Basin from 1979 to 2022 and simultaneously atmospheric circulation ((<b>a</b>) for H200, (<b>b</b>) for H500, (<b>c</b>) for H850, dotted regions are statistically significant at 90% confidence level; (<b>d</b>) for UV200H200, (<b>e</b>) for UV850H500, (<b>f</b>) for UV1000H850, green ‘+’ and black vectors are statistically significant at 90% confidence level).</p>
Full article ">Figure 4
<p>Correlation coefficients between July–August PC3 in Yangtze River Basin from 1979 to 2022 and simultaneously SST (dotted regions are statistically significant at 90% confidence level).</p>
Full article ">Figure 5
<p>Correlation coefficients between May–June AAO and simultaneously to following SST from 1979 to 2022 (As <a href="#atmosphere-15-00998-f003" class="html-fig">Figure 3</a>, (<b>a</b>) for May–June SST, (<b>b</b>) for July–August SST).</p>
Full article ">Figure 6
<p>(<b>a</b>) TN flux at 250 hPa induced by the variation of JA SOSST (vectors denote wave flux, colors denote streamfunction (STRF)). (<b>b</b>) correlation between JA SOSST and JA STRF500 (color shaded regions are significant at 90% confidence level). (<b>c</b>,<b>d</b>) Meridional wave tunnels induced by the variation of JA SOSST ((<b>c</b>) for northward tunnel, (<b>d</b>) for southward tunnel).</p>
Full article ">Figure 7
<p>Correlation coefficients July–August SOSSTI from 1979 to 2022 and simultaneously atmospheric circulation ((<b>a</b>) for H200, (<b>b</b>) for H500, (<b>c</b>) for H850, dotted regions are statistically significant at 90% confidence level; (<b>d</b>) for UV200H200, (<b>e</b>) for UV850H500, (<b>f</b>) for UV1000H850, green ‘+’ and black vectors are statistically significant at 90% confidence level).</p>
Full article ">Figure 8
<p>Correlation between JA SOSSTI and simultaneous water vapor flux ((<b>a</b>) 500 hPa; (<b>b</b>) 850 hPa, black vectors are significant at 90% confide level).</p>
Full article ">Figure 9
<p>PLS leading modes for the prediction of JA HDW in the YRB by JA SST in Southern Ocean ((<b>a</b>) for mode 1, (<b>b</b>) for mode 2, (<b>c</b>) for mode 3, (<b>d</b>) for mode 4, dotted regions are significant at 90% confidence level).</p>
Full article ">Figure 10
<p>PLS predictive result of JA HDW in the YRB during 2022.</p>
Full article ">Figure 11
<p>The 3rd heterogeneous mode of global JA SST and northeast hemispheric JA H500 ((<b>a</b>) SST; (<b>b</b>) H500).</p>
Full article ">
18 pages, 20146 KiB  
Article
Changed Relationship between the Spring North Atlantic Tripole Sea Surface Temperature Anomalies and the Summer Meridional Shift of the Asian Westerly Jet
by Lin Chen, Gen Li and Jiaqi Duan
Atmosphere 2024, 15(8), 922; https://doi.org/10.3390/atmos15080922 - 1 Aug 2024
Viewed by 640
Abstract
The summer Asian westerly jet (AWJ)’s shifting in latitudes is one important characteristic of its variability and has great impact on the East Asian summer climate. Based on the observed and reanalyzed datasets from the Hadley Center Sea Ice and Sea Surface Temperature [...] Read more.
The summer Asian westerly jet (AWJ)’s shifting in latitudes is one important characteristic of its variability and has great impact on the East Asian summer climate. Based on the observed and reanalyzed datasets from the Hadley Center Sea Ice and Sea Surface Temperature dataset (HadISST), the Japanese 55-year reanalysis (JRA-55), and the fifth generation of the European Centre for Medium-Range Weather Forecasts atmospheric reanalysis (ERA5), this study investigates the relationship between the spring tripole North Atlantic SST (TNAT) anomalies and the summer meridional shift of the AWJ (MSJ) for the period of 1958–2020. Through the method of correlation analysis and regression analysis, we show that the ‘+ - +’ TNAT anomalies in spring could induce a northward shift of the AWJ in the following summer. However, such a climatic effect of the spring TNAT anomalies on the MSJ is unstable, exhibiting an evident interdecadal strengthening since the early 1990s. Further analysis reveals that this is related to a strengthened intensity of the spring TNAT anomalies in the most recent three decades. Compared to the early epoch (1958–1993), the stronger spring TNAT anomalies in the post epoch (1994–2020) could cause a stronger pan-tropical climate response until the following summer through a series of ocean–atmosphere interactions. Through Gill responses, the resultant more prominent cooling in the central Pacific in response to the ‘+ - +’ TNAT anomalies induces a pan-tropical cooling in the upper troposphere, which weakens the poleward gradient of the tropospheric temperature over subtropical Asia. As a result, the AWJ shifts northward via a thermal wind effect. By contrast, in the early epoch, the spring TNAT anomalies are relatively weaker, inducing weaker pan-tropical ocean–atmosphere interactions and thus less change in the meridional shit of the summer AWJ. Our results highlight a strengthened lagged effect of the spring TNAT anomalies on the following summer MSJ and have important implications for the seasonal climate predictability over Asia. Full article
(This article belongs to the Section Climatology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Climatology and (<b>b</b>) standard deviations of the 200 hPa zonal winds (m s<sup>−1</sup>) during boreal summer (June–July–August) for the period of 1958–2020. The solid black lines in (<b>a</b>,<b>b</b>) denote the axis of the Asian westerly jet (AWJ).</p>
Full article ">Figure 2
<p>(<b>a</b>) The leading mode of the empirical orthogonal function (EOF) analysis of the 200 hPa zonal wind anomalies over the region of 15° N–65° N, 30° E–180° E for the period of 1958–2020. (<b>b</b>) Time series of the first principal component (PC1) of the EOF analysis on the 200 hPa zonal wind anomalies over the region of 15° N–65° N, 30° E–180° E.</p>
Full article ">Figure 3
<p>Correlations of the summer meridional shift of the AWJ (MSJ) index with the spring sea surface temperature (SST) anomalies for the period of 1958–2020. The dots indicate the correlations at a significance level of <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 4
<p>Normalized time series of the spring tripole North Atlantic SST (TNAT) index (black solid line) and the summer MSJ index (black dashed line) for the period of 1958–2020. The solid blue line denotes their 21-year sliding correlation coefficients, with the dashed blue line denoting a significance level of <span class="html-italic">p</span> &lt; 0.05. The solid red line denotes the 15-year sliding correlation coefficients, with the dashed red line denoting a significance level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>(<b>a</b>) Correlations of the summer MSJ index with the spring SST anomalies for the period of 1958–1993. (<b>b</b>) Same as (<b>a</b>), but for the period of 1994–2020. The dots indicate the correlations at a significance level of <span class="html-italic">p</span> &lt; 0.1.</p>
Full article ">Figure 6
<p>The 21-year sliding standard deviation of the spring TNAT index.</p>
Full article ">Figure 7
<p>Regressions of (<b>a</b>) spring and (<b>c</b>) summer SST (shaded; °C), 500 hPa omega (contours; pa s<sup>−1</sup>; positive upward; the interval is 3 × 10<sup>−4</sup> pa s<sup>−1</sup> with the green/purple contours denoting negative/positive values) and 850 hPa wind anomalies (arrows; m s<sup>−1</sup>) onto the spring TNAT index for the period of 1958–1993. (<b>b</b>,<b>d</b>) Same as (<b>a</b>,<b>c</b>), but for the period of 1994–2020. The dots indicate the regressed SST anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. The black arrows indicate the zonal or meridional components of the wind anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. Wind speeds less than 0.25 m s<sup>−1</sup> are not shown. The absolute values of the omega anomalies less than 1.5 × 10<sup>−4</sup> pa s<sup>−1</sup> are not shown.</p>
Full article ">Figure 8
<p>Regressions of the summer 200 hPa velocity potential (shaded; 10<sup>5</sup> m<sup>2</sup> s<sup>−1</sup>), divergent wind (arrows; m s<sup>−1</sup>), and precipitation (contours; mm month<sup>−1</sup>; the interval is 7.5 mm month<sup>−1</sup> with the green/purple contours denoting negative/positive values) anomalies onto the spring TNAT index for the period of (<b>a</b>) 1958–1993 and (<b>b</b>) 1994–2020. Wind speeds less than 0.1 m s<sup>−1</sup> are not shown.</p>
Full article ">Figure 9
<p>(<b>a</b>) Regressions of the summer 200 hPa air temperature (shaded; K) and wind anomalies (arrows; m s<sup>−1</sup>) onto the spring TNAT index for the period of 1958–1993. (<b>b</b>) Same as (<b>a</b>), but for the period of 1994–2020. The dots indicate the regressed air temperature anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. The black arrows indicate the zonal or meridional components of the wind anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. Wind speeds less than 0.25 m s<sup>−1</sup> are not shown.</p>
Full article ">Figure 10
<p>The height–latitude cross-section of the summer climatological tropospheric temperature (contours; K) and its meridional gradient (shaded; 10<sup>−5</sup> K m<sup>−1</sup>) averaged between (<b>a</b>) 40° E–90° E, (<b>d</b>) 90° E–140° E. (<b>b</b>,<b>e</b>) same as (<b>a</b>,<b>d</b>), but for the summer anomalous tropospheric temperature (contours; K; the interval is 0.05 K) and its meridional gradient (shaded; 10<sup>−6</sup> K m<sup>−1</sup>) regressed onto the spring TNAT index for the period of 1958–1993. (<b>c</b>,<b>f</b>) same as (<b>b</b>,<b>e</b>), but for the period of 1994–2020. The absolute values of the tropospheric temperature anomalies less than 0.025 K are not shown.</p>
Full article ">Figure 11
<p>Regressed height–latitude cross-section of the summer zonal wind anomalies (shaded; m s<sup>−1</sup>) averaged between 40° E and 90° E onto the spring TNAT index for the period of (<b>a</b>) 1958–1993, (<b>b</b>) 1994–2020. The dots indicate the regressed zonal wind anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. The black solid lines are the zonal averaged (40° E–90° E) climatological zonal winds equal to or larger than 15 m s<sup>−1</sup> with the interval of 5 m s<sup>−1</sup>, denoting the westerly jet. (<b>c</b>,<b>d</b>) same as (<b>a</b>,<b>b</b>), but for those averaged between 90° E and 140° E.</p>
Full article ">Figure 12
<p>(<b>a</b>) Regressions of the summer 200 hPa zonal wind anomalies (m s<sup>−1</sup>) onto the spring TNAT index for the period of 1958–1993. (<b>b</b>) Same as (<b>a</b>), but for the period of 1994–2020. The dots indicate the regressed zonal wind anomalies at a significance level of <span class="html-italic">p</span> &lt; 0.1. The black line denotes the AWJ axis.</p>
Full article ">Figure 13
<p>Spring TNAT index, summer pan-tropical SST index, summer poleward gradient of the 200 hPa tropospheric temperature (TT) index, and the summer MSJ index regressed onto the spring TNAT index for the periods of 1958–1993 (blue bars) and 1994–2020 (red bars). The solid bars indicate the regressed indices at a significance level of <span class="html-italic">p</span> &lt; 0.1. All data are standardized.</p>
Full article ">
21 pages, 15118 KiB  
Article
Annual and Seasonal Variation of the Ocean Thermal Resources off the Mexican Coast
by Carlos Melecio Carmona-Cedillo, Armando Trasviña-Castro, Valeria Chávez and Rodolfo Silva
J. Mar. Sci. Eng. 2024, 12(7), 1160; https://doi.org/10.3390/jmse12071160 - 10 Jul 2024
Viewed by 1139
Abstract
A large amount of thermal energy is stored in the oceans between the tropics, available for conversion into electrical energy using OTEC technology. The aim of this study was to determine the annual and seasonal variability of the oceanic thermal resource in Mexico. [...] Read more.
A large amount of thermal energy is stored in the oceans between the tropics, available for conversion into electrical energy using OTEC technology. The aim of this study was to determine the annual and seasonal variability of the oceanic thermal resource in Mexico. Using the WOA18 database, we mapped surface temperature at a 10 m depth, deep cold water (<5 °C), vertical temperature difference (18 and 20 °C), and temperature anomalies. From the results, four areas were analyzed as being suitable for the installation of OTEC technology: Pacific (A), Los Cabos (B), Caribbean (C), and Gulf of Mexico (G). The optimal thermal resource (≥20 °C) was found between a 400 and 1000 m depth in all seasons in A and C, in spring, summer, and autumn in G, and only in summer and autumn in B. The suboptimal thermal resource (between 18 and 20 °C) was present between 400 and 800 m in all seasons in A, C, and G, and in summer and autumn in B. These results provide new information of utmost importance for future location and design considerations of OTEC plants on Mexican coasts, and the methodology can be used in other areas where there is a lack of field data and the development of OTEC technology is being considered. Full article
(This article belongs to the Section Physical Oceanography)
Show Figures

Figure 1

Figure 1
<p>Study area. The four coastal areas shown are where the continental platform is near the coastline (yellow lines, A, B, C, and G). Bathymetry from GEBCO.</p>
Full article ">Figure 2
<p>(<b>a</b>) Annual mean sea temperature at a 10 m depth (SST10). (<b>b</b>) Deep cold water (5 °C isotherm) depth, (<b>c</b>) suboptimal warm water (23 °C isotherm) depth, and (<b>d</b>) optimal warm water (25 °C isotherm) depth. Areas of interest shown: (A) Pacific, (B) Los Cabos, (C) Caribbean, and (G) the Gulf of Mexico. WOA18 data.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Annual mean sea temperature at a 10 m depth (SST10). (<b>b</b>) Deep cold water (5 °C isotherm) depth, (<b>c</b>) suboptimal warm water (23 °C isotherm) depth, and (<b>d</b>) optimal warm water (25 °C isotherm) depth. Areas of interest shown: (A) Pacific, (B) Los Cabos, (C) Caribbean, and (G) the Gulf of Mexico. WOA18 data.</p>
Full article ">Figure 3
<p>(<b>a</b>) Annual VTD at a 700 m depth (estimated from WOA18 data using Equation (1)). (<b>b</b>) Suboptimal VTD (18 °C isothermal) depth and (<b>c</b>) optimal VTD (20 °C isothermal) depth. Areas of interest shown: (A) Pacific, (B) Los Cabos, (C) Caribbean, and (G) the Gulf of Mexico. WOA18 data.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Annual VTD at a 700 m depth (estimated from WOA18 data using Equation (1)). (<b>b</b>) Suboptimal VTD (18 °C isothermal) depth and (<b>c</b>) optimal VTD (20 °C isothermal) depth. Areas of interest shown: (A) Pacific, (B) Los Cabos, (C) Caribbean, and (G) the Gulf of Mexico. WOA18 data.</p>
Full article ">Figure 4
<p>Seasonal ocean temperature at a 10 m depth: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The 23 and 25 °C isotherms (black bold lines) identify hot water that could be useful for OTEC near the coast at Los Cabos (line B), in the Pacific (line A), on the Gulf of Mexico (line G), and in the Caribbean (line C). WOA18 Data.</p>
Full article ">Figure 4 Cont.
<p>Seasonal ocean temperature at a 10 m depth: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The 23 and 25 °C isotherms (black bold lines) identify hot water that could be useful for OTEC near the coast at Los Cabos (line B), in the Pacific (line A), on the Gulf of Mexico (line G), and in the Caribbean (line C). WOA18 Data.</p>
Full article ">Figure 4 Cont.
<p>Seasonal ocean temperature at a 10 m depth: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The 23 and 25 °C isotherms (black bold lines) identify hot water that could be useful for OTEC near the coast at Los Cabos (line B), in the Pacific (line A), on the Gulf of Mexico (line G), and in the Caribbean (line C). WOA18 Data.</p>
Full article ">Figure 5
<p>Seasonal ocean temperature (°C) minus annual climatology at a 10 m depth: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The negative values indicate a trend beneath the annual average, while positive values indicate a trend over the annual average. The four zones of interest are shown with the white lines A, B, C, and G. WOA18 data.</p>
Full article ">Figure 5 Cont.
<p>Seasonal ocean temperature (°C) minus annual climatology at a 10 m depth: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The negative values indicate a trend beneath the annual average, while positive values indicate a trend over the annual average. The four zones of interest are shown with the white lines A, B, C, and G. WOA18 data.</p>
Full article ">Figure 6
<p>Seasonal depths of the 20 °C VTD isotherm, an optimal resource for OTEC plants: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. WOA18 data.</p>
Full article ">Figure 6 Cont.
<p>Seasonal depths of the 20 °C VTD isotherm, an optimal resource for OTEC plants: (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. WOA18 data.</p>
Full article ">Figure 7
<p>Seasonal depths of the suboptimal resource for the OTEC plant (18 °C VTD isotherm): (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and <b>(d</b>) autumn. WOA18 data.</p>
Full article ">Figure 7 Cont.
<p>Seasonal depths of the suboptimal resource for the OTEC plant (18 °C VTD isotherm): (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and <b>(d</b>) autumn. WOA18 data.</p>
Full article ">
17 pages, 10295 KiB  
Article
Interannual Fluctuations and Their Low-Frequency Modulation of Summertime Heavy Daily Rainfall Potential in Western Japan
by Takashi Mochizuki
Atmosphere 2024, 15(7), 814; https://doi.org/10.3390/atmos15070814 - 7 Jul 2024
Viewed by 925
Abstract
Heavy rainfall under the conditions of the changing climate has recently garnered considerable attention. The statistics on heavy daily rainfall offer vital information for assessing present and future extreme events and for clarifying the impacts of global climate variability and change, working to [...] Read more.
Heavy rainfall under the conditions of the changing climate has recently garnered considerable attention. The statistics on heavy daily rainfall offer vital information for assessing present and future extreme events and for clarifying the impacts of global climate variability and change, working to form a favorable background. By analyzing a set of large-ensemble simulations using a global atmospheric model, this study demonstrated that two different physical processes in global climate variability control the interannual fluctuations in the 99th- and 90th-percentile values of summertime daily rainfall (i.e., the potential amounts) on Kyushu Island in western Japan. The 90th-percentile values were closely related to large-scale horizontal moisture transport anomalies due to changes in the subtropical high in the northwestern Pacific, which was usually accompanied by basin-scale warming in the Indian Ocean subsequent to the wintertime El Niño events. The contributions of the sea surface temperatures over the northern Indian Ocean and the eastern tropical Pacific Ocean showed low-frequency modulations, mainly due to the influences of the global warming tendency and the interdecadal variability in the climate system, respectively. In contrast, tropical cyclone activity played a major role in changing the 99th-percentile value. The potentials of both the tropical cyclone intensity and the existence density fluctuated, largely owing to the summertime sea surface temperature over the tropical Pacific, which can be modulated by the El Niño diversity on interdecadal timescales. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The standard deviation of the observed daily rainfall in July (1981–2010) in the CPC global unified gauge-based daily precipitation dataset. (<b>B</b>) Same as panel (<b>A</b>), except for the simulated daily rainfall in the d4PDF dataset. The plotted values are the averages of the standard deviations calculated for the individual ensemble member.</p>
Full article ">Figure 2
<p>(<b>A</b>) Time series of the 99th-percentile (red) and 90th-percentile (black) values (mm/day) of the daily rainfall potential on Kyushu Island from June to August (1981–2010). The plotted values are anomalies derived from 100 ensemble simulations of the d4PDF dataset (solid lines) and those of another subset of d4PDF that was compiled using the pseudo-observed SST with the warming tendency due to anthropogenic forcing being removed (i.e., the so-called NAT simulation) (broken lines). (<b>B</b>) Scatter plots of the 99th-percentile (red) and 90th-percentile (black) values (mm/day) relative to the seasonal mean values (mm/day) derived from 100 ensemble simulations of the d4PDF dataset for the period from June to August each year.</p>
Full article ">Figure 3
<p>(<b>A</b>) The correlation coefficients of (shaded) the surface temperature anomalies and (contours) the vorticity at 850 hPa relative to the 99th-percentile values of the daily rainfall potential on Kyushu Island during the period from June to August (1981–2010). The plotted values indicate areas where the positive (warm colors) and negative (cool colors) correlation coefficients were significant at the 99, 95, and 90% confidence levels. Green vectors indicate the regression values of the horizontal moisture transport at 850 hPa in areas where the regression values of both the zonal and meridional components were significant at the 90% confidence level. The regression value is the coefficient, <span class="html-italic">regr</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>), of linear regression, <span class="html-italic">Y</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>,<span class="html-italic">t</span>) = <span class="html-italic">regr</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>) ∗ <span class="html-italic">X</span>(<span class="html-italic">t</span>) + <span class="html-italic">b</span>, where <span class="html-italic">X</span>(<span class="html-italic">t</span>) and <span class="html-italic">Y</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>,<span class="html-italic">t</span>) denote the 99th-percentile values of the daily rainfall potential on Kyushu Island and the plotted variables, respectively. (<b>B</b>) Same as in panel (<b>A</b>), except for the correlation and regression values being relative to the 90th-percentile values of the daily rainfall potential.</p>
Full article ">Figure 4
<p>(<b>A</b>) The correlation coefficients of the 90th-percentile values of the daily rainfall potential at each grid point relative to those on Kyushu Island in July (1981–2010) in the d4PDF dataset. (<b>B</b>) Same as panel (<b>A</b>), except for the correlations being relative to the fourth-heaviest values of daily rainfall observed on Kyushu Island in the CPC global unified gauge-based daily precipitation dataset.</p>
Full article ">Figure 5
<p>(<b>A</b>) The shades indicate the correlations in the maximum wind speed of the tropical cyclone center relative to the 99th-percentile values of the daily rainfall potential on Kyushu Island during the period from June to August. The plotted values indicate areas where the correlation coefficients were significant at the 99, 95, and 90% confidence levels. The contours indicate climatological values of the maximum wind speed of the tropical cyclone center (m/s). (<b>B</b>) Same as panel (<b>A</b>), except for the maximum sea level pressure of the tropical cyclone track (hPa). (<b>C</b>) Same as panel (<b>A</b>), except for the tropical cyclone track density (days/year).</p>
Full article ">Figure 6
<p>(<b>A</b>) Correlation coefficients of the 99th-percentile values of the daily rainfall potential at each grid point relative to those on Kyushu Island in July (1981–2010) in the d4PDF dataset. (<b>B</b>) Same as panel (<b>A</b>), except for correlations being relative to the heaviest values of daily rainfall observed on Kyushu Island in the CPC global unified gauge-based daily precipitation dataset.</p>
Full article ">Figure 7
<p>(<b>A</b>) The lagged correlation coefficients of 3-month mean observed SSTs relative to the 99th-percentile values of the daily rainfall potential on Kyushu Island during the period from June to August. A cosine envelope was applied to all the data for the analysis period (1981–2010). The plotted values indicate areas where the correlation coefficients are significant at the 99, 95, and 90% confidence levels. (<b>B</b>) Same as in panel (<b>A</b>), except for the application of a sine envelope to the SST and rainfall data for the analysis period. (<b>C</b>) The significant differences in the lagged SST regressions between the sine-enveloped data (see panel (<b>B</b>)) and the cosine-enveloped data (see panel (<b>A</b>)) in an <span class="html-italic">F</span>-test. The lagged regression values were calculated using 3-month mean observed SSTs relative to the 99th-percentile values of the daily rainfall potential on Kyushu Island during the period from June to August. The regression value is the coefficient, <span class="html-italic">regr</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>), of linear regression, <span class="html-italic">Y</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>,<span class="html-italic">t</span>) = <span class="html-italic">regr</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>) ∗ <span class="html-italic">X</span>(<span class="html-italic">t</span>) + <span class="html-italic">b</span>, where <span class="html-italic">X</span>(<span class="html-italic">t</span>) and <span class="html-italic">Y</span>(<span class="html-italic">x</span>,<span class="html-italic">y</span>,<span class="html-italic">t</span>) denote the 99th-percentile values of the daily rainfall potential on Kyushu Island and the plotted variables, respectively.</p>
Full article ">Figure 8
<p>Same as in <a href="#atmosphere-15-00814-f007" class="html-fig">Figure 7</a>, except that the lagged correlation and regression values are relative to the 90th-percentile values of the daily rainfall potential on Kyushu Island during the period from June to August.</p>
Full article ">Figure 9
<p>(<b>A</b>) The perfect model anomaly correlation coefficients of the 99th-percentile values of the summertime daily rainfall potential in 100 ensembles of the d4PDF dataset. The contours represent the average of the correlation coefficients of interannual fluctuations. The shades represent the areas where more than 90% of the perfect model ensembles indicated positive correlation values, as a proxy for significant levels of the correlation coefficients. (<b>B</b>) Same as in panel (<b>A</b>), except for the 90th-percentile values of the daily rainfall potential.</p>
Full article ">Figure 10
<p>The skewness of the probability density of daily rainfall on Kyushu Island, defined as the cubed deviation of the normalized values for a specific threshold. The blue and red dots indicate the skewness values of the potential chances exceeding a specific rainfall intensity (i.e., 10, 20, 30, …, 100 mm/day) and the potential amounts of daily heavy rainfall (i.e., 99th-, 95th-, and 90th-percentile values), respectively.</p>
Full article ">
Back to TopTop