[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (314)

Search Parameters:
Keywords = snow surface temperature

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 13342 KiB  
Article
Spatiotemporal Variability in Snow and Land Cover in Sefid-Rud Basin, Iran
by Hersh Entezami, Firouz Mojarrad, Himan Shahabi and Ebrahim Ghaderpour
Sustainability 2024, 16(21), 9381; https://doi.org/10.3390/su16219381 - 29 Oct 2024
Viewed by 473
Abstract
Snow cover has a key role in balancing the Earth’s surface temperature and can help in filling rivers and reservoirs. In this study, 8-day MOD10A2 images are employed to monitor the spatiotemporal changes in snow cover in the Sefid-Rud basin and its eleven [...] Read more.
Snow cover has a key role in balancing the Earth’s surface temperature and can help in filling rivers and reservoirs. In this study, 8-day MOD10A2 images are employed to monitor the spatiotemporal changes in snow cover in the Sefid-Rud basin and its eleven sub-basins during 2000–2019. The non-parametric Mann–Kendall (MK) test and its associated Sen’s slope estimator are utilized to estimate the trends in annual, seasonal, and monthly snow cover changes. The Sen’s slope results show a decrease in the snow cover for the basin, statistically significant toward the central and southern parts of the basin. In the winter season, a decreasing trend is observed, where its decreasing rate is higher than the annual rate. The trends in the calendar months are like the seasons, i.e., December, January, and February exhibit a decreasing trend, like the winter season. The Goltapeh-Zarinabad and Ghorveh-Dehgolan sub-basins show decreasing snow cover rates of 0.51 and 0.68 (%/year) during 2000–2019, respectively, the only two sub-basins whose gradients are statistically significant at the 95% confidence level. The Pearson correlation analysis between elevation and snow cover for each year shows that the highest and lowest correlations are 0.81 for 2007 and 0.59 for 2017. Finally, analysis of the MCD12Q1 land cover data shows that a significant portion of non-vegetated lands have turned into grasslands, mainly in the central part of the basin, where the significant gradual snow cover decline is observed. The results can guide stakeholders and policymakers in the development of a sustainable environment in the face of climate change. Full article
(This article belongs to the Section Environmental Sustainability and Applications)
Show Figures

Figure 1

Figure 1
<p>Map of Sefid-Rud basin and its sub-basins.</p>
Full article ">Figure 2
<p>The workflow of the present study. The numbers displayed in red follow the same order of the subsections in <a href="#sec3-sustainability-16-09381" class="html-sec">Section 3</a>.</p>
Full article ">Figure 3
<p>Graphs of temporal snow cover changes in the Sefid-Rud basin during the years 2000, 2001, 2018, and 2019.</p>
Full article ">Figure 4
<p>The percentage of snow cover in the Sefid-Rud basin for (<b>a</b>) December, (<b>b</b>) January, and (<b>c</b>) February.</p>
Full article ">Figure 5
<p>Annual snow cover time series and their estimated linear trends (dashed lines) using an MK trend analysis for the Sefid-Rud basin and Ghorveh-Dehgolan.</p>
Full article ">Figure 6
<p>The trend results for the Sefid-Rub basin for (<b>a</b>) annual, (<b>b</b>) autumn, (<b>c</b>) winter, and (<b>d</b>) spring, based on the MK test. The statistical significance of the trends at the 90% confidence level are also illustrated by smaller panels, placed at the southeast part of the basin.</p>
Full article ">Figure 7
<p>The trend results for the Sefid-Rub basin for (<b>a</b>) November, (<b>b</b>) December, (<b>c</b>) January, (<b>d</b>) February, (<b>e</b>) March, and (<b>f</b>) April based on the MK test. The statistical significance of the trends at the 90% confidence level are also illustrated by smaller panels, placed at the southeast part of the basin.</p>
Full article ">Figure 8
<p>The MODIS land cover change results for the Sefid-Rud basin.</p>
Full article ">Figure 9
<p>The time series of dominant land covers for sub-basins: (<b>a</b>) Goltapeh-Zarinabad and (<b>b</b>) Ghorveh-Dehgolan.</p>
Full article ">
19 pages, 24334 KiB  
Article
A 40-Year Time Series of Land Surface Emissivity Derived from AVHRR Sensors: A Fennoscandian Perspective
by Mira Barben, Stefan Wunderle and Sonia Dupuis
Remote Sens. 2024, 16(19), 3686; https://doi.org/10.3390/rs16193686 - 2 Oct 2024
Viewed by 594
Abstract
Accurate land surface temperature (LST) retrieval depends on precise knowledge of the land surface emissivity (LSE). Neglecting or inaccurately estimating the emissivity introduces substantial errors and uncertainty in LST measurements. The emissivity, which varies across different surfaces and land uses, reflects material composition [...] Read more.
Accurate land surface temperature (LST) retrieval depends on precise knowledge of the land surface emissivity (LSE). Neglecting or inaccurately estimating the emissivity introduces substantial errors and uncertainty in LST measurements. The emissivity, which varies across different surfaces and land uses, reflects material composition and surface roughness. Satellite data offer a robust means to determine LSE at large scales. This study utilises the Normalised Difference Vegetation Index Threshold Method (NDVITHM) to produce a novel emissivity dataset spanning the last 40 years, specifically tailored for the Fennoscandian region, including Norway, Sweden, and Finland. Leveraging the long and continuous data series from the Advanced Very High Resolution Radiometer (AVHRR) sensors aboard the NOAA and MetOp satellites, an emissivity dataset is generated for 1981–2022. This dataset incorporates snow-cover information, enabling the creation of annual emissivity time series that account for winter conditions. LSE time series were extracted for six 15 × 15 km study sites and compared against the Moderate Resolution Imaging Spectroradiometer (MODIS) MOD11A2 LSE product. The intercomparison reveals that, while both datasets generally align, significant seasonal differences exist. These disparities are attributable to differences in spectral response functions and temporal resolutions, as well as the method considering fixed values employed to calculate the emissivity. This study presents, for the first time, a 40-year time series of the emissivity for AVHRR channels 4 and 5 in Fennoscandia, highlighting the seasonal variability, land-cover influences, and wavelength-dependent emissivity differences. This dataset provides a valuable resource for future research on long-term land surface temperature and emissivity (LST&E) trends, as well as land-cover changes in the region, particularly with the use of Sentinel-3 data and upcoming missions such as EUMETSAT’s MetOp Second Generation, scheduled for launch in 2025. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spectral emissivities of different land-cover classes, as recorded in the ECOSTRESS spectral library [<a href="#B41-remotesensing-16-03686" class="html-bibr">41</a>,<a href="#B42-remotesensing-16-03686" class="html-bibr">42</a>].</p>
Full article ">Figure 2
<p>The study area across Norway, Sweden, and Finland, showing the six chosen study sites (15 × 15 km each). The abbreviations indicating the study sites stand for Low Vegetation (LV) or Forest (F) and South (S), Mid-Latitude (ML), or North (N). The base map is the ESA CCI Land-Cover Dataset [<a href="#B43-remotesensing-16-03686" class="html-bibr">43</a>].</p>
Full article ">Figure 3
<p>Schematic workflow showing the AVHRR data preparation, emissivity dataset calculation process, and incorporated auxiliary data.</p>
Full article ">Figure 4
<p>Overview of the availability of AVHRR data since 1981 in the local archive. The data used for this study are indicated in blue-grey, while the data excluded from the analysis due to quality or processing issues are indicated in orange.</p>
Full article ">Figure 5
<p>The 40-year time series of monthly mean land surface emissivities for the 15 × 15 km low-vegetation southern (LVS) study site.</p>
Full article ">Figure 6
<p>Mean annual cycle of LSE for channel 4, including the confidence interval (1 <math display="inline"><semantics> <mi>σ</mi> </semantics></math>), for the 40-year period for the 15 × 15 km low-vegetation southern (LVS) study site.</p>
Full article ">Figure 7
<p>(<b>a</b>) Land cover (ESA CCI; see <a href="#remotesensing-16-03686-f002" class="html-fig">Figure 2</a> for details) and emissivity differences (Ch5–Ch4) derived from the monthly mean emissivity values over the 40-year period (1981–2022) for an area of 1° × 1° around the FN site in February (<b>b</b>) and July (<b>c</b>).</p>
Full article ">Figure 8
<p>(<b>a</b>) Land cover (ESA CCI; see <a href="#remotesensing-16-03686-f002" class="html-fig">Figure 2</a> for details) and emissivity differences (Ch5–Ch4) derived from the monthly mean emissivity values over the 40-year period (1981–2022) for an area of 1° × 1° around the FS site in February (<b>b</b>) and July (<b>c</b>).</p>
Full article ">Figure 9
<p>Comparison of the AVHRR LAC LSE dataset and the MODIS MOD11A2 LSE dataset for the low-vegetation southern (LVS) study site (2015–2022).</p>
Full article ">
15 pages, 11836 KiB  
Article
Variation in Glacier Albedo on the Tibetan Plateau between 2001 and 2022 Based on MODIS Data
by Ping Liu, Guangjian Wu, Bo Cao, Xuanru Zhao and Yuxuan Chen
Remote Sens. 2024, 16(18), 3472; https://doi.org/10.3390/rs16183472 - 19 Sep 2024
Viewed by 517
Abstract
Albedo is a primary driver of the glacier surface energy balance and consequent melting. As glacier albedo decreases, it further accelerates glacier melting. Over the past 20 years, glaciers on the Tibetan Plateau have experienced significant melting. However, our understanding of the variations [...] Read more.
Albedo is a primary driver of the glacier surface energy balance and consequent melting. As glacier albedo decreases, it further accelerates glacier melting. Over the past 20 years, glaciers on the Tibetan Plateau have experienced significant melting. However, our understanding of the variations in glacier albedo and its driving factors in this region remains limited. This study used MOD10A1 data to examine the average characteristics and variations in glacier albedo on the Tibetan Plateau from 2001 to 2022; the MOD10A1 snow cover product, developed at the National Snow and Ice Data Center, was employed to analyze spatiotemporal variations in surface albedo. The results indicate that the albedo values of glaciers on the Tibetan Plateau predominantly range between 0.50 and 0.60, with distinctly higher albedo in spring and winter, and lower albedo in summer and autumn. Glacier albedo on the Tibetan Plateau decreased at an average linear regression rate of 0.06 × 10−2 yr−1 over the past two decades, with the fastest declines occurring in autumn at an average rate of 0.18 × 10−2 yr−1, contributing to the prolongation of the melting period. Furthermore, significant variations in albedo change rates with altitude were found near the snowline, which is attributed to the transformation of the snow and ice surface. The primary factors affecting glacier albedo on the Tibetan Plateau are temperature and snowfall, whereas in the Himalayas, black carbon and dust primarily influence glacier albedo. Our findings reveal a clear decrease in glacier albedo on the Tibetan Plateau and demonstrate that seasonal and spatial variations in albedo and temperature are the most important driving factors. These insights provide valuable information for further investigation into surface albedo and glacier melt. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Glacier distribution on the Tibetan Plateau in 12 subregions (sourced from RGI 6.0).</p>
Full article ">Figure 2
<p>Spatial distribution characteristics of glacier albedo on the Tibetan Plateau between 2001 and 2022. Each circle represents the albedo data for one glacier.</p>
Full article ">Figure 3
<p>Trends of glacier albedo change on the Tibetan Plateau and in its various subregions from 2001 to 2022. Each circle represents the albedo data for one glacier.</p>
Full article ">Figure 4
<p>Relationship between glacier albedo and its rate of change with elevation in various subregions. The red bar in each graph represents the range of glacier snowline altitude [<a href="#B44-remotesensing-16-03472" class="html-bibr">44</a>]. The blue band represents the glacier albedo values, the red band represents the rate of change of glacier albedo, and the blue dashed line indicates where albedo’s change rate equals zero.</p>
Full article ">Figure 5
<p>Correlations between annual average albedo and driving factors in the Tibetan Plateau and its various subregions are represented by the Pearson correlation coefficient |R|. Blue dots indicate positive correlations between albedo and driving factors, while red dots indicate negative correlations between albedo and driving factors. The color change from red to green represents the rate of change in glacier albedo, with red indicating a decreasing rate and green indicating an increasing rate. Pre, Sf, Temp, BC, and Dust represent total precipitation, snowfall, temperature, black carbon, and dust, respectively.</p>
Full article ">Figure 6
<p>Correlation between glacier albedo (red dots and lines) and mass balance (blue columns) in different subregions and the zero-reference line for mass balance (blue dashed lines). The mass balance values were derived for the periods 2000–2004, 2005–2009, 2010–2014, and 2015–2019, and the corresponding albedo values were obtained for those four time periods. R represents Pearson’s correlation coefficient.</p>
Full article ">
21 pages, 7794 KiB  
Article
Spatial–Temporal Variations and Driving Factors of the Albedo of the Qilian Mountains from 2001 to 2022
by Huazhu Xue, Haojie Zhang, Zhanliang Yuan, Qianqian Ma, Hao Wang and Zhi Li
Atmosphere 2024, 15(9), 1081; https://doi.org/10.3390/atmos15091081 - 6 Sep 2024
Viewed by 610
Abstract
Surface albedo plays a pivotal role in the Earth’s energy balance and climate. This study conducted an analysis of the spatial distribution patterns and temporal evolution of albedo, normalized difference vegetation index (NDVI), normalized difference snow index snow cover (NSC), and land surface [...] Read more.
Surface albedo plays a pivotal role in the Earth’s energy balance and climate. This study conducted an analysis of the spatial distribution patterns and temporal evolution of albedo, normalized difference vegetation index (NDVI), normalized difference snow index snow cover (NSC), and land surface temperature (LST) within the Qilian Mountains (QLMs) from 2001 to 2022. This study evaluated the spatiotemporal correlations of albedo with NSC, NDVI, and LST at various temporal scales. Additionally, the study quantified the driving forces and relative contributions of topographic and natural factors to the albedo variation of the QLMs using geographic detectors. The findings revealed the following insights: (1) Approximately 22.8% of the QLMs exhibited significant changes in albedo. The annual average albedo and NSC exhibited a minor decline with rates of −0.00037 and −0.05083 (Sen’s slope), respectively. Conversely, LST displayed a marginal increase at a rate of 0.00564, while NDVI experienced a notable increase at a rate of 0.00178. (2) The seasonal fluctuations of NSC, LST, and vegetation collectively influenced the overall albedo changes in the Qilian Mountains. Notably, the highly similar trends and significant correlations between albedo and NSC, whether in intra-annual monthly variations, multi-year monthly anomalies, or regional multi-year mean trends, indicate that the changes in snow albedo reflected by NSC played a major role. Additionally, the area proportion and corresponding average elevation of PSI (permanent snow and ice regions) slightly increased, potentially suggesting a slow upward shift of the high mountain snowline in the QLMs. (3) NDVI, land cover type (LCT), and the Digital Elevation Model (DEM, which means elevation) played key roles in shaping the spatial pattern of albedo. Additionally, the spatial distribution of albedo was most significantly influenced by the interaction between slope and NDVI. Full article
(This article belongs to the Special Issue Vegetation and Climate Relationships (3rd Edition))
Show Figures

Figure 1

Figure 1
<p>Geographic location and elevation of the QLMs.</p>
Full article ">Figure 2
<p>(<b>a</b>) Slope, (<b>b</b>) aspect, and (<b>c</b>) land cover types of the QLMs (the explanation of the abbreviation is included in <a href="#atmosphere-15-01081-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 3
<p>The multi-year average (<b>a</b>) albedo, (<b>b</b>) NSC, (<b>c</b>) NDVI, and (<b>d</b>) LST.</p>
Full article ">Figure 4
<p>Trends in the annual average albedo of the QLMs from 2001 to 2022 in relation to (<b>a</b>) annual average NSC, (<b>b</b>) annual average NDVI, (<b>c</b>) annual average LST; (<b>d</b>) trends in the average elevation and area percentage of PSI regions in the QLMs.</p>
Full article ">Figure 5
<p>The spatial distribution of significant or non-significant changes in (<b>a</b>) albedo, (<b>b</b>) NSC, (<b>c</b>) NDVI, and (<b>d</b>) LST in the QLMs from 2001 to 2022.</p>
Full article ">Figure 6
<p>Comparison of the multi-year monthly average values of albedo with (<b>a</b>) NSC, (<b>b</b>) NDVI, and (<b>c</b>) LST in the QLM region from 2001 to 2022.</p>
Full article ">Figure 7
<p>Monthly albedo anomalies (<b>a</b>) and monthly NSC anomalies, (<b>b</b>) and monthly NDVI anomalies, (<b>c</b>) and monthly LST anomalies in the QLMs from 2001 to 2022.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>e</b>) Changes in explanatory power (q values) in 2001, 2006, 2011, 2016, and 2021.</p>
Full article ">Figure 9
<p>Striking differences in the driving factors (At a confidence level of 95%, “Y” indicates a significant difference in the spatial distribution of albedo due to the two factors, while “N” indicates the opposite).</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>e</b>) Changes in interactive explanatory power (q values) in 2001, 2006, 2011, 2016, and 2021; (<b>f</b>) average interactive explanatory power (q values) of 5 years (Bi: Enhance, bivariate, ENL: Enhance, nonlinear. The annotations inside parentheses indicate a higher frequency of occurrence of interaction types within five years. Without annotations, it indicates that the interaction types remained consistent over the 5 years).</p>
Full article ">
14 pages, 7934 KiB  
Article
Assimilating Satellite-Derived Snow Cover and Albedo Data to Improve 3-D Weather and Photochemical Models
by Colleen Jones, Huy Tran, Trang Tran and Seth Lyman
Atmosphere 2024, 15(8), 954; https://doi.org/10.3390/atmos15080954 - 10 Aug 2024
Viewed by 625
Abstract
During wintertime temperature inversion episodes, ozone in the Uinta Basin sometimes exceeds the standard of 70 ppb set by the US Environmental Protection Agency. Since ozone formation depends on sunlight, and less sunlight is available during winter, wintertime ozone can only form if [...] Read more.
During wintertime temperature inversion episodes, ozone in the Uinta Basin sometimes exceeds the standard of 70 ppb set by the US Environmental Protection Agency. Since ozone formation depends on sunlight, and less sunlight is available during winter, wintertime ozone can only form if snow cover and albedo are high. Researchers have encountered difficulties replicating high albedo values in 3-D weather and photochemical transport model simulations for winter episodes. In this study, a process to assimilate MODIS satellite data into WRF and CAMx models was developed, streamlined, and tested to demonstrate the impacts of data assimilation on the models’ performance. Improvements to the WRF simulation of surface albedo and snow cover were substantial. However, the impact of MODIS data assimilation on WRF performance for other meteorological quantities was minimal, and it had little impact on ozone concentrations in the CAMx photochemical transport model. The contrast between the data assimilation and reference cases was greater for a period with no new snow since albedo appears to decrease too rapidly in default WRF and CAMx configurations. Overall, the improvement from MODIS data assimilation had an observed enhancement in the spatial distribution and temporal evolution of surface characteristics on meteorological quantities and ozone production. Full article
Show Figures

Figure 1

Figure 1
<p>WRF one-way nested 12-4-1.33 km domains (<b>A</b>) and details of a 1.33 km domain, including topography and location of oil and gas wells (<b>B</b>). The white rectangle is Domain 2 and the red rectangle is Domain 3 from Table 4.</p>
Full article ">Figure 2
<p>Diagram of the MODIS data assimilation into the WRF and CAMx models.</p>
Full article ">Figure 3
<p>Comparison of the surface albedo fraction obtained in simulations using the WRF default configuration (<b>left</b>) and MODIS data assimilation (<b>right</b>). (Thin black lines = county outlines; heavy black outline = Uinta Basin ozone nonattainment area; and two-letter abbreviations = air quality monitoring stations.)</p>
Full article ">Figure 4
<p>Comparison of snow cover fraction (SNOWC) obtained in simulations using the WRF default configuration (<b>left</b>) and MODIS data assimilation (<b>right</b>). (Thin black lines = county outlines; heavy black outline = Uinta Basin ozone nonattainment area; and two-letter abbreviations = air quality monitoring stations.)</p>
Full article ">Figure 5
<p>Comparison of snow cover fraction, snow water equivalent, and snow depth using the WRF default configuration (REF) and MODIS data assimilation (MODIS). Green bars show periods where WRF reinitialized snow characteristics using the SNOWDAS dataset.</p>
Full article ">Figure 6
<p>Comparison of planetary boundary layer height (P.B.L.H.) and lapse rate using the WRF default configuration (REF) and MODIS data assimilation (MODIS).</p>
Full article ">Figure 7
<p>Comparison of photolysis rates simulated by CAMx using the default configuration (REF) and MODIS data assimilation (MODIS). (Green line = a new snow event.)</p>
Full article ">Figure 8
<p>Comparison of ozone at Ouray as simulated by CAMx using the default configuration (REF) and MODIS data assimilation (MODIS). (Red dash line = EPA National Ambient Air Quality Standard (NAAQS) for ozone).</p>
Full article ">
14 pages, 15087 KiB  
Article
The Improved SBAS-InSAR Technique Reveals Three-Dimensional Glacier Collapse: A Case Study in the Qinghai–Tibet Plateau
by Xinyao Wang, Jiayi Yao, Yanbo Cao and Jiaming Yao
Land 2024, 13(8), 1126; https://doi.org/10.3390/land13081126 - 24 Jul 2024
Viewed by 724
Abstract
Many debris-covered glaciers are widely distributed on the Qinghai–Tibet Plateau. Glaciers are important freshwater resources and cause disasters such as glacier collapse and landslides. Therefore, it is of great significance to monitor the movement characteristics of large active glaciers and analyze the process [...] Read more.
Many debris-covered glaciers are widely distributed on the Qinghai–Tibet Plateau. Glaciers are important freshwater resources and cause disasters such as glacier collapse and landslides. Therefore, it is of great significance to monitor the movement characteristics of large active glaciers and analyze the process of mass migration, which may cause serious threats and damage to roads and people living in surrounding areas. In this study, we chose a glacier with strong activity in Lulang County, Tibet, as the study area. The complete 4-year time series deformation of the glacier was estimated by using an improved small-baseline subset InSAR (SBAS-InSAR) technique based on the ascending and descending Sentinel-1 datasets. Then, the three-dimensional time series deformation field of the glacier was obtained by using the 3D decomposition technique. Furthermore, the three-dimensional movement of the glacier and its material migration process were analyzed. The results showed that the velocities of the Lulang glacier in horizontal and vertical directions were up to 8.0 m/year and 0.45 m/year, and these were basically consistent with the movement rate calculated from the historical optical images. Debris on both sides of the slope accumulated in the channel after slipping, and the material loss of the three provenances reached 6–9 × 103 m3/year, while the volume of the glacier also decreased by about 76 × 103 m3/year due to snow melting and evaporation. The correlation between the precipitation, temperature, and surface velocity suggests that glacier velocity has a clear association with them, and the activity of glaciers is linked to climate change. Therefore, in the context of global warming, the glacier movement speed will gradually increase with the annual increase in temperature, resulting in debris flow disasters in the future summer high-temperature period. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area and Google Earth image.</p>
Full article ">Figure 2
<p>Workflow of improved SBAS-InSAR technique (refer to the <a href="#sec3-land-13-01126" class="html-sec">Section 3</a> for the meaning of each methodological step).</p>
Full article ">Figure 3
<p>The distributions of temporal and perpendicular baselines of ascending and descending interference pairs used in this study: (<b>a</b>,<b>b</b>) show ascending and descending Sentinel data. The horizontal and vertical axes are the temporal and perpendicular baseline, respectively.</p>
Full article ">Figure 4
<p>Workflow of 3D time series displacement decomposition (refer to the <a href="#sec3-land-13-01126" class="html-sec">Section 3</a> for the meaning of each methodological step).</p>
Full article ">Figure 5
<p>Ascending, descending, and three-dimensional deformation field of the glacier: (<b>a</b>,<b>b</b>) positive value (red) represents moving away from the satellite, subsidence; (<b>c</b>–<b>d</b>,<b>f</b>) positive value (red) represents the movement toward down east and south; (<b>e</b>) Google image map, red dashed line is the study area boundary.</p>
Full article ">Figure 6
<p>Three-dimensional displacement for glaciers from October 2017 to August 2020. Color represents the vertical displacement and black arrows represent the horizontal displacement.</p>
Full article ">Figure 7
<p>The 3D velocities and elevations of glacier section line A-A’ in <a href="#land-13-01126-f005" class="html-fig">Figure 5</a>e, where red, yellow, and blue lines represent vertical, east, and north annual velocity, respectively; the green line represents the trend line of the displacement.</p>
Full article ">Figure 8
<p>Glacial landform features and volume changes. (<b>a</b>–<b>c</b>) are Google Earth images of local areas (EA, ES, and ED), red dotted circle is a collapse; (<b>d</b>,<b>f</b>) volume changes in the glacier; (<b>e</b>) monitoring points.</p>
Full article ">Figure 9
<p>Optical images reveal historical glacier displacement information. (<b>a</b>–<b>e</b>) image maps for different periods. (<b>f</b>) maximum and minimum horizon displacement.</p>
Full article ">Figure 10
<p>The relationship between climate change and time series of surface deformation.</p>
Full article ">
22 pages, 5834 KiB  
Article
Changes in Snow Cover and Its Surface Temperature across the Tibetan Plateau Region from 2000 to 2020
by Zhihan Li, Qikang Chen, Zhuoying Deng, Minjie Yang, Qi Zhou and Hengming Zhang
Water 2024, 16(15), 2073; https://doi.org/10.3390/w16152073 - 23 Jul 2024
Viewed by 911
Abstract
Currently, the global climate system is complex and ever-changing, with multiple factors influencing climate change. The Qinghai–Tibet Plateau, known as the “Third Pole” of the Earth, is particularly sensitive to global climate change. Without timely and scientific research on the ecological environment of [...] Read more.
Currently, the global climate system is complex and ever-changing, with multiple factors influencing climate change. The Qinghai–Tibet Plateau, known as the “Third Pole” of the Earth, is particularly sensitive to global climate change. Without timely and scientific research on the ecological environment of the Qinghai–Tibet Plateau and without summarizing relevant adaptive strategies, global climate change will impact the sustainable development of the plateau. This study utilized Landsat remote sensing images from 2000 to 2020 to extract the snow cover area and snow temperature of the Qinghai–Tibet Plateau using the snow frequency threshold method. The study analyzed the spatiotemporal characteristics of snow cover and temperature over the 20-year period and investigated some of the climate and topographical driving factors influencing their changes. The results showed that from 2000 to 2020, the permanent snow cover area in the Qinghai–Tibet Plateau region showed a fluctuating decreasing trend, reducing from approximately 12.34 thousand km2 to around 9.01 thousand km2; the permanent snow temperature showed an initial increase followed by a decrease during the same period. The highest annual average snow temperature was approximately −3.478 °C, while the lowest annual average temperature was around −8.150 °C. Over the 20-year period, the snow cover area in the plateau was negatively correlated with temperature and precipitation, while snow temperature was positively correlated with temperature and precipitation. The snow cover in the weak wind areas of the plateau showed a significant reduction. Areas with higher average wind speeds, such as shaded slopes and semi-shaded slopes, had larger snow cover areas. These research findings provide important insights into the protection and management of the ecological environment of the Qinghai–Tibet Plateau. Full article
(This article belongs to the Special Issue Cold Region Hydrology and Hydraulics)
Show Figures

Figure 1

Figure 1
<p>Qinghai–Tibet Plateau region.</p>
Full article ">Figure 2
<p>Number of images.</p>
Full article ">Figure 3
<p>Comparison of snow extraction before and after water removal. (<b>a</b>) Snow extraction results before water removal. (<b>b</b>) Snow extraction results after water removal.</p>
Full article ">Figure 4
<p>Comparison between the trend of satellite image quantity and snow cover area change.</p>
Full article ">Figure 5
<p>Diagram showing changes in permanent snow distribution. (<b>a</b>) The year 2000. (<b>b</b>) The year 2005. (<b>c</b>) The year 2010. (<b>d</b>) The year 2020.</p>
Full article ">Figure 6
<p>Interannual variation of permanent snow cover area on the Tibetan Plateau.</p>
Full article ">Figure 7
<p>Interannual variation of permanent snow temperature in the Tibetan Plateau. Due to the lack of data for the years 2002 and 2012, the temperature conditions for these two years are not shown in the graph.</p>
Full article ">Figure 8
<p>Map of snow temperature in the Tibetan Plateau in 2020.</p>
Full article ">Figure 9
<p>Interannual variation of temperature in the Tibetan Plateau.</p>
Full article ">Figure 10
<p>Interannual variation of precipitation in the Tibetan Plateau.</p>
Full article ">Figure 11
<p>Statistical chart of average temperature on different aspects.</p>
Full article ">Figure 12
<p>Relationship between elevation, snow cover area, and temperature at different elevations. Note: The linear trend in the graph represents the trend of temperature.</p>
Full article ">Figure 13
<p>Changes in snow distribution in different wind speed intervals.</p>
Full article ">Figure 14
<p>Distribution map of monsoon affecting the Qinghai–Tibet Plateau.</p>
Full article ">
15 pages, 6472 KiB  
Article
Spatio-Temporal Behavior of Land Surface Temperatures (LSTs) in Central Chile, Using Terra MODIS Images
by Pedro Muñoz-Aguayo, Luis Morales-Salinas, Roberto Pizarro, Alfredo Ibáñez, Claudia Sangüesa, Guillermo Fuentes-Jaque, Cristóbal Toledo and Pablo A. Garcia-Chevesich
Hydrology 2024, 11(7), 103; https://doi.org/10.3390/hydrology11070103 - 12 Jul 2024
Viewed by 1546
Abstract
Land surface temperature (LST) is one of the most important variables in the physical processes of surface energy and water balance. The temporal behavior of LST was analyzed between the latitudes 32°00′ S and 34°24′ S (Valparaíso and Metropolitana regions of Chile) for [...] Read more.
Land surface temperature (LST) is one of the most important variables in the physical processes of surface energy and water balance. The temporal behavior of LST was analyzed between the latitudes 32°00′ S and 34°24′ S (Valparaíso and Metropolitana regions of Chile) for three summer months (December, January, and February) in the 2000–2017 period, using the Terra MODIS image information and applying the Mann–Kendall test. The results show an increase in LST in the study area, particularly in the Andes mountain range in January (5240 km2), which mainly comprises areas devoid of vegetation and eternal snow and glaciers, and are zones that act as water reserves for the capital city of Santiago. Similarly, vegetated areas such as forests, grasslands, and shrublands also show increasing trends in LST but over smaller surfaces. Because this study is regional, it is recommended to improve the spatial and temporal resolutions of the images to obtain conclusions on more local scales. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area. Source: Own elaboration.</p>
Full article ">Figure 2
<p>Methodological scheme of the study. Source: Own elaboration. (<a href="https://lpdaac.usgs.gov/dataset_discovery/modis" target="_blank">https://lpdaac.usgs.gov/dataset_discovery/modis</a>, accessed on 15 March 2024).</p>
Full article ">Figure 3
<p>Space–time distribution of areas with significant trends with a 95% confidence level in land surface temperatures (LSTs), based on the <span class="html-italic">p</span>-value of the Mann–Kendall test. (<b>1</b>): December; (<b>2</b>): January; (<b>3</b>): February. Source: Own elaboration from MODIS LST images.</p>
Full article ">Figure 4
<p>Spatial distribution of the linear regression analysis of the LST series for the month of December. (<b>1</b>): Trend values 2000–2016; (<b>2</b>): Positive and negative trends of the 2000–2016 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 5
<p>Spatial distribution of the linear regression analysis of the LST series for the month of January. (<b>1</b>): Trend values 2001–2017; (<b>2</b>): Positive and negative trends of the 2001–2017 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 6
<p>Spatial distribution of the linear regression analysis of the LST series for the month of February. (<b>1</b>): Trend values 2001–2017; (<b>2</b>): Positive and negative trends of the 2001–2017 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 7
<p>(<b>1</b>): Land use for the entire study area; (<b>2</b>): Current land use in areas with significant trends for the month of December; (<b>3</b>): Current land use in areas with significant trends for the month of January; (<b>4</b>): Current land use in areas with significant trends for the month of February. Source: Own elaboration based on the study “Inventory of native vegetational resources of Chile, for Regions V, VI and RM” CONAF [<a href="#B38-hydrology-11-00103" class="html-bibr">38</a>].</p>
Full article ">Figure 8
<p>(<b>1</b>): Geomorphology of the study area; (<b>2</b>): Geomorphology in areas with significant trends for December; (<b>3</b>): Geomorphology in areas with significant trends for January; (<b>4</b>): Geomorphology in areas with significant trends for February. Source: Own elaboration based on the study “Geomorphological units of Chile” [<a href="#B48-hydrology-11-00103" class="html-bibr">48</a>].</p>
Full article ">Figure 9
<p>Behavior of the average LST by year, month, and land use for the analyzed period within the study area. The red line represents the linear trend.</p>
Full article ">
22 pages, 33778 KiB  
Article
Synthetic Aperture Radar Monitoring of Snow in a Reindeer-Grazing Landscape
by Ida Carlsson, Gunhild Rosqvist, Jenny Marika Wennbom and Ian A. Brown
Remote Sens. 2024, 16(13), 2329; https://doi.org/10.3390/rs16132329 - 26 Jun 2024
Viewed by 1064
Abstract
Snow cover and runoff play an important role in the Arctic environment, which is increasingly affected by climate change. Over the past 30 years, winter temperatures in northern Sweden have risen by 2 °C, accompanied by an increase in precipitation. This has led [...] Read more.
Snow cover and runoff play an important role in the Arctic environment, which is increasingly affected by climate change. Over the past 30 years, winter temperatures in northern Sweden have risen by 2 °C, accompanied by an increase in precipitation. This has led to a higher incidence of thaw–freeze and rain-on-snow events. Snow properties, such as the snow depth and longevity, and the timing of snowmelt in spring significantly impact the alpine tundra vegetation. The emergent vegetation at the edge of the snow patches during spring and summer constitutes an essential nutrient supply for reindeer. We have used Sentinel-1 synthetic aperture radar (SAR) to determine the onset of the surface melt and the end of the snow cover in the core reindeer grazing area of the Laevás Sámi reindeer-herding community in northern Sweden. Using SAR data from March to August during the period 2017 to 2021, the start of the surface melt is identified by detecting the season’s backscatter minimum. The end of the snow cover is determined using a threshold approach. A comparison between the results of the analysis of the end of the snow cover from Sentinel-1 and in situ measurements, for the years 2017 to 2020, derived from an automatic weather station located in Laevásvággi reveals a 2- to 10-day difference in the snow-free ground conditions, which indicates that the method can be used to investigate when the ground is free of snow. VH data are preferred to VV data due to the former’s lower sensitivity to temporary wetting events. The outcomes from the season backscatter minimum demonstrate a distinct 25-day difference in the start of the runoff between the 5 investigated years. The backscatter minimum and threshold-based method used here serves as a valuable complement to global snowmelt monitoring. Full article
(This article belongs to the Section Ecological Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Modified illustration based on Buchelt et al. [<a href="#B36-remotesensing-16-02329" class="html-bibr">36</a>], describing the backscatter intensity during the snow melting season derived from S-1 SAR data. The start of the surface melt (SOSM) is marked by the S-1 backscatter reaching its minimum, while the end of the snowmelt (EOS) is indicated by the backscatter starting to reach a higher value after reaching the season’s minimum value [<a href="#B36-remotesensing-16-02329" class="html-bibr">36</a>].</p>
Full article ">Figure 2
<p>The area of interest for this study is the spring and summer grazing area used by reindeer of the Laevás Sámi reindeer-herding community, northern Sweden. The yellow circle marks where the automatic weather station (AWS) in Laevásvággi 18.96°E 68.04°N is located, and the area is also the calving ground for Laevás reindeers [<a href="#B42-remotesensing-16-02329" class="html-bibr">42</a>,<a href="#B44-remotesensing-16-02329" class="html-bibr">44</a>].</p>
Full article ">Figure 3
<p>The acquisitions of S-1 (blue) single-look complex data, in interferometric wide-swath mode, in ascending orbit from the Alaska Satellite Facility [<a href="#B45-remotesensing-16-02329" class="html-bibr">45</a>] downloaded on the 12th of September, 14th of December of 2022 and 6th of May 2024.</p>
Full article ">Figure 4
<p>Workflow for preprocessing the S-1 images from 2017–2021 using the Sentinel Application Platform (SNAP). This process involved obtaining the latest orbit file, splitting swaths, and debursting images, followed by merging them into a single coherent image. Calibration to the backscatter coefficient (<span class="html-italic">β</span><sup>0</sup>) was conducted according to Small (2011). Subsequent steps included multilooking, terrain flattening for pixel location rectification, and Lee Sigma speckle filtering [<a href="#B36-remotesensing-16-02329" class="html-bibr">36</a>] for noise removal. Radiometric terrain correction utilized the range-Doppler technique with a 2 m DEM [<a href="#B44-remotesensing-16-02329" class="html-bibr">44</a>].</p>
Full article ">Figure 5
<p>Simplification (the season backscatter is smoothed by the average backscattering during the season) of the seasonal backscatter in decibels from Sentinel-1 for all five years in both polarisations (<b>a</b>,<b>b</b>). VV polarisation (<b>a</b>) consistently exhibits higher backscatter values throughout the season compared to VH polarisation (<b>b</b>). VV polarisation is preferred for surface features and roughness, while VH polarisation is more suitable for detecting internal structure and volume scattering within targets like vegetation. Notably, in March 2017, there is a period characterized by lower backscatter values in the VV polarisation.</p>
Full article ">Figure 6
<p>Overview of monthly SOSM<sub><span class="html-italic">S</span>-1</sub> in the VV polarisation from 2017 to 2021 in the spring and summer grazing area of Laevás reindeer. In 2017 (<b>a</b>), a substantial SOSM<sub><span class="html-italic">S</span>-1</sub> was detected in the VV polarisation mode, covering 56% of the area in March. In 2018 (<b>b</b>), the SOSM<sub><span class="html-italic">S</span>-1</sub> started in April (red) and May (light yellow) and in May (light yellow) during 2019 (<b>c</b>). In 2020 (<b>d</b>), the SOSM<sub><span class="html-italic">S</span>-1</sub> started in May (light yellow) and June (light blue). During the year 2021 (<b>e</b>), there were large SOSM<sub><span class="html-italic">S</span>-1</sub> in May (light yellow) and in July (blue). These findings suggest that the SOSM<sub><span class="html-italic">S</span>-1</sub> varies significantly across the years, with the VV polarisation mode consistently exhibiting an earlier SOSM than the VH.</p>
Full article ">Figure 7
<p>Overview of the monthly SOSM<sub><span class="html-italic">S</span>-1</sub> in the VH polarisation over the observed years 2017 to 2021. In 2017 (<b>a</b>), the VH exhibited the largest snowmelt in May (light yellow) and June (light blue). In 2018 (<b>b</b>), a significant SOSM<sub><span class="html-italic">S</span>-1</sub> was displayed in April (red) and May (light yellow), as well in May (light yellow) in 2019 (<b>c</b>). Noteworthily, 2020 (<b>d</b>) exhibited pronounced SOSM<sub><span class="html-italic">S</span>-1</sub> peaks in May (light yellow) and June (light blue). In 2021 (<b>e</b>), the highest SOSM was recorded in May (light yellow). These findings highlight the variability in the seasonal snowmelt dynamics captured in the data.</p>
Full article ">Figure 8
<p>Monthly end of snowmelt (EOS<sub><span class="html-italic">S</span>-1</sub>) in the VV polarisation between the years 2017 and 2021 are shown in the figure. In the VV polarisation for 2017 (<b>a</b>), the deviation is evident, with the largest amount of EOS<sub><span class="html-italic">S</span>-1</sub> occurring in March, a pattern not observed in the VH polarisation (<a href="#remotesensing-16-02329-f009" class="html-fig">Figure 9</a>). The year 2018 (<b>b</b>) exhibits an early EOS<sub><span class="html-italic">S</span>-1</sub> in the VV polarisation, indicating bare ground in a significant portion of the area as early as May (light yellow). Moreover, 2019 (<b>c</b>) and 2020 (<b>d</b>) show similar EOS<sub><span class="html-italic">S</span>-1</sub> in June (light blue) and July (blue), while 2019 exhibits some earlier melting, particularly in April (red). In 2021 (<b>e</b>), the EOS<sub><span class="html-italic">S</span>-1</sub> occurrence was notable in May, with a more substantial presence observed in July. Moreover, there are areas within the region where data are not available, as evidenced across all the years.</p>
Full article ">Figure 9
<p>End of season (EOS) observations from 2017 to 2021, as depicted by the VH polarisation. The data reveal significant variations in the EOS percentages across different years and months, with certain trends standing out prominently. For instance, noticeable spikes in the snowmelt are observed in May and June across multiple years, indicating periods of accelerated melting. In 2017 (<b>a</b>), the EOS percentages remained consistently low throughout the observed months, with minimal snowmelt recorded with the largest EOS in June (light blue). In 2018 (<b>b</b>), snowmelt began to appear in April (red) and increased notably in May (light blue). In 2019 (<b>c</b>), the trend of increasing snowmelt continued into 2019, with May (light yellow) showcasing substantial melting percentages. Moreover, 2020 (<b>d</b>) witnessed a pronounced increase in snowmelt compared to previous years, particularly notable in May (light yellow) and June (light blue). In 2021 (<b>e</b>), the EOS percentages displayed a remarkable spike in May (light yellow), indicating a notably accelerated snowmelt compared to previous years.</p>
Full article ">
13 pages, 4156 KiB  
Article
Advancing Insights into Runway De-Icing: Combining Infrared Thermography and Raman Spectroscopy to Assess Ice Melt
by Claire Charpentier, Jean-Denis Brassard, Mario Marchetti and Gelareh Momen
Appl. Sci. 2024, 14(12), 5096; https://doi.org/10.3390/app14125096 - 12 Jun 2024
Viewed by 816
Abstract
The “bare runway” principle aims to ensure passenger and employee safety by making runways more usable during winter conditions, allowing for easier removal of contaminants like snow and ice. Maintaining runway operations in winter is essential, but it involves considerable cost and environmental [...] Read more.
The “bare runway” principle aims to ensure passenger and employee safety by making runways more usable during winter conditions, allowing for easier removal of contaminants like snow and ice. Maintaining runway operations in winter is essential, but it involves considerable cost and environmental impacts. Greater knowledge about the de-icing and anti-icing performance of runway de-icing products (RDPs) optimizes operations. The ice melting test, as per the AS6170 standard, gauges the rate at which an RDP dissolves an ice mass to determine RDP effectiveness. Here, we introduce a novel integrated methodology for assessing RDP-related ice melting. We combine laboratory-based procedures with infrared thermography and Raman spectroscopy to monitor the condition of RDPs placed on ice. The plateau of maximum efficiency, marked by the most significant Raman peak intensity, corresponds to the peak minimum temperature, indicating optimal RDP performance. Beyond this point, RDP efficacy declines, and the system temperature, including melted contaminants and RDP, approaches ambient temperature. Effective RDP performance persists when the ambient temperature exceeds the mixture’s freezing point; otherwise, a freezing risk remains. The initial phases of RDP–ice contact involve exothermic reactions that generate brine, followed by heat exchange with surrounding ice to encourage melting. The final phase is complete ice melt, leaving only brine with reduced heat exchange on the surface. By quantifying these thermal and chemical changes, we gain a deeper understanding of RDP-related ice melting, and a more robust assessment can be provided to airports using RDPs. Full article
(This article belongs to the Section Aerospace Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>The ice melting test for a runway de-icing product (RDP) following the AS6170 standard. Panel (<b>A</b>) depicts the initial ice sample in a Petri dish and the placing of the RDP (sodium formate) onto the ice sample. In panel (<b>B</b>), the RDP and melted ice are removed using compressed air [<a href="#B20-applsci-14-05096" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Sodium formate (NaFo) over a concrete pavement runway surface, illustrating the application concentration used at Mirabel airport, Québec.</p>
Full article ">Figure 3
<p>Experimental setup for thermometric imaging and Raman spectroscopy.</p>
Full article ">Figure 4
<p>Ice melting rate of the 5 g and 0.5 g sodium formate (NaFo) samples at −10 °C.</p>
Full article ">Figure 5
<p>Minimum temperature of the sodium formate (NaFo) at the ice surface for the 0.5 g (<b>top</b>) and 5 g tests (<b>bottom</b>) and the corresponding thermal photographs of the ice surface at −10 °C during the ice melting test at 0, 1, and 10 min, as well as at the min.</p>
Full article ">Figure 6
<p>Raman spectra of the runway de-icing product obtained after 125, 293, 608, and 1196 s (laser at 785 nm, integration time of 20 s, average of four spectra), with the identification of the main Raman peaks.</p>
Full article ">Figure 7
<p>Thermal pictures during Raman spectroscopy of the RDP (sodium formate, NaFo) with a Raman probe at (<b>A</b>) 125 s, (<b>B</b>) 293 s, (<b>C</b>) 608 s, and (<b>D</b>) 1196 s.</p>
Full article ">Figure 8
<p>Coupled thermographic and spectral analyses; (<b>A</b>) maximum intensity of a Raman peak at 930 cm<sup>−1</sup> and minimum temperature during the test with 0.05 mL of the RDP sodium formate (NaFo) at −10 °C; (<b>B</b>) RDP with rhodamine B melting on ice at −10 °C; and (<b>C</b>) a 2D representation of the RDP de-icing phenomenon.</p>
Full article ">
23 pages, 16889 KiB  
Article
Mapping Debris-Covered Glaciers Using High-Resolution Imagery (GF-2) and Deep Learning Algorithms
by Xin Yang, Fuming Xie, Shiyin Liu, Yu Zhu, Jinghui Fan, Hongli Zhao, Yuying Fu, Yunpeng Duan, Rong Fu and Siyang Guo
Remote Sens. 2024, 16(12), 2062; https://doi.org/10.3390/rs16122062 - 7 Jun 2024
Viewed by 941
Abstract
Glacier inventories are fundamental in understanding glacier dynamics and glacier-related environmental processes. High-resolution mapping of glacier outlines is lacking, although high-resolution satellite images have become available in recent decades. Challenges in development of glacier inventories have always included accurate delineation of boundaries of [...] Read more.
Glacier inventories are fundamental in understanding glacier dynamics and glacier-related environmental processes. High-resolution mapping of glacier outlines is lacking, although high-resolution satellite images have become available in recent decades. Challenges in development of glacier inventories have always included accurate delineation of boundaries of debris-covered glaciers, which is particularly true for high-resolution satellite images due to their limited spectral bands. To address this issue, we introduced an automated, high-precision method in this study for mapping debris-covered glaciers based on 1 m resolution Gaofen-2 (GF-2) imagery. By integrating GF-2 reflectance, topographic features, and land surface temperature (LST), we used an attention mechanism to improve the performance of several deep learning network models (the U-Net network, a fully convolutional neural network (FCNN), and DeepLabV3+). The trained models were then applied to map the outlines of debris-covered glaciers, at 1 m resolution, in the central Karakoram regions. The results indicated that the U-Net model enhanced with the Convolutional Block Attention Module (CBAM) outperforms other deep learning models (e.g., FCNN, DeepLabV3+, and U-Net model without CBAM) in terms of precision for supraglacial debris identification. On the testing dataset, the CBAM-enhanced U-Net model achieved notable performance metrics, with its accuracy, F1 score, mean intersection over union (MIoU), and kappa coefficient reaching 0.93, 0.74, 0.79, and 0.88. When applied at the regional scale, the model even exhibits heightened precision (accuracies = 0.94, F1 = 0.94, MIoU = 0.86, kappa = 0.91) in mapping debris-covered glaciers. The experimental glacier outlines were accurately extracted, enabling the distinction of supraglacial debris, clean ice, and other features on glaciers in central Karakoram using this trained model. The results for our method revealed differences of 0.14% for bare ice and 10.36% against the manually interpreted glacier boundary for supraglacial debris. Comparison with previous glacier inventories revealed raised precisions of 8.74% and 4.78% in extracting clean ice and with supraglacial debris, respectively. Additionally, our model demonstrates exceptionally high exclusion for bare rock outside glaciers and could reduce the influence of non-glacial snow on glacier delineation, showing substantial promise in mapping debris-covered glaciers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The central Karakoram study area. The training dataset was obtained from the area in the red boxes; the green boxes mark the test data region. (<b>a</b>) The study area of the High Mountain Asia DEM. (<b>b</b>) The Landsat 8 RGB image. (<b>c</b>) The GF-2 RGB image.</p>
Full article ">Figure 2
<p>Framework of debris-covered glacier mapping in this study.</p>
Full article ">Figure 3
<p>The procession of Sample production. Sample subset data on the composite image of GF-2 (<b>a</b>), LST (<b>b</b>) from Landsat 8, and the slope (<b>c</b>) from ASTER GDEM. The procedure for preprocessing is visually displayed for cropping, labeling, and augmentation based on a sample chip size of 512 × 512.</p>
Full article ">Figure 4
<p>Partial training samples. The RGB image is in the top half and manually labeled true values are in the bottom half. In the labeled sections, gray sections represent supraglacial debris, white sections represent glaciers, and black sections represent non-glacial areas.</p>
Full article ">Figure 5
<p>Different model training accuracy and loss changes. The upper half is the training accuracy and the lower half is the training loss. The horizontal coordinate represents the epoch, with a total of 100 epochs. The extent of convergence of the model can be deduced by observing the fluctuation in accuracy.</p>
Full article ">Figure 6
<p>Different model validation accuracy and loss changes. The upper half is the validation accuracy and the lower half is the validation loss. The horizontal axis represents the number of training iterations.</p>
Full article ">Figure 7
<p>Visualization of the prediction results of the different models on the test set. (<b>a</b>,<b>b</b>) Bare ice boundaries are clearly visible. (<b>c</b>) Covering seasonal snow. (<b>d</b>,<b>e</b>) The sample contains cryoconite with an unclear boundary with the supraglacial debris; Xception and ResNet-34 are representative of DeepLabV3+ for different feature extraction networks, respectively. CBAM denotes the introduction of the Convolutional Block Attention Module.</p>
Full article ">Figure 8
<p>Visualization of debris-covered glacier extractions at glacier terminus. (<b>a</b>) The GF-2 image, (<b>b</b>) the ground truth, (<b>c</b>) DeepLabV3+(Xception), (<b>d</b>) DeepLabV3+(Xception)+ CBAM, (<b>e</b>) DeepLabV3+(ResNet-34)+ CBAM, (<b>f</b>) FCNN, (<b>g</b>) U-Net, (<b>h</b>) U-Net with CBAM. The green and purple boxes represent high-reflectivity debris and glaciers with cryoconite or weak fouling, respectively.</p>
Full article ">Figure 9
<p>Visualization results of debris-covered glacier extraction in the predicted region. (<b>a</b>) The RGB image, (<b>b</b>) the ground truth, (<b>c</b>) DeepLabV3+(Xception), (<b>d</b>) DeepLabV3+(Xception)+ CBAM, (<b>e</b>) DeepLabV3+(ResNet-34)+ CBAM, (<b>f</b>) FCNN, (<b>g</b>) U-Net, (<b>h</b>) U-Net with CBAM. The red and blue boxes represent low-reflectivity debris and mixed-type supraglacial debris, respectively.</p>
Full article ">Figure 10
<p>Identification of supraglacial debris with different reflectance characteristics. (<b>a</b>) Mixed types of supraglacial debris, (<b>b</b>) low-reflectivity debris composed predominantly of mineral or clay particles, (<b>c</b>) glaciers with cryoconite or weak fouling, (<b>d</b>) high-reflectivity debris with a significant presence of ice crystals or flakes.</p>
Full article ">Figure 11
<p>Comparison of area statistics extracted by various models on typical glaciers. It includes areas of individual and summed bare ice and supraglacial debris, differences in total area extracted by distinct models, and corresponding percentages. (<b>a</b>–<b>c</b>) Glacier units segmented by ridge lines.</p>
Full article ">Figure 12
<p>The results of our extraction on typical glaciers are compared with the KGIs. (<b>a</b>,<b>b</b>) An RGB image of a typical glacier, (<b>c</b>,<b>d</b>) our extraction results, (<b>e</b>,<b>f</b>) the KGI glacier range.</p>
Full article ">Figure 13
<p>Comparison of our results with the KGI at the end of the glacier. KGI refers to the Karakoram glacier inventories produced in 2023.</p>
Full article ">Figure 14
<p>Spatial distribution characteristics of LST and slope in glaciated and non-glaciated areas. (<b>a</b>,<b>b</b>) The GF-2 images of the test area. (<b>c</b>,<b>d</b>) LST: the LST of glacial regions is primarily below 13 °C, with most bare ice areas having an LST below 0 °C. (<b>e</b>,<b>f</b>) Slope: the glacial area has a predominant concentration of slopes between 0 and 50°.</p>
Full article ">Figure 15
<p>Extraction results of different metrics based on U-Net with CBAM. (<b>a</b>) GF-2 images of the test area. (<b>b</b>) Manually revised true values. (<b>c</b>) Results of metric extraction, combining spectrum with slope and LST. (<b>d</b>) The results of the extraction are derived from a combination of spectral and slope features. (<b>e</b>) The results of the extraction are derived from a combination of spectral and LST features. (<b>f</b>) Extraction results from the combination of spectral features.</p>
Full article ">
22 pages, 25403 KiB  
Article
A Comprehensive Assessment of Climate Change and Anthropogenic Effects on Surface Water Resources in the Lake Urmia Basin, Iran
by Mohammad Kazemi Garajeh, Rojin Akbari, Sepide Aghaei Chaleshtori, Mohammad Shenavaei Abbasi, Valerio Tramutoli, Samsung Lim and Amin Sadeqi
Remote Sens. 2024, 16(11), 1960; https://doi.org/10.3390/rs16111960 - 29 May 2024
Cited by 3 | Viewed by 1425
Abstract
In recent decades, the depletion of surface water resources within the Lake Urmia Basin (LUB), Iran, has emerged as a significant environmental concern. Both anthropogenic activities and climate change have influenced the availability and distribution of surface water resources in this area. This [...] Read more.
In recent decades, the depletion of surface water resources within the Lake Urmia Basin (LUB), Iran, has emerged as a significant environmental concern. Both anthropogenic activities and climate change have influenced the availability and distribution of surface water resources in this area. This research endeavors to provide a comprehensive evaluation of the impacts of climate change and anthropogenic activities on surface water resources across the LUB. Various critical climatic and anthropogenic factors affecting surface water bodies, such as air temperature (AT), cropland (CL), potential evapotranspiration (PET), snow cover, precipitation, built-up areas, and groundwater salinity, were analyzed from 2000 to 2021 using the Google Earth Engine (GEE) cloud platform. The JRC-Global surface water mapping layers V1.4, with a spatial resolution of 30 m, were employed to monitor surface water patterns. Additionally, the Mann–Kendall (MK) non-parametric trend test was utilized to identify statistically significant trends in the time series data. The results reveal negative correlations of −0.56, −0.89, −0.09, −0.99, and −0.79 between AT, CL, snow cover, built-up areas, and groundwater salinity with surface water resources, respectively. Conversely, positive correlations of 0.07 and 0.12 were observed between precipitation and PET and surface water resources, respectively. Notably, the findings indicate that approximately 40% of the surface water bodies in the LUB have remained permanent over the past four decades. However, there has been a loss of around 30% of permanent water resources, transitioning into seasonal water bodies, which now account for nearly 13% of the total. The results of our research also indicate that December and January are the months with the most water presence over the LUB from 1984 to 2021. This is because these months align with winter in the LUB, during which there is no water consumption for the agriculture sector. The driest months in the study area are August, September, and October, with the presence of water almost at zero percent. These months coincide with the summer and autumn seasons in the study area. In summary, the results underscore the significant impact of human activities on surface water resources compared to climatic variables. Full article
(This article belongs to the Special Issue Natural Hazard Mapping with Google Earth Engine)
Show Figures

Figure 1

Figure 1
<p>Location of the study area in (<b>a</b>) Iran basins and (<b>b</b>) ULB.</p>
Full article ">Figure 2
<p>Various predisposing variables for monitoring climate change and anthropogenic effects on surface water resources, including: (<b>a</b>) AT for the years 2000, 2010, 2015, and 2020; (<b>b</b>) LC for the years 2000, 2010, 2015, and 2020; (<b>c</b>) PET for the years 2000, 2010, 2015, and 2020; (<b>d</b>) snow cover for the years 2000, 2010, 2015, and 2020; (<b>e</b>) precipitation for the years 2000, 2010, 2015, and 2020; (<b>f</b>) built-up area for the years 2000, 2010, 2015, and 2020; and (<b>g</b>) groundwater salinity for the years 2000, 2010, 2015, and 2020.</p>
Full article ">Figure 2 Cont.
<p>Various predisposing variables for monitoring climate change and anthropogenic effects on surface water resources, including: (<b>a</b>) AT for the years 2000, 2010, 2015, and 2020; (<b>b</b>) LC for the years 2000, 2010, 2015, and 2020; (<b>c</b>) PET for the years 2000, 2010, 2015, and 2020; (<b>d</b>) snow cover for the years 2000, 2010, 2015, and 2020; (<b>e</b>) precipitation for the years 2000, 2010, 2015, and 2020; (<b>f</b>) built-up area for the years 2000, 2010, 2015, and 2020; and (<b>g</b>) groundwater salinity for the years 2000, 2010, 2015, and 2020.</p>
Full article ">Figure 3
<p>Changes observed in various surface water resources from 1984 to 2021 depicted in the JRC Global Surface Water Mapping Layers, version 1.4.</p>
Full article ">Figure 4
<p>Changes observed in various surface water resources in different months from 1984 to 2021 depicted in the JRC Global Surface Water Mapping Layers, version 1.4.</p>
Full article ">Figure 5
<p>An overview of the applied methodology.</p>
Full article ">Figure 6
<p>Correlation between different climatic and anthropogenic variables and surface water resources across the LUB, analyzed using the Pearson correlation heatmap.</p>
Full article ">Figure 7
<p>Time series analysis depicting the annual variations and trends in surface water bodies and climatic and anthropogenic variables, including AT, CL, PET, snow cover, precipitation, built-up areas, and groundwater salinity, throughout the LUB. The significance of the trend lines was determined using the Mann–Kendall test. The trend line slope is based on Sen’s slope estimator. In cases where a time series experienced a significant abrupt shift, identified by Buishand’s test, the change year is indicated. For these instances, “mu” denotes the average of the sub-series.</p>
Full article ">Figure 7 Cont.
<p>Time series analysis depicting the annual variations and trends in surface water bodies and climatic and anthropogenic variables, including AT, CL, PET, snow cover, precipitation, built-up areas, and groundwater salinity, throughout the LUB. The significance of the trend lines was determined using the Mann–Kendall test. The trend line slope is based on Sen’s slope estimator. In cases where a time series experienced a significant abrupt shift, identified by Buishand’s test, the change year is indicated. For these instances, “mu” denotes the average of the sub-series.</p>
Full article ">Figure 8
<p>Surface water changes in various surface water resources from 1984 and 2000 to 2022 obtained from Landsat 5, 7, and 8 series images.</p>
Full article ">Figure 9
<p>Changes in surface water body transitions from 1984 to 2021 in the LUB [<a href="#B62-remotesensing-16-01960" class="html-bibr">62</a>].</p>
Full article ">
22 pages, 5382 KiB  
Article
Development and Application of the Snow, Soil Water and Water Balance Model (SNOSWAB), an Online Model for Daily Estimation of Snowpack Processes, Soil Water Content and Soil Water Balance
by Serban Danielescu
Water 2024, 16(11), 1503; https://doi.org/10.3390/w16111503 - 24 May 2024
Viewed by 1544
Abstract
SNOSWAB (Snow, Soil Water and Water Balance) is a unique online deterministic model built using tipping-bucket approaches that allows for the daily estimation of (i) snowpack processes; (ii) soil water content; and (iii) soil water budget. SNOSWAB is most suitable for modeling field-scale [...] Read more.
SNOSWAB (Snow, Soil Water and Water Balance) is a unique online deterministic model built using tipping-bucket approaches that allows for the daily estimation of (i) snowpack processes; (ii) soil water content; and (iii) soil water budget. SNOSWAB is most suitable for modeling field-scale processes for vertically and horizontally homogeneous soils, and its applicability is not limited to specific climate zones or geographical areas. The model is freely available, and its streamlined online interface integrates powerful calibration, visualization and data export routines. In this study, SNOSWAB development and a conceptual model, as well as an example of its application using data collected during a 12-year (2008–2019) field study conducted at the Agriculture and Agri-Food Canada Harrington Experimental Farm (HEF) on Prince Edward Island (PEI), Canada, are presented. Input data consisting of daily air temperature, total precipitation, rainfall and evapotranspiration were used in conjunction with soil properties and daily soil water content, snowpack thickness, surface runoff and groundwater recharge to calibrate (2010–2014) and validate (2015–2019) the model. For both the calibration and validation simulations, the statistical indicators used for evaluating model performance indicated, in most cases, high model fitness (i.e., R2 > 0.5, NRMSE < 50% and −25% < PBIAS < 25%) for the various time intervals and parameters analyzed. SNOSWAB fills an existing gap in the online environment and, due to its ease of use, robustness and flexibility, shows promise to be adopted as an alternative for more complex, standalone models that might require extensive resources and expertise. Full article
(This article belongs to the Section Soil and Water)
Show Figures

Figure 1

Figure 1
<p>Simplified workflow diagram of the SNOSWAB model.</p>
Full article ">Figure 2
<p>Simplified workflow diagram for the SNOSWAB SNOW module (green boxes—input data, orange boxes—module coefficients, red boxes—output variables, yellow box—connection with a different calculation module; TOTPP—daily total precipitation input data, ETA—daily actual evapotranspiration input data, RAIN—daily rain input data, SNOF—snowfall amount, CFets—correction factor for portion of evapotranspiration occurring in the soil, THRets—soil water content threshold for stopping soil evapotranspiration when the soil is dry, THRrs—air temperature threshold for rain to be accumulated as snow, CFRsm—correction factor for snowmelt due to rain, SNMT—snowmelt due to temperature, THRsm—air temperature threshold for initiating snowmelt, CFTsm—correction factor for snowmelt due to air temperature, SNMR—snowmelt due to rain, Etas—above-soil ET, ETicds—soil ET, SNTF—snow layer thickness, SNMF—[net] snowmelt, WATisrf—water available for infiltration or surface runoff; WBM—WATER BALANCE module; dashed lines indicate amounts transferred to WBM).</p>
Full article ">Figure 3
<p>Simplified workflow diagram for the SNOSWAB WATER BALANCE module (yellow box—connection with a different calculation module, orange boxes—module coefficients, red boxes—output variables; WATisrf—water available for infiltration or surface runoff, ETicds—soil ET, THRinfLH—SWC threshold for switching between low and high infiltration rates, INFlr—infiltration rate at low SWC, INFhr—infiltration rate at high SWC, THRdraLH—SWC threshold for switching drainage from high to low rates, THRets—SWC threshold for stopping soil evapotranspiration, THRswstd—SWC threshold for stopping drainage, DRAlr—drainage rate at low SWC, DRAhr—drainage rate at high SWC, THRstd—air temperature threshold for stopping drainage; INFact—actual infiltration, SWCfin—soil water content, DRAact—total drainage, PORe—soil effective porosity, INFcap—infiltration capacity, DRAcap—drainage capacity, SRact—total surface runoff, CFeidr—drainage correction factor for excess infiltration, CFosdr—drainage correction factor for soil oversaturation; dashed lines indicate amounts transferred from the SNOW module).</p>
Full article ">Figure 4
<p>Examples of comparisons between UCD and SNOSWAB model outputs for the calibration (left) and validation (right) simulations. (<b>a</b>) Daily snow layer thickness (winter 2010/11); (<b>b</b>) daily snow layer thickness (winter 2014/15); (<b>c</b>) daily soil water content (2010); (<b>d</b>) daily soil water content (2017); (<b>e</b>) monthly surface runoff (2010–2014); (<b>f</b>) monthly surface runoff (2015–2019); (<b>g</b>) monthly drainage (2010–2014); (<b>h</b>) monthly drainage (2015–2019).</p>
Full article ">Figure 5
<p>Multi-year daily averages for snow layer thickness and net snow layer gains and losses.</p>
Full article ">Figure 6
<p>Multi-year monthly averages for total precipitation (TP), snowmelt (SNMF), actual evapotranspiration (ETA) and soil water content (SWC).</p>
Full article ">Figure 7
<p>Multi-year monthly averages for number of days with soil water content (SWC) in low, high and optimal ranges for potato crop.</p>
Full article ">Figure 8
<p>Multi-year monthly averages of key soil water budget terms (INFact—infiltration; DRAact—drainage; SRact—surface runoff; ETcds—soil evapotranspiration; SWC fin—soil water content).</p>
Full article ">
19 pages, 4258 KiB  
Article
Investigating Road Ice Formation Mechanisms Using Road Weather Information System (RWIS) Observations
by Menglin Jin and Douglas G. McBroom
Climate 2024, 12(5), 63; https://doi.org/10.3390/cli12050063 - 2 May 2024
Viewed by 1645
Abstract
Ice formation on roads leads to a higher incidence of accidents and increases winter de-icing/anti-icing costs. This study analyzed 3 years (2019–2021) of Road Weather Information System (RWIS) sub-hourly measurements collected by the Montana Department of Transportation (MDT) to understand the first-order factors [...] Read more.
Ice formation on roads leads to a higher incidence of accidents and increases winter de-icing/anti-icing costs. This study analyzed 3 years (2019–2021) of Road Weather Information System (RWIS) sub-hourly measurements collected by the Montana Department of Transportation (MDT) to understand the first-order factors of road ice formation and its mechanisms. First, road ice is formed only when the road pavement surface temperature is equal to or below the freezing point (i.e., 32 °F (i.e., 0 °C)), while the corresponding 2 m air temperature could be above 32 °F. Nevertheless, when the road pavement was below 32 °F ice often did not form on the roads. Therefore, one challenge is to know under what conditions road ice forms. Second, the pavement surface temperature is critical for road ice formation. The clear road (i.e., with no ice or snow) surface pavement temperature is generally warmer than the air temperature during both day and night. This feature is different from a natural land surface, where the land skin temperature is lower than the air temperature on cloud-free nights due to radiative cooling. Third, subsurface temperature, measured using a RWIS subsurface sensor below a road surface, did not vary as much as the pavement temperature and, thus, may not be a good index for road ice formation. Fourth, urban heat island effects lead to black ice formation more frequently than roads located in other regions. Fifth, evaporative cooling from the water surface near a road segment further reduces the outlying air temperature, a mechanism that increases heat loss for bridges or lake-side roads in addition to radiative cooling. Additionally, mechanical lifting via mountains and hills is also an efficient mechanism that makes the air condense and, consequently, form ice on the roads. Forecasting road ice formation is in high demand for road safety. These observed features may help to develop a road ice physical model consisting of functions of hyper-local weather conditions, local domain knowledge, the road texture, and geographical environment. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>). Concept profile of cross section of highway. (Modified after <a href="https://pubs.usgs.gov/fs/2006/3127/2006-3127.pdf" target="_blank">https://pubs.usgs.gov/fs/2006/3127/2006-3127.pdf</a> (accessed on 24 April 2024)). (<b>b</b>). A typical section or roadway for two-lane road with shoulders (Source: <a href="https://www.mdt.mt.gov/business/contracting/detailed-drawings.aspx" target="_blank">https://www.mdt.mt.gov/business/contracting/detailed-drawings.aspx</a> (accessed on 24 April 2024)).</p>
Full article ">Figure 2
<p>(<b>a</b>) Shows the locations of the 73 RWIS sites with the selected site shown in red. (<b>b</b>) Presents the observed air temperature (in red) measured using the RWIS sensor and the North American Mesoscale Forecast System (NAM) model forecasted surface 2 m air temperature (in green). Selecting different RWIS sites from the window, (<b>a</b>) displays corresponding observations measured at the selected site and corresponding NAM model forecast. (<b>a</b>) was built based using Google Maps. The blue dashed line in (<b>b</b>) is 32 °F (i.e., 0 °C).</p>
Full article ">Figure 3
<p>RWIS observations on (<b>a</b>) MacDonald Pass, a mountain site, January 2020; the dashed black line is 32 °F (i.e., 0 °C). (<b>b</b>) Bozeman Pass, MT, a site at canyon 10 miles east of city Bozeman for March 2020; (<b>c</b>) is the same as (<b>b</b>) Bozeman Pass, MT, except occurring in March 2019; and (<b>d</b>) is Gray Cooper Bridge, near the city of Great Falls for January 2020. The <span class="html-italic">x</span>-axis gives the temperature data entries; the <span class="html-italic">y</span>-axis is the 2 m air temperature (unit in °F). The green star marks precipitation. Only data when the road status was reposted as “ice warning/ice watch” were analyzed here. RWIS data have intervals of about 5–10 min. The data are available at <a href="https://rwis.mdt.mt.gov" target="_blank">https://rwis.mdt.mt.gov</a> (URL accessible on 30 April 2024).</p>
Full article ">Figure 4
<p>RWIS observations at Avon North site of Montana. (<b>a</b>) Time series for all ice reported times with the road pavement surface temperature, 2 m air temperature, and precipitation time for January 2020. The green star denotes when precipitation occurred. (<b>b</b>) The road pavement surface temperature vs. the 2 m air temperature, with blue representing all road ice without precipitation occurrence and green for road ice with precipitation. The vertical and horizontal dashed lines mark 32 °F (i.e., 0 °C) on both axes.</p>
Full article ">Figure 5
<p>The road pavement surface temperature and the 2 m air temperature relations in March 2020 for icy roads on Bozeman Pass site, Montana. The sub-hourly measurements from the RWIS site road-embedded sensor are analyzed. Dashed line marks 32 °F (i.e., 0 °C).</p>
Full article ">Figure 6
<p>The road pavement surface temperature (blue line), 2 m air temperature (purple dash line), and precipitation (green marker) for no ice road during March 2019 for Bozeman Pass, MT. The data are RWIS sub-hourly observations. Only no ice times are sampled and analyzed.</p>
Full article ">Figure 7
<p>Gary Cooper Bridge for ice-occurrence cases in (<b>a</b>) February 2020 and (<b>b</b>) March 2020. Data are from RWIS site Gary Cooper Bridge, Montana. Only icy road cases were analyzed. The green markers depict precipitation-occurring times and blue markers depict no precipitation occurring at the same time.</p>
Full article ">Figure 8
<p>Subsurface temperature, pavement surface temperature, air temperature, and precipitation for (<b>a</b>) MacDonald Pass, MT in January 2020 and (<b>b</b>) Bozeman Pass, MT in January 2020. Observations were from RWIS sites for only road-ice-occurring times.</p>
Full article ">Figure 9
<p>Ice cases for MacDonald Pass site, MT, in January 2020. The road pavement surface temperature, 2 m air temperature, dew point, RH, and precipitation presence are shown. Dashed black line depicts 32 °F (i.e., 0 °C). This figure is similar to <a href="#climate-12-00063-f002" class="html-fig">Figure 2</a>a, except that RH and dew point data are displayed.</p>
Full article ">Figure 10
<p>Georgetown Lake RWIS site (Site ID: 267009) at Montana Highway 1 MP 25.4 and Highway 1. (<b>a</b>) No ice: the lake is clear at the right side of the road. (<b>b</b>) 15 April 2020: the lake is covered by ice and snow.</p>
Full article ">Figure 11
<p>The 2 m air temperature and road pavement surface temperature relationships for 1 November 2020 to 31 March 2021, for Georgetown Lake (263000). The blue dots represent no ice and the red dot represents ice. The green line represents the least-square regression for all ice and no ice cases (1.31 × T<sub>air</sub> for day time (<b>a</b>) and 1.20 × T<sub>air</sub> for nighttime, (<b>b</b>)).</p>
Full article ">
18 pages, 9425 KiB  
Article
Two-Decadal Glacier Changes in the Astak, a Tributary Catchment of the Upper Indus River in Northern Pakistan
by Muzaffar Ali, Qiao Liu and Wajid Hassan
Remote Sens. 2024, 16(9), 1558; https://doi.org/10.3390/rs16091558 - 27 Apr 2024
Viewed by 1385
Abstract
Snow and ice melting in the Upper Indus Basin (UIB) is crucial for regional water availability for mountainous communities. We analyzed glacier changes in the Astak catchment, UIB, from 2000 to 2020 using remote sensing techniques based on optical satellite images from Landsat [...] Read more.
Snow and ice melting in the Upper Indus Basin (UIB) is crucial for regional water availability for mountainous communities. We analyzed glacier changes in the Astak catchment, UIB, from 2000 to 2020 using remote sensing techniques based on optical satellite images from Landsat and ASTER digital elevation models. We used a surface feature-tracking technique to estimate glacier velocity. To assess the impact of climate variations, we examined temperature and precipitation anomalies using ERA5 Land climate data. Over the past two decades, the Astak catchment experienced a slight decrease in glacier area (−1.8 km2) and the overall specific mass balance was −0.02 ± 0.1 m w.e. a−1. The most negative mass balance of −0.09 ± 0.06 m w.e. a−1 occurred at elevations between 2810 to 3220 m a.s.l., with a lesser rate of −0.015 ± 0.12 m w.e. a−1 above 5500 m a.s.l. This variation in glacier mass balance can be attributed to temperature and precipitation gradients, as well as debris cover. Recent glacier mass loss can be linked to seasonal temperature anomalies at higher elevations during winter and autumn. Given the reliance of mountain populations on glacier melt, seasonal temperature trends can disturb water security and the well-being of dependent communities. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The Astak catchment location with red dot in northern Pakistan. (<b>b</b>) The boundary of the study area (Astak catchment) is highlighted with the black line, and glaciers in the catchment are highlighted with a blue color; the stream network is shown with a blue line. Major glaciers in the Astak catchment are highlighted in panel (<b>b</b>).</p>
Full article ">Figure 2
<p>Altitudinal variations in glacier area change for clean glacier ice are shown in panel (<b>a</b>) and for debris cover, in panel (<b>b</b>). Total glacier area during 2020 is highlighted with a blue line. The total change from 2000 to 2020 is highlighted with a black line in panel (<b>b</b>).</p>
Full article ">Figure 3
<p>Glacier area changes in the Astak catchment from 2000 to 2020. Area changes in the terminus of the glacier are highlighted in the panels (<b>a</b>–<b>c</b>) for Kutiah Lungma Glacier, Tuklah Glacier, and Goropah Glacier, respectively.</p>
Full article ">Figure 4
<p>Variation in glacier mass balance for all the glaciers in the Astak catchment. Subplots in the first column show the spatially distributed glacier mass balance, and the second column shows altitudinal variations in glacier mass balance for each time period from 2000 to 2020. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) show glacier mass balance for the periods of 2000–2004, 2005–2009, 2010–2014, and 2015–2020, respectively. The color variation corresponds to the color bar in each subplot. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) altitudinal variations in glacier mass balance, with the shaded area in each subplot indicating the range of error. (<b>i</b>,<b>j</b>) Mean annual mass balance and altitudinal variations in mean annual mass balance from 2000–2020, respectively.</p>
Full article ">Figure 5
<p>(<b>a</b>) Variation in glacier mass elevation change (m) corresponds to the color bar for all the glaciers in the Astak catchment. (<b>b</b>) Altitudinal variations in glacier mass balance (m w.e. a<sup>−1</sup>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Variations in mean annual surface velocity of Kutiah Lungma Glacier from 2000 to 2020. Distance from the terminus of the glacier is shown on the horizontal x-axis. (<b>b</b>) Variation in mean glacier surface velocity from 2000 to 2020 is shown as the brown line over the corresponding glacier area, with 200 m internal elevation bins shown with blue bars. The range of error in the estimated velocity is highlighted with the brown shaded line.</p>
Full article ">Figure 7
<p>Seasonal temperature and precipitation anomalies for the Astak catchment from 2000 to 2020 compared to the baseline period from 1990 to 2000. (<b>a</b>–<b>d</b>) Temperature anomalies for winter, spring, summer, and autumn months and (<b>e</b>–<b>h</b>) precipitation anomalies for the winter, spring, summer, and autumn months. The boundary of the Astak catchment is shown in black, and glacier boundaries are shown in blue.</p>
Full article ">Figure 8
<p>(<b>a</b>) Areas affected by a flood that originated from glaciers are highlighted with red and black dotted lines, and the settlements of the Astak catchment are shown by the yellow line. (<b>b</b>) Settlements in the Astak catchment are located along the stream from glaciers. (<b>c</b>) Location of the area affected by a previous flood from the Tuklah Glacier catchment during 1999 is highlighted with the red line. (<b>d</b>) Area affected by a flood from the upstream glacier during 1998 is highlighted with the red line. (<b>e</b>) Flood-affected areas during a recent flood in 2020 are highlighted with the red dotted line.</p>
Full article ">
Back to TopTop