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

Search Results (425)

Search Parameters:
Keywords = MODIS-Aqua

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 12829 KiB  
Article
Evaluating the Relationship Between Vegetation Status and Soil Moisture in Semi-Arid Woodlands, Central Australia, Using Daily Thermal, Vegetation Index, and Reflectance Data
by Mauro Holzman, Ankur Srivastava, Raúl Rivas and Alfredo Huete
Remote Sens. 2025, 17(4), 635; https://doi.org/10.3390/rs17040635 - 13 Feb 2025
Viewed by 423
Abstract
Wet rainfall pulses control vegetation growth through evapotranspiration in most dryland areas. This topic has not been extensively analyzed with respect to the vast semi-arid ecosystems of Central Australia. In this study, we investigated vegetation water responses to in situ root zone soil [...] Read more.
Wet rainfall pulses control vegetation growth through evapotranspiration in most dryland areas. This topic has not been extensively analyzed with respect to the vast semi-arid ecosystems of Central Australia. In this study, we investigated vegetation water responses to in situ root zone soil moisture (SM) variations in savanna woodlands (Mulga) in Central Australia using satellite-based optical and thermal data. Specifically, we used the Land Surface Water Index (LSWI) derived from the Advanced Himawari Imager on board the Himawari 8 (AHI) satellite, alongside Land Surface Temperature (LST) from MODIS Terra and Aqua (MOD/MYD11A1), as indicators of vegetation water status and surface energy balance, respectively. The analysis covered the period from 2016 to 2021. The LSWI increased with the magnitude of wet pulses and showed significant lags in the temporal response to SM, with behavior similar to that of the Enhanced Vegetation Index (EVI). By contrast, LST temporal responses were quicker and correlated with daily in situ SM at different depths. These results were consistent with in situ relationships between LST and SM, with the decreases in LST being coherent with wet pulse magnitude. Daily LSWI and EVI scores were best related to subsurface SM through quadratic relationships that accounted for the lag in vegetation response. Tower flux measures of gross primary production (GPP) were also related to the magnitude of wet pulses, being more correlated with the LSWI and EVI than LST. The results indicated that the vegetation response varied with SM depths. We propose a conceptual model for the relationship between LST and SM in the soil profile, which is useful for the monitoring/forecasting of wet pulse impacts on vegetation. Understanding the temporal changes in rainfall-driven vegetation in the thermal/optical spectra associated with increases in SM can allow us to predict the spatial impact of wet pulses on vegetation dynamics in extensive drylands. Full article
Show Figures

Figure 1

Figure 1
<p>Map of major vegetation groups showing the location of Alice Springs Mulga (ASM) and Ti Tree Ozflux sites (data source: Dynamic Land Cover Dataset Version 2.1).</p>
Full article ">Figure 2
<p>Workflow diagram of satellite data (AHI and MODIS) and field data. Both data sources were considered to obtain LST and spectral indices and analyze vegetation response to SM during rainfall wet pulses.</p>
Full article ">Figure 3
<p>Study periods, data from the ASM OzFlux station. (<b>a</b>) Daily SM at different depths and GPP; (<b>b</b>) rainfall; (<b>c</b>) in situ SM, EVI, and LSWI values from the AHI; (<b>d</b>) maximum GPP and LSWI values in function of magnitude of wet pulses expressed as m<sup>3</sup>/m<sup>3</sup>. Vertical lines show the 5 analyzed wet pulses during late spring and summer: 2016–2017 and 2020–2021, the wettest seasons, and 2018–2019, the driest season.</p>
Full article ">Figure 3 Cont.
<p>Study periods, data from the ASM OzFlux station. (<b>a</b>) Daily SM at different depths and GPP; (<b>b</b>) rainfall; (<b>c</b>) in situ SM, EVI, and LSWI values from the AHI; (<b>d</b>) maximum GPP and LSWI values in function of magnitude of wet pulses expressed as m<sup>3</sup>/m<sup>3</sup>. Vertical lines show the 5 analyzed wet pulses during late spring and summer: 2016–2017 and 2020–2021, the wettest seasons, and 2018–2019, the driest season.</p>
Full article ">Figure 4
<p>Detailed temporal series of the EVI and LSW from the AHI and SM in ASM for each analyzed season: (<b>a</b>) 2017–2018 (normal), (<b>b</b>) 2018–2019 (moderately dry), (<b>c</b>) 2019–2020 (normal), (<b>d</b>) 2020–2021 (extremely wet). Lags between peaks in SM and spectral indices are included.</p>
Full article ">Figure 5
<p>MYD/MOD11A1, in situ daily LST and actual evapotranspiration during the study period in ASM. In situ LST was calculated from upwelling longwave radiances measured by the pyrgeometer CNR1.</p>
Full article ">Figure 6
<p>Detailed temporal series of MODIS LST, daily in situ LST and SM in ASM during the 5 analyzed seasons: (<b>a</b>) 2016–2017 (extremely wet), (<b>b</b>) 2017–2018 (normal), (<b>c</b>) 2018–2019 (moderately dry), (<b>d</b>) 2019–2020 (normal), and (<b>e</b>) 2020–2021 (extremely wet).</p>
Full article ">Figure 7
<p>Relationship between daily in situ SM at different depths, LSWI (<b>left</b>) and EVI (<b>right</b>) from AHI in ASM (<span class="html-italic">n</span> = 356).</p>
Full article ">Figure 8
<p>Relationship between average in situ LST and SM at different depths (the best correlation up to 4 days is included) in ASM (<span class="html-italic">n</span> = 475). Although correlation at 100 cm depth is included, most of the time LST fluctuates according to shallower SM.</p>
Full article ">Figure 9
<p>Relationship between daily in situ SM at different depths, MOD11A1 (<span class="html-italic">n</span> = 287) and MYD11A1 (n = 274).</p>
Full article ">Figure 10
<p>Relationship between daily in situ GPP (gC/m<sup>2</sup>), LSWI (<b>left</b>), and EVI (<b>right</b>) values from the AHI in ASM (n = 145). Note that 2018–2019 was not included, as there was no evident growing season.</p>
Full article ">Figure 11
<p>Relationship between daily in situ GPP (gC/m<sup>2</sup>), MOD LST (<b>left</b>), and MYD LST (<b>right</b>) in ASM (n = 232).</p>
Full article ">Figure 12
<p>Conceptual model of satellite-derived LST from MODIS (<b>left</b>) and daily in situ GPP (<b>right</b>) as a function of daily SM for the Mulga woodland area. For the GPP plot, the maximum GPP values and average of the maximum values of SM in the soil profile for each analyzed pulse were considered. GPP versus maximum LSWI from the AHI is included. Note that on the left plot, SM values correspond to ASM data (a similar pattern was observed in Ti Tree).</p>
Full article ">Figure A1
<p>Study periods: data from the Ti Tree OzFlux station. (<b>a</b>) Daily SM at different depths and GPP; (<b>b</b>) rainfall; (<b>c</b>) in situ SM, EVI, and LSWI values from the AHI; (<b>d</b>) maximum GPP and LSWI values in function of magnitude of wet pulses expressed as m<sup>3</sup>/m<sup>3</sup>. Vertical lines show the analyzed wet pulses during late spring and summer. SM at a 60 cm depth was considered under spinifex, given the lack of data under Mulga.</p>
Full article ">Figure A2
<p>MYD/MOD11A1, in situ daily LST and actual evapotranspiration during the study period in the Ti Tree station. In situ LST was calculated from upwelling longwave radiances measured by the pyrgeometer CNR1.</p>
Full article ">Figure A3
<p>Relationship between daily in situ SM at different depths: the LSWI (<b>left</b>) and EVI (<b>right</b>) from the AHI in the Ti Tree station (n = 236).</p>
Full article ">Figure A4
<p>Relationship between average in situ LST and SM at different depths (the best correlation up to 4 days is included) in the Ti Tree station (n = 377).</p>
Full article ">Figure A5
<p>Relationship between daily in situ SM at different depths: MOD11A1 (<b>left</b>, n = 251) and MYD11A1 (<b>right</b>, n = 239) in the Ti Tree station.</p>
Full article ">Figure A6
<p>Relationship between daily in situ GPP (gC/m<sup>2</sup>), LSWI (<b>left</b>), and EVI (<b>right</b>) values from the AHI in the Ti Tree station (n = 51). Note that 2018–2019 was not included, as there was no evident growing season.</p>
Full article ">Figure A7
<p>Relationship between daily in situ GPP (gC/m<sup>2</sup>), MOD LST (<b>left</b>, n = 177), and MYD LST (<b>right</b>, n = 169) in the Ti Tree station.</p>
Full article ">
19 pages, 7468 KiB  
Article
Spatial–Temporal Changes in Air Pollutants in Four Provinces of Sumatra Island, Indonesia: Insights from Sentinel-5P Satellite Imagery
by Zarah Arwieny Hanami, Muhammad Amin, Muralia Hustim, Rahmi Mulia Putri, Sayed Esmatullah Torabi, Andi Annisa Tenri Ramadhani and Isra Suryati
Urban Sci. 2025, 9(2), 42; https://doi.org/10.3390/urbansci9020042 - 12 Feb 2025
Viewed by 455
Abstract
This study examined spatial–temporal variations in air pollutant levels across four provinces on Sumatra Island, Indonesia, utilizing data from the Sentinel-5P satellite equipped with TROPOMI and MODIS aboard NASA’s Terra and Aqua satellites from 2019 to 2021. Sentinel-5P data, with a spatial resolution [...] Read more.
This study examined spatial–temporal variations in air pollutant levels across four provinces on Sumatra Island, Indonesia, utilizing data from the Sentinel-5P satellite equipped with TROPOMI and MODIS aboard NASA’s Terra and Aqua satellites from 2019 to 2021. Sentinel-5P data, with a spatial resolution of 3.5 × 5.5 km2 and near-daily temporal coverage, were used to analyze the nitrogen dioxide (NO2), carbon monoxide (CO), and Aerosol Optical Depth (AOD) in North Sumatra, West Sumatra, Jambi, and Riau—regions selected for their distinct industrial, agricultural, and urban characteristics. The purpose of this study was to investigate seasonal trends, regional differences, and the impact of the COVID-19 pandemic on air pollution, aiming to provide insights for improved air quality management and policy development. The satellite data were validated using zonal statistics to ensure consistency and reliability. The findings revealed significant seasonal fluctuations in pollution, with elevated levels during the dry season, primarily due to land clearing and forest fires. Urban and industrial areas such as Medan, Pekanbaru, Jambi, and Padang consistently exhibited high levels of NO2, primarily due to vehicular and industrial emissions. The regions affected by biomass burning and agriculture, particularly Jambi and Riau, displayed notably higher CO and AOD levels during the dry season. The COVID-19 pandemic provided a unique opportunity to observe potential improvements in air quality, with significant reductions in NO2, CO, and AOD levels during the 2020 lockdowns. The NO2 levels in urban centers decreased by over 20%, while the reductions in CO and AOD reached up to 29% and 64%, respectively, reflecting diminished human activities and biomass burning. This study underscores the need for enhanced air quality monitoring and targeted management strategies in Sumatra, Indonesia. Future research should aim to improve the resolution and validation of data with ground-based measurements and broaden the number of pollutants studied to better understand air quality dynamics and support effective policy development. Full article
Show Figures

Figure 1

Figure 1
<p>Maps of (<b>a</b>) Indonesian (<b>b</b>) study area in four provinces (<b>c</b>) and the 65 cities or districts in those four provinces. (<b>d</b>) West Sumatra Province, Indonesia. The numbers indicate city or district codes, and the arrows represent zoomed-in views of the study area.</p>
Full article ">Figure 2
<p>Flowchart of methodological phases for air pollution analysis on Sumatra Island, Indonesia.</p>
Full article ">Figure 3
<p>Annual average NO<sub>2</sub> concentration in (<b>a</b>) Jambi (<b>b</b>) Riau (<b>c</b>) West Sumatra, and (<b>d</b>) North Sumatra Provinces in Indonesia from 2019 to 2021.</p>
Full article ">Figure 4
<p>Annual average CO concentration in (<b>a</b>) Jambi (<b>b</b>) Riau (<b>c</b>) West Sumatra, and (<b>d</b>) North Sumatra Provinces in Indonesia from 2019 to 2021.</p>
Full article ">Figure 5
<p>Annual average AOD concentration in (<b>a</b>) Jambi (<b>b</b>) Riau (<b>c</b>) West Sumatra, and (<b>d</b>) North Sumatra Provinces in Indonesia from 2019 to 2021.</p>
Full article ">Figure 6
<p>Land use and landcover overlap with the hotspot maps during (<b>a</b>) Rainy, 2019; (<b>b</b>) Rainy 2020; (<b>c</b>) Rainy 2021; (<b>d</b>) Dry, 2019; (<b>e</b>) Dry, 2020; and (<b>f</b>) Dry, 2021 on Sumatra Island, Indonesia.</p>
Full article ">Figure 7
<p>Spatial distribution of NO<sub>2</sub> in four provinces of Sumatra Island, Indonesia: (<b>a</b>) Rainy, 2019; (<b>b</b>) Rainy 2020; (<b>c</b>) Rainy 2021; (<b>d</b>) Dry, 2019; (<b>e</b>) Dry, 2020; and (<b>f</b>) Dry, 2021.</p>
Full article ">Figure 8
<p>Spatial distribution of CO in four provinces of Sumatra Island, Indonesia: (<b>a</b>) Rainy, 2019; (<b>b</b>) Rainy 2020; (<b>c</b>) Rainy 2021; (<b>d</b>) Dry, 2019; (<b>e</b>) Dry, 2020; and (<b>f</b>) Dry, 2021.</p>
Full article ">Figure 9
<p>Spatial distribution of AOD in four provinces of Sumatra Island, Indonesia: (<b>a</b>) Rainy, 2019; (<b>b</b>) Rainy 2020; (<b>c</b>) Rainy 2021; (<b>d</b>) Dry, 2019; (<b>e</b>) Dry, 2020; and (<b>f</b>) Dry, 2021.</p>
Full article ">
33 pages, 13410 KiB  
Article
Near-Time Measurement of Aerosol Optical Depth and Black Carbon Concentration at Socheongcho Ocean Research Station: Aerosol Episode Case Analysis
by Soi Ahn, Meehye Lee, Hyeon-Su Kim, Eun-ha Sohn and Jin-Yong Jeong
Remote Sens. 2025, 17(3), 382; https://doi.org/10.3390/rs17030382 - 23 Jan 2025
Viewed by 561
Abstract
This study examined the seasonal variations and influencing factors for black carbon (BC) concentrations and aerosol optical depth (AOD) at the Socheongcho Ocean Research Station (SORS) on the Korean Peninsula from July 2019 to December 2020. An AOD algorithm was developed and validated [...] Read more.
This study examined the seasonal variations and influencing factors for black carbon (BC) concentrations and aerosol optical depth (AOD) at the Socheongcho Ocean Research Station (SORS) on the Korean Peninsula from July 2019 to December 2020. An AOD algorithm was developed and validated using the Geo-KOMPSAT-2A (GK-2A) satellite. The GK-2A AOD demonstrated comparable performance to that of Low Earth Orbit satellites, including the Terra/MODIS (R2 = 0.86), Aqua/MODIS (R2 = 0.83), and AERONET AODs (R2 = 0.85). Multi-angle absorption photometry revealed that seasonal average BC concentrations were the highest in winter (0.91 ± 0.80 µg·m−3), followed by fall (0.80 ± 0.66 µg·m−3), wet summer (0.75 ± 0.55 µg·m−3), and dry summer (0.52 ± 0.20 µg·m−3). The seasonal average GK-2A AOD was higher in wet summer (0.45 ± 0.37 µg·m−3) than in winter. The effects of meteorological parameters, AERONET AOD wavelength, and gaseous substances on GK-2A AOD and BC were investigated. The SHapley Additive exPlanations-based feature importance analysis for GK-2A AOD identified temperature, relative humidity (RH), and evaporation as major contributors. BC concentrations were increased, along with PM2.5 and CO levels, due to the effects of combustion processes during fall and winter. Analysis of high-aerosol-loading cases revealed an increase in the fine-mode fraction, emphasizing the meteorological effects on GK-2A AOD. Thus, long-range transport and local BC sources played a critical role at the SORS. Full article
(This article belongs to the Special Issue Air Quality Mapping via Satellite Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The location of the Socheongcho Ocean Research Station (SORS), marked with a red star, is shown in the Yellow Sea.</p>
Full article ">Figure 2
<p>Flow chart of Geo-KOMPSAT-2A Advanced Meteorological Imager (GK-2A) aerosol optical depth (AOD) algorithm.</p>
Full article ">Figure 3
<p>Aerosol optical depth (AOD) scatterplot of (<b>a</b>) GK-2A/AMI, (<b>b</b>) Terra/MODIS, and (<b>c</b>) Aqua/MODIS against ground-based reference AERONET data for 40 sites from July 2019 to December 2020. Dotted line shows linear regression, and black line is 1:1 line.</p>
Full article ">Figure 4
<p>Seasonal scatterplots of GK-2A/AMI aerosol optical depth (AOD) against ground-based reference AERONET data for 40 sites from July 2019 to December 2020. Dotted line shows linear regression, and black line is 1:1 line.</p>
Full article ">Figure 5
<p>Seasonal (<b>a</b>,<b>b</b>) and monthly (<b>c</b>,<b>d</b>) variations in (a,<b>c</b>) GK-2A aerosol optical depth (AOD) and (<b>b</b>,<b>d</b>) GK-2A Angstrom Exponent (AE) at Socheongcho Ocean Research Station (SORS).</p>
Full article ">Figure 6
<p>Seasonal (<b>a</b>) and monthly (<b>b</b>) variation in BC at the Socheongcho Ocean Research Station (SORS).</p>
Full article ">Figure 7
<p>Pollutant distribution by wind direction and speed in Socheongcho Ocean Research Station (SORS): (<b>a</b>) GK-2A aerosol optical depth (AOD) and (<b>b</b>) BC. Seasonal distribution of (<b>c</b>) GK-2A AODs and (<b>d</b>) black carbon (BC).</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>h</b>) Weighted potential source contribution function (WPSCF) results of 3-day back-trajectory of HYSPLIT Model for GK-2A aerosol optical depth (AOD) &gt; 0.5 at Socheongcho Ocean Research Station (SORS) (standard).</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>h</b>) Weighted potential source contribution function (WPSCF) result of 3-day back-trajectory of HYSPLIT Model for GK-2A aerosol optical depth (AOD) &gt; 1.0 at Socheongcho Ocean Research Station (SORS) (standard).</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>h</b>) Weighted potential source contribution function (WPSCF) result of 3-day back-trajectory of HYSPLIT Model for BC &gt; 1.0 μg m<sup>−3</sup> at Socheongcho Ocean Research Station (SORS) (standard).</p>
Full article ">Figure 11
<p>(<b>a</b>–<b>h</b>) Weighted potential source contribution function (WPSCF) result of 3-day back-trajectory of HYSPLIT Model for BC &gt; 2.0 μg m<sup>−3</sup> at Socheongcho Ocean Research Station (SORS) (standard).</p>
Full article ">Figure 12
<p>Time series distribution of BC (blue circle) and aerosol optical depth (AOD) from ground measurements (AERONET: red triangle) and satellites (GK-2A/AMI: pink circle, Terra/MODIS: emerald star, Aqua/MODIS: yellow circle) from July 2019 to December 2020.</p>
Full article ">Figure 13
<p>Scatter plot of GK-2A aerosol optical depth (AOD) and BC at Socheongcho Ocean Research Station (SORS) by season. Green triangle: wet summer (July, August), red square: fall (September, October), lavender triangle: winter (December, January, February), green rhombus: spring (March, April), and blue circle: dry summer (May, June).</p>
Full article ">Figure 14
<p>Time series of aerosol episode on 12–15 July 2019 and images of high-aerosol event obtained on 04:00 UTC, 14 July 2019, using (<b>a</b>) GK-2A True = RGB, (<b>b</b>) GK-2A/AMI aerosol optical depth (AOD), (<b>c</b>) Terra/MODIS AOD, (<b>d</b>) Aqua/MODIS AOD, (<b>e</b>) Suomi-NPP/VIIRS AOD, and (<b>f</b>) Suomi-NPP/VIIRS FMF.</p>
Full article ">Figure 15
<p>Time series of aerosol episode on 30 October–2 November 2019 and images of fall aerosol event obtained at 04:00 UTC on 30 October 2019, using (<b>a</b>) GK-2A True RGB, (<b>b</b>) GK-2A/AMI AOD, (<b>c</b>) Terra/MODIS AOD, (<b>d</b>) Aqua/MODIS AOD, (<b>e</b>) Suomi-NPP/VIIRS AOD, and (<b>f</b>) Suomi-NPP/VIIRS FMF.</p>
Full article ">Figure 16
<p>Time series of aerosol episode on 7–10 December 2019 and images of winter aerosol event obtained at 04:00 UTC, on 9 December 2019, using (<b>a</b>) GK-2A True RGB, (<b>b</b>) GK-2A/AMI aerosol optical depth (AOD), (<b>c</b>) Terra/MODIS AOD, (<b>d</b>) Aqua/MODIS AOD, (<b>e</b>) Suomi-NPP/VIIRS AOD, and (<b>f</b>) Suomi-NPP/VIIRS FMF.</p>
Full article ">Figure 17
<p>SHapley Additive exPlanations-based feature importance results for (<b>a</b>) GK-2A aerosol optical depth (AOD) and (<b>b</b>) black carbon (BC) during analysis period.</p>
Full article ">Figure 18
<p>Results for PHEATMAP-based feature importance analysis by (<b>a</b>) season and (<b>b</b>) month.</p>
Full article ">
20 pages, 18304 KiB  
Article
Assessment of Radiometric Calibration Consistency of Thermal Emissive Bands Between Terra and Aqua Moderate-Resolution Imaging Spectroradiometers
by Tiejun Chang, Xiaoxiong Xiong, Carlos Perez Diaz, Aisheng Wu and Hanzhi Lin
Remote Sens. 2025, 17(2), 182; https://doi.org/10.3390/rs17020182 - 7 Jan 2025
Viewed by 461
Abstract
Moderate-Resolution Imaging Spectroradiometer (MODIS) sensors onboard the Terra and Aqua spacecraft have been in orbit for over 24 and 22 years, respectively, providing continuous observations of the Earth’s surface. Among the instrument’s 36 bands, 16 of them are thermal emissive bands (TEBs) with [...] Read more.
Moderate-Resolution Imaging Spectroradiometer (MODIS) sensors onboard the Terra and Aqua spacecraft have been in orbit for over 24 and 22 years, respectively, providing continuous observations of the Earth’s surface. Among the instrument’s 36 bands, 16 of them are thermal emissive bands (TEBs) with wavelengths that range from 3.75 to 14.24 μm. Routine post-launch calibrations are performed using the sensor’s onboard blackbody and space view port, the moon, and vicarious targets that include the ocean, Dome Concordia (Dome C) in Antarctica, and quasi-deep convective clouds (DCC). The calibration consistency between the satellite measurements from the two instruments is essential in generating a multi-year data record for the long-term monitoring of the Earth’s Level 1B (L1B) data. This paper presents the Terra and Aqua MODIS TEB comparison for the upcoming Collection 7 (C7) L1B products using measurements over Dome C and the ocean, as well as the double difference via simultaneous nadir overpasses with the Infrared Atmospheric Sounding Interferometer (IASI) sensor. The mission-long trending of the Terra and Aqua MODIS TEB is presented, and their cross-comparison is also presented and discussed. Results show that the calibration of the two MODIS sensors and their respective Earth measurements are generally consistent and within their design specifications. Due to the electronic crosstalk contamination, the PV LWIR bands show slightly larger drifts for both MODIS instruments across different Earth measurements. These drifts also have an impact on the Terra-to-Aqua calibration consistency. This thorough assessment serves as a robust record containing a summary of the MODIS calibration performance and the consistency between the two MODIS sensors over Earth view retrievals. Full article
Show Figures

Figure 1

Figure 1
<p>MODTRAN profile over an ocean scene simulated using MODIS Atmospheric Profile product as input. MODIS RSR are superimposed over the MODTRAN simulation.</p>
Full article ">Figure 2
<p>Aqua (<b>left</b>) and Terra (<b>right</b>) brightness temperature series over Dome C for MODIS C7 bands 20, 25, 29, 30, 31, and 33. All bands are referenced to AWS. Results are monthly averaged.</p>
Full article ">Figure 3
<p>Aqua minus Terra brightness temperature series over Dome C for MODIS C7 bands 20, 25, 29, 30, 31, and 33. All bands are referenced to AWS. Red dashed horizontal line defines average Aqua minus Terra BT differences. Results are monthly averaged.</p>
Full article ">Figure 4
<p>Aqua (<b>left</b>) and Terra (<b>right</b>) brightness temperature series over ocean for MODIS C7 bands 20, 25, 29, 30, 31, and 33. All bands are normalized (BT (band) = BT (band)—BT (band 31) + avg BT (band 31)) to band 31, except for band 31. Results are monthly averaged.</p>
Full article ">Figure 5
<p>Aqua minus Terra brightness temperature series over ocean for MODIS C7 bands 20, 25, 29, 30, 31, and 33. All bands are normalized (BT (band) = BT (band)—BT (band 31) + avg BT (band 31)) to band 31, except for band 31. Red dashed horizontal line defines average Aqua minus Terra BT difference. Results are monthly averaged.</p>
Full article ">Figure 6
<p>Aqua-IASI and Terra-IASI BT difference time series for MODIS sample bands 20, 25, 29, 30, 31, and 33. Average value for every SNO crossover between MODIS and IASI shown. Empty markers represent difference with IASI-A, while filled makers are used to denote IASI-C.</p>
Full article ">Figure 7
<p>Aqua-IASI minus Terra-IASI BT difference for MODIS sample bands 20, 25, 29, 30, 31, and 33. Red dashed horizontal line defines average Aqua-IASI minus Terra-IASI BT difference. Results are monthly averaged.</p>
Full article ">Figure 8
<p>Terra-IASI BT difference as a function of Terra MODIS BT for MODIS sample bands 20, 25, 29, 30, 31, and 33.</p>
Full article ">Figure 9
<p>Aqua-IASI BT difference as a function of Aqua MODIS BT for MODIS sample bands 20, 25, 29, 30, 31, and 33.</p>
Full article ">
19 pages, 7527 KiB  
Article
Satellite Signatures of Pre-Seismic Atmospheric Anomalies of 6 February 2023 Türkiye Earthquakes
by Maria Zoran, Dan Savastru and Marina Tautan
Atmosphere 2024, 15(12), 1514; https://doi.org/10.3390/atmos15121514 - 18 Dec 2024
Viewed by 623
Abstract
Time series satellite data, coupled with available ground-based observations, enable geophysicists to survey earthquake precursors in areas of strong geotectonic activity. This paper is focused on pre-seismic atmospheric disturbances resulting from the stress accumulated during the seismogenic process related to the 6 February [...] Read more.
Time series satellite data, coupled with available ground-based observations, enable geophysicists to survey earthquake precursors in areas of strong geotectonic activity. This paper is focused on pre-seismic atmospheric disturbances resulting from the stress accumulated during the seismogenic process related to the 6 February 2023 Kahramanmaras doublet earthquake sequence in Türkiye. We investigated the pre- and post-seismic anomalies of multiple precursors of different spatiotemporal patterns from MODIS Terra/Aqua and NOAA-AVHRR satellite data (air temperature at 2 m height—AT, air relative humidity—RH, and air pressure—AP, surface outgoing long-wave radiation—OLR, and land surface temperature—LST). Pre-seismic recorded anomalies of AT within seven months and OLR within one month before the main shocks suggested the existence of the preparatory process of the Kahramanmaras doublet earthquake. The 8-Day LST_Day and LST_night data evidenced pre-seismic and post-seismic thermal anomalies for both the Pazarcik and Elbistan earthquakes. The results of this study highlight that the spatiotemporal evolution of earthquake precursors can be important information for updating the seismic hazard in geotectonic active areas. Full article
(This article belongs to the Special Issue Ionospheric Sounding for Identification of Pre-seismic Activity)
Show Figures

Figure 1

Figure 1
<p>Kahramanmaras earthquakes epicenters (red stars), main fault zones, plates, impacted area (dash box), and their movement direction (yellow arrows) in and around Türkiye.</p>
Full article ">Figure 2
<p>NOAA positive composite anomaly map of air temperature at 2 m height AT for Pazarcik (EQ1) and Elbistan (EQ2) earthquakes in Türkiye between July 2022 and 6 February 2023. (Red stars represent focal points of the earthquakes).</p>
Full article ">Figure 3
<p>NOAA negative composite anomaly map of air temperature at 2 m height AT after Pazarcik (EQ1) and Elbistan (EQ2) earthquakes in Türkiye during 7 February 2023 and 15 February 2023 (white stars represent focal points).</p>
Full article ">Figure 4
<p>Temporal patterns of air temperature at 2 m height AT for Pazarcik and Elbistan earthquakes (The pink arrow represents the day when the EQ1 and EQ2 happened).</p>
Full article ">Figure 5
<p>Temporal patterns of air relative humidity RH for EQ1 (Pazarcik) and EQ2 (Elbistan) earthquakes. (The pink arrow represents the day when the EQ1 and EQ2 happened).</p>
Full article ">Figure 6
<p>NOAA positive composite anomaly map of surface air relative humidity for Pazarcik (EQ1) and Elbistan (EQ2) earthquakes in Türkiye between 1 February and 6 February 2023.</p>
Full article ">Figure 7
<p>Temporal patterns of surface air pressure for Pazarcik and Elbistan earthquakes. (The pink arrows represent the day when the EQ1 and EQ2 occurred).</p>
Full article ">Figure 8
<p>NOAA negative composite anomaly map of surface air pressure AP during pre-earthquake period 1 February–6 February 2023 over Pazarcik (EQ1) and Elbistan (EQ2) areas in Türkiye.</p>
Full article ">Figure 9
<p>NOAA negative composite anomaly map of surface geopotential map over Türkiye and Pazarcik (EQ1) and Elbistan (EQ2) areas between 1 February 2023 and 6 February 2023.</p>
Full article ">Figure 10
<p>NOAA positive composite anomaly map of surface OLR over Türkiye and Pazarcik (EQ1) and Elbistan (EQ2) areas between 1 July 2022 and 6 February 2023.</p>
Full article ">Figure 11
<p>NOAA negative composite anomaly map of surface OLR centered on Pazarcik (EQ1) and Elbistan (EQ2) areas between 28 January 2023 and 31 January 2023.</p>
Full article ">Figure 12
<p>NOAA negative composite anomaly map of surface OLR centered on Pazarcik (EQ1) and Elbistan (EQ2) areas between 1 February 2023 and 6 February 2023.</p>
Full article ">Figure 13
<p>NOAA positive composite anomaly map of surface OLR centered on Pazarcik (EQ1) and Elbistan (EQ2) areas between 7 February 2023 and 28 February 2023.</p>
Full article ">Figure 14
<p>LST MODIS/Terra_8 Day_Night variation for the period 1 January to 30 March 2023 for two test areas: (<b>a</b>) EQ1 centered on Pazarcik with 7 km radius; (<b>b</b>) EQ2 centered on Elbistan with 7 km radius.</p>
Full article ">
29 pages, 11518 KiB  
Article
Evaluating the Two-Source Energy Balance Model Using MODIS Data for Estimating Evapotranspiration Time Series on a Regional Scale
by Mahsa Bozorgi, Jordi Cristóbal and Magí Pàmies-Sans
Remote Sens. 2024, 16(23), 4587; https://doi.org/10.3390/rs16234587 - 6 Dec 2024
Viewed by 845
Abstract
Estimating daily continuous evapotranspiration (ET) can significantly enhance the monitoring of crop stress and drought on regional scales, as well as benefit the design of agricultural drought early warning systems. However, there is a need to verify the models’ performance in estimating the [...] Read more.
Estimating daily continuous evapotranspiration (ET) can significantly enhance the monitoring of crop stress and drought on regional scales, as well as benefit the design of agricultural drought early warning systems. However, there is a need to verify the models’ performance in estimating the spatiotemporal continuity of long-term daily evapotranspiration (ETd) on regional scales due to uncertainties in satellite measurements. In this study, a thermal-based two-surface energy balance (TSEB) model was used concurrently with Terra/Aqua MODIS data and the ERA5 atmospheric reanalysis dataset to calculate the surface energy balance of the soil–canopy–atmosphere continuum and estimate ET at a 1 km spatial resolution from 2000 to 2022. The performance of the model was evaluated using 11 eddy covariance flux towers in various land cover types (i.e., savannas, woody savannas, croplands, evergreen broadleaf forests, and open shrublands), correcting for the energy balance closure (EBC). The Bowen ratio (BR) and residual (RES) methods were used for enforcing the EBC in the EC observations. The modeled ET was evaluated against unclosed ET and closed ET (ETBR and ETRES) under clear-sky and all-sky observations as well as gap-filled data. The results showed that the modeled ET presented a better agreement with closed ET compared to unclosed ET in both Terra and Aqua datasets. Additionally, although the model overestimated ETd across all different land cover types, it successfully captured the spatiotemporal variability in ET. After the gap-filling, the total number of days compared with flux measurements increased substantially, from 13,761 to 19,265 for Terra and from 13,329 to 19,265 for Aqua. The overall mean results including clear-sky and all-sky observations as well as gap-filled data with the Aqua dataset showed the lowest errors with ETRES, by a mean bias error (MBE) of 0.96 mm.day−1, an average mean root square (RMSE) of 1.47 mm.day−1, and a correlation (r) value of 0.51. The equivalent figures for Terra were about 1.06 mm.day−1, 1.60 mm.day−1, and 0.52. Additionally, the result from the gap-filling model indicated small changes compared with the all-sky observations, which demonstrated that the modeling framework remained robust, even with the expanded days. Hence, the presented modeling framework can serve as a pathway for estimating daily remote sensing-based ET on regional scales. Furthermore, in terms of temporal trends, the intra-annual and inter-annual variability in ET can be used as indicators for monitoring crop stress and drought. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area and the 11 selected flux towers. Projection system in UTM-30N WGS-84.</p>
Full article ">Figure 2
<p>Flowchart of the modeling framework estimating ET<sub>d</sub> time series. The orange rectangle denotes the pre-processing of MODIS vegetation indices through TIMESAT.</p>
Full article ">Figure 3
<p>TSEB modelling scheme (adapted from [<a href="#B61-remotesensing-16-04587" class="html-bibr">61</a>]).</p>
Full article ">Figure 4
<p>Scatterplots of the EBC calculated using the EC with the statistical metrics for the entire study period (<b>top left</b>) and the days compared with the mode (<b>bottom right</b>) at each flux tower.</p>
Full article ">Figure 5
<p>Temporal variations in the modeled ET<sub>d</sub> estimated using Terra and Aqua datasets compared to in situ-measured ET at flux towers.</p>
Full article ">Figure 6
<p>Scatterplots of the modeled ET<sub>d</sub> estimated using Terra against unclosed and closed ET (ET<sub>BR</sub> and ET<sub>RES</sub>).</p>
Full article ">Figure 7
<p>Scatterplots of the modeled ET<sub>d</sub> estimated using Aqua against unclosed and closed ET (ET<sub>BR</sub> and ET<sub>RES</sub>).</p>
Full article ">Figure 8
<p>Scatterplots of the LST obtained from MODSI Terra dataset and the LST calculated using half-hourly flux tower data at satellite overpass time.</p>
Full article ">Figure 9
<p>Scatterplots of the LST obtained from MODSI Aqua dataset and the LST calculated using half-hourly flux tower data at satellite overpass time.</p>
Full article ">Figure 10
<p>Mean monthly variability in estimated ET using Terra dataset.</p>
Full article ">Figure 11
<p>Mean monthly variability in estimated ET using Aqua dataset.</p>
Full article ">Figure 12
<p>Seasonal variation in estimated ET from 2000 to 2022 in the study area using Terra dataset.</p>
Full article ">Figure 13
<p>Seasonal variation in estimated ET from 2002 to 2022 in the study area using Aqua dataset.</p>
Full article ">Figure 14
<p>Temporal variation in annual cumulative of estimated <span class="html-italic">ET</span> from 2000 to 2022 in the study area using Aqua (<b>left</b>) and Terra (<b>right</b>) dataset.</p>
Full article ">Figure 15
<p>Spatiotemporal variability in annual cumulative of estimated ET from 2000 to 2022 in the study area using Terra dataset.</p>
Full article ">Figure 16
<p>Spatiotemporal variability in annual cumulative of estimated ET from 2000 to 2022 in the study area using Aqua dataset.</p>
Full article ">
17 pages, 12533 KiB  
Article
Potential Impact of Sea Surface Temperature Variability on the 2007 Sudden Bloom of Ulva prolifera in the Southern Yellow Sea
by Yufeng Pan, Pin Li, Jiaxuan Sun, Siyu Liu, Lvyang Xing, Di Yu and Qi Feng
Remote Sens. 2024, 16(23), 4407; https://doi.org/10.3390/rs16234407 - 25 Nov 2024
Viewed by 645
Abstract
Since 2007, Ulva prolifera (U. prolifera) originating in northern Jiangsu (NJ) has consistently expanded to the southern coast of the Shandong Peninsula. However, the underlying reasons for the 2007 sudden bloom of U. prolifera on a large scale remain unknown. This [...] Read more.
Since 2007, Ulva prolifera (U. prolifera) originating in northern Jiangsu (NJ) has consistently expanded to the southern coast of the Shandong Peninsula. However, the underlying reasons for the 2007 sudden bloom of U. prolifera on a large scale remain unknown. This study uses remote sensing data from MODIS/AQUA spanning the period 2003–2022 to investigate the sea surface temperature (SST) structure changes in the southern Yellow Sea (SYS) over the past 20 years. The results demonstrate the following. (1) Since 2007, the NJ northward current and the Yangtze estuary warm current have exhibited higher temperatures, earlier northward intrusions, and larger influence areas, leading to a faster warming rate in NJ before mid-May. This rapid increase in SST to a level suitable for early U. prolifera growth triggers large-scale blooms. (2) The change in temperature structure is primarily induced by a prolonged and intense La Niña event in 2007–2008. However, since 2016, under stable global climate conditions, the temperature structure of the SYS has returned to the pre-2007 state, corresponding to a decrease in the scale of U. prolifera blooms. Full article
(This article belongs to the Special Issue Satellite Remote Sensing for Ocean and Coastal Environment Monitoring)
Show Figures

Figure 1

Figure 1
<p>Schematic map of the study area. The black single-arrow dashed curves depict the primary flow paths, as primarily determined by previous studies [<a href="#B27-remotesensing-16-04407" class="html-bibr">27</a>,<a href="#B28-remotesensing-16-04407" class="html-bibr">28</a>,<a href="#B29-remotesensing-16-04407" class="html-bibr">29</a>] and the temperature structure in Chapter 4.2. The red and blue squares indicate the characteristic regions associated with each respective current. The yellow dashed square is the tongue-shaped topography area where <span class="html-italic">Ulva prolifera</span> (<span class="html-italic">U. prolifera</span>) originates. The visual representation is rendered using the Surfer16 software.</p>
Full article ">Figure 2
<p>(<b>a</b>) Distribution of in-situ data survey stations. (<b>b</b>) Correlation analysis between the sea surface temperature (SST) derived from MODIS/AQUA satellite imagery and in-situ data. (<b>c</b>) Correlation analysis between the suspended sediment concentration (SSC) derived from MODIS/AQUA satellite imagery and in-situ data. (<b>d</b>) Standardized calculation for the classification of the <span class="html-italic">U. prolifera</span>.</p>
Full article ">Figure 3
<p>Diagram illustrating nutrient concentration and Suspended Sediment Concentration (SSC) from different months. Blue box indicates the period of the sudden large-scale bloom of <span class="html-italic">U. prolifera</span> in 2007–2008. (<b>a</b>) April. (<b>b</b>) May. (<b>c</b>) June.</p>
Full article ">Figure 4
<p>Spatial pattern of the Pearson product-moment correlation coefficient (PPMCC) between SST and the coverage area of <span class="html-italic">U. prolifera</span> from mid-May to late July.</p>
Full article ">Figure 5
<p>Spatial pattern of SST structure in the Southern Yellow Sea (SYS) from April–July. The black single-arrow dashed curves represent the primary current paths within the SST structure spatial pattern.</p>
Full article ">Figure 6
<p>Spatial pattern of SST structure characteristics. Case 1: Period from 2003–2006, characterized by the absence of <span class="html-italic">U. prolifera</span> blooms. Case 2: The years of 2012 and 2020, characterized by small-scale blooms of <span class="html-italic">U. prolifera</span>. Case 3: The years of 2009 and 2021, characterized by large-scale blooms of <span class="html-italic">U. prolifera</span>. The warming rate anomaly (WRA) is calculated by subtracting the warming rate (WR) index from early April to mid-May from that of mid-May to late June for each respective case.</p>
Full article ">Figure 7
<p>Black curve represents the continuous change in the three-month moving average of the NINO3.4 index from 2003–2022, which is an indicator of El Niño-Southern Oscillation (ENSO) activity. Green curve represents the maximum coverage area (MCA) of U. prolifera, indicating the extent of U. prolifera blooms. Other curves represent the WRA from early April to mid-May and from mid-May to late June of each year, respectively, as calculated using (3). Blue box indicates an El Niño event, red box indicates a La Niña event.</p>
Full article ">
24 pages, 6494 KiB  
Article
Reconstruction of Fine-Spatial-Resolution FY-3D-Based Vegetation Indices to Achieve Farmland-Scale Winter Wheat Yield Estimation via Fusion with Sentinel-2 Data
by Xijia Zhou, Tao Wang, Wei Zheng, Mingwei Zhang and Yuanyuan Wang
Remote Sens. 2024, 16(22), 4143; https://doi.org/10.3390/rs16224143 - 6 Nov 2024
Viewed by 906
Abstract
The spatial resolution (250–1000 m) of the FY-3D MERSI is too coarse for agricultural monitoring at the farmland scale (20–30 m). To achieve the winter wheat yield (WWY) at the farmland scale, based on FY-3D, a method framework is developed in this work. [...] Read more.
The spatial resolution (250–1000 m) of the FY-3D MERSI is too coarse for agricultural monitoring at the farmland scale (20–30 m). To achieve the winter wheat yield (WWY) at the farmland scale, based on FY-3D, a method framework is developed in this work. The enhanced deep convolutional spatiotemporal fusion network (EDCSTFN) was used to perform a spatiotemporal fusion on the 10 day interval FY-3D and Sentinel-2 vegetation indices (VIs), which were compared with the enhanced spatial and temporal adaptive reflectance fusion model (ESTARFM). In addition, a BP neural network was built to calculate the farmland-scale WWY based on the fused VIs, and the Aqua MODIS gross primary productivity product was used as ancillary data for WWY estimation. The results reveal that both the EDCSTFN and ESTARFM achieve satisfactory precision in the fusion of the Sentinel-2 and FY-3D VIs; however, when the period of spatiotemporal data fusion is relatively long, the EDCSTFN can achieve greater precision than ESTARFM. Finally, the WWY estimation results based on the fused VIs show remarkable correlations with the WWY data at the county scale and provide abundant spatial distribution details about the WWY, displaying great potential for accurate farmland-scale WWY estimations based on reconstructed fine-spatial-temporal-resolution FY-3D data. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study region: (<b>a</b>) location of the Weihe Plain; (<b>b</b>) FY-3D false colour composite image for 3 May 2020; and (<b>c</b>) locations of the county-scale WWY data points used in the WWY estimation.</p>
Full article ">Figure 2
<p>Flowchart of the 10 day interval VI imagery reconstruction and farmland-scale WWY estimation.</p>
Full article ">Figure 3
<p>Flowchart of farmland-scale WWY estimation: (<b>a</b>) Y estimation model based on the cumulative GPP; and (<b>b</b>) farmland-scale Y estimation model based on multiple parameters.</p>
Full article ">Figure 4
<p>Results of the consistency analysis of the Sentinel-2 and FY-3D VIs: (<b>a</b>) R<sup>2</sup> values between the aggregated Sentinel-2 VI imagery and the FY-3D VI imagery at an SR of 250 m; and (<b>b</b>) average deviations and RMSE values of the fitting results between the aggregated Sentinel-2 VI imagery and FY-3D VI imagery. The error line in (<b>b</b>) denotes the RMSE of the fitting results.</p>
Full article ">Figure 5
<p>EVI at each WW growing stage from 2020 to 2022.</p>
Full article ">Figure 6
<p>Y estimation model based on the cumulative GPP for the main WW growing period and the Y estimation precision evaluation results in 2020 and 2021. The dotted lines in the figures denote the fitted linear functions, which are close to the diagonal solid lines, indicating that the systematic deviation in the Y estimation results is small. (<b>a</b>) Linear regression model between the cumulative GPP data for the main WW growing period and the county-scale WWY from 2014 to 2018, (<b>b</b>) linear regression results between the WWY estimation results from 2020 based on the cumulative GPP and county-scale Y statistical data, and (<b>c</b>) linear regression results between the WWY estimation results in 2021 based on the cumulative GPP and county-scale Y statistical data.</p>
Full article ">Figure 7
<p>WWY estimation results for 2020 to 2022 based on the MODIS cumulative GPP data.</p>
Full article ">Figure 8
<p>Farmland-scale WWY estimation results for the Weihe Plain from 2020 to 2022 based on multiple parameters.</p>
Full article ">Figure 9
<p>Linear regression results between the farmland-scale WWY estimation results and the Y statistical data in 2020 and 2021. The dotted lines in the figures denote the fitted linear functions, which are close to the diagonal solid lines, indicating that the systematic deviation of the Y estimation results is small.</p>
Full article ">
28 pages, 14472 KiB  
Article
Characteristics of R2019 Processing of MODIS Sea Surface Temperature at High Latitudes
by Chong Jia, Peter J. Minnett and Malgorzata Szczodrak
Remote Sens. 2024, 16(21), 4102; https://doi.org/10.3390/rs16214102 - 2 Nov 2024
Cited by 1 | Viewed by 673
Abstract
Satellite remote sensing is the best way to derive sea surface skin temperature (SSTskin) in the Arctic. However, as surface temperature retrieval algorithms in the infrared (IR) part of the electromagnetic spectrum are designed to compensate for atmospheric effects mainly due [...] Read more.
Satellite remote sensing is the best way to derive sea surface skin temperature (SSTskin) in the Arctic. However, as surface temperature retrieval algorithms in the infrared (IR) part of the electromagnetic spectrum are designed to compensate for atmospheric effects mainly due to water vapor, MODIS SSTskin retrievals have larger uncertainties at high latitudes where the atmosphere is very dry and cold, which is an extreme in the distribution of global conditions. MODIS R2019 SSTskin fields are currently derived using latitudinally and monthly dependent algorithm coefficients, including an additional band above 60°N to better represent the effects of Arctic atmospheres. However, the R2019 processing of MODIS SSTskin still has some unrevealed error characteristics. This study uses 21 years (2002–2022) of collocated, simultaneous satellite brightness temperature (BT) data from Aqua MODIS and in situ buoy-measured subsurface temperature data from iQuam for validation. Unlike elsewhere over the oceans, the 11 μm and 12 μm BT differences are poorly related to the column water vapor at high latitudes, resulting in poor atmospheric water vapor correction. Anomalous BT difference signals are identified, caused by the temperature and humidity inversions in the lower troposphere, which are especially significant during the summer. Although the existence of negative BT differences is physically reasonable, this makes the retrieval algorithm lose its effectiveness. Moreover, the statistics of the MODIS SSTskin data when compared with the iQuam buoy temperature data show large differences (in terms of mean and standard deviation) for the matchups at the Northern Atlantic and Pacific sides of the Arctic due to the disparity of in situ measurements and distinct surface and vertical atmospheric conditions. Therefore, it is necessary to further improve the retrieval algorithms to obtain more accurate MODIS SSTskin data to study surface ocean processes and climate change in the Arctic. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Distribution of global drifters, moored buoys, and some other platforms in June 2024 from <a href="https://www.ocean-ops.org/dbcp/network/maps.html" target="_blank">https://www.ocean-ops.org/dbcp/network/maps.html</a> (accessed on 2 September 2024).</p>
Full article ">Figure 2
<p>(<b>a</b>) Time series of the Aqua MODIS-iQuam SST difference (ΔSST) above 60°N from 2002 to 2022, including both QL = 0 and QL = 1 data. The black dashed horizontal line indicates the value of −0.17 K. (<b>b</b>) Histogram of the ΔSST with fitted normal distribution as blue dashed curve. (<b>c</b>) Scatter plot of Aqua MODIS SST<sub>skin</sub> and in situ buoy SST measurements from iQuam colored according to data density. (<b>d</b>) Map of Aqua MODIS ΔSST. (<b>e</b>) Map of Aqua MODIS ΔSST &gt; 2 K or &lt;−2 K.</p>
Full article ">Figure 3
<p>Aqua MODIS ΔSST as a function of time difference between the satellite and in situ buoy measurements for the (<b>a</b>) daytime and (<b>b</b>) nighttime data. The scatter plots are colored according to the data density. The dashed black horizontal lines are at −0.17 K, and the linear regressions are plotted as black solid lines with the functional expressions given in the plots.</p>
Full article ">Figure 4
<p>Similar to <a href="#remotesensing-16-04102-f003" class="html-fig">Figure 3</a>, but for the Aqua MODIS ΔSST as a function of spatial distance between the satellite and in situ buoy measurements.</p>
Full article ">Figure 5
<p>BT difference (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="normal">B</mi> <mi mathvariant="normal">T</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mi mathvariant="normal">B</mi> <mi mathvariant="normal">T</mi> </mrow> <mrow> <mn>12</mn> </mrow> </msub> </mrow> </semantics></math>) as a function of total column water vapor in the MUDB during the (<b>a</b>) day and (<b>b</b>) at night. The scatter plots are colored according to data density, and the black dashed lines indicate a BT difference of zero. The correlation coefficients are also given on the upper right.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>l</b>) Monthly scatter plots of Aqua MODIS ΔSST as a function of BT difference colored according to data density and fitted by linear regression in black dashed lines. (<b>m</b>) Monthly regression slope (blue) and intercept (red) variation, with the error bars indicating the uncertainties.</p>
Full article ">Figure 7
<p>(<b>a</b>) Map distribution of the anomalous BT differences in the R2019 MUDB above 60°N. Histograms of (<b>b</b>) latitude and (<b>c</b>) longitude for Aqua MODIS measurements in the MUDB when the BT differences are &gt;0.036 K (light blue) and are ≤0.036 K (light red).</p>
Full article ">Figure 8
<p>MERRA-2 reanalysis atmospheric profiles matched with Aqua MODIS R2019 MUDB north of 60°N including (<b>a</b>) air temperature and (<b>b</b>) specific humidity, plotted as the mean (dashed line) ± 1 STD (envelope) beneath the 500 hPa level for the measurements of BT difference &gt; 0.036 K (blue) and BT difference ≤ 0.036 K (red).</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>l</b>) Monthly distributions (mean ± 1 STD) of the MERRA-2 air temperature profiles beneath the 500 hPa level for the normal and abnormal BT difference cases. The number of data points in each dataset is given in the lower left section of each panel.</p>
Full article ">Figure 10
<p>Similar to <a href="#remotesensing-16-04102-f009" class="html-fig">Figure 9</a>a–1, but for the MERRA-2 specific humidity profiles beneath the 500 hPa level. The number of each dataset is given in the upper right section of each panel.</p>
Full article ">Figure 11
<p>Monthly occurrence frequency of temperature inversion (TI) only (blue), humidity inversion (HI) only (red), simultaneous TI and HI (yellow), and neither TI nor HI (green) for (<b>a</b>) BT difference &gt; 0.036 K and (<b>b</b>) BT difference ≤ 0.036 K. The highest frequency among the TI/HI situations in each month is shown in the corresponding part of the column.</p>
Full article ">Figure 12
<p>Monthly variations of (<b>a</b>) TI strength, (<b>b</b>) HI strength, and (<b>c</b>) BT difference between 11 μm and 12 μm for BT difference &gt; 0.036 K (blue) and BT difference ≤ 0.036 K (red) only when TI or HI exists. For the boxplots, the ends of the boxes, the ends of the whiskers, and the line in the box represent the 25th and 75th percentiles, the minimum and maximum values that are not outliers, and the median, respectively. Outliers are values beyond 75th percentile +1.5* (interquartile range) or less than 25th percentile −1.5* (interquartile range), which are not shown here. (<b>d</b>) Scatter plot of BT difference as a function of TI strength colored according to the HI strength. The black dotted line indicates the BT difference value of 0.036 K, the threshold of positive BT differences greater than the NEΔT effects. The red dots and error bars are the mean and STD of the BT difference, calculated at 1 K intervals of TI strength.</p>
Full article ">Figure 13
<p>Monthly boxplots of the Aqua MODIS ΔSST in the Atlantic Sector (blue) and the Pacific Sector (red) of the Arctic. Outliers are not shown.</p>
Full article ">Figure 14
<p>(<b>a</b>–<b>l</b>) Monthly distributions (mean ± 1 STD) of the MERRA-2 air temperature profiles beneath the 500 hPa level matched with the data in the MUDB at the Atlantic and Pacific Sectors of the Arctic. The number of each dataset is given in the lower left section of each panel.</p>
Full article ">Figure 15
<p>Similar to <a href="#remotesensing-16-04102-f014" class="html-fig">Figure 14</a>a–l, but for the MERRA-2 specific humidity profiles beneath the 500 hPa level. The number of each dataset is given in the upper right section of each panel.</p>
Full article ">Figure 16
<p>Monthly occurrence frequency of TI only (blue), HI only (red), simultaneous TI and HI (yellow), and neither TI nor HI (green) for the data in MUDB for the (<b>a</b>) Atlantic Sector and (<b>b</b>) Pacific Sector of the Arctic. The highest frequency among the inversion situations in each month is shown in the corresponding part of the column. Monthly variations of (<b>c</b>) TI strength and (<b>d</b>) HI strength for the Atlantic Sector (blue) and Pacific Sector (red) are plotted only when TI or HI exists. Monthly boxplots of (<b>e</b>) BT difference between 11 μm and 12 μm and (<b>f</b>) total column water vapor at the Atlantic side (blue) and Pacific side (red) are also given. The black dotted line in (<b>e</b>) is the horizon of 0.036 K. Outliers are not shown in the boxplots.</p>
Full article ">Figure 17
<p>Maps of monthly MERRA-2 total column water vapor north of 60°N in (<b>a</b>) January, (<b>b</b>) April, (<b>c</b>) July, and (<b>d</b>) October. The data are averaged from 2002 to 2022.</p>
Full article ">Figure 18
<p>Monthly boxplots of (<b>a</b>) latitude of MODIS pixels and (<b>b</b>) in situ subsurface SST in the MUDB for the Atlantic (blue) and Pacific (red) Sectors. Outliers are not shown here.</p>
Full article ">
23 pages, 32897 KiB  
Article
On the Suitability of Different Satellite Land Surface Temperature Products to Study Surface Urban Heat Islands
by Alexandra Hurduc, Sofia L. Ermida and Carlos C. DaCamara
Remote Sens. 2024, 16(20), 3765; https://doi.org/10.3390/rs16203765 - 10 Oct 2024
Cited by 2 | Viewed by 1452
Abstract
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are [...] Read more.
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are considerable advantages that arise from these different characteristics (spatiotemporal resolution, time of observation, etc.), it also means that there is a need for understanding the ability of sensors in capturing spatial and temporal SUHI patterns. For this, several land surface temperature products are compared for the cities of Madrid and Paris, retrieved from five sensors: the Spinning Enhanced Visible and InfraRed Imager onboard Meteosat Second Generation, the Advanced Very-High-Resolution Radiometer onboard Metop, the Moderate-resolution Imaging Spectroradiometer onboard both Aqua and Terra, and the Thermal Infrared Sensor onboard Landsat 8 and 9. These products span a wide range of LST algorithms, including split-window, single-channel, and temperature–emissivity separation methods. Results show that the diurnal amplitude of SUHI may not be well represented when considering daytime and nighttime polar orbiting platforms. Also, significant differences arise in SUHI intensity and spatial and temporal variability due to the different methods implemented for LST retrieval. Full article
(This article belongs to the Section AI Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Land cover resampled for the three projections of LST products (<b>a</b>–<b>f</b>) along with the percentage of urban pixels (<b>g</b>–<b>l</b>).</p>
Full article ">Figure 2
<p>Time of observation of each sensor: (<b>a</b>) for Madrid during daytime and time of minimum SUHI (SUHI<sub>min</sub>), (<b>b</b>) for Madrid during nighttime and time of maximum SUHI (SUHI<sub>max</sub>), (<b>c</b>) for Paris during daytime and SUHI<sub>max</sub>, (<b>d</b>) for Paris during nighttime and SUHI<sub>min</sub>. Colored bins are sampled every 15 min.</p>
Full article ">Figure 3
<p>Mean DJF (December, January, February) LST for all products considered. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 4
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for JJA (June, July, and August). (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 5
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 6
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris and DJF; (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime, an extension of the histogram in (<b>d6</b>) is seen in (<b>d8</b>). Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 7
<p>Diurnal cycle of SUHI for Madrid: (<b>a</b>) DJF, (<b>b</b>) MAM, (<b>c</b>) JJA, (<b>d</b>) SON.</p>
Full article ">Figure 8
<p>As <a href="#remotesensing-16-03765-f007" class="html-fig">Figure 7</a> but for Paris.</p>
Full article ">Figure 9
<p>Correlation of monthly SUHI anomalies between all products considered: (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</p>
Full article ">Figure 10
<p>As <a href="#remotesensing-16-03765-f009" class="html-fig">Figure 9</a> but for Paris. (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</p>
Full article ">
20 pages, 16133 KiB  
Article
Changes in Vegetation Cover and the Relationship with Surface Temperature in the Cananéia–Iguape Coastal System, São Paulo, Brazil
by Jakeline Baratto, Paulo Miguel de Bodas Terassi and Emerson Galvani
Remote Sens. 2024, 16(18), 3460; https://doi.org/10.3390/rs16183460 - 18 Sep 2024
Viewed by 1085
Abstract
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized [...] Read more.
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized Difference Vegetation Index (NDVI) and the Enhanced Vegetation Index (EVI) were acquired from the MODIS/Aqua sensor (MYD13Q1) and the leaf area index (LAI) from the MODIS/Terra (MOD15A2H). Surface temperature data were acquired from MODIS/Aqua (MYD11A2). The data were processed using Google Earth Engine and Google Colab. The data were collected, and spatial and temporal correlations were applied. Correlations were applied in the annual and seasonal period. The annual temporal correlation between vegetation indices and surface temperature was positive, but statistically significant for the LAI, with r = 0.43 (90% significance). In the seasonal period, positive correlations occurred in JFM for all indices (95% significance). Spatially, the results of this research indicate that the largest area showed a positive correlation between VI and LST. The hottest and rainiest periods (OND and JFM) had clearer and more significant correlations. In some regions, significant and clear correlations were observed, such as in some areas in the north, south and close to the city of Iguape. This highlights the complexity of the interactions between vegetation indices and climatic attributes, and highlights the importance of considering other environmental variables and processes when interpreting changes in vegetation. However, this research has significantly progressed the field, by establishing new correlations and demonstrating the importance of considering climate variability, for a more accurate understanding of the impacts on vegetation indices. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area (<b>A</b>,<b>B</b>) and land use mapping (<b>C</b>).</p>
Full article ">Figure 2
<p>Variation in surface temperature and monthly (<b>A</b>) and annual (<b>B</b>) rainfall for the Cananéia-Iguape Coastal System for the 20032022 period. Source: MODIS/Aqua and CHIRPS, 2024.</p>
Full article ">Figure 3
<p>Annual variation in vegetation indices for the 2003–2022 period in the Cananéia–Iguape Coastal System.</p>
Full article ">Figure 4
<p>Scatter plot of annual NDVI (<b>a</b>), EVI (<b>b</b>) and LAI (<b>c</b>) values and surface temperature from 2003 to 2022.</p>
Full article ">Figure 5
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and surface temperature for the JFM (<b>a</b>–<b>c</b>) and AMJ (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 6
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and climate variables for the JAS (<b>a</b>–<b>c</b>) and OND (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 7
<p>Annual linear correlation between surface temperature and NDVI (<b>A</b>), EVI (<b>B</b>) and LAI (<b>C</b>) between 2003 and 2022.</p>
Full article ">Figure 8
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JFM (<b>A</b>–<b>C</b>) and AMJ (<b>D</b>–<b>F</b>) periods.</p>
Full article ">Figure 9
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JAS (<b>A</b>–<b>C</b>) and OND (<b>D</b>–<b>F</b>) periods.</p>
Full article ">
17 pages, 3602 KiB  
Article
Understanding Two Decades of Turbidity Dynamics in a Coral Triangle Hotspot: The Berau Coastal Shelf
by Faruq Khadami, Ayi Tarya, Ivonne Milichristi Radjawane, Totok Suprijo, Karina Aprilia Sujatmiko, Iwan Pramesti Anwar, Muhamad Faqih Hidayatullah and Muhamad Fauzan Rizky Adisty Erlangga
Water 2024, 16(16), 2300; https://doi.org/10.3390/w16162300 - 15 Aug 2024
Viewed by 1344
Abstract
Turbidity serves as a crucial indicator of coastal water health and productivity. Twenty years of remote sensing data (2003–2022) from the Aqua Moderate Resolution Imaging Spectroradiometer (MODIS) satellite were used to analyze the spatial and temporal variations in turbidity, as measured by total [...] Read more.
Turbidity serves as a crucial indicator of coastal water health and productivity. Twenty years of remote sensing data (2003–2022) from the Aqua Moderate Resolution Imaging Spectroradiometer (MODIS) satellite were used to analyze the spatial and temporal variations in turbidity, as measured by total suspended matter (TSM), in the Berau Coastal Shelf (BCS), East Kalimantan, Indonesia. The BCS encompasses the estuary of the Berau River and is an integral part of the Coral Triangle, renowned for its rich marine and coastal habitats, including coral reefs, mangroves, and seagrasses. The aim of this research is to comprehend the seasonal and interannual patterns of turbidity and their associations with met-ocean parameters, such as wind, rainfall, and climate variations like the El Niño–Southern Oscillation (ENSO) and the Indian Ocean Dipole (IOD). The research findings indicate that the seasonal spatial pattern of turbidity is strongly influenced by monsoon winds, while its temporal patterns are closely related to river discharge and rainfall. The ENSO and IOD climate cycles exert an influence on the interannual turbidity variations, with turbidity values decreasing during La Niña and negative IOD events and conversely increasing during El Niño and positive IOD events. Furthermore, the elevated turbidity during negative IOD and La Niña coincides with rising temperatures, potentially acting as a compound stressor on marine habitats. These findings significantly enhance our understanding of turbidity dynamics in the BCS, thereby supporting the management of marine and coastal ecosystems in the face of changing climatic and environmental conditions. Full article
(This article belongs to the Section Oceans and Coastal Zones)
Show Figures

Figure 1

Figure 1
<p>Berau Coastal Shelf. The red boxes mark regions selected for spatial averaging, which encompass the north, river mouth, and south regions. The triangle symbol shows the river discharge station. The color shading in the image represents the varying bathymetry depths.</p>
Full article ">Figure 2
<p>Mean climatological TSM (mg/L) in the BCS.</p>
Full article ">Figure 3
<p>Monthly climatological TSM and wind rose.</p>
Full article ">Figure 4
<p>TSM variability at three locations in the BCS, showing median, interquartile ranges, max/min, and outliers. The black dots indicating outliers.</p>
Full article ">Figure 5
<p>Monthly differences in average TSM (mg/L): average TSM in the (<b>a</b>) north, (<b>b</b>) river mouth, and (<b>c</b>) south areas of the BCS from 2003 to 2022. The black dots indicating outliers.</p>
Full article ">Figure 6
<p>Power spectral density of average TSM in (<b>a</b>) north, (<b>b</b>) river mouth, and (<b>c</b>) south areas of the BCS.</p>
Full article ">Figure 7
<p>Berau River discharge data from May 2006 to January 2008. The black line represents the observed discharge values. The red line shows the data smoothed by a 30-day low-pass filter.</p>
Full article ">Figure 8
<p>The variation in the 15-month moving average of (<b>a</b>) the TSM at river mouth, (<b>b</b>) the precipitation, the (<b>c</b>) Dipole Mode Index, (<b>d</b>) El Niño–Southern Oscillation Index (NINO 3.4), and (<b>e</b>) Sea Surface Temperature (SST) anomaly. The red (blue) color indicating positive (negative) anomaly.</p>
Full article ">
19 pages, 9008 KiB  
Article
The Carpathian Agriculture in Poland in Relation to Other EU Countries, Ukraine and the Environmental Goals of the EU CAP 2023–2027
by Marek Zieliński, Artur Łopatka, Piotr Koza and Barbara Gołębiewska
Agriculture 2024, 14(8), 1325; https://doi.org/10.3390/agriculture14081325 - 9 Aug 2024
Cited by 1 | Viewed by 1235
Abstract
This study discusses the issue of determining the direction and strength of changes taking place in the structure of agricultural land in the mountain and foothill areas of the Carpathians in Poland in comparison with Slovakia, Romania and Ukraine. The most important financial [...] Read more.
This study discusses the issue of determining the direction and strength of changes taking place in the structure of agricultural land in the mountain and foothill areas of the Carpathians in Poland in comparison with Slovakia, Romania and Ukraine. The most important financial institutional measures dedicated to the protection of the natural environment in Polish agriculture in the Areas facing Natural and other specific Constraints (ANCs) mountain and foothill in the first year of the CAP 2023–2027 were also established. Satellite data from 2001 to 2022 were used. The analyses used the land use classification MCD12Q1 provided by NASA and were made on the basis of satellite imagery collections from the MODIS sensor placed on two satellites: TERRA and AQUA. In EU countries, a decreasing trend in agricultural areas has been observed in areas below 350 m above sea level. In areas above 350 m, this trend weakened or even turned into an upward trend. Only in Ukraine was a different trend observed. It was found that in Poland, the degree of involvement of farmers from mountain and foothill areas in implementing financial institutional measures dedicated to protecting the natural environment during the study period was not satisfactory. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the analysis of agriculture within separate groups of communes due to the fact and nuisance of ANCs mountain and foothill in Poland. Source: own study.</p>
Full article ">Figure 2
<p>Distribution of communes with different shares of ANCs mountain and foothill in Poland. Source: own study ISSPC SRI; IAFE NRI.</p>
Full article ">Figure 3
<p>Land use in the Carpathians in 2001 and 2022. Source: own study based on MODIS.</p>
Full article ">Figure 4
<p>Trends in the percentage share [%] of the total agricultural area and cropland in the total area of land in the Carpathians in 2001–2022. Source: own study based on MODIS.</p>
Full article ">Figure 5
<p>Number of farms participating in practices under eco-schemes, in organic and agri–environment–climate measures in communes with different shares of ANCs mountain and foothill in Poland in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 6
<p>Share of [%] farms with eco-schemes in total number of farms in communes with ANCs mountain and foothill in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 7
<p>Share of [%] farms with organic and agri–environmental–climate measure in total number of farms in communes with ANCs mountain and foothill in 2023.</p>
Full article ">Figure 8
<p>Agricultural area covered by practices under eco-schemes, ecological and agri–environment–climate measures in communes with different shares of ANCs mountain and foothill in Poland in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 9
<p>Share [%] of UAA in farms with eco-schemes in total UAA in communes with ANCs mountain and foothill in 2023.</p>
Full article ">Figure 10
<p>Share [%] of UAA covered by organic and agri–environmental–climate measures in total UAA in communes with ANCs mountain and foothill in 2023.</p>
Full article ">
25 pages, 6444 KiB  
Article
Long-Term Evaluation of Aerosol Optical Properties in the Levantine Region: A Comparative Analysis of AERONET and Aqua/MODIS
by Ayse Gokcen Isik, S. Yeşer Aslanoğlu and Gülen Güllü
Remote Sens. 2024, 16(14), 2651; https://doi.org/10.3390/rs16142651 - 20 Jul 2024
Viewed by 1226
Abstract
The focus on aerosol analysis in the Levantine Region is driven by climate-change impacts, the region’s increasing urban development and industrial activities, and its geographical proximity to major dust-source areas. This study conducts a comparative analysis of aerosol optical depth data from Aqua/MODIS [...] Read more.
The focus on aerosol analysis in the Levantine Region is driven by climate-change impacts, the region’s increasing urban development and industrial activities, and its geographical proximity to major dust-source areas. This study conducts a comparative analysis of aerosol optical depth data from Aqua/MODIS and AERONET during different periods between 2003 and 2023 at four stations: IMS-METU-ERDEMLI (Mersin/Türkiye) (2004–2019), CUT-TEPAK (Limassol/Cyprus) (2010–2023), Cairo_EMA_2 (Cairo/Egypt) (2010–2023), and SEDE_BOKER (Sede Boker/Israel) (2003–2023). The objective is to evaluate the variability and reliability of AOD measurements between satellite and ground-based observations and to determine how well they represent regional climatology. The highest percentage of measurements within the expected error envelope was observed at the IMS-METU-ERDEMLI station, indicating the best agreement between MODIS and AERONET data at this location. The Seasonal-Trend Decomposition using Loess (STL) method revealed consistent spring and summer peaks influenced by dust transport from the Sahara and the Middle East, with lower values in winter. The study also considers the influence of cloud fraction on MODIS measurements and includes aerosol classification. A statistically significant slight positive trend in AOD values was identified at the IMS-METU-ERDEMLI station. Conversely, no significant trends were detected at the other stations. The results of this study agree with those of previous research on the impact of long-range dust transport on regional aerosol loadings, emphasizing the importance of integrating satellite and ground-based observations. Full article
Show Figures

Figure 1

Figure 1
<p>Levantine Region with the AERONET stations used for the study.</p>
Full article ">Figure 2
<p>MODIS AOD versus AERONET AOD at Erdemli Station from 2004 to 2019. The red line is the regression line, and the green dashed lines define the envelope of expected error (EE). Colors of the points represent the cloud-fraction range associated with MODIS measurements.</p>
Full article ">Figure 3
<p>MODIS AOD versus AERONET AOD at Cyprus station from 2010 to 2023. The red line is the regression line, and the green dashed lines define the envelope of expected error (EE). Colors of the points represent the cloud-fraction range associated with MODIS measurements.</p>
Full article ">Figure 4
<p>MODIS AOD versus AERONET AOD at Cairo station from 2010 to 2023. The red line is the regression line, and the green dashed lines define the envelope of expected error (EE). Colors of the points represent the cloud-fraction range associated with MODIS measurements.</p>
Full article ">Figure 5
<p>MODIS AOD versus AERONET AOD at Sede Boker station from 2003 to 2023. The red line is the regression line, and the green dashed lines define the envelope of expected error (EE). Colors of the points represent the cloud-fraction range associated with MODIS measurements.</p>
Full article ">Figure 6
<p>Comparison of AOD measurements from MODIS to those from AERONET at IMS-METU-ERDEMLI, CUT-TEPAK, Cairo_EMA_2, and SEDE_BOKER stations over a decade, from 2010 to 2019.</p>
Full article ">Figure 7
<p>Seasonal comparison between MODIS and AERONET AOD measurements across four locations. The seasons are abbreviated as (<b>a</b>) DJF (December, January, February), (<b>b</b>) MAM (March, April, May), (<b>c</b>) JJA (June, July, August), and (<b>d</b>) SON (September, October, November).</p>
Full article ">Figure 8
<p>Histograms of the MODIS and AERONET measurements at 550 nm Across Different AERONET Stations: (<b>a</b>) IMS-METU-ERDEMLI, (<b>b</b>) CUT-TEPAK, (<b>c</b>) Cairo_EMA_2, (<b>d</b>) SEDE_BOKER.</p>
Full article ">Figure 9
<p>Aerosol classification for the IMS-METU-ERDEMLI station in Erdemli/Mersin (Time period: 2004–2019).</p>
Full article ">Figure 10
<p>Aerosol classification for the CUT-TEPAK station in Limassol/Cyprus (Time period: 2010–2023).</p>
Full article ">Figure 11
<p>Aerosol classification for the Cairo_EMA_2 station in Cairo/Egypt (Time period: 2010–2023).</p>
Full article ">Figure 12
<p>Aerosol classification for the SEDE_BOKER station in Sede Boker/Israel (Time period: 2003–2023).</p>
Full article ">Figure 13
<p>Trends in MODIS and AERONET AOD at 550 nm for four different stations: (<b>a</b>) IMS-METU-ERDEMLI (2004–2019), (<b>b</b>) CUT-TEPAK (2010–2023), (<b>c</b>) Cairo_EMA_2 (2010–2023), and (<b>d</b>) SEDE_BOKER (2003–2023).</p>
Full article ">Figure A1
<p>Scatterplot of MODIS AOD versus AERONET AOD for different cloud-fraction intervals (CF) across four different sites: (<b>a</b>–<b>c</b>) IMS-METU-ERDEMLI, (<b>d</b>–<b>f</b>) CUT-TEPAK, (<b>g</b>–<b>i</b>) Cairo_EMA, and (<b>j</b>–<b>l</b>) SEDE_BOKER. Each station was analyzed for three cloud-fraction intervals: CF (0–0.3), CF (0.3–0.6), and CF (0.6–1.0).</p>
Full article ">Figure A2
<p>STL decomposition of MODIS AOD time-series data at 550 nm for four different sites: (<b>a</b>) IMS-METU-ERDEMLI, (<b>b</b>) CUT-TEPAK, (<b>c</b>) Cairo_EMA_2, and (<b>d</b>) SEDE_BOKER. Each subplot includes the original data, trend component, seasonal component, and residuals from 2003 to 2023.</p>
Full article ">
16 pages, 1280 KiB  
Article
Are Harmful Algal Blooms Increasing in the Great Lakes?
by Karl R. Bosse, Gary L. Fahnenstiel, Cal D. Buelo, Matthew B. Pawlowski, Anne E. Scofield, Elizabeth K. Hinchey and Michael J. Sayers
Water 2024, 16(14), 1944; https://doi.org/10.3390/w16141944 - 10 Jul 2024
Cited by 1 | Viewed by 1416
Abstract
This study used satellite remote sensing to investigate trends in harmful algal blooms (HABs) over the last 21 years, focusing on four regions within the Laurentian Great Lakes: western Lake Erie, Green Bay, Saginaw Bay, and western Lake Superior. HABs in the water [...] Read more.
This study used satellite remote sensing to investigate trends in harmful algal blooms (HABs) over the last 21 years, focusing on four regions within the Laurentian Great Lakes: western Lake Erie, Green Bay, Saginaw Bay, and western Lake Superior. HABs in the water column were identified from remote sensing-derived chlorophyll concentrations, and surface HAB scums were classified based on the Normalized Difference Vegetation Index (NDVI) band ratio index. Using imagery from the Moderate Resolution Imaging Spectroradiometer sensor on the Aqua satellite (MODIS-Aqua) from 2002 to 2022, we generated daily estimates of the HAB and surface scum extents for each region, which were then averaged to generate mean annual extents. We observed a significant decline in the Saginaw Bay mean annual HAB extents over the 21-year study period. Otherwise, no significant changes were observed over this period in any region for either the HAB or surface scum mean annual extents, thus suggesting that HABs are not increasing in the Great Lakes. Despite the lack of increasing trends, the blooms are still recurring annually and causing a negative impact on the nearby communities; thus, we believe that it is crucial to continue studying Great Lakes HABs to monitor the impact of current and future abatement strategies. Full article
Show Figures

Figure 1

Figure 1
<p>The regions studied for historical HAB extents include western Lake Erie (WLE), Saginaw Bay, Green Bay, and western Lake Superior.</p>
Full article ">Figure 2
<p>Mean annual HAB extents for WLE (panel (<b>A</b>)), Saginaw Bay (panel (<b>B</b>)), and Green Bay (panel (<b>C</b>)). Error bars represent the standard error of the mean. Only Saginaw Bay had a significant trend over the data record (slope = −2.1 km<sup>2</sup>/yr; <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>&lt;</mo> <mn>0.05</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 3
<p>Mean annual surface HAB scum extents for WLE (panel (<b>A</b>)), Saginaw Bay (panel (<b>B</b>)), and Green Bay (panel (<b>C</b>)). Error bars represent the standard error of the mean.</p>
Full article ">Figure 4
<p>Mean annual HAB extent plotted against mean annual surface HAB scum extent for WLE (panel (<b>A</b>)), Saginaw Bay (panel (<b>B</b>)), and Green Bay (panel (<b>C</b>)). Only Green Bay showed a significant correlation between these indices (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Mean annual algal bloom extent for Lake Superior. Error bars represent the standard error of the mean.</p>
Full article ">
Back to TopTop