[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 (355)

Search Parameters:
Keywords = cryosphere

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 6210 KiB  
Article
Research on Glacier Changes and Their Influencing Factors in the Yigong Zangbo River Basin of the Tibetan Plateau, China, Based on ICESat-2 Data
by Wei Nie, Qiqi Du, Xuepeng Zhang, Kunxin Wang, Yang Liu, Yongjie Wang, Peng Gou, Qi Luo and Tianyu Zhou
Water 2024, 16(18), 2617; https://doi.org/10.3390/w16182617 - 15 Sep 2024
Viewed by 274
Abstract
The intense changes in glaciers in the southeastern Tibetan Plateau (SETP) have essential impacts on regional water resource management. In order to study the seasonal fluctuations of glaciers in this region and their relationship with climate change, we focus on the Yigong Zangbo [...] Read more.
The intense changes in glaciers in the southeastern Tibetan Plateau (SETP) have essential impacts on regional water resource management. In order to study the seasonal fluctuations of glaciers in this region and their relationship with climate change, we focus on the Yigong Zangbo River Basin in the SETP, extract the annual and seasonal variations of glaciers in the basin during 2018–2023, and analyze their spatio-temporal characteristics through the seasonal-trend decomposition using the LOESS (STL) method. Finally, combining the Extreme Gradient Boosting (XGBoost) model and the Shapley additive explanations (SHAP) model, we assess the comprehensive impact of meteorological factors such as temperature and snowfall on glacier changes. The results indicate that glaciers in the Yigong Zangbo River Basin experienced remarkable mass loss during 2018–2023, with an average annual melting rate of −0.83 ± 0.12 m w.e.∙yr−1. The glacier mass exhibits marked seasonal fluctuations, with increases in January–March (JFM) and April–June (AMJ) and noticeable melting in July–September (JAS) and October–December (OND). The changes over these four periods are 2.12 ± 0.04 m w.e., 0.93 ± 0.15 m w.e., −1.58 ± 0.19 m w.e., and −1.32 ± 0.17 m w.e., respectively. Temperature has been identified as the primary meteorological driver of glacier changes in the study area, surpassing the impact of snowfall. This study uses advanced altimetry data and meteorological data to monitor and analyze glacier changes, which provides valuable data for cryosphere research and also validates a set of replicable research methods, which provides support for future research in related fields. Full article
(This article belongs to the Section Water and Climate Change)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Overview of the study area, (<b>b</b>) distribution of the ICESat-2 points in the study area, and (<b>c</b>) the location of the study area on the Tibetan Plateau.</p>
Full article ">Figure 2
<p>Technical flowchart. “GMC” denotes glacier mass changes, “DEM” the digital elevation model, “RGI” the Randolph Glacier Inventory data, “STL” the seasonal-trend decomposition using LOESS, and “SHAP” the Shapley additive explanations.</p>
Full article ">Figure 3
<p>A schematic diagram of the XGBoost model. The red arrow denotes the selected direction of the tree, the yellow circle denotes the selected node, the green circle denotes the unselected node.</p>
Full article ">Figure 4
<p>Glacier mass change (GMC), seasonal mass difference (<math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi mathvariant="normal">M</mi> </mrow> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">L</mi> </mrow> </msub> </mrow> </semantics></math>), and trend and seasonal components after the STL decomposition.</p>
Full article ">Figure 5
<p>Deseasonalized GMC series X <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mrow> <mi mathvariant="normal">X</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>=</mo> <msub> <mrow> <mi mathvariant="normal">T</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>+</mo> <msub> <mrow> <mi mathvariant="sans-serif">ε</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Average seasonal GMC relative to ALOS DEM (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi mathvariant="normal">G</mi> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">C</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) and glacial area with altitude, and (<b>b</b>) average seasonal variation in glacier mass (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mo>∆</mo> <mi mathvariant="normal">M</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) at different altitudes.</p>
Full article ">Figure 7
<p>Seasonal variation in glacier mass (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mo>∆</mo> <mi mathvariant="normal">M</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) in (<b>a</b>) JFM, (<b>b</b>) AMJ, (<b>c</b>) JAS, and (<b>d</b>) OND.</p>
Full article ">Figure 8
<p>SHAP values of the meteorological drivers for the GMC.</p>
Full article ">Figure 9
<p>Scatter plots of the SHAP values of (<b>a</b>) temperature and (<b>b</b>) snowfall and their functional relationships with the GMC.</p>
Full article ">Figure 10
<p>(<b>a</b>) Monthly average temperature and accumulated precipitation within the study region (2018–2023); seasonal average (<b>b</b>) temperature and (<b>c</b>) precipitation. “Tem” denotes temperature, “TP” denotes total precipitation, and “SF” denotes snowfall.</p>
Full article ">Figure 11
<p>Seasonal averages of (<b>a</b>–<b>d</b>) temperature and (<b>e</b>–<b>h</b>) snowfall from 2018 to 2023.</p>
Full article ">Figure 12
<p>Variations in the standard deviations of the (<b>a</b>) temperature (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">m</mi> </mrow> </msub> </mrow> </semantics></math>), snowfall (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">F</mi> </mrow> </msub> </mrow> </semantics></math>), and (<b>b</b>) seasonal variation in glaciers (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi mathvariant="normal">M</mi> </mrow> </msub> </mrow> </semantics></math>) with altitude.</p>
Full article ">
12 pages, 5379 KiB  
Article
Snow Leopard (Panthera uncia) Activity Patterns Using Camera Traps in the Qilian Mountain National Park (Qinghai Area), China
by Hu Ma, Bading Qiuying, Zhanlei Rong, Jinhu Zhang, Guozhu Liang, Shuguang Ma, Yayue Gao and Shengyun Chen
Animals 2024, 14(18), 2680; https://doi.org/10.3390/ani14182680 - 14 Sep 2024
Viewed by 209
Abstract
In recent years, there has been growing concern about the condition of snow leopards. The snow leopard (Panthera uncia), an apex predator of alpine ecosystems, is essential for the structural and functional stability of ecosystems. Monitoring of snow leopards’ activity patterns [...] Read more.
In recent years, there has been growing concern about the condition of snow leopards. The snow leopard (Panthera uncia), an apex predator of alpine ecosystems, is essential for the structural and functional stability of ecosystems. Monitoring of snow leopards’ activity patterns based on camera traps in the Qilian Mountain National Park (Qinghai area) between August 2020 to October 2023 was performed. The results showed that autumn is the peak period of snow leopard activity, especially in September when the frequency of activity is the highest, and there is one peak in the frequency of snow leopard daily activity in the time period of 18:00–22:00, while the highest overlap of the daily activity curves of snow leopards in different months was from spring to autumn (Δ = 0.97), and there were significant differences in diurnal activity rhythm between spring and autumn (p = 0.002). Snow leopards prefer sunny days, and they tend to be active at temperatures of −10–9 °C. Our research aimed to uncover the activity patterns of snow leopards at different scales within the study area and provide data for further studies on snow leopards and other wildlife by researchers. This study can be used to gain a comprehensive understanding of the ecological characteristics of snow leopards and to assess their habitats, and it will also serve as a reference for the local wildlife management authorities in formulating snow leopard conservation measures. Full article
(This article belongs to the Special Issue Ecology and Conservation of Large Carnivores)
Show Figures

Figure 1

Figure 1
<p>A map of the study area showing the distribution of camera traps for monitoring snow leopard activity.</p>
Full article ">Figure 2
<p>Relative activity intensity of snow leopard daily activity and seasonal variation.</p>
Full article ">Figure 3
<p>Snow leopard daily activity and daily activity curves across different seasons.</p>
Full article ">Figure 4
<p>Relative activity intensity of monthly snow leopard activity.</p>
Full article ">Figure 5
<p>Relative activity intensity of seasonal snow leopard activity.</p>
Full article ">Figure 6
<p>Relative activity intensity of snow leopard activity at different temperature intervals.</p>
Full article ">Figure 7
<p>Percentages of valid snow leopard photographs in all weather types. Note: Sunny day—bright sunlight visible (includes visible sunrise and sunset light); cloudy day—sunlight not visible and weather is overcast; snowy day—weather in which snow is falling (includes nighttime when snow is falling).</p>
Full article ">
23 pages, 4759 KiB  
Article
Crafting Glacial Narratives: Virtual Exploration of Alpine Glacial and Periglacial Features in Preston Park, Glacier National Park, Montana, USA
by Jacquelyn Kelly, Dianna Gielstra, Lynn Moorman, Uwe Schulze, Niccole V. Cerveny, Johan Gielstra, Rohana J. Swihart, Scott Ramsey, Tomáš J. Oberding, David R. Butler and Karen Guerrero
Glacies 2024, 1(1), 57-79; https://doi.org/10.3390/glacies1010005 - 6 Sep 2024
Viewed by 690
Abstract
Virtual learning environments (VLEs) in physical geography education offer significant potential to aid students in acquiring the essential skills for the environmental interpretation of glacial and periglacial environments for geoscience careers. Simulated real-world field experiences aim to help the student evaluate landscapes for [...] Read more.
Virtual learning environments (VLEs) in physical geography education offer significant potential to aid students in acquiring the essential skills for the environmental interpretation of glacial and periglacial environments for geoscience careers. Simulated real-world field experiences aim to help the student evaluate landscapes for natural hazards, assess their intensity, and translate and communicate this information to various stakeholders in human systems. The TREE-PG framework and VRUI model provide a philosophical and practical foundation for VLE architects, aiming to cultivate students’ knowledge, skills, and identity as geoscientists, specifically as physical geographers and geomorphologists. These frameworks emphasize the importance of translating scientific knowledge from physical features into engaging, accessible online lessons, exemplified by landscapes like those in Glacier National Park, Montana. Open-source software and open educational resources (OERs) can broaden access and incorporate diverse perspectives in these experiences, which are necessary to address the impacts of vulnerable communities to global deglaciation. Designing and creating virtual proxies of field-based education may help address issues associated with inclusion and belonging within geoscience disciplines to connect all students with dynamic physical environments beyond the classroom. Ethical AI approaches and discipline-specific repositories are needed to ensure high-quality, contextually accurate VLEs. AI’s tendency to produce output necessitates using domain-specific guardrails to maintain relevance and precision in virtual educational content. Full article
Show Figures

Figure 1

Figure 1
<p>Visualization of the Virtual Reality User Interface (VRUI) orders for VLE architects.</p>
Full article ">Figure 2
<p>Preston Park site location in eastern Glacier National Park, Montana, U.S.A., within the International Peace Park.</p>
Full article ">Figure 3
<p>Strabo AI output to the prompt “You are a lesson planner at the college level and the query “How can I refine the learning objective to be more targeted for glacier environments, “identify environmental patterns of physical features?”. The yellow box with the arrow highlights where the seasoned teachers’ lessons are recalled from Strabo AI’s repository.</p>
Full article ">Figure 4
<p>The result of clicking “Show Citations” in the previous figure is the presentation of the actual lessons that Strabo AI used to generate that output [<a href="#B54-glacies-01-00005" class="html-bibr">54</a>].</p>
Full article ">Figure 5
<p>360-degree capture at Preston Park site for photo sphere resource to build the virtual environment for the VRUI 1st Order.</p>
Full article ">Figure 6
<p>Glacial geomorphology visual design considerations adapted for ergonomics and VR user comfort [<a href="#B50-glacies-01-00005" class="html-bibr">50</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Piegan Glacier is a cirque glacier perched on the lip edge. The concavity at the mountaintop was carved by glaciers that were located on a more protected slope. The depression allows snow and ice to accumulate, further hollowing the bowl. (<b>b</b>) Siyeh Pass is the low-relief topography where the contour descends from the two mountain peaks. The front-facing landform further descends into the Preston Park valley. (<b>c</b>) Source material for the talus slopes, slides and debris flows. (<b>d</b>) Mt. Reynolds is a classic horn feature that exhibits evidence of the vertical scraping by glaciers on three sides of the mountaintop. (<b>e</b>) Debris flow deposits with debris flow levees modifying the talus slope [<a href="#B56-glacies-01-00005" class="html-bibr">56</a>]. (<b>f</b>) Side view of talus slopes modified by avalanches into boulder tongues and the boulder tongue deposition area. The margins of the tongues can be viewed just above the two lower-positioned small snow fields on the far right of the image. Avalanches transport material down the slope and smooth the slope angle into a concave fan. (<b>g</b>) Cryoturbation shows frost churning at the site with eroded material moving downslope [<a href="#B57-glacies-01-00005" class="html-bibr">57</a>]. (<b>h</b>) Incipient solifluction with the formation of the turf-banked terrace and risers is consistent with a wind-swept environment that is exposed to solar radiation [<a href="#B57-glacies-01-00005" class="html-bibr">57</a>,<a href="#B58-glacies-01-00005" class="html-bibr">58</a>,<a href="#B59-glacies-01-00005" class="html-bibr">59</a>].</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) Piegan Glacier is a cirque glacier perched on the lip edge. The concavity at the mountaintop was carved by glaciers that were located on a more protected slope. The depression allows snow and ice to accumulate, further hollowing the bowl. (<b>b</b>) Siyeh Pass is the low-relief topography where the contour descends from the two mountain peaks. The front-facing landform further descends into the Preston Park valley. (<b>c</b>) Source material for the talus slopes, slides and debris flows. (<b>d</b>) Mt. Reynolds is a classic horn feature that exhibits evidence of the vertical scraping by glaciers on three sides of the mountaintop. (<b>e</b>) Debris flow deposits with debris flow levees modifying the talus slope [<a href="#B56-glacies-01-00005" class="html-bibr">56</a>]. (<b>f</b>) Side view of talus slopes modified by avalanches into boulder tongues and the boulder tongue deposition area. The margins of the tongues can be viewed just above the two lower-positioned small snow fields on the far right of the image. Avalanches transport material down the slope and smooth the slope angle into a concave fan. (<b>g</b>) Cryoturbation shows frost churning at the site with eroded material moving downslope [<a href="#B57-glacies-01-00005" class="html-bibr">57</a>]. (<b>h</b>) Incipient solifluction with the formation of the turf-banked terrace and risers is consistent with a wind-swept environment that is exposed to solar radiation [<a href="#B57-glacies-01-00005" class="html-bibr">57</a>,<a href="#B58-glacies-01-00005" class="html-bibr">58</a>,<a href="#B59-glacies-01-00005" class="html-bibr">59</a>].</p>
Full article ">
16 pages, 5926 KiB  
Article
Ecological Status Assessment of Permafrost-Affected Soils in the Nadym Region, Yamalo-Nenets Autonomous District, Russian Arctic
by Wenjuan Wang, Timur Nizamutdinov, Aleksander Pechkin, Eugeniya Morgun, Gensheng Li, Xiaodong Wu, Sizhong Yang and Evgeny Abakumov
Land 2024, 13(9), 1406; https://doi.org/10.3390/land13091406 - 1 Sep 2024
Viewed by 341
Abstract
Permafrost-affected regions in the Russian Arctic are a critical study area for studying the sources of metal elements (MEs) in soils originating from geological/pedogenic processes or from anthropogenic sources via atmospheric transport. In the Nadym region of the Yamalo-Nenets Autonomous District, we investigated [...] Read more.
Permafrost-affected regions in the Russian Arctic are a critical study area for studying the sources of metal elements (MEs) in soils originating from geological/pedogenic processes or from anthropogenic sources via atmospheric transport. In the Nadym region of the Yamalo-Nenets Autonomous District, we investigated the contents of soil organic carbon (SOC), total nitrogen (TN), and MEs across different soil types and horizons, explored the source apportionment of MEs, and assessed local ecological risks of potentially toxic elements (PTEs). The results showed that (1) the contents of SOC and TN in Histic Cryosols (8.59% and 0.27%) were significantly higher than in Plaggic Podzols (Arenic, Gelic, and Turbic) (2.28% and 0.15%) and in Ekranic Technosols (Umbric) (1.32% and 0.09%); (2) the concentrations of MEs in the Nadym region were lower than in other Arctic regions; (3) the primary sources of MEs were identified as geological processes (36%), atmospheric transport (23%), agricultural activities (21%), and transportation (20%); and (4) the permafrost-affected soils in the Nadym region exhibited low ecological risks from PTEs. These results underscore the critical role of geological and anthropogenic factors in shaping soil conditions and highlight the relatively low ecological risk from PTEs, providing a valuable benchmark for future environmental assessments and policy development in Yamal permafrost regions. Full article
Show Figures

Figure 1

Figure 1
<p>The location of the study areas. (<b>a</b>) Global map highlighting the study region. (<b>b</b>) Detailed map of the Nadym region in the Yamalo-Nenets Autonomous District. (<b>c</b>–<b>e</b>) Photographs of the three sampling sites in the Nadym region (TD-tundra, AF-abandoned farmland, and UA-urban area). (<b>f</b>–<b>h</b>) Soil profiles from the sampling sites in the Nadym region.</p>
Full article ">Figure 2
<p>Physicochemical properties in permafrost-affected soils of the Nadym region.TD—tundra; AF—abandoned farmland; UA—urban area; (<b>a</b>) SOC—soil organic carbon. (<b>b</b>) TN—soil total nitrogen. (<b>c</b>) C/N—the mass ratio of SOC to TN. (<b>d</b>) pH; (<b>e</b>) Clay (&lt;0.002 mm). (<b>f</b>) Silt (0.002–0.05 mm). (<b>g</b>) Sand (&gt;0.05 mm).</p>
Full article ">Figure 3
<p>The concentrations (mg kg<sup>−1</sup>) of metal elements (MEs) in permafrost-affected soils of the Nadym region. TD—tundra; AF—abandoned farmland; UA—urban area. (<b>a</b>–<b>h</b>) The concentrations of eight MEs (Fe, Mn, Zn, As, Cr, Ni, Cu, and Pb) in soil depths and horizons.</p>
Full article ">Figure 4
<p>Regression models of metal elements (MEs). TD—tundra; AF—abandoned farmland; UA—urban area. (<b>a</b>) Clay (&lt;0.002 mm). (<b>b</b>) Silt (0.002–0.05 mm). (<b>c</b>) Sand (&gt;0.05 mm). (<b>d</b>,<b>g</b>) SOC—soil organic carbon. (<b>e</b>,<b>h</b>). TN—soil total nitrogen. (<b>f</b>,<b>i</b>) C/N—the mass ratio of SOC to TN. **: significance level of <span class="html-italic">p</span> &lt; 0.01; *: significance level of <span class="html-italic">p</span> &lt; 0.05. The grey shadowed areas represent the 95% confidence interval. Only statistically significant results are shown here.</p>
Full article ">Figure 5
<p>Source apportionment of metal elements (MEs) in the Nadym region. (<b>a</b>) The percentage of contribution for each factor by PMF model. (<b>b</b>) Factor profiles of MEs in permafrost-affected soils derived from PMF model. (<b>c</b>) The correlations of MEs by combining Pearson analysis and PMF model, **: significance level of <span class="html-italic">p</span> &lt; 0.01 and *: significance level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Ecological state of potentially toxic elements (PTEs) in permafrost-affected soils of the Nadym region. (<b>a</b>) Geoaccumulation index (I<sub>geo</sub>), Class 0: I<sub>geo</sub> ≤ 0 (no pollution) and Class 1: 0 &lt; I<sub>geo</sub> ≤ 1 (no contamination to slight pollution). (<b>b</b>) Enrichment factor (EF), Class 1: EF &lt; 2 (no enrichment) and Class 2: 2 ≤ EF &lt; 5 (moderate enrichment).</p>
Full article ">Figure 7
<p>Potential ecological risk index in permafrost-affected soils of the Nadym region. E<sub>r</sub>—potential ecological risk index of the i-th element. RI—potential ecological risk index for all potentially toxic elements (PTEs), including Ni, As, Cu, Pb, Cr, and Zn.</p>
Full article ">
16 pages, 5236 KiB  
Article
Effects of Organic Fertilizer and Biochar on Carbon Release and Microbial Communities in Saline–Alkaline Soil
by Pengfei Zhang, Ziwei Jiang, Xiaodong Wu, Nannan Zhang, Jiawei Zhang, Siyuan Zou, Jifu Wang and Shuying Zang
Agronomy 2024, 14(9), 1967; https://doi.org/10.3390/agronomy14091967 - 31 Aug 2024
Viewed by 662
Abstract
Climate change and aridification have increased the risk of salinization and organic carbon loss in dryland soils. Enrichment using biochar and organic fertilizers has the potential to reduce salt toxicity and soil carbon loss. However, the effects of biochar and organic fertilizers on [...] Read more.
Climate change and aridification have increased the risk of salinization and organic carbon loss in dryland soils. Enrichment using biochar and organic fertilizers has the potential to reduce salt toxicity and soil carbon loss. However, the effects of biochar and organic fertilizers on CO2 and CH4 emissions from saline soils in dryland areas, as well as their microbial mechanisms, remain unelucidated. To clarify these issues, we performed a 5-month incubation experiment on typical soda-type saline soil from the western part of the Songnen Plain using five treatments: control treatment (CK), 5% urea (U), straw + 5% urea (SU), straw + 5% urea + microbial agent (SUH), and straw + 5% urea + biochar (SUB). Compared with the SU treatment, the SUH and SUB treatments reduced cumulative CO2 emissions by 14.85% and 35.19%, respectively. The addition of a microbiological agent to the SU treatment reduced the cumulative CH4 emissions by 19.55%, whereas the addition of biochar to the SU treatment increased the cumulative CH4 emissions by 4.12%. These additions also increased the relative abundances of Proteobacteria, Planctomycetes, and Ascomycota. Overall, the addition of biochar and organic fertilizer promoted CO2 emissions and CH4 uptake. This was mainly attributed to an improved soil gas diffusion rate due to the addition of organic materials and enhanced microbial stress due to soil salinity and alkalinity from the release of alkaline substances under closed-culture conditions. Our findings have positive implications for enhancing carbon storage in saline soils in arid regions. Full article
(This article belongs to the Section Agricultural Biosystem and Biological Engineering)
Show Figures

Figure 1

Figure 1
<p>Cumulative CO<sub>2</sub> emissions (<b>a</b>) and cumulative CH<sub>4</sub> emissions (<b>b</b>) in different treatment groups (n = 3). CK, control treatment; U, urea; SU, straw + urea; SUH, straw + urea + microbial agent; SUB, straw + urea +biochar.</p>
Full article ">Figure 2
<p>Relative abundances of bacteria (<b>a</b>) and fungi (<b>c</b>) at the phylum level (&gt;1%) for all treatment groups on day 150. Nonmetric multidimensional scaling (NMDS) analysis of bacteria (<b>b</b>) and fungi (<b>d</b>). Each graph is grouped and connected based on the samples from each treatment group. CK, control treatment; U, urea; SU, straw + urea; SUH, straw + urea + microbial agent; SUB, straw + urea + biochar.</p>
Full article ">Figure 3
<p>Normalized heatmap analysis of predicted abundances of carbon degradation and CH₄ oxidation functional enzymes derived from soil bacterial (<b>a</b>) and fungal (<b>b</b>) sequencing data following cultivation experiments. CK, control treatment; U, urea; SU, straw + urea; SUH, straw + urea + microbial agent; SUB, straw + urea + biochar. Enzymes include α-amylase (EC 3.2.1.1), glucoamylase (EC 3.2.1.3), α-glucosidase (EC 3.2.1.20), isoamylase (EC 3.2.1.68), glycogen phosphorylase (EC 2.4.1.1), pullulanase (EC 3.2.1.41), cyclodextrin glycosyltransferase (EC 2.4.1.19), exocellobiohydrolase (EC 3.2.1.91), β-glucosidase (BG; EC 3.2.1.21), cellulase (EC 3.2.1.4), xylanase (EC 3.2.1.8), β-mannosidase (EC 3.2.1.25), α-L-arabinosidase (EC 3.2.1.55), β-xylosidase (EC 3.2.1.37), hemicellulase (EC 3.1.1.73), chitinase (EC 3.2.1.14), chitobiase (EC 3.2.1.132), α-N-acetylglucosaminidase (EC 3.2.1.50), particulate methane monooxygenase (pMMO; EC 1.14.18.3), laccase (LA; EC 1.10.3.2), and α-D-glucuronidase (EC 3.2.1.20).</p>
Full article ">Figure 4
<p>Heatmap of Spearman’s correlation analysis between CO<sub>2</sub> emissions (<b>a</b>) and CH₄ emissions (<b>b</b>) with microbial diversity indices and soil physicochemical properties at the phylum-level SAC, soil additives with different characteristics; CCO<sub>2</sub>, cumulative CO<sub>2</sub> emissions; CCH<sub>4</sub>, cumulative CH<sub>4</sub> emissions; CN<sub>2</sub>O, cumulative N<sub>2</sub>O emissions; BOS, bacterial observed species index; FChao1, fungal Chao1 index; FOS, fungal observed species index. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Structural equation modeling (SEM) based on the effects of SAC, soil physicochemical properties, and fungal alpha diversity index on cumulative CO<sub>2</sub> emissions (<b>a</b>) and cumulative CH<sub>4</sub> emissions (<b>b</b>) in saline–alkaline soil samples. SEM-based standardized total effect on cumulative CO<sub>2</sub> emissions (<b>c</b>) and cumulative CH<sub>4</sub> emissions (<b>d</b>). Blue and red lines indicate significant positive and negative correlations, respectively (<span class="html-italic">p</span> &lt; 0.05), and dashed lines indicate a potential nonsignificant path. Numbers on the arrows indicate standardized path coefficients (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001). Black double arrows indicate the covariance between the exogenous variables. R<sup>2</sup> denotes the total variance of the dependent variables explained by the model. SAC, soil additives with different characteristics; FC, fungal Chao1 index; FOS, observed fungal species index; FP, fungal Pielou evenness index.</p>
Full article ">
15 pages, 6347 KiB  
Article
Distribution Characteristics and Driving Factors of the Bacterial Community Structure in the Soil Profile of a Discontinuous Permafrost Region
by Qilong Liu, Liquan Song, Siyuan Zou, Xiaodong Wu and Shuying Zang
Forests 2024, 15(8), 1456; https://doi.org/10.3390/f15081456 - 18 Aug 2024
Viewed by 680
Abstract
Global warming leads to the melting of permafrost, affects changes in soil microbial community structures and related functions, and contributes to the soil carbon cycle in permafrost areas. Located at the southern edge of Eurasia’s permafrost region, the Greater Khingan Mountains are very [...] Read more.
Global warming leads to the melting of permafrost, affects changes in soil microbial community structures and related functions, and contributes to the soil carbon cycle in permafrost areas. Located at the southern edge of Eurasia’s permafrost region, the Greater Khingan Mountains are very sensitive to climate change. Therefore, by analyzing the bacterial community structure, diversity characteristics, and driving factors of soil profiles (active surface layer, active deep layer, transition layer, and permafrost layer) in this discontinuous permafrost region, this research provides support for the study of the carbon cycling process in permafrost regions. The results show that the microbial diversity (Shannon index (4.81)) was the highest at 0–20 cm, and the Shannon index of the surface soil of the active layer was significantly higher than that of the other soil layers. Acidobacteria and Proteobacteria were the dominant bacteria in the active layer soil of the permafrost area, and Chloroflexi, Actinobacteria, and Firmicutes were the dominant bacteria in the permafrost layer. Chloroflexi made the greatest contribution to the bacterial community in the permafrost soil, and Bacteroidota made the greatest contribution to the bacterial community in the active surface soil. The structure, richness, and diversity of the soil bacterial community significantly differed between the active layer and the permafrost layer. The number of bacterial species was the highest in the active layer surface soil and the active layer bottom soil. The difference in the structure of the bacterial community in the permafrost soil was mainly caused by changes in electrical conductivity and soil–water content, while that in the active layer soil was mainly affected by pH and soil nutrient indices. Soil temperature, NO3-N, and pH had significant effects on the structure of the bacterial community. The active layer and permafrost soils were susceptible to environmental information processing and genetic information processing. Infectious disease: the number of bacterial species was significantly higher in the surface and permafrost layers than in the other layers of the soil. In conclusion, changes in the microbial community structure in soil profiles in discontinuous permafrost areas sensitive to climate change are the key to soil carbon cycle research. Full article
(This article belongs to the Section Forest Soil)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Study area location map. (<b>b</b>) Soil profile in permafrost area (0–20 cm: upper active layer; 20–60 cm: lower active layer; 60–80 cm: transition layer; 80–120 cm: permafrost layer).</p>
Full article ">Figure 2
<p>Venn diagram of the structure of the soil bacterial community at different OTU levels in the permafrost region.</p>
Full article ">Figure 3
<p>Principal coordinate analysis (PCoA) of soil bacteria in different soil layers.</p>
Full article ">Figure 4
<p>Composition of soil microbial communities in different soil layers: (<b>a</b>) analysis at phylum level; (<b>b</b>) analysis at genus level.</p>
Full article ">Figure 5
<p>Difference analysis of dominant soil microbial species in different soil layers: (<b>a</b>) analysis at phylum level; (<b>b</b>) analysis at genus level. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Correlation analysis of soil physicochemical properties and relative abundance of the major bacterial phyla. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Analysis of the prediction function of microbial communities in different soil layers: (<b>a</b>) grade 1 functional group; (<b>b</b>) grade 2 functional group. a, b, c: If there is one identical marking letter, the difference is not significant, and if there is different marking letter, the difference is significant.</p>
Full article ">Figure 8
<p>(<b>a</b>) Analysis of the metabolic pathways of soil bacteria regulated by thawing permafrost. The red line represents the positive path, and the blue line represents the negative path. Insignificant effects are indicated by dotted arrows. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. (<b>b</b>) Normalization effects between variables in different permafrost regions (direct and indirect normalization effects).</p>
Full article ">
40 pages, 19379 KiB  
Article
Evaluation of Sentinel-5P TROPOMI Methane Observations at Northern High Latitudes
by Hannakaisa Lindqvist, Ella Kivimäki, Tuomas Häkkilä, Aki Tsuruta, Oliver Schneising, Michael Buchwitz, Alba Lorente, Mari Martinez Velarte, Tobias Borsdorff, Carlos Alberti, Leif Backman, Matthias Buschmann, Huilin Chen, Darko Dubravica, Frank Hase, Pauli Heikkinen, Tomi Karppinen, Rigel Kivi, Erin McGee, Justus Notholt, Kimmo Rautiainen, Sébastien Roche, William Simpson, Kimberly Strong, Qiansi Tu, Debra Wunch, Tuula Aalto and Johanna Tamminenadd Show full author list remove Hide full author list
Remote Sens. 2024, 16(16), 2979; https://doi.org/10.3390/rs16162979 - 14 Aug 2024
Viewed by 679
Abstract
The Arctic and boreal regions are experiencing a rapid increase in temperature, resulting in a changing cryosphere, increasing human activity, and potentially increasing high-latitude methane emissions. Satellite observations from Sentinel-5P TROPOMI provide an unprecedented coverage of a column-averaged dry-air mole fraction of methane [...] Read more.
The Arctic and boreal regions are experiencing a rapid increase in temperature, resulting in a changing cryosphere, increasing human activity, and potentially increasing high-latitude methane emissions. Satellite observations from Sentinel-5P TROPOMI provide an unprecedented coverage of a column-averaged dry-air mole fraction of methane (XCH4) in the Arctic, compared to previous missions or in situ measurements. The purpose of this study is to support and enhance the data used for high-latitude research through presenting a systematic evaluation of TROPOMI methane products derived from two different processing algorithms: the operational product (OPER) and the scientific product (WFMD), including the comparison of recent version changes of the products (OPER, OPER rpro, WFMD v1.2, and WFMD v1.8). One finding is that OPER rpro yields lower XCH4 than WFMD v1.8, the difference increasing towards the highest latitudes. TROPOMI product differences were evaluated with respect to ground-based high-latitude references, including four Fourier Transform Spectrometer in the Total Carbon Column Observing Network (TCCON) and five EM27/SUN instruments in the Collaborative Carbon Column Observing Network (COCCON). The mean TROPOMI–TCCON GGG2020 daily median XCH4 difference was site-dependent and varied for OPER rpro from −0.47 ppb to 22.4 ppb, and for WFMD v1.8 from 1.2 ppb to 19.4 ppb with standard deviations between 13.0 and 20.4 ppb and 12.5–15.0 ppb, respectively. The TROPOMI–COCCON daily median XCH4 difference varied from −26.5 ppb to 5.6 ppb for OPER rpro, with a standard deviation of 14.0–28.7 ppb, and from −5.0 ppb to 17.2 ppb for WFMD v1.8, with a standard deviation of 11.5–13.0 ppb. Although the accuracy and precision of both TROPOMI products are, on average, good compared to the TCCON and COCCON, a persistent seasonal bias in TROPOMI XCH4 (high values in spring; low values in autumn) is found for OPER rpro and is reflected in the higher standard deviation values. A systematic decrease of about 7 ppb was found between TCCON GGG2014 and GGG2020 product update highlighting the importance of also ensuring the reliability of ground-based retrievals. Comparisons to atmospheric profile measurements with AirCore carried out in Sodankylä, Northern Finland, resulted in XCH4 differences comparable to or smaller than those from ground-based remote sensing. Full article
(This article belongs to the Section Biogeosciences Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Locations of the high-latitude TCCON, COCCON, and AirCore sites overlaid on a map of the regional permafrost extent re-gridded from ESA CCI Permafrost data. Regions with &gt;90% permafrost extent are considered continuous permafrost.</p>
Full article ">Figure 2
<p>Effect of the COCCON prior correction on TROPOMI OPER rpro (blue) and WFMD v1.8 (red) daily median XCH<sub>4</sub> at (<b>a</b>) Kiruna, Sweden, (<b>b</b>) SN039 and (<b>c</b>) SN122 in Sodankylä, Finland, (<b>d</b>) Fairbanks, Alaska, USA, and (<b>e</b>) St. Petersburg, Russia.</p>
Full article ">Figure 3
<p>Spatial variability of monthly-averaged total column methane for TROPOMI OPER rpro (<b>left</b>) and WFMD v1.8 (<b>right</b>) for April, August, and October in 2020. The grid size is 0.25° × 0.2°.</p>
Full article ">Figure 4
<p>Spatial variability of the difference in monthly-averaged total column methane for TROPOMI OPER rpro and WFMD v1.8 in 2020. The grid size is 0.25° × 0.2°.</p>
Full article ">Figure 5
<p>Temporal dependence of (<b>a</b>) TROPOMI OPER rpro–WFMD v1.8 XCH<sub>4</sub> difference, (<b>b</b>) TROPOMI OPER–OPER rpro XCH<sub>4</sub> difference, and (<b>c</b>) TROPOMI WFMD v1.2–WFMD v1.8 XCH<sub>4</sub> difference, colored based on 5-degree latitude bands north of 50°N. Gaps represent missing data.</p>
Full article ">Figure 6
<p>Averages of (<b>a</b>) TROPOMI OPER rpro–WFMD v1.8 XCH<sub>4</sub> difference, (<b>b</b>) TROPOMI OPER–OPER rpro XCH<sub>4</sub> difference, and (<b>c</b>) TROPOMI WFMD v1.2–WFMD v1.8 XCH<sub>4</sub> difference for different latitude bands. The error bars denote the standard deviation. Colored circles denote the mean difference of the satellite product to the ground-based TCCON instrument (GGG2020 retrieval) located in that latitude band.</p>
Full article ">Figure 7
<p>Seasonal dependence of the number of TROPOMI observations over different permafrost regions for (<b>a</b>) OPER, (<b>b</b>) OPER rpro, (<b>c</b>) WFMD v1.2, and (<b>d</b>) WFMD v1.8. The colors refer to different permafrost classes, determined using the ESA Permafrost CCI Level 4 product.</p>
Full article ">Figure 8
<p>Daily median XCH<sub>4</sub> from TCCON GGG2020 (blue) and co-located TROPOMI (red) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–TCCON differences (<b>b</b>,<b>d</b>) at Eureka, Canada. The dashed line in (<b>b</b>,<b>d</b>) shows the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure 9
<p>Daily median XCH<sub>4</sub> from TCCON GGG2020 (blue) and co-located TROPOMI (red) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–TCCON differences (<b>b</b>,<b>d</b>) at Ny-Ålesund, Norway. The dashed line in (<b>b</b>,<b>d</b>) shows the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure 10
<p>Daily median XCH<sub>4</sub> from TCCON GGG2020 (blue) and co-located TROPOMI (red) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–TCCON differences (<b>b</b>,<b>d</b>) at Sodankylä, Finland. The dashed line in (<b>b</b>,<b>d</b>) shows the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure 11
<p>Daily median XCH<sub>4</sub> from TCCON GGG2020 (blue) and co-located TROPOMI (red) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–TCCON differences (<b>b</b>,<b>d</b>) at East Trout Lake, Canada. The dashed line in (<b>b</b>,<b>d</b>) shows the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure 12
<p>Comparison of TROPOMI–TCCON mean XCH<sub>4</sub> differences for TCCON GGG2014 and TCCON GGG2020, considering TROPOMI (<b>a</b>) OPER rpro and (<b>b</b>) WFMD v1.8 products. The TCCON data have been selected to include only the days for which both processing versions yield good-quality retrievals, so their temporal sampling is identical.</p>
Full article ">Figure 13
<p>Daily median XCH<sub>4</sub> from COCCON (red) and co-located TROPOMI (blue) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–COCCON differences (<b>b</b>,<b>d</b>) at Kiruna, Sweden. The dashed line in (<b>b</b>,<b>d</b>) shows the mean differences and the shaded area denotes one standard deviation from the mean.</p>
Full article ">Figure 14
<p>Daily median XCH<sub>4</sub> from COCCON (red) and co-located TROPOMI (blue) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–COCCON differences (<b>b</b>,<b>d</b>) at Sodankylä, Finland (SN039). The dashed line in (<b>b</b>,<b>d</b>) shows the mean differences and the shaded area denotes one standard deviation from the mean.</p>
Full article ">Figure 15
<p>Daily median XCH<sub>4</sub> from COCCON (red) and co-located TROPOMI (blue) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–COCCON differences (<b>b</b>,<b>d</b>) at Sodankylä, Finland (SN122). The dashed line in (<b>b</b>,<b>d</b>) shows the mean differences and the shaded area denotes one standard deviation from the mean.</p>
Full article ">Figure 16
<p>Daily median XCH<sub>4</sub> from COCCON (green) and co-located TROPOMI (orange) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–COCCON differences (<b>b</b>,<b>d</b>) at Fairbanks, Alaska, USA. The dashed line in (<b>b</b>,<b>d</b>) shows the mean differences and the shaded area denotes one standard deviation from the mean.</p>
Full article ">Figure 17
<p>Daily median XCH<sub>4</sub> from COCCON (ref) and co-located TROPOMI (blue) (<b>a</b>) OPER rpro and (<b>c</b>) WFMD v1.8 retrievals, and TROPOMI–COCCON differences (<b>b</b>,<b>d</b>) at St. Petersburg, Russia. The dashed line in (<b>b</b>,<b>d</b>) shows the mean differences and the shaded area denotes one standard deviation from the mean.</p>
Full article ">Figure 18
<p>Comparison of measured AirCore profiles (black curves), TCCON GGG2020 prior profiles (red), and TROPOMI prior profiles: (<b>a</b>) TROPOMI OPER rpro (blue), and (<b>c</b>) TROPOMI WFMD v1.8 (orange) prior profiles. The TROPOMI prior profiles are also scaled (dashed lines) so that the profile corresponds to each retrieved XCH<sub>4</sub>. The difference of the scaled profile compared to the measured concentration is also shown: (<b>b</b>) OPER rpro–AirCore, and (<b>d</b>) WFMD v1.8–AirCore.</p>
Full article ">Figure 19
<p>Collection of the AirCore XCH<sub>4</sub> results with co-located TCCON GGG2020 and TROPOMI observations.</p>
Full article ">Figure 20
<p>Collection of the TROPOMI high-latitude evaluation results obtained in this paper. The bar plots show the average difference of TROPOMI (OPER rpro or WFMD v1.8 product) and ground-based references, and the error bars depict the standard deviation. The references from left to right are TCCON stations at East Trout Lake, Sodankylä, Ny-Ålesund, and Eureka, then COCCON stations at St. Petersburg, Fairbanks, Sodankylä (SN039 and SN122), and Kiruna, and AirCore measurements at Sodankylä. It should be noted that the temporal sampling differs between the references.</p>
Full article ">Figure A1
<p>TCCON GGG2020 (blue) and co-located TROPOMI daily median XCH<sub>4</sub> (red) with (<b>a</b>) OPER and (<b>c</b>) WFMD v1.2 retrievals, and TROPOMI−TCCON differences (<b>b</b>,<b>d</b>) at Eureka (Canada) TCCON site. The dashed line in (<b>b</b>,<b>d</b>) show the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure A2
<p>TCCON GGG2020 (blue) and co-located TROPOMI daily median XCH<sub>4</sub> (red) with (<b>a</b>) OPER and (<b>c</b>) WFMD v1.2 retrievals, and TROPOMI−TCCON differences (<b>b</b>,<b>d</b>) at Ny-Ålesund (Norway) TCCON site. The dashed line in (<b>b</b>,<b>d</b>) show the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure A3
<p>TCCON GGG2020 (blue) and co-located TROPOMI daily median XCH<sub>4</sub> (red) with (<b>a</b>) OPER and (<b>c</b>) WFMD v1.2 retrievals, and TROPOMI−TCCON differences (<b>b</b>,<b>d</b>) at Sodankylä (Finland) TCCON site. The dashed line in (<b>b</b>,<b>d</b>) show the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">Figure A4
<p>TCCON GGG2020 (blue) and co-located TROPOMI daily median XCH<sub>4</sub> (red) with (<b>a</b>) OPER and (<b>c</b>) WFMD v1.2 retrievals, and TROPOMI−TCCON differences (<b>b</b>,<b>d</b>) at East Trout Lake (Canada) TCCON site. The dashed line in (<b>b</b>,<b>d</b>) show the mean difference at the site and shaded area the standard deviation of the mean.</p>
Full article ">
19 pages, 10843 KiB  
Article
Development of a Daily Cloud-Free Snow-Cover Dataset Using MODIS-Based Snow-Cover Probability for High Mountain Asia during 2000–2020
by Dajiang Yan, Yinsheng Zhang and Haifeng Gao
Remote Sens. 2024, 16(16), 2956; https://doi.org/10.3390/rs16162956 - 12 Aug 2024
Viewed by 626
Abstract
Investigating the changes in snow cover caused by climate change is extremely important and has attracted increasing attention in cryosphere and climate research. Optimal remote sensing-based snow datasets can provide long-term daily and global spatial-temporal snow-cover distribution at regional and global scales. However, [...] Read more.
Investigating the changes in snow cover caused by climate change is extremely important and has attracted increasing attention in cryosphere and climate research. Optimal remote sensing-based snow datasets can provide long-term daily and global spatial-temporal snow-cover distribution at regional and global scales. However, the application of these snow-cover products is inevitably limited because of the space–time discontinuities caused by cloud obscuration, which poses a significant challenge in snowpack-related studies, especially in High Mountain Asia (HMA), an area that has high-elevation mountains, complex terrain, and harsh environments and has fewer observation stations. To address this issue, we developed an improved five-step hybrid cloud removal strategy by integrating the daily merged snow-cover probability (SCP) algorithm, eight-day merged SCP algorithm, decision tree algorithm, temporal downscaling algorithm, and optimal threshold segmentation algorithm to produce a 21-year, daily cloud-free snow-cover dataset using two daily MODIS snow-cover products over the HMA. The accuracy assessment demonstrated that the newly developed cloud-free snow-cover product achieved a mean overall accuracy of 93.80%, based on daily classified snow depth observations from 86 meteorological stations over 10 years. The time series of the daily percentage of binary snow-cover over HMA was analyzed during this period, indicating that the maximum snow cover tended to change more dramatically than the minimum snow cover. The annual snow-cover duration (SCD) experienced an insignificantly increasing trend over most of the northeastern and southwestern HMA (e.g., Qilian, eastern Kun Lun, the east of Inner Tibet, the western Himalayas, the central Himalayas, and the Hindu Kush) and an insignificant declining trend over most of the northwestern and southeastern HMA (e.g., the eastern Himalayas, Hengduan, the west of Inner Tibet, Pamir, Hissar Alay, and Tien). This new high-quality snow-cover dataset will promote studies on climate systems, hydrological modeling, and water resource management in this remote and cold region. Full article
Show Figures

Figure 1

Figure 1
<p>Study region. In the above figure, W, C, E, and S indicate western, central, eastern, and southern, respectively.</p>
Full article ">Figure 2
<p>Schematic illustration of the gap-filling method.</p>
Full article ">Figure 3
<p>Cloud removal results for each step on 1 January 2004 (the white color indicates clouds or missing gaps; the image marked SCP-8day is snow-cover probability from 1 January 2004 to 8 January 2004).</p>
Full article ">Figure 4
<p>Comparison of the original binary snow cover and cloud-free binary snow cover from 1 January 2004 to 8 January 2004. (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>,<b>s</b>,<b>v</b>) are the MOD10C1 snow-cover product; (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>,<b>t</b>,<b>w</b>) are the MYD10C1 snow-cover product; and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>,<b>u</b>,<b>x</b>) are the cloud-free snow-cover product. (The white color refers to clouds; the blue color refers to snow cover, and the light blue color refers to non-snow cover.).</p>
Full article ">Figure 5
<p>Spatial distributions of the overall accuracy, underestimation error, and overestimation error of the new cloud-free snow-cover product over these 86 stations.</p>
Full article ">Figure 6
<p>The overall accuracy, underestimation error, and overestimation error of the newly developed snow-cover product under different weather conditions.</p>
Full article ">Figure 7
<p>Monthly snow-cover distribution in 2004 (blue indicates snow cover, and light blue indicates non-snow cover).</p>
Full article ">Figure 8
<p>The snow-cover fraction over the whole HMA region and its subregions. Each image shows the multiyear mean, minimum, and maximum daily snow-cover fractions in each snow year (1 August to 31 July of the following year) during 2000–2020.</p>
Full article ">Figure 9
<p>Time series of daily snow-cover fractions from 2000–2020 in snow years (1 August to 31 July of the following year).</p>
Full article ">Figure 10
<p>The trend of the annual SCD during 2000–2020 and its significance test: (<b>a</b>) total annual SCD; (<b>b</b>) annual SCD in autumn; (<b>c</b>) annual SCD in winter; (<b>d</b>) annual SCD in spring.</p>
Full article ">
24 pages, 2241 KiB  
Article
Measurement of Tourism Ecological Efficiency and Analysis of Influencing Factors under the Background of Climate Change: A Case Study of Three Provinces in China’s Cryosphere
by Yubin Wu, Feiyang He, Zhujun Sun and Yongyu Wang
Sustainability 2024, 16(14), 6085; https://doi.org/10.3390/su16146085 - 16 Jul 2024
Viewed by 699
Abstract
Against the backdrop of climate change and the “dual carbon” goals, enhancing the ecological efficiency of cryospheric tourism is crucial not only for the high-quality development of the tourism industry itself but also for the protection of the ecological environment and the promotion [...] Read more.
Against the backdrop of climate change and the “dual carbon” goals, enhancing the ecological efficiency of cryospheric tourism is crucial not only for the high-quality development of the tourism industry itself but also for the protection of the ecological environment and the promotion of green sustainable development in the cryospheric region. In light of this, this study, taking climate change as its background and based on the perspective of carbon emission constraints, integrates multidimensional factors such as “climate change, carbon emission constraints, and cryospheric resources” into a unified measurement framework to construct a model for evaluating the ecological efficiency of tourism in the cryosphere. Specifically, the model considers inputs, expected outputs, and unexpected outputs. Subsequently, employing the super-efficiency slack-based measure (SBM) model, this study measures the tourism ecological efficiency (TEE) of three provinces (Xinjiang, Qinghai, Tibet) in the cryosphere from 2013 to 2021 and utilizes the Malmquist–Luenberger index and gray correlation model to reveal their dynamic changes, efficiency decomposition, and influencing factors. The results indicate that: (1) The overall mean of TEE in the cryosphere is between 0.2428 and 1.2142, Over the study period, the average annual growth rate and corresponding confidence interval were 14.74%, (−8.61%, 64.23%), showing a significant fluctuating growth trend. Among them, Xinjiang stands out, with its mean scores ranging from 0.2418 to 1.6229, surpassing the overall average level of the cryosphere. (2) During the study period, the overall dynamic efficiency of tourism ecology in the cryosphere increased by 21.54%, driven by the synergy of technological progress (TC), pure technical efficiency (PET), and scale efficiency (SE). For each province, the dynamic efficiency of tourism ecology has improved, but to varying degrees. (3) Regarding the driving factors of TEE in the cryosphere, each driving factor is closely related to TEE, ranked from large to small as follows: carbon emission structure, level of economic development, infrastructure, intensity of technological input, industrial structure, resource endowment, and environmental regulation. This article holds theoretical and practical significance for promoting the high-quality development of polar tourism and achieving synergistic progress between the economy and environment. Full article
(This article belongs to the Special Issue Climate Change Impacts and Sustainable Tourism)
Show Figures

Figure 1

Figure 1
<p>Number of domestic and foreign tourists received. Data source: statistics for each province for the 2021 National Economic and Social Development (<a href="https://www.xzxw.com/" target="_blank">https://www.xzxw.com/</a>, <a href="https://www.xinjiang.gov.cn/" target="_blank">https://www.xinjiang.gov.cn/</a>, <a href="http://tjj.qinghai.gov.cn/" target="_blank">http://tjj.qinghai.gov.cn/</a>, all accessed on 3 November 2023).</p>
Full article ">Figure 2
<p>Distribution of glaciers and ski resorts in the cryosphere. Note: The standard map No. GS (2023) 2767 downloaded from the standard map service website of the National Bureau of Surveying, Mapping and Geographic Information is made, and the base map is not modified.</p>
Full article ">Figure 3
<p>Trends in tourism carbon emissions and energy consumption.</p>
Full article ">Figure 4
<p>Dynamic changes of ML index in each province of the cryosphere.</p>
Full article ">Figure 5
<p>Dynamic change of decomposition efficiency in the cryosphere.</p>
Full article ">Figure 6
<p>Differences in the correlation degree of driving factors of tourism eco-efficiency.</p>
Full article ">
22 pages, 23824 KiB  
Article
DEDNet: Dual-Encoder DeeplabV3+ Network for Rock Glacier Recognition Based on Multispectral Remote Sensing Image
by Lujun Lin, Lei Liu, Ming Liu, Qunjia Zhang, Min Feng, Yasir Shaheen Khalil and Fang Yin
Remote Sens. 2024, 16(14), 2603; https://doi.org/10.3390/rs16142603 - 16 Jul 2024
Viewed by 496
Abstract
Understanding the distribution of rock glaciers provides key information for investigating and recognizing the status and changes of the cryosphere environment. Deep learning algorithms and red–green–blue (RGB) bands from high-resolution satellite images have been extensively employed to map rock glaciers. However, the near-infrared [...] Read more.
Understanding the distribution of rock glaciers provides key information for investigating and recognizing the status and changes of the cryosphere environment. Deep learning algorithms and red–green–blue (RGB) bands from high-resolution satellite images have been extensively employed to map rock glaciers. However, the near-infrared (NIR) band offers rich spectral information and sharp edge features that could significantly contribute to semantic segmentation tasks, but it is rarely utilized in constructing rock glacier identification models due to the limitation of three input bands for classical semantic segmentation networks, like DeeplabV3+. In this study, a dual-encoder DeeplabV3+ network (DEDNet) was designed to overcome the flaws of the classical DeeplabV3+ network (CDNet) when identifying rock glaciers using multispectral remote sensing images by extracting spatial and spectral features from RGB and NIR bands, respectively. This network, trained with manually labeled rock glacier samples from the Qilian Mountains, established a model with accuracy, precision, recall, specificity, and mIoU (mean intersection over union) of 0.9131, 0.9130, 0.9270, 0.9195, and 0.8601, respectively. The well-trained model was applied to identify new rock glaciers in a test region, achieving a producer’s accuracy of 93.68% and a user’s accuracy of 94.18%. Furthermore, the model was employed in two study areas in northern Tien Shan (Kazakhstan) and Daxue Shan (Hengduan Shan, China) with high accuracy, which proved that the DEDNet offers an innovative solution to more accurately map rock glaciers on a larger scale due to its robustness across diverse geographic regions. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the QLMs. (<b>b</b>) The location of the study area in the QLMs. (<b>c</b>) Location of VIAs (visual interpretation areas), MTA (model test area), field investigation, model training, model validation, and model test rock glaciers in the study area. In VIAs, the centroids of subareas A, B, C, and D are [38.314, 100.424], [38.116, 100.088], [37.526, 101.723], and [37.367, 101.315], respectively. The centroid of MTA is [37.758, 101.400].</p>
Full article ">Figure 2
<p>Flowchart of the methodology. RG_mm and RG_man represent the rock glacier model mapped and rock glacier manually delineated, respectively.</p>
Full article ">Figure 3
<p>The characteristics of four rock glaciers: (<b>a<sub>1</sub></b>–<b>d<sub>1</sub></b>) field photos, (<b>a<sub>2</sub></b>–<b>d<sub>2</sub></b>) GF1/6 images, and (<b>a<sub>3</sub></b>–<b>d<sub>3</sub></b>) Google Earth 3D oblique view. The dLL locations in the first row images represent the centroid of four rock glaciers. The red lines represent the boundary of rock glaciers.</p>
Full article ">Figure 4
<p>The DEDNet with the backbone of HRNetV2 and the block 1 and block 2 we designed.</p>
Full article ">Figure 5
<p>The flowchart of mapping rock glaciers in MTA.</p>
Full article ">Figure 6
<p>Three area-extracting methods, where the area of (<b>A</b>) corresponds to that of a; the area of (<b>B</b>) corresponds to the combined area of b<sub>1</sub>, b<sub>2</sub>, b<sub>3</sub>, b<sub>4</sub>, b<sub>5</sub>, and b<sub>6</sub>; the area of (<b>C</b>) corresponds to the combined area of c<sub>1</sub>, c<sub>2</sub>, and c<sub>3</sub>. The centroids of polygon (<b>A</b>–<b>C</b>) are [38.345, 100.334], [38.337, 100.327], and [38.355, 100.326], respectively.</p>
Full article ">Figure 7
<p>The evaluation metrics on validation set during training and validating the DEDNet.</p>
Full article ">Figure 8
<p>Comparison of boundaries between RG_mm and RG_man in VIAs (<b>a</b>,<b>b</b>). The panels (<b>c</b>–<b>f</b>) correspond to the yellow box in subareas A, B, C, and D. The background images of (<b>a</b>,<b>b</b>) are the true color image composed of a combination of RGB bands of GF1/6. The background images of (<b>c</b>–<b>f</b>) are a combination of NIR-G-B bands of GF1/6.</p>
Full article ">Figure 9
<p>Scatterplots illustrating the areas of RG_mm polygons compared to RG_man polygons on the training and validation datasets (<b>a</b>), and boxplots displaying the deviations in area across various scales on the training and validation datasets (<b>b</b>). In boxplots, “small” denotes areas less than 0.10 km<sup>2</sup>, “medium_s” refers to areas between 0.10 and 0.50 km<sup>2</sup>, “medium_l” signifies areas between 0.50 and 1.00 km<sup>2</sup>, and “large” indicates areas larger than 1.00 km<sup>2</sup>.</p>
Full article ">Figure 10
<p>Comparison of RG_man’s boundaries with RG_mm’s boundaries in MTA (<b>a</b>), along with two typical subregions (<b>b</b>,<b>c</b>), and their corresponding probability heatmap (<b>d</b>,<b>e</b>) on GF1/6 imagery. NRG_mm represents non-rock glacier model mapped. Background image is a true color image composed of a combination of RGB bands of GF1/6.</p>
Full article ">Figure 11
<p>Scatterplots illustrating the areas of RG_mm compared to RG_man on the test datasets (<b>a</b>), and boxplots displaying the deviations in area across various scales on the test datasets (<b>b</b>). Here, “small”, “medium_s”, “medium_l”, and “large” correspond to the same meanings as the terms used in <a href="#remotesensing-16-02603-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 12
<p>Boundaries of RG_man and four models outlined in three subregions in MTA. (<b>a</b>–<b>c</b>) represent three distinct subregions. 1, 2, 3, and 4 represent four different models. The four RG_models represent the rock glaciers delineated by corresponding models. For example, RG_CDNet_Positive represents rock glaciers delineated by CDNet_Positive. Black arrows represent adjacent rock glaciers mapped by one model, which were delineated as a single union rock glacier by another model. White arrows indicate that DEDNet_Positive_Negative identifies these rock glaciers better than CDNet_Positive_Negative.</p>
Full article ">Figure 13
<p>Locations of northern Tien Shan (Kazakhstan) (<b>a</b>) and Daxue Shan (<b>d</b>). Boundaries of RG_man and RG_inventory in northern Tien Shan (<b>b</b>) and Daxue Shan (<b>e</b>). Rock glaciers were model mapped but not included in the inventory of northern Tien Shan (<b>c</b>) and Daxue Shan (<b>f</b>). The background images of (<b>b</b>,<b>e</b>) are the true color image composed of a combination of RGB bands of GF1/6. The background images of (<b>c</b>,<b>f</b>) are a combination of NIR-G-B bands of GF1/6.</p>
Full article ">
25 pages, 18712 KiB  
Article
Spatial Distribution and Variation in Debris Cover and Flow Velocities of Glaciers during 1989–2022 in Tomur Peak Region, Tianshan Mountains
by Weiyong Zhou, Min Xu and Haidong Han
Remote Sens. 2024, 16(14), 2587; https://doi.org/10.3390/rs16142587 - 15 Jul 2024
Viewed by 728
Abstract
In this study, we utilized a feature optimization method combining texture and topographical factors with the random forest (RF) approach to identify changes in the extent of the debris cover around the Tianshan Tomur Peak between 1989 and 2022. Based on Sentinel-1 image [...] Read more.
In this study, we utilized a feature optimization method combining texture and topographical factors with the random forest (RF) approach to identify changes in the extent of the debris cover around the Tianshan Tomur Peak between 1989 and 2022. Based on Sentinel-1 image data, we extracted glacier flow velocities using an offset tracking method and conducted a long-term analysis of flow velocities in combination with existing datasets. The debris identification results for 2022 showed that the debris-covered area in the study region was 409.2 km2, constituting 22.8% of the total glacier area. Over 34 years, the area of debris cover expanded by 69.4 km2, reflecting a growth rate of 20.0%. Analysis revealed that glaciers in the Tomur Peak area have been decelerating at an overall rate of −4.0% per decade, with the complexity of the glacier bed environment and the instability of the glacier’s internal structure contributing to significant seasonal and interannual variability in the movement speeds of individual glaciers. Full article
(This article belongs to the Special Issue Remote Sensing of Cryosphere and Related Processes)
Show Figures

Figure 1

Figure 1
<p>The geographical location and glacier distribution in the study area. Glaciers exceeding 40 km<sup>2</sup> are emphasized with FDC (fractional debris cover from linear spectral unmixing) color scales.</p>
Full article ">Figure 2
<p>Workflow schematic for the semi-automatic delineation of the extent of debris cover and methodology for mapping glacier flow velocities. NC: network common data format; LST: land surface temperature.</p>
Full article ">Figure 3
<p>Normalized importance ranking for 30 feature variables, with No. corresponding to <a href="#remotesensing-16-02587-t003" class="html-table">Table 3</a>. This curve illustrates how the accuracy metrics varied as the number of feature variables increased.</p>
Full article ">Figure 4
<p>Area–altitude distributions of the debris-covered and clean ice/snow in the study region.</p>
Full article ">Figure 5
<p>Histogram showing the normal distribution of the elevation and slope for the debris-covered area, with dashed lines marking the upper and lower quartiles. (<b>a</b>) Southwest region (KB, TM, QT in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>); (<b>b</b>) northwest region (KI, NI, SI in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>); (<b>c</b>) northeast region (TG, WK in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>); (<b>d</b>) southeast region (QK, KQ in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 6
<p>Changes in the extent of debris-cover on glaciers (larger than 40 km<sup>2</sup>) in Tomur Peak Region from 1989 to 2022.</p>
Full article ">Figure 7
<p>An example frequency distribution of surface displacement over stable terrain areas.</p>
Full article ">Figure 8
<p>Seasonal variations in the average flow velocity along the central flow line from January 2021 to December 2022. The orange region covers the melt season of glaciers, and interval symbols indicate the 95% confidence interval of the mean flow velocity. The abbreviations of glacier names are shown in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 9
<p>(<b>a</b>) Spatial distribution of average glacier flow velocities in the northwest region from 1989 to 2022; (<b>b</b>–<b>d</b>) spatial distributions of the flow velocity change trend of the Kaindy Glacier, Northern Inylchek Glacier, and Southern Inylchek Glacier. Slope, slope of linear regression; (<b>e</b>) changes in average flow velocities along the central flow line over the past 34 years.</p>
Full article ">Figure 10
<p>(<b>a</b>) Same as <a href="#remotesensing-16-02587-f009" class="html-fig">Figure 9</a>a, but for the southwest region; (<b>b</b>–<b>d</b>) spatial distributions of the flow velocity change trend of the Koxqar Baqi Glacier, Tomur Glacier, and Qiongtailan Glacier. Slope, slope of linear regression; (<b>e</b>) same as <a href="#remotesensing-16-02587-f009" class="html-fig">Figure 9</a>e, but for the southwest region.</p>
Full article ">Figure 11
<p>(<b>a</b>) Same as <a href="#remotesensing-16-02587-f009" class="html-fig">Figure 9</a>a, but for the east region; (<b>b</b>–<b>e</b>) spatial distributions of the flow velocity change trend of the Wukuer Glacier, Tugebieliqi Glacier, Qiongkuziwayi Glacier, and Keqiketieliekesu Glacier. Slope, slope of linear regression; (<b>f</b>) same as <a href="#remotesensing-16-02587-f009" class="html-fig">Figure 9</a>e, but for the east region.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>) Temporal variation and linear regression of regional average land surface temperature and cumulative precipitation; (<b>c</b>,<b>d</b>) Mann–Kendall test for regional average land surface temperature and cumulative precipitation.</p>
Full article ">Figure 13
<p>(<b>a</b>–<b>e</b>) Upward evolution in debris cover and debris-covered terminus retreat on Mushketova Glacier from 1989 to 2022.</p>
Full article ">Figure 14
<p>Diagram of climate–debris–glacier system interactions and feedback mechanisms.</p>
Full article ">Figure 15
<p>Spatial distribution of thermal resistances of debris layers in the study area. The abbreviations of glacier names are detailed in <a href="#remotesensing-16-02587-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 16
<p>The relationship between debris thickness and elevation along the central flowline of the Koxkar Glacier, with debris thickness estimates derived from Rounce et al. [<a href="#B89-remotesensing-16-02587" class="html-bibr">89</a>] is accessible at <a href="https://nsidc.org/data/hma_dte/versions/1" target="_blank">https://nsidc.org/data/hma_dte/versions/1</a> (accessed on 1 July 2024).</p>
Full article ">
22 pages, 16146 KiB  
Article
The Impact of Glacial Shrinkage on Future Streamflow in the Urumqi River Source Region of Eastern Tien Shan, Central Asia
by Weibo Zhao, Zhongqin Li, Hongliang Li, Chunhai Xu, Jianxin Mu and Yefei Yang
Remote Sens. 2024, 16(14), 2546; https://doi.org/10.3390/rs16142546 - 11 Jul 2024
Viewed by 681
Abstract
Understanding changes in runoff due to climate variations in glacier-dominated headwaters is key to managing water resources and dryland watersheds effectively and rationally. The continuous glacier shrinkage caused by climate warming has significantly impacted the water supply and ecological systems in the vast [...] Read more.
Understanding changes in runoff due to climate variations in glacier-dominated headwaters is key to managing water resources and dryland watersheds effectively and rationally. The continuous glacier shrinkage caused by climate warming has significantly impacted the water supply and ecological systems in the vast arid regions of Central Asia, attracting extensive public concern. The study results indicate an increase in total runoff at the Urumqi River source region during both the baseline (1997–2016) and mid-century (2040–2059) periods, encompassing rain, glacier meltwater, and snowmelt components. Compared to the baseline period, the temperature increases by the mid-century under the three climate scenarios (SSP1−26, SSP2−45, and SSP5−85) range from 0.98 to 1.48 °C. In this region, during the period from 1997 to 2016, glacier meltwater was the dominant component of runoff, comprising 42.10–43.79% of the total, followed by snowmelt at 29.64–30.40% and rainfall contributions of 26.56–27.49%. Additionally, glacier storage in this typical catchment responds quickly to temperature fluctuations, significantly impacting runoff. The Urumqi River source region’s runoff exhibits heightened sensitivity to these temperature shifts compared to precipitation effects. We hypothesized three glacier coverage scenarios: unchanged at 100% glaciation, reduced by half to 50%, and fully retreated to 0% glaciation. Analysis of these scenarios demonstrated that glaciers are pivotal in runoff formation. Under the SSP1−26, SSP2−45, and SSP5−85 climate scenarios, glaciers contributed additional runoff increases of 51.61%, 57.64%, and 62.07%, respectively. Generally, glaciers play a critical role in supplying water in dry areas. Thus, accurately forecasting future water resource shifts in high-altitude glacier regions is crucial for downstream water resource management and utilization. Full article
(This article belongs to the Special Issue Advances in Remote Sensing in Glacial and Periglacial Geomorphology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The Urumqi River source region, marked by a solid red circle, is situated within the Xinjiang Uygur Autonomous Region of China. (<b>b</b>) This figure details the study area, highlighting the locations of Daxigou Meteorological Station and Zongkong Hydrometeorological Stations, along with glacier distribution.</p>
Full article ">Figure 2
<p>Performance after percentile correction of the CMIP6 precipitation output. (<b>a</b>) The calibration period (2007–2011); (<b>b</b>) the validation period (2012–2016).</p>
Full article ">Figure 3
<p>Comparison of temperature (<b>a</b>) and precipitation (<b>b</b>) after statistical downscaling between observed and CMIP6 output results from 2012−2016.</p>
Full article ">Figure 4
<p>Comparison of average monthly temperature (<b>a</b>) and precipitation (<b>b</b>) between the baseline (1997–2016) and mid-century period (2040–2059).</p>
Full article ">Figure 5
<p>Monthly changes in temperature (<b>a</b>–<b>c</b>) and precipitation (<b>d</b>–<b>f</b>) from 2040 to 2059.</p>
Full article ">Figure 6
<p>Multi-year monthly temperature (<b>a</b>) and precipitation (<b>b</b>) changes from 2040–2059 under the SSP1−26, SSP2−45, and SSP5−85 scenarios in comparison to the period 1997–2016.</p>
Full article ">Figure 7
<p>Daily runoff from observations and simulations in the calibration period (1997–2006) (<b>a</b>) and validation period (2007–2016) (<b>b</b>) are shown by red and black lines, and precipitation is shown by gray lines. The performance of the HBV-Light model during calibration and validation was assessed using the Nash–Sutcliffe efficiency and the coefficient of determination. Both metrics ranged from 0.68 to 0.79 for the respective periods. Specifically, the Nash–Sutcliffe efficiency values were 0.69 and 0.68 during calibration and 0.79 and 0.78 during validation. Detailed results are shown in <a href="#remotesensing-16-02546-t005" class="html-table">Table 5</a>. These results indicate that the model can effectively simulate daily runoff variations in the Urumqi River source region during both the calibration and validation periods, demonstrating good applicability of the HBV-Light model in glacier areas of small watersheds.</p>
Full article ">Figure 8
<p>Comparison of runoff components for validation period (1997–2006, (<b>a</b>)) and calibration period (2007–2016, (<b>b</b>)).</p>
Full article ">Figure 9
<p>Comparison of spring, summer, and autumn glacier runoff from 1997 to 2016 (winter runoff is 0 in the Urumqi River source region), including total runoff (<b>a</b>), rain runoff (<b>b</b>), snowmelt runoff (<b>c</b>), and glacier meltwater runoff (<b>d</b>).</p>
Full article ">Figure 10
<p>Total monthly average simulated runoff relative to baseline (1997–2016) in future period (2040–2059). (<b>a</b>–<b>c</b>) represent the comparisons of monthly runoff under different glaciation scenarios for SSP1−26, SSP2−45, and SSP5−85 climate scenarios, respectively.</p>
Full article ">Figure 11
<p>The annual variations in total runoff, rain runoff, snowmelt runoff, and glacier meltwater runoff under the three climate scenarios (SSP1−26, SSP2−45, and SSP5−85) from 2040 to 2059.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) Comparison of runoff modeling results for the Urumqi River source region in mid-century period (2040–2059) (this study and Rounce 2022 [<a href="#B64-remotesensing-16-02546" class="html-bibr">64</a>]).</p>
Full article ">Figure 13
<p>Sensitivity test of the HBV-Light for changes in temperature (<b>a</b>), precipitation (<b>b</b>), and their combination (<b>c</b>) during 1997–2016.</p>
Full article ">
13 pages, 2277 KiB  
Technical Note
Early Radiometric Assessment of NOAA-21 Visible Infrared Imaging Radiometer Suite Reflective Solar Bands Using Vicarious Techniques
by Aisheng Wu, Xiaoxiong Xiong, Qiaozhen Mu, Amit Angal, Rajendra Bhatt and Yolanda Shea
Remote Sens. 2024, 16(14), 2528; https://doi.org/10.3390/rs16142528 - 10 Jul 2024
Viewed by 518
Abstract
The VIIRS instrument on the JPSS-2 (renamed NOAA-21 upon reaching orbit) spacecraft was launched in November 2022, making it the third sensor in the VIIRS series, following those onboard the SNPP and NOAA-20 spacecrafts, which are operating nominally. As a multi-disciplinary instrument, the [...] Read more.
The VIIRS instrument on the JPSS-2 (renamed NOAA-21 upon reaching orbit) spacecraft was launched in November 2022, making it the third sensor in the VIIRS series, following those onboard the SNPP and NOAA-20 spacecrafts, which are operating nominally. As a multi-disciplinary instrument, the VIIRS provides the worldwide user community with high-quality imagery and radiometric measurements of the land, atmosphere, cryosphere, and oceans. This study provides an early assessment of the calibration stability and radiometric consistency of the NOAA-21 VIIRS RSBs using the latest NASA SIPS C2 L1B products. Vicarious approaches are employed, relying on reflectance data obtained from the Libya-4 desert and Dome C sites, deep convective clouds, and simultaneous nadir overpasses, as well as intercomparison with the first two VIIRS instruments using MODIS as a transfer radiometer. The impact of existing band spectral differences on sensor-to-sensor comparison is corrected using scene-specific a priori hyperspectral observations from the SCIAMACHY sensor onboard the ENVISAT platform. The results indicate that the overall radiometric performance of the newly launched NOAA-21 VIIRS is quantitatively comparable to the NOAA-20 for the VIS and NIR bands. For some SWIR bands, the measured reflectances are lower by more than 2%. An upward adjustment of 6.1% in the gain of band M11 (2.25 µm), based on lunar intercomparison results, generates more consistent results with the NOAA-20 VIIRS. Full article
(This article belongs to the Section Satellite Missions for Earth and Planetary Exploration)
Show Figures

Figure 1

Figure 1
<p>Trends of the BRDF normalized reflectances before SBAF correction for SNPP and NOAA-20 and -21 VIIRS M1, M3, M7, and M10 over the Libya-4 site.</p>
Full article ">Figure 2
<p>Trends of the reflectances normalized by the mean of SNPP before SBAF correction for SNPP and NOAA-20 and -21 VIIRS M1, M3, M7, and M8 from DCCs.</p>
Full article ">Figure 3
<p>Reflectances as a function of solar zenith angle for SNPP and NOAA-20 and -21 VIIRS M1, M3, I1, and I2 over the Dome C site.</p>
Full article ">Figure 4
<p>VIIRS-to-MODIS reflectance ratios versus reflectance for SNPP and NOAA-20 and -21 VIIRS M2, M4, M7, and M10 from SNOs. Ratios at the pixel level are binned at a reflectance interval of 0.02.</p>
Full article ">Figure 5
<p>Following the establishment of the absolute calibration of NOAA-20 VIIRS via CPF, it can serve as a bridge between CPF and other reflective solar instruments.</p>
Full article ">
14 pages, 9927 KiB  
Article
Assessment of Soil Temperature and Its Change Trends in the Permafrost Regions of the Northern Hemisphere
by Yifan Wu, Guojie Hu, Lin Zhao, Defu Zou, Xiaofan Zhu, Yao Xiao, Tonghua Wu, Xiaodong Wu, Youqi Su and Rui Zhang
Land 2024, 13(7), 1029; https://doi.org/10.3390/land13071029 - 9 Jul 2024
Viewed by 500
Abstract
In this paper, we used data from 42 soil temperature observation sites in permafrost regions throughout the Northern Hemisphere to analyze the characteristics and variability in soil temperature. The observation data were used to evaluate soil temperature simulations at different depths from 10 [...] Read more.
In this paper, we used data from 42 soil temperature observation sites in permafrost regions throughout the Northern Hemisphere to analyze the characteristics and variability in soil temperature. The observation data were used to evaluate soil temperature simulations at different depths from 10 CMIP6 models in the permafrost region of the Northern Hemisphere. The results showed that the annual average soil temperature in the permafrost regions in the Northern Hemisphere gradually decreased with increasing latitude, and the soil temperature gradually decreased with depth. The average soil temperatures at different depths were mainly concentrated around 0 °C. The 10 CMIP6 models performed well in simulating soil temperature, but most models tended to underestimate temperatures compared to the measured values. Overall, the CESM2 model yielded the best simulation results, whereas the CNRM-CM6-1 model performed the worst. The change trends in annual average soil temperature across the 42 sites ranged from −0.17 °C/10a to 0.41 °C/10a from 1900 to 2014, the closer to the Arctic, the faster the soil warming rate. The rate of soil temperature change also varied at different depths between 1900–2014 and 1980–2014. The rate of soil temperature change from 1980 to 2014 was approximately three times greater than that from 1900 to 2014. Full article
Show Figures

Figure 1

Figure 1
<p>Spatial distribution diagram of temperature observation stations in permafrost areas in the Northern Hemisphere.</p>
Full article ">Figure 2
<p>Spatial distribution of annual average soil temperature among observation stations in permafrost areas of the Northern Hemisphere. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, and 200–300 cm, respectively.</p>
Full article ">Figure 3
<p>Taylor plot of soil temperature simulated and observed for different CMIP 6 modes from 2003 to 2012. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, and 200–300 cm, respectively.</p>
Full article ">Figure 4
<p>Observed and simulated soil temperature trends in the Northern Hemisphere from 2003 to 2012. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, 200–300 cm, respectively.</p>
Full article ">Figure 5
<p>Rate of soil temperature change in the permafrost region of the Northern Hemisphere from 1900 to 2014. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, 200–300 cm, respectively.</p>
Full article ">Figure 5 Cont.
<p>Rate of soil temperature change in the permafrost region of the Northern Hemisphere from 1900 to 2014. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, 200–300 cm, respectively.</p>
Full article ">Figure 6
<p>Trend of soil temperature in Northern Hemisphere permafrost from 1900 to 2014 and 1950 to 2014. (<b>a</b>), (<b>b</b>), (<b>c</b>) and (<b>d</b>) represent 0–50 cm, 50–100 cm, 100–200 cm, 200–300 cm, respectively.</p>
Full article ">
24 pages, 22139 KiB  
Article
Improving the Estimation of Lake Ice Thickness with High-Resolution Radar Altimetry Data
by Anna Mangilli, Claude R. Duguay, Justin Murfitt, Thomas Moreau, Samira Amraoui, Jaya Sree Mugunthan, Pierre Thibaut and Craig Donlon
Remote Sens. 2024, 16(14), 2510; https://doi.org/10.3390/rs16142510 - 9 Jul 2024
Viewed by 708
Abstract
Lake ice thickness (LIT) is a sensitive indicator of climate change, identified as a thematic variable of Lakes as an Essential Climate Variable (ECV) by the Global Climate Observing System (GCOS). Here, we present a novel and efficient analytically based retracking approach for [...] Read more.
Lake ice thickness (LIT) is a sensitive indicator of climate change, identified as a thematic variable of Lakes as an Essential Climate Variable (ECV) by the Global Climate Observing System (GCOS). Here, we present a novel and efficient analytically based retracking approach for estimating LIT from high-resolution Ku-band (13.6 GHz) synthetic-aperture radar (SAR) altimetry data. The retracker method is based on the analytical modeling of the SAR radar echoes over ice-covered lakes that show a characteristic double-peak feature attributed to the reflection of the Ku-band radar waves at the snow–ice and ice–water interfaces. The method is applied to Sentinel-6 Unfocused SAR (UFSAR) and Fully Focused SAR (FFSAR) data, with their corresponding tailored waveform model, referred to as the SAR_LIT and FFSAR_LIT retracker, respectively. We found that LIT retrievals from Sentinel-6 high-resolution SAR data at different posting rates are fully consistent with the LIT estimations obtained from thermodynamic lake ice model simulations and from low-resolution mode (LRM) Sentinel-6 and Jason-3 data over two ice seasons during the tandem phase of the two satellites, demonstrating the continuity between LRM and SAR LIT retrievals. By comparing the Sentinel-6 SAR LIT estimates to optical/radar images, we found that the Sentinel-6 LIT measurements are fully consistent with the evolution of the lake surface conditions, accurately capturing the seasonal transitions of ice formation and melt. The uncertainty in the LIT estimates obtained with Sentinel-6 UFSAR data at 20 Hz is in the order of 5 cm, meeting the GCOS requirements for LIT measurements. This uncertainty is significantly smaller, by a factor of 2 to 3 times, than the uncertainty obtained with LRM data. The FFSAR processing at 140 Hz provides even better LIT estimates, with 20% smaller uncertainties. The LIT retracker analysis performed on data at the higher posting rate (140 Hz) shows increased performance in comparison to the 20 Hz data, especially during the melt transition period, due to the increased statistics. The LIT analysis has been performed over two representative lakes, Great Slave Lake and Baker Lake (Canada), demonstrating that the results are robust and hold for lake targets that differ in terms of size, bathymetry, snow/ice properties, and seasonal evolution of LIT. The SAR LIT retrackers presented are promising tools for monitoring the inter-annual variability and trends in LIT from current and future SAR altimetry missions. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere II)
Show Figures

Figure 1

Figure 1
<p>Illustration of the evolution of the bimodal lake ice thickness signature in Sentinel-6 UFSAR radargrams (<b>left column</b>) and the normalized waveforms (<b>right column</b>) at 20 Hz resolution at Great Slave Lake in December 2021 (<b>top</b>), February 2021 (<b>second row</b>), end of April 2021 (<b>third row</b>), and May 2021 (<b>bottom</b>). The black line in the plots of the right column corresponds to the mean waveform in the selected region of the lake.</p>
Full article ">Figure 2
<p>Illustration of the evolution of the bimodal lake ice thickness signature in Sentinel-6 FFSAR radargrams (<b>left column</b>) and the normalized waveforms (<b>right column</b>) at 140 Hz posting rate at Great Slave Lake in December 2021 (<b>top</b>), February 2021 (<b>second row</b>), end of April 2021 (<b>third row</b>), and May 2021 (<b>bottom</b>). The black line in the plots of the right column corresponds to the mean waveform in the selected region of the lake.</p>
Full article ">Figure 3
<p>The target lakes of the LIT analysis, Great Slave Lake and Baker Lake, Canada, are shown on the map (<b>bottom</b>) and with the satellite ground tracks superimposed on the lakes (<b>upper left</b> and <b>right</b>, respectively).</p>
Full article ">Figure 4
<p>Examples of Sentinel-6 UFSAR waveform with, in blue, the <tt>SAR_LIT</tt> fit (<b>left column</b>) and LIT histograms, with the corresponding Gaussian fits (<b>right column</b>), in the RoI of Great Slave Lake at the end of December 2020 (<b>top row</b>), in February 2021 (<b>second row</b>), in April 2021 (<b>third row</b>), and mid-May 2021 (<b>bottom row</b>).</p>
Full article ">Figure 5
<p>Examples of Sentinel-6 FFSAR waveforms with, in blue, the <tt>FFSAR_LIT</tt> fit (<b>left column</b>) and LIT histograms, with the corresponding Gaussian fits (<b>right column</b>), in the RoI of Great Slave Lake at the end of December 2020 (<b>top row</b>), in February 2021 (<b>second row</b>), in April 2021 (<b>third row</b>), and mid-May 2021 (<b>bottom row</b>).</p>
Full article ">Figure 6
<p>Example of the spatial evolution of the LIT estimates at Great Slave Lake (<b>left column</b>) and Baker Lake (<b>right column</b>) in February 2021. The top row plots show the results for the Sentinel-6 UFSAR data at 20 Hz, while the bottom row plots for the Sentinel-6 FFSAR data at 140 Hz. The gray lines in the bottom panels of the figures show the evolution of the reduced <math display="inline"><semantics> <msup> <mi>χ</mi> <mn>2</mn> </msup> </semantics></math> goodness of fit metric.</p>
Full article ">Figure 7
<p>Comparison of the LIT estimates obtained with Sentinel-6 high-resolution SAR data for one ice season at Great Slave Lake (<b>left</b>) and Baker Lake (<b>right</b>). The curves refer to UFSAR at 20 Hz (red), UFSAR at 140 Hz (purple), and FFSAR at 140 Hz (cyan).</p>
Full article ">Figure 8
<p>Evolution of LIT estimates at Great Slave Lake obtained with Sentinel-6 UFSAR at 20 Hz data (red), Sentinel-6 LRM data (green) and Jason-3 data (blue) for the 2020–2021 and 2021–2022 ice seasons (upper panel). The shaded regions of the corresponding colors refer to the LIT error envelopes at 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> for each case. The orange shaded area shows the evolution of LIT obtained from CLIMo thermodynamic simulations with different on-ice snow scenarios (see text in <a href="#sec4dot3-remotesensing-16-02510" class="html-sec">Section 4.3</a> for details). The middle panel shows the evolution of the mean 2 m air temperature (black) with the minimum and maximum values (gray shading) extracted from ERA5 data. The bottom panel shows the evolution of the 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> LIT uncertainties for the three datasets.</p>
Full article ">Figure 9
<p>Evolution of the LIT estimates at Baker Lake obtained with different datasets (see the caption of <a href="#remotesensing-16-02510-f008" class="html-fig">Figure 8</a> for details).</p>
Full article ">Figure 10
<p>Sentinel–6 20 Hz UFSAR LIT estimates superimposed on radar/optical images taken on the same dates for Great Slave Lake (<b>left</b>) and Baker Lake (<b>right</b>) on the lake area shown in the red boxes in the top row panels.</p>
Full article ">
Back to TopTop