[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

Journals

Article Types

Countries / Regions

Search Results (121)

Search Parameters:
Keywords = snow coverage

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 5433 KiB  
Article
A Novel Detection Algorithm for the Icing Status of Transmission Lines
by Dongxu Dai, Yan Hu, Hao Qian, Guoqiang Qi and Yan Wang
Symmetry 2024, 16(10), 1264; https://doi.org/10.3390/sym16101264 - 25 Sep 2024
Viewed by 388
Abstract
As more and more transmission lines need to pass through areas with heavy icing, the problem of transmission line faults caused by ice and snow disasters frequently occurs. Existing ice coverage monitoring methods have defects such as the use of a single monitoring [...] Read more.
As more and more transmission lines need to pass through areas with heavy icing, the problem of transmission line faults caused by ice and snow disasters frequently occurs. Existing ice coverage monitoring methods have defects such as the use of a single monitoring type, low accuracy of monitoring results, and an inability to obtain ice coverage data over time. Therefore, this study proposes a new algorithm for detecting the icing status of transmission lines. The algorithm uses two-dimensional multifractal detrended fluctuation analysis (2D MF-DFA) to determine the optimal sliding-window size and wave function and accurately segment and extract local feature areas. Based on the local Hurst exponent (Lh(z)) and the power-law relationship between the fluctuation function and the scale at multiple continuous scales, the ice-covered area of a transmission conductor was accurately detected. By analyzing and calculating the key target pixels, the icing thickness was accurately measured, achieving accurate detection of the icing status of the transmission lines. The experimental results show that this method can accurately detect ice-covered areas and the icing thickness of transmission lines under various working conditions, providing a strong guarantee for the safe and reliable operation of transmission lines under severe weather conditions. Full article
(This article belongs to the Special Issue Symmetry and Fractals: Theory and Applications)
Show Figures

Figure 1

Figure 1
<p>Common types of conductor icing.</p>
Full article ">Figure 2
<p>Illustration of a sliding window with a size of <math display="inline"><semantics> <mrow> <mn>7</mn> <mo>×</mo> <mn>7</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Original grayscale icing image.</p>
Full article ">Figure 4
<p>(<b>a</b>) Initial state of a sliding window with a size of <math display="inline"><semantics> <mrow> <mn>11</mn> <mo>×</mo> <mn>11</mn> </mrow> </semantics></math> in an icing image; (<b>b</b>) route passing points of the sliding window in the retinal icing image.</p>
Full article ">Figure 5
<p>Sub-regions of the sub-images of (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Double-log plots of the sub-image.</p>
Full article ">Figure 7
<p>Values of the local Hurst exponent as the sliding window moves.</p>
Full article ">Figure 8
<p><math display="inline"><semantics> <mrow> <mi>q</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics></math>. Experimenting with a synthetic image. (<b>a</b>) Original image; (<b>b</b>–<b>d</b>) plots of segmentation results with <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math> parameters of (13,16), (3,6), and (9,26), respectively.</p>
Full article ">Figure 9
<p>Multifractal analysis of (<b>a</b>) the original image, (<b>b</b>) double-logarithm plots, (<b>c</b>) the generalized Hurst exponent, and (<b>d</b>) the multifractal spectrum.</p>
Full article ">Figure 10
<p>Segmentation results of the synthesized image: (<b>a</b>) original image; (<b>b</b>) original pixels; (<b>c</b>) local fractal feature values; (<b>d</b>) contour edge state; and (<b>e</b>) segmentation results. Parameters: <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>[</mo> <mn>8</mn> <mo>,</mo> <mn>60</mn> <mo>]</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p><span class="html-italic">q</span> = −10, <span class="html-italic">w</span> = 5. Experimenting with simple images of ice coverage. (<b>a</b>) Original image; (<b>b</b>–<b>d</b>) plots of the segmentation results with <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math> parameters of (2,8), (12,23), and (15,30), respectively.</p>
Full article ">Figure 11 Cont.
<p><span class="html-italic">q</span> = −10, <span class="html-italic">w</span> = 5. Experimenting with simple images of ice coverage. (<b>a</b>) Original image; (<b>b</b>–<b>d</b>) plots of the segmentation results with <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math> parameters of (2,8), (12,23), and (15,30), respectively.</p>
Full article ">Figure 12
<p><span class="html-italic">q</span> = 10, <span class="html-italic">w</span> = 5. Experimenting with complex images of ice cover. (<b>a</b>) Original image; (<b>b</b>–<b>d</b>) plots of the segmentation results with <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </semantics></math> parameters of (25,30), (6,22), and (6,24), respectively.</p>
Full article ">Figure 13
<p>(<b>a</b>) The segmented image; (<b>b</b>) a mesh representation of the segmented image; (<b>c</b>) an enlarged view of the mesh in (<b>b</b>); (<b>d</b>) the thickness of the icing area.</p>
Full article ">Figure 14
<p>The 20 selected thickness data points.</p>
Full article ">Figure 15
<p>(<b>a</b>) The original icing image; (<b>b</b>) a mesh representation of the segmented image; (<b>c</b>) the thickness of the icing area.</p>
Full article ">
21 pages, 15871 KiB  
Article
Tracking Forest Disturbance in Northeast China’s Cold-Temperate Forests Using a Temporal Sequence of Landsat Data
by Yueting Wang, Xiang Jia, Xiaoli Zhang, Lingting Lei, Guoqi Chai, Zongqi Yao, Shike Qiu, Jun Du, Jingxu Wang, Zheng Wang and Ran Wang
Remote Sens. 2024, 16(17), 3238; https://doi.org/10.3390/rs16173238 - 1 Sep 2024
Viewed by 753
Abstract
Cold-temperate forests (CTFs) are not only an important source of wood but also provide significant carbon storage in China. However, under the increasing pressure of human activities and climate change, CTFs are experiencing severe disturbances, such as logging, fires, and pest infestations, leading [...] Read more.
Cold-temperate forests (CTFs) are not only an important source of wood but also provide significant carbon storage in China. However, under the increasing pressure of human activities and climate change, CTFs are experiencing severe disturbances, such as logging, fires, and pest infestations, leading to evident degradation trends. Though these disturbances impact both regional and global carbon budgets and their assessments, the disturbance patterns in CTFs in northern China remain poorly understood. In this paper, the Genhe forest area, which is a typical CTF region located in the Inner Mongolia Autonomous Region, Northeast China (with an area of about 2.001 × 104 km2), was selected as the study area. Based on Landsat historical archived data on the Google Earth Engine (GEE) platform, we used the continuous change detection and classification (CCDC) algorithm and considered seasonal features to detect forest disturbances over nearly 30 years. First, we created six inter-annual time series seasonal vegetation index datasets to map forest coverage using the maximum between-class variance algorithm (OTSU). Second, we used the CCDC algorithm to extract disturbance information. Finally, by using the ECMWF climate reanalysis dataset, MODIS C6, the snow phenology dataset, and forestry department records, we evaluated how disturbances relate to climate and human activities. The results showed that the disturbance map generated using summer (June–August) imagery and the enhanced vegetation index (EVI) had the highest overall accuracy (88%). Forests have been disturbed to the extent of 12.65% (2137.31 km2) over the last 30 years, and the disturbed area generally showed a trend toward reduction, especially after commercial logging activities were banned in 2015. However, there was an unusual increase in the number of disturbed areas in 2002 and 2003 due to large fires. The monitoring of potential widespread forest disturbance due to extreme drought and fire events in the context of climate change should be strengthened in the future, and preventive and salvage measures should be taken in a timely manner. Our results demonstrate that CTF disturbance can be robustly mapped by using the CCDC algorithm based on Landsat time series seasonal imagery in areas with complex meteorological conditions and spatial heterogeneity, which is essential for understanding forest change processes. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>Forest disturbance analysis workflow.</p>
Full article ">Figure 3
<p>Mapping the forest. (<b>a</b>) OTSU value and forest area every year (Summer &amp; EVI); (<b>b</b>) forest cover synthesis map from 1990 to 2021 (Summer &amp; EVI).</p>
Full article ">Figure 4
<p>Schematic diagram of typical disturbance types and disturbance processes. (<b>a</b>) Logging in 1999; (<b>b</b>) anthropogenic fire in 2003; (<b>c</b>) wildfires in 2003 and 2010, respectively; (<b>d</b>) logging in 1990 and anthropogenic fire in 2003; (<b>e</b>) EVI curves for the sample points in disturbance areas from (<b>a</b>–<b>d</b>), with red boxes indicating disturbance events.</p>
Full article ">Figure 5
<p>Forest disturbance extraction. (<b>a</b>) Forest disturbance zone; (<b>b</b>) disturbance caused by logging after 1990; (<b>c</b>) disturbance caused by man-made fire in 2003; (<b>d</b>) other disturbances caused by multiple factors such as wildfire, etc.; <b>b1</b>–<b>d1</b> show the results of extracting forest disturbance information, <b>b2</b>–<b>d2</b> display the satellite images that correspond to these areas after the disturbance has occurred; (<b>e</b>) forest disturbance zone after fire-induced disturbances have been removed; (<b>f</b>) annual forest disturbance area caused by fires and other factors; the lines in the plot are the univariate linear trendlines.</p>
Full article ">Figure 6
<p>Distance of forest disturbance patches from roads and rivers. (<b>a</b>,<b>c</b>) represent the distance of disturbance patches from roads; (<b>b</b>,<b>d</b>) represent the distance of disturbance patches from rivers.</p>
Full article ">Figure 7
<p>Number of forest disturbance events. (<b>a</b>) Number of forest disturbance events in Genhe; (<b>b</b>) number of forest disturbance events in each administrative unit.</p>
Full article ">Figure 8
<p>The relationship between forest disturbance and its influencing factors. a<sub>i</sub>, b<sub>i</sub>, c<sub>i</sub>, d<sub>i</sub>, and e<sub>i</sub> are models of the area of disturbance and its influencing factors (annual precipitation, annual average temperature, annual snow cover days, the annual number of fires, and annual commercial logging output, respectively) for every year; the pink circles are for the anomalous years (2002 and 2003); the period of (<b>a<sub>1</sub></b>–<b>d<sub>1</sub></b>) is from 1991 to 2020; the period of (<b>a<sub>2</sub></b>–<b>d<sub>2</sub></b>) is from 2011 to 2020; the period of (<b>a<sub>3</sub></b>–<b>c<sub>3</sub></b>) is from 1991 to 2020, the disturbance area of (<b>a<sub>3</sub></b>–<b>c<sub>3</sub></b>) is the disturbance caused by factors other than fire; (<b>d<sub>3</sub></b>) is the model of the disturbance area and burned area for every year from 1991 to 2020; (<b>e<sub>1</sub></b>) annual commercial logging output; the period of (<b>e<sub>2</sub></b>,<b>e<sub>3</sub></b>) is from 1991 to 2015; the disturbance area of (<b>e<sub>3</sub></b>) is the disturbance caused by factors other than fire; the red line in the figure is the univariate linear trendline.</p>
Full article ">Figure 9
<p>Time and location of forest disturbance and fires. Note: the blue, pink, and purple areas are the areas where disturbances and fires occurred. It should be noted that the purple areas are areas where disturbances and fires occurred in the same year, and other areas with colors (except white and gray) are all where disturbances were detected but the region of fire did not exist in the auxiliary dataset.</p>
Full article ">
21 pages, 4625 KiB  
Article
Research on an Evaluation Method of Snowdrift Hazard for Railway Subgrades
by Shumao Qiu, Mingzhou Bai, Daming Lin, Yufang Zhang, Haoying Xia, Jiawei Fan, Wenjiao Zhou and Zhenyu Tang
Appl. Sci. 2024, 14(16), 7247; https://doi.org/10.3390/app14167247 - 17 Aug 2024
Viewed by 593
Abstract
The objective of this study is to investigate the potential risks posed by snowdrifts, a prevalent cause of natural disasters in northern China, on railway subgrades, and to assess their risk level. As a wind-driven process of snow migration and redeposition, snowdrifts pose [...] Read more.
The objective of this study is to investigate the potential risks posed by snowdrifts, a prevalent cause of natural disasters in northern China, on railway subgrades, and to assess their risk level. As a wind-driven process of snow migration and redeposition, snowdrifts pose a significant threat to the safety of transportation infrastructures. This study focuses on the Afu Railway in Xinjiang, situated on the northern slopes of the Eastern Tianshan Mountains, where it experiences periodic snowdrifts. We employed a combination of the Analytic Hierarchy Process (AHP) and fuzzy comprehensive evaluation (FCE) to construct an integrated evaluation system for assessing the risk of snowdrift to railway subgrades. The results indicate that subgrade design parameters and regional snowfield conditions are two key metrics affecting the extent of snowdrift disasters, with topography, vegetation coverage, and wind speed also exerting certain impacts. The evaluation method of this study aligns with the results of on-site observations, verifying its accuracy and practicality, thereby providing a solid risk assessment framework for snowdrifts along the railway. The scientific and systematic hazard assessment method of railway subgrades developed in this research provides basic data and theoretical support for future research, and provides a scientific basis for relevant departments to formulate countermeasures, so as to improve the safety and reliability of railway operations. Full article
Show Figures

Figure 1

Figure 1
<p>Map of days of snow cover on railways.</p>
Full article ">Figure 2
<p>Analytic Hierarchy Process (AHP) model.</p>
Full article ">Figure 3
<p>Ridge distribution or semi-ridge distribution. (<b>a</b>) right-skewed; (<b>b</b>) left-skewed; (<b>c</b>) symmetric.</p>
Full article ">Figure 4
<p>Trapezoidal distribution or semi-trapezoidal distribution. (<b>a</b>) right-skewed; (<b>b</b>) left-skewed; (<b>c</b>) symmetric.</p>
Full article ">Figure 5
<p>Monitoring sites and test section areas of weather stations.</p>
Full article ">Figure 6
<p>Winter wind speed distribution along the railway.</p>
Full article ">Figure 7
<p>Rose wind map of each monitoring point.</p>
Full article ">Figure 8
<p>Measurement of snow density.</p>
Full article ">Figure 9
<p>Humidity changes during snowfall along the railway.</p>
Full article ">Figure 10
<p>Indicator weights. (<b>a</b>) Level 1 index weight; (<b>b</b>) Level 2 index weight.</p>
Full article ">Figure 11
<p>ROC curve.</p>
Full article ">
21 pages, 4101 KiB  
Article
Two Decades of Arctic Sea-Ice Thickness from Satellite Altimeters: Retrieval Approaches and Record of Changes (2003–2023)
by Sahra Kacimi and Ron Kwok
Remote Sens. 2024, 16(16), 2983; https://doi.org/10.3390/rs16162983 - 14 Aug 2024
Viewed by 1261
Abstract
There now exists two decades of basin-wide coverage of Arctic sea ice from three dedicated polar-orbiting altimetry missions (ICESat, CryoSat-2, and ICESat-2) launched by NASA and ESA. Here, we review our retrieval approaches and discuss the composite record of Arctic ice thickness (2003–2023) [...] Read more.
There now exists two decades of basin-wide coverage of Arctic sea ice from three dedicated polar-orbiting altimetry missions (ICESat, CryoSat-2, and ICESat-2) launched by NASA and ESA. Here, we review our retrieval approaches and discuss the composite record of Arctic ice thickness (2003–2023) after appending two more years (2022–2023) to our earlier records. The present availability of five years of snow depth estimates—from differencing lidar (ICESat-2) and radar (CryoSat-2) freeboards—have benefited from the concurrent operation of two altimetry missions. Broadly, the dramatic volume loss (5500 km3) and Arctic-wide thinning (0.6 m) captured by ICESat (2003–2009), primarily due to the decline in old ice coverage between 2003 and 2007, has slowed. In the central Arctic, away from the coasts, the CryoSat-2 and shorter ICESat-2 records show near-negligible thickness trends since 2007, where the winter and fall ice thicknesses now hover around 2 m and 1.3 m, from a peak of 3.6 m and 2.7 m in 1980. Ice volume production has doubled between the fall and winter with the faster-growing seasonal ice cover occupying more than half of the Arctic Ocean at the end of summer. Seasonal ice behavior dominates the Arctic Sea ice’s interannual thickness and volume signatures. Full article
(This article belongs to the Special Issue Monitoring Sea Ice Loss with Remote Sensing Techniques)
Show Figures

Figure 1

Figure 1
<p>Two-layer model of sea ice assumed in thickness calculations.</p>
Full article ">Figure 2
<p>Arctic sea ice thickness composites from ICESat (IS), CryoSat-2 (CS-2) and ICESat-2 between 2003 and 2023. These 25 km-gridded composites are February and March averages of thickness estimates in SSM/I polar stereographic projection. For comparison, the 2-month averages are aligned with the winter campaigns (durations of ~33 days) of IS, which was not operated continuously as the altimeters on CS-2 and IS-2 observatories. There is a gap of about a year between the completion of the IS mission and the launch of CS-2. The third and fourth rows show the separate thickness retrievals in a 5-year overlap (2018–2023) between CS-2 and IS-2, highlighting the spatial differences between retrievals using snow depth from two approaches: modified climatology for CS-2 and snow depth calculated using differences between radar (CS-2) and lidar (IS-2) freeboards. Thicknesses are calculated within the Arctic basin (of ~7 × 10<sup>6</sup> km<sup>2</sup>) bounded by the gateways into the Pacific (Bering Strait), the Canadian Arctic Archipelago (CAA), and the Greenland (Fram Strait) and Barents Seas.</p>
Full article ">Figure 3
<p>Decline in ice sea thickness and multiyear ice (MYI) coverage. (<b>a</b>) Changes in area-averaged basin-wide, multiyear ice and first-year ice thickness in winter between 2003 and 2023 from IS, CS-2, and IS-2. (<b>b</b>) Declines in MYI coverage and September sea-ice extent and increases in first-year ice (FYI) coverage over the same period. Area and thickness computed within the same bounds as in <a href="#remotesensing-16-02983-f002" class="html-fig">Figure 2</a>. The corresponding September ice extent behavior is for comparison.</p>
Full article ">Figure 4
<p>Arctic sea ice volume and ice production since ICESat. (<b>a</b>) Decline in sea ice volume calculated from IS, CS-2 and IS-2 thickness fields. Volume is computed within the same bounds as in <a href="#remotesensing-16-02983-f002" class="html-fig">Figure 2</a>. (<b>b</b>) Increase in ice production between the fall (Oct-Nov) and winter (Feb-Mar) calculated by differencing the winter and fall ice volume. Note that ice volume export is not accounted for here.</p>
Full article ">Figure 5
<p>Seasonal (October-April) evolution of snow depth over the Arctic ice cover from (<b>a</b>) <span class="html-italic">mW99</span> (dashed line) and satellite-derived snow depths (solid line with symbols). (<b>b</b>) Monthly differences between the <span class="html-italic">mW99</span> and satellite-derived snow depths. Their impact on ice thickness and volume can be seen in earlier figures.</p>
Full article ">Figure 6
<p>Interannual changes in mean winter and fall ice thickness (1975–2003), within the data release area, from regression analysis of the submarine record, ICESat, CryoSat-2, and ICESat-2 retrievals. Inset shows the data release area (irregular polygon) of submarine data from U.S. Navy cruises, which covers ~38% of the Arctic Ocean. Sampling of winter and summer are centered on the dates of the ICESat campaigns. Shadings (blue and red) show expected residuals in the regression analysis. Thickness estimates from more localized airborne and ground EM surveys near the North Pole (diamonds) and from Operation IceBridge (circles) are shown within the context of the larger-scale changes in the submarine and satellite records. The corresponding September ice extent behavior is shown as a backdrop.</p>
Full article ">
23 pages, 18685 KiB  
Article
Simulation of Spectral Albedo and Bidirectional Reflectance over Snow-Covered Urban Canyon: Model Development and Factor Analysis
by Qi-Xiang Chen, Zi-Yi Gao, Chun-Lin Huang, Shi-Kui Dong and Kai-Feng Lin
Remote Sens. 2024, 16(13), 2340; https://doi.org/10.3390/rs16132340 - 27 Jun 2024
Viewed by 883
Abstract
A critical comprehension of the impact of snow cover on urban bidirectional reflectance is pivotal for precise assessments of energy budgets, radiative forcing, and urban climate change. This study develops a numerical model that employs the Monte Carlo ray-tracing technique and a snow [...] Read more.
A critical comprehension of the impact of snow cover on urban bidirectional reflectance is pivotal for precise assessments of energy budgets, radiative forcing, and urban climate change. This study develops a numerical model that employs the Monte Carlo ray-tracing technique and a snow anisotropic reflectance model (ART) to simulate spectral albedo and bidirectional reflectance, accounting for urban structure and snow anisotropy. Validation using three flat surfaces and MODIS data (snow-free, fresh snow, and melting snow scenarios) revealed minimal errors: the maximum domain-averaged BRDF bias was 0.01% for flat surfaces, and the overall model-MODIS deviation was less than 0.05. The model’s performance confirmed its accuracy in reproducing the reflectance spectrum. A thorough investigation of key factors affecting bidirectional reflectance in snow-covered urban canyons ensued, with snow coverage found to be the dominant influence. Urban coverage, building height, and soot pollutant concentration significantly impact visible and infrared reflectance, while snow grain size has the greatest effect on shortwave infrared. The bidirectional reflectance at backward scattering angles (0.5–0.6) at 645 nm is lower than forward scattering (around 0.8) in the principal plane as snow grain size increases. These findings contribute to a deeper understanding of snow-covered urban canyons’ reflectance characteristics and facilitate the quantification of radiation interactions, cloud-snow discrimination, and satellite-based retrieval of aerosol and snow parameters. Full article
(This article belongs to the Topic Advances in Earth Observation and Geosciences)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The location of Harbin, situated in northeastern China, and (<b>b</b>) image taken from the EOSDIS Worldview website (<a href="https://worldview.earthdata.nasa.gov/" target="_blank">https://worldview.earthdata.nasa.gov/</a>, accessed 17 June 2024), which showcases a snapshot of the urban structure. Distribution of spectral albedo (<b>c</b>–<b>e</b>) within the study area (126.5°E to 126.8°E, 45.6°N to 45.9°N), with data derived from the MODIS Terra MOD09A1 product on 13 October 2019 (autumn season).</p>
Full article ">Figure 2
<p>The topographical representation of the modeled urban canyons configured with standard parameters.</p>
Full article ">Figure 3
<p>Viewing geometry and plane definitions.</p>
Full article ">Figure 4
<p>The spectral reflectance of snow as simulated by the ART model for MODIS Band 1 (469 nm), with a snow grain size of 3600 µm and a soot pollutant concentration of 500 ppb.</p>
Full article ">Figure 5
<p>Spectral albedo of selected man-made and natural materials.</p>
Full article ">Figure 6
<p>Flowchart of the proposed snow-covered urban BRDF model.</p>
Full article ">Figure 7
<p>Variations of domain-averaged relative error with incident photon numbers.</p>
Full article ">Figure 8
<p>Comparisons of simulated BRDF with truth values over flat Lambertian, RTLSR, and ART surfaces. Solar zenith angle is 45°; Lambertian albedo is 0.2; RTLSR model coefficients are <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>i</mi> <mi>s</mi> <mi>o</mi> </mrow> </msub> <mo>=</mo> <mn>0.091</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>v</mi> <mi>o</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>0.032</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>g</mi> <mi>e</mi> <mi>o</mi> </mrow> </msub> <mo>=</mo> <mn>0.012</mn> </mrow> </semantics></math>; and the ART model has a snow grain size of 500 µm and pollutant concentration of 3600 ppb.</p>
Full article ">Figure 9
<p>Time series of spectral albedos from the MOD09/MYD09 dataset over Harbin between 2018 and 2022.</p>
Full article ">Figure 10
<p>Spectral variations and deviations of MOD, MYD, and modeled albedos under snow-free, fresh snow, and snow melt scenarios.</p>
Full article ">Figure 10 Cont.
<p>Spectral variations and deviations of MOD, MYD, and modeled albedos under snow-free, fresh snow, and snow melt scenarios.</p>
Full article ">Figure 11
<p>Spectral albedo and BRDF variations over urban canyons with building coverage varying from 10% to 50%. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with BC = 50%. (<b>c</b>) BRDF distribution at Band 6 with BC = 50%. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 11 Cont.
<p>Spectral albedo and BRDF variations over urban canyons with building coverage varying from 10% to 50%. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with BC = 50%. (<b>c</b>) BRDF distribution at Band 6 with BC = 50%. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 12
<p>Spectral albedo and BRDF variations over urban canyons with building height varying from 10 m to 90 m. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with BH = 90 m. (<b>c</b>) BRDF distribution at Band 6 with BH = 90 m. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 13
<p>Spectral albedo and BRDF variations over urban canyons with snow coverage varying from 20% to 100%. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with SC = 20%. (<b>c</b>) BRDF distribution at Band 6 with SC = 20%. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 14
<p>Spectral albedo and BRDF variations over urban canyons with snow grain size varying from 100 to 5000 µm. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with SGS = 5000 µm. (<b>c</b>) BRDF distribution at Band 6 with SGS = 5000 µm. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 14 Cont.
<p>Spectral albedo and BRDF variations over urban canyons with snow grain size varying from 100 to 5000 µm. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with SGS = 5000 µm. (<b>c</b>) BRDF distribution at Band 6 with SGS = 5000 µm. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 15
<p>Spectral albedo and BRDF variations over urban canyons with pollutant concentration varying from <math display="inline"><semantics> <msup> <mn>10</mn> <mn>2</mn> </msup> </semantics></math> ppb to <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> ppb. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with PC =<math display="inline"><semantics> <msup> <mn>10</mn> <mn>2</mn> </msup> </semantics></math> ppb. (<b>c</b>) BRDF distribution at Band 6 with PC =<math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> ppb. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 15 Cont.
<p>Spectral albedo and BRDF variations over urban canyons with pollutant concentration varying from <math display="inline"><semantics> <msup> <mn>10</mn> <mn>2</mn> </msup> </semantics></math> ppb to <math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> ppb. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with PC =<math display="inline"><semantics> <msup> <mn>10</mn> <mn>2</mn> </msup> </semantics></math> ppb. (<b>c</b>) BRDF distribution at Band 6 with PC =<math display="inline"><semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics></math> ppb. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 16
<p>Spectral albedo and BRDF variations over urban canyons with solar zenith angle varying from 40°to 80°. The negative VZA in the PP represents the direction of the relative azimuth angle of 180° (forward scattering direction), and the negative VZA in the CP represents the direction of the relative azimuth angle of 270°. (<b>a</b>) Spectral albedo. (<b>b</b>) BRDF distribution at Band 1 with SZA = 80°. (<b>c</b>) BRDF distribution at Band 6 with SZA = 80°. (<b>d</b>) ALM reflectance at Band 1. (<b>e</b>) PP reflectance at Band 1. (<b>f</b>) CP reflectance at Band 1. (<b>g</b>) ALM reflectance at Band 6. (<b>h</b>) PP reflectance at Band 6. (<b>i</b>) CP reflectance at Band 6.</p>
Full article ">Figure 17
<p>Spectral albedo with different settings of building heights and snow coverage (<b>a</b>–<b>c</b>) and of snow grain size and concentration of soot pollution (<b>d</b>–<b>f</b>).</p>
Full article ">
11 pages, 1737 KiB  
Article
The Effect of Foliar Application of Oligogalacturonides on the Functional Value of Turfgrass
by Adam Radkowski, Iwona Radkowska, Michał Kozdęba, Karen Khachatryan, Karol Wolski and Henryk Bujak
Agriculture 2024, 14(3), 369; https://doi.org/10.3390/agriculture14030369 - 25 Feb 2024
Cited by 1 | Viewed by 961
Abstract
Turf grasses play a crucial role in enhancing the beauty and usability of landscapes, gardens, parks, and sports facilities due to their functional and aesthetic properties. However, various unfavourable conditions, such as plant disorders and environmental pressures, can compromise their amenity value. Ongoing [...] Read more.
Turf grasses play a crucial role in enhancing the beauty and usability of landscapes, gardens, parks, and sports facilities due to their functional and aesthetic properties. However, various unfavourable conditions, such as plant disorders and environmental pressures, can compromise their amenity value. Ongoing research aims to identify natural remedies that improve the quality and resilience of these grasses. A study was conducted at the Experimental Station of the Agricultural University of Krakow (50°07′ N, 20°05′ E) to evaluate the practical value of the turf produced by seeding of the ‘Super Lawn’ grass mixture. The experiment involved applying a spray containing oligogalacturonides at two doses: 1.0 and 2.0 dm3∙ha−1, along with a commercial fungicide. The traits were analysed using a 9-point scale. Plants in variant III (treated with the higher dose of oligogalacturonides) and variant IV (treated with the commercial fungicide) exhibited the highest aesthetic and functional values. The application of oligogalacturonides and a commercial fungicide resulted in a decrease in plant diseases. The treatment area showed a reduction in pink snow mould (Microdochium nivale) and leaf spot incidence compared to the control area. Variant II showed enhanced outcomes with the application of 1.0 dm3∙ha−1 of the preparation. In this area, the plant canopy had greater coverage, and the plants demonstrated increased resistance to pink snow mould and leaf spot compared to the plants in the control area. The use of commercial fungicide was found to be more effective than applying oligogalacturonides. Additionally, the plants that were protected with the fungicide displayed the highest values for the analysed parameters. Full article
(This article belongs to the Section Crop Protection, Diseases, Pests and Weeds)
Show Figures

Figure 1

Figure 1
<p>LAI index for options variants I, II, III, and IV. The same letters demonstrate a lack of significant difference between values (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>NDVI index for variants I, II, III, and IV. The same letters demonstrate a lack of significant difference between values (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>SPAD index for variants I, II, III, and IV. The same letters demonstrate a lack of significant difference between values (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 27731 KiB  
Article
Yearly Elevation Change and Surface Velocity Revealed from Two UAV Surveys at Baishui River Glacier No. 1, Yulong Snow Mountain
by Leiyu Li, Yuande Yang, Shijin Wang, Chuya Wang, Qihua Wang, Yuqiao Chen, Junhao Wang, Songtao Ai and Yanjun Che
Atmosphere 2024, 15(2), 231; https://doi.org/10.3390/atmos15020231 - 14 Feb 2024
Cited by 2 | Viewed by 1259
Abstract
Glaciers play an important role in understanding the climate, water resources, and surrounding natural change. Baishui River Glacier No. 1, a temperate glacier in the monsoon-influenced Southeastern Qinghai–Tibet Plateau, has experienced significant ablation due to regional warming during the past few decades. However, [...] Read more.
Glaciers play an important role in understanding the climate, water resources, and surrounding natural change. Baishui River Glacier No. 1, a temperate glacier in the monsoon-influenced Southeastern Qinghai–Tibet Plateau, has experienced significant ablation due to regional warming during the past few decades. However, little is known about the yearly changes in Baishui River Glacier No. 1. To investigate how Baishui River Glacier No. 1 has changed in recent years, digital orthophoto maps and digital elevation models were obtained from an unmanned aerial vehicle on 20 October 2018 and 22 July 2021, covering 84% and 47% of the total area, respectively. The results of the Baishui River Glacier No. 1 changes were obtained by differencing the digital elevation models, manual tracking, and terminus-retreat calculation methods. Our results showed that the surveyed area had a mean elevation change of −4.26 m during 2018 and 2021, and the lower area lost more ice than other areas. The terminus of Baishui River Glacier No. 1 has retreated by 16.35 m/a on average, exhibiting spatial variation with latitude. Moreover, we initially found that there was a high correlation between surface velocity and elevation gradient in this high-speed glacier. The surface velocity of Baishui River Glacier No. 1 was derived with the manual feature tracking method and ranged from 10.48 to 32.00 m/a, which is slightly smaller than the seasonal average. However, the snow coverage and ice melting of the two epochs led to the underestimation of our elevation change and velocity results, which need further investigation. Full article
(This article belongs to the Section Climatology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Yulong Snow Mountain is located in the Southeastern Qinghai–Tibetan Plateau and the southern end of the Hengduan Mountains. (<b>b</b>) The location of Baishui River Glacier No. 1 on Yulong Snow Mountain. (<b>c</b>) Baishui River Glacier No. 1, the areas of different unmanned aerial vehicle surveys (the 2018 survey is shown as a red rectangle, and the 2021 survey is shown as a yellow rectangle), and the position where unmanned aerial vehicles were launched.</p>
Full article ">Figure 2
<p>(<b>a</b>) Terminus of Baishui River Glacier No. 1 in 2018 and 2021; (<b>b</b>) elevation changes in Baishui River Glacier No. 1 from 20 October 2018 to 22 July 2021—Area1 is enclosed by the blue line, and Area2 is enclosed by the black line; (<b>c</b>) elevation change in Area1; (<b>d</b>) 8 selected feature points from glacier terminus.</p>
Full article ">Figure 3
<p>Digital orthophoto maps acquired in (<b>a</b>) 2018 and (<b>b</b>) 2021.</p>
Full article ">Figure 4
<p>(<b>a</b>) Gradient of study area; (<b>b</b>) relationship between surface velocity and gradient; (<b>c</b>) derived surface velocity; (<b>d</b>) the surface velocity along the mainstream line.</p>
Full article ">Figure 5
<p>(<b>a</b>) The precipitation and temperature data collected between 22 July and 20 October 2018 from the meteorological station. (<b>b</b>) Surface elevation data between 22 July and 20 October 2021 from a real-time platform.</p>
Full article ">Figure 6
<p>(<b>a</b>) Relationship between elevation and elevation change; (<b>b</b>) elevation change distribution along the elevation in different regions.</p>
Full article ">Figure A1
<p>Glacier surface velocity extracted by ImGRAFT.</p>
Full article ">Figure A2
<p>(<b>a</b>) The 151 homologous points selected; (<b>b</b>) relationship between velocity and elevation gradient.</p>
Full article ">Figure A3
<p>(<b>a</b>) Relationship between velocity and elevation gradient of 40 points selected from <a href="#atmosphere-15-00231-f0A2" class="html-fig">Figure A2</a>; (<b>b</b>) the selected homologous points: red points represent 2021, and blue points represent 2018.</p>
Full article ">
17 pages, 4036 KiB  
Article
Ecological and Public Advantages of a Dual Flagship Strategy: Giant Panda and Snow Leopard
by Ying Yue, Yihong Wang, Ziyi Ye, Chengcheng Zhang, Lan Qiu, Qiang Xu, Xin He, Chendi Ma, Biao Yang, Zhisong Yang and Qiang Dai
Diversity 2024, 16(2), 76; https://doi.org/10.3390/d16020076 - 25 Jan 2024
Viewed by 1920
Abstract
Flagship species’ conservation strategies hold significant prominence in biodiversity preservation. The giant panda, a globally recognized species, has drawn attention to its benefits and constraints as a flagship species. This study aimed to assess the potential benefits of a dual flagship strategy using [...] Read more.
Flagship species’ conservation strategies hold significant prominence in biodiversity preservation. The giant panda, a globally recognized species, has drawn attention to its benefits and constraints as a flagship species. This study aimed to assess the potential benefits of a dual flagship strategy using both the giant panda and snow leopard, compared to an approach solely using the giant panda. We identified the number of potential beneficiary species based on their habitat overlap with the giant panda and snow leopard in Sichuan and Gansu, China. Subsequently, we examined public preferences for these two flagships and their influencing factors through questionnaire surveys within and outside China. The dual flagship strategy covered the habitats of more species and amplified existing protection for those species already benefiting from giant panda conservation efforts. The giant panda was commonly perceived as “Adorable”, “Innocent”, and “Rare”, while perceptions of the snow leopard leaned towards “Mighty”, “Mysterious”, and “Rare”. Though the giant panda is widely favored, the survey indicates a notable preference for snow leopards among a proportion of respondents. The dual flagship strategy offers expanded wildlife habitat coverage and benefits a broader range of species. Moreover, the combined appeal of the snow leopard and giant panda, each possessing unique charm and symbolism, holds the potential to garner broader societal interest and support. This study may serve as a reference for policy decisions in the Giant Panda National Park and other similar protected areas, optimizing conservation management and outreach initiatives for flagship species strategies. It may also benefit conservation strategies centered on other flagship species. Full article
(This article belongs to the Special Issue Ecology, Conservation and Restoration of Threatened Animal)
Show Figures

Figure 1

Figure 1
<p>Study area.</p>
Full article ">Figure 2
<p>Immediate impressions from respondents to the photos for the giant panda (<b>A</b>) and the snow leopard (<b>B</b>) in an independent pilot survey.</p>
Full article ">Figure 3
<p>The coverage of potential beneficiary species for the giant panda and snow leopard across varying <span class="html-italic">P</span><sub>1</sub> and <span class="html-italic">P</span><sub>2</sub> threshold levels. Panel (<b>A</b>) shows the count of potential beneficiary species. Red indicates the number of species effectively covered only by the giant panda’s habitat, blue represents those covered only by the snow leopard’s habitat, and yellow are species effectively covered by both habitats. (<b>B</b>) is the count of potential beneficiary species with larger habitat coverage by the snow leopard than the giant panda.</p>
Full article ">Figure 4
<p>Respondents’ level of affection (<b>A</b>–<b>C</b>) and self-perceived familiarity (<b>D</b>–<b>F</b>) towards the giant panda and snow leopard. Panels (<b>A</b>,<b>D</b>) represent the giant panda; (<b>B</b>,<b>E</b>) denote the snow leopard; (<b>C</b>,<b>F</b>) illustrate the difference in affection and familiarity between the two species. Blue bars indicate respondents from within China, while orange bars are those outside China. Differences in affection and familiarity levels for the two species are derived by subtracting the Likert scale values for the giant panda from those for the snow leopard. Species illustrations were generated by DALL-E.</p>
Full article ">Figure 5
<p>Influence of affection levels toward the giant panda (<b>A</b>,<b>B</b>) and snow leopard (<b>C</b>,<b>D</b>) on the species preference choice. Panels (<b>A</b>,<b>C</b>) show the effects of affection levels among Chinese respondents, while panels (<b>B</b>,<b>D</b>) display the same for respondents outside China. Orange indicates preferring the giant panda, blue for the snow leopard, and gray denotes opting for neither. Stacked areas show respondent counts, and lines represent percentage distributions of preferences.</p>
Full article ">Figure 6
<p>Perceptions of the giant panda (<b>A</b>) and snow leopard (<b>B</b>) among respondents. Blue represents responses from China, while orange represents those outside China. Bars indicate the counts of respondents choosing each impression, while the lines with circles present the weighted percentage of selections for each impression.</p>
Full article ">Figure 7
<p>Chord diagrams show the associations between respondents’ perceptions of the species and their preferred photo. The upper half of the diagram indicates the weighted number of respondents selecting each impression, while the lower half illustrates the photo they chose. “SI” represents the perception of the giant panda or the snow leopard as “Silly”, “IN” for Innocent, “AD” for Adorable, “MY” for Mysterious, “MI” for Mighty, and “RA” for Rare. “MI P” corresponds to the Mighty Photo of the giant panda or the snow leopard, “AD P” to the Adorable Photo, and “MY P” to the Mysterious Photo.</p>
Full article ">
16 pages, 11157 KiB  
Article
Distance to a River Modifies Climate Legacy on Vegetation Growth in a Boreal Riparian Forest
by Yingyu Li, Qiaoqi Sun, Hongfei Zou and Petra Marschner
Remote Sens. 2023, 15(23), 5582; https://doi.org/10.3390/rs15235582 - 30 Nov 2023
Viewed by 1035
Abstract
Inter-annual variability in growing season temperature and precipitation, together with snow coverage duration, determine vegetation growth in boreal ecosystems. However, little is known about the impact of concurrent and antecedent climate, particularly snow cover duration, on vegetation growth in a boreal riparian forest. [...] Read more.
Inter-annual variability in growing season temperature and precipitation, together with snow coverage duration, determine vegetation growth in boreal ecosystems. However, little is known about the impact of concurrent and antecedent climate, particularly snow cover duration, on vegetation growth in a boreal riparian forest. Additionally, significant uncertainty exists regarding whether the distance to a river (as a proxy of groundwater availability) further modifies these climatic legacy effects on vegetation growth. To fill this knowledge gap, we quantified the responses of different vegetation types (shrub, deciduous coniferous and broadleaf forests) to concurrent and antecedent climate variables in a boreal riparian forest, and further determined the magnitude and duration of climate legacies in relation to distance to a river, using MODIS-derived NDVI time series with gridded climate data from 2001 to 2020. Results showed that higher temperature and precipitation and longer snow cover duration increased vegetation growth. For deciduous coniferous forests and broadleaf forests, the duration of temperature legacy was about one year, precipitation legacy about two years and snow cover duration legacy was 3 to 4 years. Further, distance to a river modified the concurrent and antecedent temperature and snow cover duration legacy effects on vegetation growth, but not that of precipitation. Specifically, temperature and snow cover duration legacies were shorter at the sites near a river compared to sites at greater distance to a river. Our research highlights the importance of snow cover duration on vegetation growth and that closeness to a river can buffer adverse climate impacts by shortening the strength and duration of climate legacies in a boreal riparian forest. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Nanweng River National Nature Reserve and vegetation types (inset). Three types of vegetation (i.e., shrub, deciduous coniferous forests and deciduous broadleaf forests) obtained from the 2019 global land cover map (CGLS-LC100 Collection 3, with a spatial resolution of 100 m). (<b>b</b>) Photo of shrub at study area. (<b>c</b>) Photo of dominant species <span class="html-italic">Larix gmelinii</span> in the de-ciduous coniferous forest. (<b>d</b>) Photo of dominant species <span class="html-italic">Betula platyphylla</span> in the deciduous broadleaf forest.</p>
Full article ">Figure 2
<p>Flowchart for analyzing climate legacy on vegetation growth used in this study.</p>
Full article ">Figure 3
<p>Posterior means values (points) and 95% credible intervals (CIs; horizontal lines) for antecedent climatic variables (i.e., temperature, precipitation and snow cover duration) and their interaction (temperature × precipitation) for three vegetation types (different colors represent different vegetation types, with red for shrub, green for deciduous coniferous forest (DCF), and blue for deciduous broadleaf forest (DBF)). The vertical dotted line indicates the zero line, and 95% CIs that overlap the zero line denote parameters that are not statistically different from zero.</p>
Full article ">Figure 4
<p>Cumulative importance of antecedent temperature (<b>a</b>), precipitation (<b>b</b>) and snow cover duration (<b>c</b>) on different vegetation types. Equivalent to ‘cumulative probabilities’, climate effects experienced over the concurrent year and past five years (<span class="html-italic">x</span>-axis) account for cumulative importance (<span class="html-italic">y</span>-axis) of the climate covariate to NDVI. The posterior mean cumulative importance is shown for each antecedent climatic variable, with different colored lines representing different vegetation types. Only vegetation types responding to climate variables in <a href="#remotesensing-15-05582-f003" class="html-fig">Figure 3</a> are included. The dashed line at a cumulative importance of 0.5 indicates the threshold for the critical lag period. The time when the cumulative importance crosses this line is considered the timescale of the memory effect for each vegetation type and for the climatic driver. DCF and DBF are deciduous coniferous forest and deciduous broadleaf forest, respectively.</p>
Full article ">Figure 5
<p>Posterior mean coefficient values (and 95% credible intervals) for antecedent climate effects (e.g., temperature, precipitation, snow cover days) and their interactions (temperature × precipitation) for different vegetation types by distance to a river (close represents close to a river (&lt;1 km); far represents 2–3 km away from a river). The vertical dashed line at estimate = 0 is depicted (i.e., no significant effect).</p>
Full article ">Figure 6
<p>The cumulative importance for vegetation types at different distances from rivers of antecedent climate effects (temperature (<b>a</b>), precipitation (<b>b</b>) and snow cover days (<b>c</b>)). Only vegetation types responding to climate variables in <a href="#remotesensing-15-05582-f005" class="html-fig">Figure 5</a> are included. The timescale (legacy ‘length’) is based on when the cumulative importance weights reach 50% (dashed horizontal line).</p>
Full article ">
19 pages, 5655 KiB  
Article
Implications of Accuracy of Global Glacier Inventories in Hydrological Modeling: A Case Study of the Western Himalayan Mountain Range
by Haleema Attaullah, Asif Khan, Mujahid Khan, Hadia Atta and Muhammad Shahid Iqbal
Water 2023, 15(22), 3887; https://doi.org/10.3390/w15223887 - 8 Nov 2023
Viewed by 1434
Abstract
Alpine glaciers are a fundamental component of the cryosphere and are significantly sensitive to climate change. One such region is the Hindukush Karakoram Himalaya (HKH) and Tibetan Plateau (TP) region, which contains more than 40,000 glaciers. There are more than 12 glacier inventories [...] Read more.
Alpine glaciers are a fundamental component of the cryosphere and are significantly sensitive to climate change. One such region is the Hindukush Karakoram Himalaya (HKH) and Tibetan Plateau (TP) region, which contains more than 40,000 glaciers. There are more than 12 glacier inventories available covering parts of (or the entire) HKH region, but these show significant uncertainties regarding the extent of glaciers. Researchers have used different glacier inventories without assessing their accuracy. This study, therefore, assessed the implications of the accuracy of global glacier inventories in hydrological modeling and future water resource planning. The accuracy assessment of most commonly used two global glacier inventories (Global Land Ice Monitoring from Space-GLIMS v 2.0 and Randolph Glacier Inventory-RGI v 6.0) has been carried out for three sub-basins of the Upper Indus Basin—the Swat, the Chitral, and the Kabul River basins (combined, this is referred to as the Great Kabul River Basin)—with a total basin area of 94,552.86 km2. Glacier outlines have been compared with various Landsat 7 ETM+, Landsat 8, high-resolution Google Earth images, and manually digitized debris-covered glacier outlines during different years. The total glacier area for the Great Kabul River Basin derived from RGI and GLIMS is estimated to be 2120.35 km2 and 1789.94 km2, respectively, which was a difference of 16.9%. Despite being sub-basins of the Great Kabul River Basin, the Swat, and the Chitral River basins were different by 54.74% and 19.71%, respectively, between the two inventories, with a greater glacierized area provided by RGI, whereas the Kabul River basin was different by 54.72%, with greater glacierized area provided by GLIMS. The results and analysis show that GLIMS underestimates glacier outlines in the Swat and the Chitral basins and overestimates glacier extents in the Kabul River basin. The underestimation is mainly due to the non-representation of debris-covered glaciers. The overestimation in GLIMS data is due to the digitization of seasonal snow as part of the glaciers. The use of underestimated GLIMS outlines may result in 5–10% underestimation of glacier-melt contribution to flows in the Swat River basin, while an underestimation of 7% to 15% is expected in the Chitral River Basin, all compared to RGI v 6.0 outlines. The overestimation of glacier-melt contribution to flows in the Kabul River basin is insignificant (1% to 2%) using GLIMS data. In summary, the use of the GLIMS inventory will lead to underestimated flows and show that the Great Kabul River Basin (particularly the Chitral River Basin) is less sensitive to climate change effects. Thus, the current study recommends the use of RGI v 6.0 (best glacier inventory) to revisit the existing biased hydro-climate studies and to improve future hydro-climate studies with the concomitant rectification of the MODIS snow coverage data. The use of the best glacier inventory will provide the best estimates of flow sensitivity to climate change and will result in well-informed decision-making, precise and accurate policies, and sustainable water resource management in the study area. The methodology adopted in the current study may also be used in nearby areas with similar hydro-climate conditions, as well as for the most recently released RGI v 7.0 data. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>The study area is shown as a transboundary basin between Pakistan and Afghanistan. The lower figure shows enlarged basins in various colors. The Great Kabul River Basin composed of three sub-basins, namely the Swat River basin (purple), the Chitral River Basin (green), and the Kabul River basin (red). The junction point of the Himalayas–Karakoram and Hindukush is also demarcated.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>d</b>) The Landsat images for the dates 22-09-2002, 28-09-2013, 15-09-2014, and 18-09-2015 showing compound valley-based large glaciers with debris-covered ice at the terminus. The presence of debris-covered ice has also been confirmed from the Google Earth image in <a href="#water-15-03887-f003" class="html-fig">Figure 3</a>, showing the same glaciers.</p>
Full article ">Figure 3
<p>Google Earth image of 08/30/2010. The inset shows the Kabul River basin boundary along with the area of interest. Red represents the GLIMS outlines, and yellow represents the RGI outlines. The image clearly shows that RGI has underestimated the glacier areas by ignoring debris-covered ice.</p>
Full article ">Figure 4
<p>The locations of glaciers in the basin are shown in (<b>a</b>). Comparison of Landsat images for the dates 22-09-2002 and 15-09-2014 (<b>d</b>,<b>e</b>) with Google Earth images (<b>b</b>,<b>c</b>) of 08/30/2010 providing evidence that RGI has underestimated the glaciers (see within blue rectangles) by not considering debris-covered ice. However, GLIMS is showing consistency with Google Earth images for these medium glaciers.</p>
Full article ">Figure 5
<p>Google Earth image of 08/30/2010 showing little difference between the two glacier inventories for small glaciers.</p>
Full article ">Figure 6
<p>Comparison of a Landsat image from 15-09-2014 with a Google Earth image in <a href="#water-15-03887-f007" class="html-fig">Figure 7</a> showing small glaciers (within the blue polygon) along the Badakhshan and Nuristan boundary near the Kohe–Khrebek mountain range. The inset shows the Kabul River basin boundary and area of interest.</p>
Full article ">Figure 7
<p>The Landsat image for 28-09-2003 (<b>a</b>) and Google Earth image (<b>b</b>) show small glaciers covered with debris in the region of Mir Samir (Afghanistan). The area of interest is shown in (<b>c</b>).</p>
Full article ">Figure 8
<p>Landsat images obtained for 22-09-2002 (<b>a</b>), 28-09-2013 (<b>b</b>), 15-09-2014 (<b>c</b>), and 07-09-2017 (<b>d</b>) showing very small and very very small glaciers across the Kabul River basin. Seasonal snow patches can be observed in the image of 15-09-2014 (<b>c</b>).</p>
Full article ">Figure 9
<p>Landsat images obtained for 22-09-2002 (<b>a</b>), 28-09-2013 (<b>b</b>), 15-09-2014 (<b>c</b>), and 07-09-2017 (<b>d</b>) confirming that GLIMS has digitized the glacier boundaries for most of the very small and very very small glaciers, which are neither seasonal snow nor debris-covered ice.</p>
Full article ">Figure 10
<p>Glacier-melt contribution from the Kabul River basin to Kabul River flow for ablation rates 0.5 m/year, 0.75 m/year, and 1 m/year for 100 m intervals using GLIMS, RGI, and Best Glacier Outlines. The Best Glacier Outlines for the Kabul River basin (GLIMS for &gt;1 km<sup>2</sup> glacier sizes and RGI for &lt;1 km<sup>2</sup> glacier sizes) are shown in gray.</p>
Full article ">
24 pages, 6098 KiB  
Article
Reconstruction of Snow Cover in Kaidu River Basin via Snow Grain Size Gap-Filling Based on Machine Learning
by Linglong Zhu, Guangyi Ma, Yonghong Zhang, Jiangeng Wang and Xi Kan
Water 2023, 15(21), 3726; https://doi.org/10.3390/w15213726 - 25 Oct 2023
Viewed by 1460
Abstract
Fine spatiotemporal resolution snow monitoring at the watershed scale is crucial for the management of snow water resources. This research proposes a cloud removal algorithm via snow grain size (SGS) gap-filling based on a space–time extra tree, which aims to address the issue [...] Read more.
Fine spatiotemporal resolution snow monitoring at the watershed scale is crucial for the management of snow water resources. This research proposes a cloud removal algorithm via snow grain size (SGS) gap-filling based on a space–time extra tree, which aims to address the issue of cloud occlusion that limits the coverage and time resolution of long-time series snow products. To fully characterize the geomorphic characteristics and snow duration time of the Kaidu River Basin (KRB), we designed dimensional data that incorporate spatiotemporal information. Combining other geographic and snow phenological information as input for estimating SGS. A spatiotemporal extreme tree model was constructed and trained to simulate the nonlinear mapping relationship between multidimensional inputs and SGS. The estimation results of SGS can characterize the snow cover under clouds. This study found that when the cloud cover is less than 70%, the model’s estimation of SGS meets expectations, and snow cover reconstruction achieves good results. In specific cloud removal cases, compared to traditional spatiotemporal filtering and multi-sensor fusion, the proposed method has better detail characterization ability and exhibits better performance in snow cover reconstruction and cloud removal in complex mountainous environments. Overall, from 2000 to 2020, 66.75% of snow products successfully removed cloud coverage. This resulted in a decrease in the annual average cloud coverage rate from 52.46% to 34.41% when compared with the MOD10A1 snow product. Additionally, there was an increase in snow coverage rate from 21.52% to 33.84%. This improvement in cloud removal greatly enhanced the time resolution of snow cover data without compromising the accuracy of snow identification. Full article
(This article belongs to the Special Issue Cold Regions Ice/Snow Actions in Hydrology, Ecology and Engineering)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The monthly mean precipitation and temperature at the Bayanbulak meteorological station from January 2000 to December 2019; (<b>b</b>) Bayanbulak meteorological station time series of daily snow depth from January 2000 to December 2019, with a monthly histogram of mean snow depth from January 2000 to December 2019 in the inset figure; (<b>c</b>) geographical location and topographic relief of the KRB. The different background colors indicate different elevations. The location of the Bayanbulak meteorological station and the river system are also shown.</p>
Full article ">Figure 2
<p>Slope (<b>a</b>), aspect (<b>b</b>), and land cover (<b>c</b>) in the KRB.</p>
Full article ">Figure 3
<p>Flow chart of space–time extra tree model.</p>
Full article ">Figure 4
<p>Construction of the spatial distance factor; the weighted sum of the four green segments is <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mn>1</mn> </msub> </mrow> </semantics></math>, and the weighted sum of the three pink segments is <math display="inline"><semantics> <mrow> <msub> <mi>D</mi> <mn>2</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Construction of snow duration index. White indicates snow-free at this time point, blue indicates snow-covered at this time point, and the constructed index values are shown in square brackets.</p>
Full article ">Figure 6
<p>Loss variation of the model training under cloud coverage change (i.e., Data Loss Rate).</p>
Full article ">Figure 7
<p>Importance scores of input factors of the space–time extra tree model.</p>
Full article ">Figure 8
<p>Comparison of cloud removal effect based on SGS filling, (<b>a</b>,<b>f</b>,<b>k</b>): Landsat true color map; (<b>b</b>,<b>g</b>,<b>l</b>): Landsat snow cover map; (<b>c</b>,<b>h</b>,<b>m</b>): MODIS snow cover product; (<b>d</b>,<b>i</b>,<b>n</b>): snow cover by SGS filling; and (<b>e</b>,<b>j</b>,<b>o</b>): snow map by Hao.</p>
Full article ">Figure 9
<p>Comparison of snow cover days between MODIS snow products and snow cover data based on SGS filling.</p>
Full article ">Figure 10
<p>Spatial distribution of differences in snow cover days among different hydrological years.</p>
Full article ">Figure 11
<p>Variation in the proportion of weekly average cloud coverage and snow coverage before and after cloud removal in the KRB.</p>
Full article ">Figure 12
<p>Validation results: (<b>a</b>) original snow cover for 8 October 2015, (<b>b</b>) original snow cover for 9 October 2015, (<b>c</b>) snow cover data from 8 October 2015 filled with cloud values of 9 October 2015 snow cover data, and (<b>d</b>) snow cover data from (<b>c</b>) after cloud removal; (<b>e</b>) snow cover data from (<b>c</b>) by Hao’s method; (<b>f</b>) original snow cover for 20 April 2016, (<b>g</b>) original snow cover for 21 April 2016, (<b>h</b>) snow cover data from 20 April 2016 filled with cloud values of 21 April 2016 snow cover data, (<b>i</b>) snow cover data from (<b>h</b>) after cloud removal, and (<b>j</b>) snow cover data from (<b>h</b>) by Hao’s method; (<b>k</b>) original snow cover for 25 April 2018, (<b>l</b>) original snow cover for 24 April 2018, (<b>m</b>) snow cover data from 25 April 2018 filled with cloud values of 24 April 2018 snow cover data, (<b>n</b>) snow cover data from (<b>m</b>) data after cloud removal, and (<b>o</b>) snow cover data from (<b>h</b>) by Hao’s method; and (<b>p</b>) original snow cover for 1 June 2018, (<b>q</b>) original snow cover for 2 June 2018, (<b>r</b>) snow cover data from 1 June 2018 filled with cloud values of 2 June 2018 snow cover data, (<b>s</b>) snow cover data from (<b>r</b>) data after cloud removal, and (<b>t</b>) snow cover data from (<b>h</b>) by Hao’s method.</p>
Full article ">
17 pages, 3201 KiB  
Technical Note
Evaluating the Effects of UAS Flight Speed on Lidar Snow Depth Estimation in a Heterogeneous Landscape
by Franklin B. Sullivan, Adam G. Hunsaker, Michael W. Palace and Jennifer M. Jacobs
Remote Sens. 2023, 15(21), 5091; https://doi.org/10.3390/rs15215091 - 24 Oct 2023
Viewed by 1107
Abstract
Recently, sensors deployed on unpiloted aerial systems (UAS) have provided snow depth estimates with high spatial resolution over watershed scales. While light detection and ranging (LiDAR) produces precise snow depth estimates for areas without vegetation cover, there has generally been poorer precision in [...] Read more.
Recently, sensors deployed on unpiloted aerial systems (UAS) have provided snow depth estimates with high spatial resolution over watershed scales. While light detection and ranging (LiDAR) produces precise snow depth estimates for areas without vegetation cover, there has generally been poorer precision in forested areas. At a constant flight speed, the poorest precision within forests is observed beneath tree canopies that retain foliage into or through winter. The precision of lidar-derived elevation products is improved by increasing the sample size of ground returns but doing so reduces the spatial coverage of a mission due to limitations of battery power. We address the influence of flight speed on ground return density for baseline and snow-covered conditions and the subsequent effect on precision of snow depth estimates across a mixed landscape, while evaluating trade-offs between precision and bias. Prior to and following a snow event in December 2020, UAS flights were conducted at four different flight speeds over a region consisting of three contrasting land types: (1) open field, (2) deciduous forest, (3) conifer forest. For all cover types, we observed significant improvements in precision as flight speeds were reduced to 2 m s−1, as well as increases in the area over which a 2 cm snow depth precision was achieved. On the other hand, snow depth estimate differences were minimized at baseline flight speeds of 2 m s−1 and 4 m s−1 and snow-on flight speeds of 6 m s−1 over open fields and between 2 and 4 m s−1 over forest areas. Here, with consideration to precision and estimate bias within each cover type, we make recommendations for ideal flight speeds based on survey ground conditions and vegetation cover. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere II)
Show Figures

Figure 1

Figure 1
<p>An optical orthomosaic of the study area (<b>a</b>) was used to produce a Green Leaf Index map (<b>b</b>) to derive vegetation classification. Any Green Leaf Index values &lt; 0.2 were classified as deciduous canopy for our analysis, except for the manually classified field cells.</p>
Full article ">Figure 2
<p>Violin plot of 150% of the interquantile range of confidence intervals for each cover type.</p>
Full article ">Figure 3
<p>Maps of the study area colored by the maximum snow-on flight speed per pixel required to produce 2 cm estimate precision at baseline speeds of (<b>a</b>) 2 m s<sup>−1</sup>, (<b>b</b>) 4 m s<sup>−1</sup>, (<b>c</b>) 6 m s<sup>−1</sup>, and (<b>d</b>) 8 m s<sup>−1</sup>. The areas in white did not produce a 2 cm estimate precision at any flight speed.</p>
Full article ">Figure 4
<p>Box plots of the differences between lidar-derived snow depth and in situ snow depth for all flight speed combinations by field (F), deciduous (D), and coniferous (C) vegetation cover types. The box is framed by lower and upper quartile values with its center line indicating the median value. The box height is the interquartile range (IQR). The lower whisker extends to 1.5 times the IQR below the lower quartile or the minimum value in the dataset. The upper whisker extends to 1.5 times the IQR above the upper quartile or the maximum value in the dataset. Outliers are also shown.</p>
Full article ">Figure 5
<p>Examples of areas of overlap metrics between baseline and snow-off flights by land use. Each 1 × 1 m grid cell was divided into nine subplots. The total area of overlap was calculated for the entire grid cell. For consistency, colors used here are for snow-on flight speeds of 2 m s<sup>−1</sup> (red), 4 m s<sup>−1</sup> (blue), 6 m s<sup>−1</sup> (light blue), and 8 m s<sup>−1</sup> (gold).</p>
Full article ">Figure 6
<p>Snow depth difference between in situ and lidar observations vs. percent overlap for conifer and deciduous study plots for all baseline flight speeds, colored by snow-on flight speeds of 2 m s<sup>−1</sup> (red), 4 m s<sup>−1</sup> (blue), 6 m s<sup>−1</sup> (light blue), and 8 m s<sup>−1</sup> (gold). In each figure, there are four points for each individual grid cell corresponding to the overlap and bias determined for each of the four snow-on flight speeds.</p>
Full article ">Figure 7
<p>Scatter plot of lidar-derived and in situ snow depth estimates with 95% confidence intervals shown for each measurement, using flight speed combinations that minimized the mean differences between methods, with a model forced to zero intercept (solid black line) and a model with a free intercept (dashed line).</p>
Full article ">
19 pages, 20699 KiB  
Article
Applying High-Resolution Satellite and UAS Imagery for Detecting Coldwater Inputs in Temperate Streams of the Iowa Driftless Region
by Niti B. Mishra, Michael J. Siepker and Greg Simmons
Remote Sens. 2023, 15(18), 4445; https://doi.org/10.3390/rs15184445 - 9 Sep 2023
Viewed by 1868
Abstract
Coldwater streams are crucial habitats for many biota including Salmonidae and Cottidae species that are unable to tolerate warmer water temperatures. Accurate classification of coldwater streams is essential for their conservation, restoration, and management, especially in light of increasing human disturbance and climate [...] Read more.
Coldwater streams are crucial habitats for many biota including Salmonidae and Cottidae species that are unable to tolerate warmer water temperatures. Accurate classification of coldwater streams is essential for their conservation, restoration, and management, especially in light of increasing human disturbance and climate change. Coldwater streams receive cooler groundwater inputs and, as a result, typically remain ice-free during the winter. Based on this empirical thermal evidence, we examined the potential of very high-resolution (VHR) satellite and uncrewed aerial system (UAS) imagery to (i) detect coldwater streams using semi-automatic classification versus visual interpretation approaches, (ii) examine the physical factors that contribute to inaccuracies in detecting coldwater habitats, and (iii) use the results to identify inaccuracies in existing thermal stream classification datasets and recommend coverage updates. Due to complex site conditions, semi-automated classification was time consuming and produced low mapping accuracy, while visual interpretation produced better results. VHR imagery detected only the highest quality coldwater streams while lower quality streams that still met the thermal and biological criteria to be classified as coldwater remained undetected. Complex stream and site variables (narrow stream width, canopy cover, terrain shadow, stream covered by ice and drifting snow), image quality (spatial resolution, solar elevation angle), and environmental conditions (ambient temperature prior to image acquisition) make coldwater detection challenging; however, UAS imagery is uniquely suited for mapping very narrow streams and can bridge the gap between field data and satellite imagery. Field-collected water temperatures and stream habitat and fish community inventories may be necessary to overcome these challenges and allow validation of remote sensing results. We detected >30 km of coldwater streams that are currently misclassified as warmwater. Overall, visual interpretation of VHR imagery it is a relatively quick and inexpensive approach to detect the location and extent of coldwater stream resources and could be used to develop field monitoring programs to confirm location and extent of coldwater aquatic resources. Full article
Show Figures

Figure 1

Figure 1
<p>Satellite imagery of a northeast Iowa stream taken during spring (<b>a</b>)<b>,</b> winter (<b>b</b>), and summer (<b>c</b>) showing the location of groundwater input into the stream. Regardless of thermal condition, both stream segments appear identical in the spring and summer imagery (<b>a</b>,<b>c</b>). The winter imagery (<b>b</b>) shows the segment of stream influenced by groundwater. The stream segment downstream of the groundwater input remains open while the segment upstream of the groundwater input is ice- and snow-covered. (Image copyright 2023 Maxar).</p>
Full article ">Figure 2
<p>(<b>a</b>) Location of the Canoe Creek watershed within the Driftless Region of Iowa; (<b>b</b>) streams in the study watershed and locations where continuous water temperature data (black dot) was available. The watershed is shown using hillshade coverage. Site IDs (A–W) for each water temperature monitoring location are also shown.</p>
Full article ">Figure 3
<p>(<b>a</b>) Panchromatic imagery cropped to the 200 m buffer generated using the stream polygon vector (shown in <a href="#remotesensing-15-04445-f001" class="html-fig">Figure 1</a>); the red polygon indicates the area where imagery from only 2014 was available; fish community survey locations are shown as green dots. (<b>b</b>,<b>c</b>) Comparison of winter panchromatic imagery and summer false color multispectral imagery for Site 10 (<b>d</b>) and Site 3 (<b>e</b>). Site IDs (1–21) of each fish community surveys are also shown.</p>
Full article ">Figure 4
<p>Illustration of the workflow used to develop and interpret the winter imagery for locating coldwater streams in the Canoe Creek watershed.</p>
Full article ">Figure 5
<p>For Site 1: (<b>a</b>) multispectral imagery (2 m pixel size), (<b>b</b>) panchromatic imagery (0.5 m pixel size), (<b>c</b>) UAS-acquired imagery (0.03 m pixel size), and (<b>d</b>) oblique field photo.</p>
Full article ">Figure 6
<p>Results of GEOBIA for two selected stream reaches: (<b>a</b>) shows an upstream reach with a very narrow stream, (<b>b</b>) GEOBIA classification result and (<b>c</b>) classification accuracy for the upstream reach, (<b>d</b>) represents a much wider downstream reach, (<b>e</b>) shows the GEOBIA classification result and (<b>f</b>) shows the classification accuracy for this area.</p>
Full article ">Figure 7
<p>(<b>a</b>) Coldwater stream reaches identified by visual interpretation of VHR winter imagery combined with fish community and stream temperature data in this study; (<b>b</b>) coldwater stream reaches currently classified by IA-DNR as coldwater; (<b>c</b>) stream reaches that are not currently classified as coldwater but were identified as coldwater in our analysis; and (<b>d</b>) stream reaches currently classified as coldwater by IA-DNR but which were not identified as coldwater in our analysis.</p>
Full article ">Figure 8
<p>(<b>a</b>) Multispectral imagery (2 m pixel), (<b>b</b>) panchromatic imagery (0.5 m pixel), (<b>c</b>) UAS-acquired imagery (0.03 m pixel size), and (<b>d</b>) oblique field photo for Site 2.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Dense canopy cover limiting coldwater stream identification from winter imagery; (<b>c</b>) summer imagery limiting stream detection due to dense overhead canopy; and (<b>d</b>) oblique field photo of Site 4 where trout were sampled, supporting classification as a coldwater stream.</p>
Full article ">Figure 10
<p>(<b>a</b>) Example of spatial mismatch/disagreement between the DNR’s and this study’s interpretations of a coldwater stream and (<b>b</b>) terrain shadow impacting coldwater stream visualization/interpretation in winter panchromatic imagery.</p>
Full article ">
23 pages, 41454 KiB  
Article
2chADCNN: A Template Matching Network for Season-Changing UAV Aerial Images and Satellite Imagery
by Yan Ren, Yuwei Liu, Zhenjia Huang, Wanquan Liu and Weina Wang
Drones 2023, 7(9), 558; https://doi.org/10.3390/drones7090558 - 30 Aug 2023
Cited by 1 | Viewed by 2435
Abstract
Visual navigation based on image matching has become one of the most important research fields for UAVs to achieve autonomous navigation, because of its low cost, strong anti-jamming ability, and high performance. Currently, numerous positioning and navigation methods based on visual information have [...] Read more.
Visual navigation based on image matching has become one of the most important research fields for UAVs to achieve autonomous navigation, because of its low cost, strong anti-jamming ability, and high performance. Currently, numerous positioning and navigation methods based on visual information have been proposed for UAV navigation. However, the appearance, shape, color, and texture of objects can change significantly due to different lighting conditions, shadows, and surface coverage during different seasons, such as vegetation cover in summer or ice and snow cover in winter. These changes pose greater challenges for feature-based image matching methods. This encouraged us to overcome the limitations of previous works, which did not consider significant seasonal changes such as snow-covered UAV aerial images, by proposing an image matching method using season-changing UAV aerial images and satellite imagery. Following the pipeline of a two-channel deep convolutional neural network, we first pre-scaled the UAV aerial images, ensuring that the UAV aerial images and satellite imagery had the same ground sampling distance. Then, we introduced attention mechanisms to provide additional supervision for both low-level local features and high-level global features, resulting in a new season-specific feature representation. The similarity between image patches was calculated using a similarity measurement layer composed of two fully connected layers. Subsequently, we conducted template matching to estimate the UAV matching position with the highest similarity. Finally, we validated our proposed method on both synthetic and real UAV aerial image datasets, and conducted direct comparisons with previous popular works. The experimental results demonstrated that our method achieved the highest matching accuracy on multi-temporal and multi-season images. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of different seasonal UAV aerial images and the corresponding satellite imagery. (<b>a</b>) UAV aerial image in summer. (<b>b</b>) UAV aerial image in winter. (<b>c</b>) satellite imagery in spring.</p>
Full article ">Figure 2
<p>Flowchart of the proposed method.</p>
Full article ">Figure 3
<p>Overview of the two-channel attention deep convolutional neural network.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) part of the original UAV aerial images. (<b>e</b>–<b>h</b>) part of the synthetic images.</p>
Full article ">Figure 5
<p>Part of the season-changing UAV aerial images. (<b>a</b>) UAV aerial images in summer. (<b>b</b>) UAV aerial images in winter.</p>
Full article ">Figure 6
<p>Matching results on the Summer dataset. (<b>a</b>) A summer UAV aerial image at 252 m high and the corresponding spring satellite imagery, Accuracy was 0.9197. (<b>b</b>) Summer UAV aerial image at 304 m high, and the corresponding spring satellite imagery, Accuracy was 0.9522. (<b>c</b>) Summer UAV aerial image at 498 m high and the corresponding spring satellite imagery, Accuracy was 0.9884. (<b>d</b>) Summer aerial image at 500 m high and the corresponding spring satellite imagery, Accuracy was 0.9716.</p>
Full article ">Figure 7
<p>Matching results on the Syn-Winter dataset. (<b>a</b>) A synthetic snow-covered UAV aerial image at 252 m high and the corresponding spring satellite imagery, Accuracy was 0.9522. (<b>b</b>) A synthetic snow-covered UAV aerial image at 304 m high and the corresponding spring satellite imagery, Accuracy was 0.9227. (<b>c</b>) A synthetic snow-covered UAV aerial image at 498 m high and the corresponding spring satellite imagery, Accuracy was 0.8247. (<b>d</b>) A synthetic snow-covered aerial image at 500 m high and the corresponding spring satellite imagery, Accuracy was 0.9197.</p>
Full article ">Figure 8
<p>Matching results on the Winter dataset. (<b>a</b>) A winter UAV aerial image at 190 m high and the corresponding spring satellite imagery, Accuracy was 0.8223. (<b>b</b>) A winter UAV aerial image at 270 m high and the corresponding spring satellite imagery, Accuracy was 0.9273. (<b>c</b>) A winter UAV aerial image at 292 m high and the corresponding spring satellite imagery, Accuracy was 0.9226. (<b>d</b>) A winter UAV aerial image at 320 m high and the corresponding spring satellite imagery, Accuracy was 0.9238.</p>
Full article ">Figure 9
<p>Example results on synthetic UAV aerial images: (<b>a</b>) Two synthetic snow-covered UAV aerial images; (<b>b</b>) SIFT; (<b>c</b>) SSD; (<b>d</b>) SAD; (<b>e</b>) NCC; (<b>f</b>) SuperGlue; (<b>g</b>) LoFTR; (<b>h</b>) MatchNet; (<b>i</b>) RORD; (<b>j</b>) LightGlue; (<b>k</b>) 2chDCNN; (<b>l</b>) the proposed method.</p>
Full article ">Figure 10
<p>Example matching results of UAV aerial images in summer: (<b>a</b>) Two synthetic snow-covered UAV aerial images; (<b>b</b>) SIFT; (<b>c</b>) SSD; (<b>d</b>) SAD; (<b>e</b>) NCC; (<b>f</b>) SuperGlue; (<b>g</b>) LoFTR; (<b>h</b>) MatchNet; (<b>i</b>) RORD; (<b>j</b>) LightGlue; (<b>k</b>) 2chDCNN; (<b>l</b>) the proposed method.</p>
Full article ">Figure 11
<p>Example matching results of snow-covered UAV aerial images in winter: (<b>a</b>) Two synthetic snow-covered UAV aerial images; (<b>b</b>) SIFT; (<b>c</b>) SSD; (<b>d</b>) SAD; (<b>e</b>) NCC; (<b>f</b>) SuperGlue; (<b>g</b>) LoFTR; (<b>h</b>) MatchNet; (<b>i</b>) RORD; (<b>j</b>) LightGlue; (<b>k</b>) 2chDCNN; (<b>l</b>) the proposed method.</p>
Full article ">Figure 12
<p>ROC curves of various methods on the Syn-Winter dataset.</p>
Full article ">Figure 13
<p>ROC curves of various methods on the Summer dataset.</p>
Full article ">Figure 14
<p>ROC curves of various methods on the Winter dataset.</p>
Full article ">Figure 15
<p>Comparison on the features of the “BASE + CBAM(after the first conv) + CBAM(after the last conv)” model in the different layers on the Winter and Summer datasets. Visualization of features on the Winter dataset at (<b>a</b>) the first conv; (<b>b</b>) CBAM(after the first conv); (<b>e</b>) the last conv; (<b>f</b>) CBAM(after the last conv). Visualization of features on the Summer dataset at (<b>c</b>) the first conv; (<b>d</b>) CBAM(after the first conv); (<b>g</b>) the last conv; (<b>h</b>) CBAM(after the last conv).</p>
Full article ">
18 pages, 6031 KiB  
Article
UCTNet with Dual-Flow Architecture: Snow Coverage Mapping with Sentinel-2 Satellite Imagery
by Jinge Ma, Haoran Shen, Yuanxiu Cai, Tianxiang Zhang, Jinya Su, Wen-Hua Chen and Jiangyun Li
Remote Sens. 2023, 15(17), 4213; https://doi.org/10.3390/rs15174213 - 27 Aug 2023
Cited by 3 | Viewed by 1284
Abstract
Satellite remote sensing (RS) has been drawing considerable research interest in land-cover classification due to its low price, short revisit time, and large coverage. However, clouds pose a significant challenge, occluding the objects on satellite RS images. In addition, snow coverage mapping plays [...] Read more.
Satellite remote sensing (RS) has been drawing considerable research interest in land-cover classification due to its low price, short revisit time, and large coverage. However, clouds pose a significant challenge, occluding the objects on satellite RS images. In addition, snow coverage mapping plays a vital role in studying hydrology and climatology and investigating crop disease overwintering for smart agriculture. Distinguishing snow from clouds is challenging since they share similar color and reflection characteristics. Conventional approaches with manual thresholding and machine learning algorithms (e.g., SVM and Random Forest) could not fully extract useful information, while current deep-learning methods, e.g., CNNs or Transformer models, still have limitations in fully exploiting abundant spatial/spectral information of RS images. Therefore, this work aims to develop an efficient snow and cloud classification algorithm using satellite multispectral RS images. In particular, we propose an innovative algorithm entitled UCTNet by adopting a dual-flow structure to integrate information extracted via Transformer and CNN branches. Particularly, CNN and Transformer integration Module (CTIM) is designed to maximally integrate the information extracted via two branches. Meanwhile, Final Information Fusion Module and Auxiliary Information Fusion Head are designed for better performance. The four-band satellite multispectral RS dataset for snow coverage mapping is adopted for performance evaluation. Compared with previous methods (e.g., U-Net, Swin, and CSDNet), the experimental results show that the proposed UCTNet achieves the best performance in terms of accuracy (95.72%) and mean IoU score (91.21%) while with the smallest model size (3.93 M). The confirmed efficiency of UCTNet shows great potential for dual-flow architecture on snow and cloud classification. Full article
Show Figures

Figure 1

Figure 1
<p>Framework of snow/cloud classification research in this study.</p>
Full article ">Figure 2
<p>Visualization of all 40 scenes captured using Sentinel-2 satellite with scene captured date [<a href="#B12-remotesensing-15-04213" class="html-bibr">12</a>].</p>
Full article ">Figure 3
<p>Labeled classification maps of all 40 collected scenes. The three labeled classes are in different colors: black denotes background, red denotes cloud, and cyan denotes snow [<a href="#B12-remotesensing-15-04213" class="html-bibr">12</a>].</p>
Full article ">Figure 4
<p>Illustration of the vanilla Transformer architecture.</p>
Full article ">Figure 5
<p>Architecture of the proposed UCTNet by dual–flow approach.</p>
Full article ">Figure 6
<p>Structure of CTIM module to leverage local features and global representations.</p>
Full article ">Figure 7
<p>Structure of FIFM module to finally fuse CNN–Transformer branch information.</p>
Full article ">Figure 8
<p>The structure of AIFH module to enhance feature representations of stage Dec2.</p>
Full article ">Figure 9
<p>Visualization of segmentation results by different models.</p>
Full article ">
Back to TopTop