[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

Search Results (171)

Search Parameters:
Keywords = snow classification

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 1731 KiB  
Article
A Study on the Identification of the Water Army to Improve the Helpfulness of Online Product Reviews
by Chuyang Li, Shijia Zhang and Xiangdong Liu
Mathematics 2024, 12(20), 3234; https://doi.org/10.3390/math12203234 (registering DOI) - 16 Oct 2024
Viewed by 280
Abstract
Based on the perspective of identifying the water army, this paper uses the methods of machine learning and data visualization to analyze the helpfulness of online produce reviews, portray product portraits, and provide real and helpful product reviews. In order to identify and [...] Read more.
Based on the perspective of identifying the water army, this paper uses the methods of machine learning and data visualization to analyze the helpfulness of online produce reviews, portray product portraits, and provide real and helpful product reviews. In order to identify and eliminate the water army, the Term Frequency-Inverse Document Frequency Model (TF-IDF) and Latent Semantic Index Model (LSI) are used. After eliminating the water army, three classification methods were selected to perform sentimental analysis, including logistics, SnowNLP, and Convolutional Neural Network for text(TextCNN). The TextCNN has the highest F1 score among the three classification methods. At the same time, the Latent Dirichlet Allocation mode (LDA) is used to extract the topics of various reviews. Finally, targeted countermeasures are proposed to manufacturers, consumers, and regulators. Full article
(This article belongs to the Special Issue Big Data Mining and Analytics with Applications)
Show Figures

Figure 1

Figure 1
<p>Sentiment analysis process.</p>
Full article ">Figure 2
<p>CNN training process.</p>
Full article ">Figure 3
<p>Sentiment analysis score chart. The horizontal axis represents the ID of 4561 users, and the vertical axis represents the score.</p>
Full article ">Figure 4
<p>Cosine similarity for both positive and negative reviews. The horizontal axis measures a number of topics, and the vertical axis measures cosine similarity.</p>
Full article ">
17 pages, 9390 KiB  
Article
Applicability of Relatively Low-Cost Multispectral Uncrewed Aerial Systems for Surface Characterization of the Cryosphere
by Colby F. Rand and Alia L. Khan
Remote Sens. 2024, 16(19), 3662; https://doi.org/10.3390/rs16193662 - 1 Oct 2024
Viewed by 519
Abstract
This paper investigates the ability of a relatively low cost, commercially available uncrewed aerial vehicle (UAV), the DJI Mavic 3 Multispectral, to perform cryospheric research. The performance of this UAV, where applicable, is compared to a similar but higher cost system, the DJI [...] Read more.
This paper investigates the ability of a relatively low cost, commercially available uncrewed aerial vehicle (UAV), the DJI Mavic 3 Multispectral, to perform cryospheric research. The performance of this UAV, where applicable, is compared to a similar but higher cost system, the DJI Matrice 350, equipped with a Micasense RedEdge-MX Multispectral dual-camera system. The Mavic 3 Multispectral was tested at three field sites: the Lemon Creek Glacier, Juneau Icefield, AK; the Easton Glacier, Mt. Baker, WA; and Bagley Basin, Mt. Baker, WA. This UAV proved capable of mapping the spatial distribution of red snow algae on the surface of the Lemon Creek Glacier using both spectral indices and a random forest supervised classification method. The UAV was able to assess the timing of snowmelt and changes in suncup morphology on snow-covered areas within the Bagley Basin. Finally, the UAV was able to classify glacier surface features using a random forest algorithm with an overall accuracy of 68%. The major advantages of this UAV are its low weight, which allows it to be easily transported into the field, its low cost compared to other alternatives, and its ease of use. One limitation would be the omission of a blue multispectral band, which would have allowed it to more easily classify glacial ice and snow features. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere II)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SkySat image of the Lemon Creek Glacier (© Planet Labs). The locations of the ground control points used during this drone survey are shown by the black crosses. (<b>b</b>) The Lemon Creek glacier is located in the southernmost extent of the Juneau Icefield. Glaciated areas are shown in white (RGI Consortium [<a href="#B26-remotesensing-16-03662" class="html-bibr">26</a>]). (<b>c</b>) The Juneau Icefield is located on the border of southeast Alaska, USA and British Columbia, Canada, as shown by the red star. (Map projection: WGS 1984 UTM Zone 8N).</p>
Full article ">Figure 2
<p>(<b>a</b>) Topographic reference map of Mt. Baker, Washington, USA, showing glaciated areas in white (RGI Consortium [<a href="#B26-remotesensing-16-03662" class="html-bibr">26</a>]). (<b>b</b>) SkySat image of the Easton glacier on the southern slope of Mt. Baker (© Planet Labs). The outline of the glacier is shown by the black polygon (RGI Consortium [<a href="#B26-remotesensing-16-03662" class="html-bibr">26</a>]). (<b>c</b>) SkySat image of Bagley Basin, an alpine basin to the northeast of Mt. Baker, adjacent to the Mt. Baker Ski Area (© Planet Labs). (Map projection: WGS 1984 UTM Zone 10N).</p>
Full article ">Figure 3
<p>Drone platforms and sensors. (<b>a</b>) DJI Mavic 3 Multispectral (Mavic 3M), (<b>b</b>) DJI Matrice 350 RTK (Matrice 350), (<b>c</b>) Micasense RedEdge-MX Multispectral Dual-Camera System (Micasense) and downwelling light sensor (DLS).</p>
Full article ">Figure 4
<p>Orthomosaic images of the Lemon Creek Glacier, Juneau Icefield, Alaska, derived from images captured by the RGB bands (Band 5: Red-668, Band 3: Green-560, and Band 2: Blue-475) of the Micasense RedEdge-MX dual camera system (<b>left</b>) and the RGB camera on a DJI Mavic 3 Multispectral (<b>right</b>) collected from 21 August through 23 August 2023. The black rectangle toward the southern end of the Lemon Creek glacier depicts the high density algae area used for analysis in <a href="#remotesensing-16-03662-f005" class="html-fig">Figure 5</a>. (Map projection: WGS 1984 UTM Zone 8N).</p>
Full article ">Figure 5
<p>High density algae area in the southern region of the Lemon Creek Glacier, captured by the DJI Mavic 3 Multispectral, as shown in the left column, and Micasense Red-Edge MX camera system, as shown in the right column. (<b>a</b>) Mavic 3M multispectral composite orthomosaic, symbolized by Red: Band 2, Green: Band 1, and Blue: Band 1. Since there is no blue multispectral band on the Mavic 3M, the green band was used for both the green and blue symbology. (<b>b</b>) Micasense multispectral composite orthomosaic, symbolized by Red: Band 5, Green: Band 3, and Blue: Band 2. (<b>c</b>) ORG index applied to Mavic 3M orthomosaic. (<b>d</b>) ORG index applied to Micasense orthomosaic. (<b>e</b>) random forest classification of Mavic 3M orthomosaic. (<b>f</b>) random forest classification of the Micasense orthomosaic. Algae sample locations are shown by the black and white circles. (Map projection: WGS 1984 UTM Zone 8N).</p>
Full article ">Figure 6
<p>RGB orthomosaic of Bagley Basin, captured by the Mavic 3M on 8 June 2023. The inset map shows meltwater filled suncups on the surface of Upper Bagley Lake. (Map projection: WGS 1984 UTM Zone 10N).</p>
Full article ">Figure 7
<p>RGB orthomosaics and DEMs of a snow surface to the west of Upper Bagley Lake, within Bagley Basin, derived from DJI Mavic 3M imagery captured on (<b>a</b>) 8 June 2023, (<b>b</b>) 26 June 2023, (<b>c</b>) 5 July 2023, (<b>d</b>) 10 July 2023, (<b>e</b>) 13 June 2023, and (<b>f</b>) 17 July 2023. The red circle denotes an individual who’s diameter was measured to assess changes to suncup morphology.</p>
Full article ">Figure 8
<p>Digital elevation models of Easton glacier, captured by Mavic 3M mapping surveys on 20 July, 1 August, 13 August, and 8 September 2023. (Map projection: WGS 1984 UTM Zone 10N).</p>
Full article ">Figure 9
<p>RGB orthomosaic and random forest supervised classification results of the 13 August 2023, Mavic 3M flight of the Easton glacier.</p>
Full article ">
17 pages, 14569 KiB  
Article
Cross-Country Ski Skating Style Sub-Technique Detection and Skiing Characteristic Analysis on Snow Using High-Precision GNSS
by Shunya Uda, Naoto Miyamoto, Kiyoshi Hirose, Hiroshi Nakano, Thomas Stöggl, Vesa Linnamo, Stefan Lindinger and Masaki Takeda
Sensors 2024, 24(18), 6073; https://doi.org/10.3390/s24186073 - 19 Sep 2024
Viewed by 550
Abstract
A comprehensive analysis of cross-country skiing races is a pivotal step in establishing effective training objectives and tactical strategies. This study aimed to develop a method of classifying sub-techniques and analyzing skiing characteristics during cross-country skiing skating style timed races on snow using [...] Read more.
A comprehensive analysis of cross-country skiing races is a pivotal step in establishing effective training objectives and tactical strategies. This study aimed to develop a method of classifying sub-techniques and analyzing skiing characteristics during cross-country skiing skating style timed races on snow using high-precision kinematic GNSS devices. The study involved attaching GNSS devices to the heads of two athletes during skating style timed races on cross-country ski courses. These devices provided precise positional data and recorded vertical and horizontal head movements and velocity over ground (VOG). Based on these data, sub-techniques were classified by defining waveform patterns for G2, G3, G4, and G6P (G6 with poling action). The validity of the classification was verified by comparing the GNSS data with video analysis, a process that yielded classification accuracies ranging from 95.0% to 98.8% for G2, G3, G4, and G6P. Notably, G4 emerged as the fastest technique, with sub-technique selection varying among skiers and being influenced by skiing velocity and course inclination. The study’s findings have practical implications for athletes and coaches as they demonstrate that high-precision kinematic GNSS devices can accurately classify sub-techniques and detect skiing characteristics during skating style cross-country skiing races, thereby providing valuable insights for training and strategy development. Full article
(This article belongs to the Special Issue Sensors and Wearable Technologies in Sport Biomechanics)
Show Figures

Figure 1

Figure 1
<p>The figure shows the Ikenotaira cross-country ski course, Japan, used in this study. The plotted data were obtained from the study subject, covering one lap of 0.8 km. The figure shows the course profile’s plan view data (<b>a</b>) and course inclination data (<b>b</b>).</p>
Full article ">Figure 2
<p>This picture and image show the experimental setup. The GNSS antenna was attached to the skier’s head, and the receiver and mobile router were stored in a small bag at the skier’s waist. This setup obtained head positioning data (latitude, longitude, altitude, and VOG) during the timed race.</p>
Full article ">Figure 3
<p>The figure shows the typical waveform patterns of subject A (<b>a</b>) and subject B (<b>b</b>) for G2, G3, G4, and G6P. The black dashed lines indicate the points where the net vertical head movement reaches a peak. The interval between two black lines represents one cycle. The green lines indicate the VOG. The blue waveform shows the trajectory of the net vertical head movement. The red waveform shows the trajectory of the net horizontal head movement. The red bars indicate the amplitude of the net horizontal head movement.</p>
Full article ">Figure 4
<p>The figure shows the quality of the positional data obtained from the RTK GNSS devices for subject A and subject B. The green color indicates the fix solution, the orange color indicates the float solution, and the blue color indicates the dGNSS solution.</p>
Full article ">Figure 5
<p>This figure shows the usage ratio over time (<b>a</b>) and the ratio over distance (<b>b</b>) for each sub-technique during the timed race.</p>
Full article ">Figure 6
<p>The distribution of sub-techniques used by two subjects during the second lap of the timed race is shown on the course profile’s plan view data.</p>
Full article ">Figure 7
<p>The course inclination data show the distribution of sub-techniques used by two subjects during the second lap of the timed race.</p>
Full article ">Figure 8
<p>The distribution of sub-techniques used by two subjects during the second lap of the timed race. The <span class="html-italic">X</span>-axis indicates the distance traveled, and the <span class="html-italic">Y</span>-axis indicates the VOG of the skier’s head.</p>
Full article ">Figure 9
<p>This figure shows the CL, CT, skiing velocity, and course inclination data for subjects A and B’s sub-techniques during the timed race. Each sub-technique cycle was defined from the vertical movement peak at the waveform data’s head to the next peak. The horizontal line within each box represents the median value of the dataset, while the “x” symbol denotes the mean value. ** indicates a significance level of <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 10
<p>The distribution of four sub-techniques used by two subjects during the timed race is shown with skiing velocity (<span class="html-italic">X</span>-axis) and course inclination (<span class="html-italic">Y</span>-axis).</p>
Full article ">Figure 11
<p>The distribution of four sub-techniques used by two subjects during the timed race is shown as a histogram of skiing velocity frequencies.</p>
Full article ">Figure 12
<p>The distribution of four sub-techniques used by two subjects during the timed race is shown as a histogram of course inclination frequencies.</p>
Full article ">
19 pages, 4666 KiB  
Article
Quantifying Qiyi Glacier Surface Dirtiness Using UAV and Sentinel-2 Imagery
by Jiangtao Chen, Ninglian Wang, Yuwei Wu, Anan Chen, Chenlie Shi, Mingjie Zhao and Longjiang Xie
Remote Sens. 2024, 16(17), 3351; https://doi.org/10.3390/rs16173351 - 9 Sep 2024
Viewed by 426
Abstract
The glacier surface is composed not only of ice or snow but also of a heterogeneous mixture of various materials. The presence of light-absorbing impurities darkens the glacier surface, reducing local reflectance and thereby accelerating the glacier melting process. However, our understanding of [...] Read more.
The glacier surface is composed not only of ice or snow but also of a heterogeneous mixture of various materials. The presence of light-absorbing impurities darkens the glacier surface, reducing local reflectance and thereby accelerating the glacier melting process. However, our understanding of the spatial distribution of these impurities remains limited, and there is a lack of studies on quantifying the dirty degree of glacier surfaces. During the Sentinel satellite overpass on 21 August 2023, we used an ASD FieldSpec3 spectrometer to measure the reflectance spectra of glacier surfaces with varying degrees of dirtiness on the Qiyi glacier, Qinghai–Tibet Plateau. Using Multiple Endmember Spectral Mixture Analysis (MESMA), the Sentinel imagery was decomposed to generate fraction images of five primary ice surface materials as follows: coarse-grained snow, slightly dirty ice, moderately dirty ice, extremely dirty ice, and debris. Using unmanned aerial vehicle (UAV) imagery with a 0.05 m resolution, the primary ice surface was delineated and utilized as reference data to validate the fraction images. The findings revealed a strong correlation between the fraction images and the reference data (R2 ≥ 0.66, RMSE ≤ 0.21). Based on pixel-based classification from the UAV imagery, approximately 80% of the glacier surface is covered by slightly dirty ice (19.2%), moderately dirty ice (33.3%), extremely dirty ice (26.3%), and debris (1.2%), which significantly contributes to its darkening. Our study demonstrates the effectiveness of using Sentinel imagery in conjunction with MESMA to map the degree of glacier surface dirtiness accurately. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Qiyi glacier (red star). (<b>b</b>) A true-color RGB image (10 m resolution) of the glacier, with the blue curve outlining its boundary. Red circles represent spectral sampling points, yellow triangles indicate UAV ground control points, and pink rectangles delineate the validation areas. (<b>c</b>,<b>d</b>) are images of the glacier terminus taken on 31 July 2013, and 15 August 2023, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) Spectral measurements were collected with a fiber optic probe ~1 m above the ice surface. (<b>b</b>) The actual measured spectral curves are depicted with solid black lines, while colored circles represent the reflectance values at the central wavelengths of Sentinel-2B bands (B2-B8A bands correspond to red to pink hues on the graph).</p>
Full article ">Figure 3
<p>Spectral curves after SRF conversion, where solid lines represent mean values and shaded areas denote standard deviations (<b>a</b>). Photographs of the following categories of ice are shown: (<b>b</b>) coarse-grained snow; (<b>c</b>) slightly dirty ice; (<b>d</b>) moderately dirty ice; (<b>e</b>) extremely dirty ice; and (<b>f</b>) supraglacial rivers. The spectrometer’s field of view is a ~50 cm diameter circle; a pen is placed for scale, aiming to provide readers with a sense of proportion for better comprehension.</p>
Full article ">Figure 4
<p>A comparison of measured reflectance and satellite products, where red pentagrams denote the sampling positions of the spectrometer. (<b>a</b>,<b>b</b>) represent relatively clean glacier surfaces, while (<b>c</b>,<b>d</b>) depict relatively dirty glacier surfaces. L2A denotes products produced by the ESA, FLAASH (10 m) signifies atmospheric correction through FLAASH, and L2A (Sen2cor) indicates correction via the Sen2cor plugin. SRF refers to spectral response function conversion, the green line represents the measured spectra, and L1C denotes ESA L1C products.</p>
Full article ">Figure 5
<p>(<b>a</b>) The UAV image and (<b>b</b>) the SVM-classified image.</p>
Full article ">Figure 6
<p>The final spectral endmembers for the following different glacier surface types: (<b>a</b>) coarse-grained snow; (<b>b</b>) slightly dirty ice; (<b>c</b>) moderately dirty ice; (<b>d</b>) extremely dirty ice; and (<b>e</b>) debris.</p>
Full article ">Figure 7
<p>Fraction images for the following five distinct ice surface types are presented: (<b>a</b>) coarse-grained snow; (<b>b</b>) slightly dirty ice; (<b>c</b>) moderately dirty ice; (<b>d</b>) extremely dirty ice; and (<b>e</b>) debris.</p>
Full article ">Figure 8
<p>A regression model was constructed to examine the relationship between MESMA fraction images and reference fraction (UAV images). The solid line illustrates the degree of fitting, while the shaded area represents the 95% confidence interval. The determination coefficient (R<sup>2</sup>) and root mean square error (RMSE) are presented, <span class="html-italic">n</span> = 330.</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 760
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 ">
25 pages, 12201 KiB  
Article
Spatial Consistency and Accuracy Analysis of Multi-Source Land Cover Products on the Southeastern Tibetan Plateau, China
by Binghua Zhang, Linshan Liu, Yili Zhang, Bo Wei, Dianqing Gong and Lanhui Li
Remote Sens. 2024, 16(17), 3219; https://doi.org/10.3390/rs16173219 - 30 Aug 2024
Viewed by 641
Abstract
Land cover products provide the key inputs for terrestrial change monitoring and modeling. Numerous land cover products have been generated in the past few decades, but their performance on the southeastern Tibetan Plateau remains unclear. This study analyzed 15 land cover products for [...] Read more.
Land cover products provide the key inputs for terrestrial change monitoring and modeling. Numerous land cover products have been generated in the past few decades, but their performance on the southeastern Tibetan Plateau remains unclear. This study analyzed 15 land cover products for consistency through compositional similarity and overlay analyses. Additionally, 1305 validation samples from four datasets were employed to construct confusion matrices to evaluate their accuracy. The results indicate the following: (1) Land cover products exhibit relatively high consistency in 62.92% of the region. (2) Land cover products are strongly influenced by terrain fluctuations, showing lower consistency at elevation below 200 m and instability in land cover classification with increasing elevation, particularly between 2800–4400 m and 4800–5400 m. (3) The accuracy for forest, water, and snow/ice is relatively high. However, there is a relatively lower accuracy for wetland and shrubland, necessitating more field samples for reference to improve classification. (4) The average values of the four validation datasets show that the overall accuracy of the 15 products ranges from 50.97% to 73.50%. For broad-scale studies with lower resolution requirements, the CGLS-LC100 product can be considered. For studies requiring a finer scale, a combination of multiple land cover products should be utilized. ESRI is recommended as a reference for built-up land, while FROM-GLC30 can be used for cropland, although misclassification issues should be noted. This study provides valuable insights for analyzing land cover types on plateaus to refine classification. It also offers guidance for selecting suitable land cover products for future research in this region. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area, showing: (<b>a</b>) the distribution of the validation sample sets; (<b>b</b>) the distribution of the study area on the TP; and (<b>c</b>) the distribution of the two prefecture-level cities. Red color in (<b>a</b>,<b>b</b>) indicate study area. The boundary of the TP was obtained from the study of Zhang et al. [<a href="#B35-remotesensing-16-03219" class="html-bibr">35</a>].</p>
Full article ">Figure 2
<p>Flowchart of this study.</p>
Full article ">Figure 3
<p>Number of validation points for different land cover types.</p>
Full article ">Figure 4
<p>Comparison of proportions of land cover types for 15 products.</p>
Full article ">Figure 5
<p>Comparative analysis of spatial distribution for 15 land cover products.</p>
Full article ">Figure 6
<p>Compositional similarity analysis of 15 products.</p>
Full article ">Figure 7
<p>Spatial consistency analysis of 15 land cover products on the southeastern TP. Note: For number of datasets showing agreement, full agreement corresponds to a value of 15 (all products indicate the same land cover types), while lowest agreement corresponds to a value of 2 (only 2 products show the same land cover type).</p>
Full article ">Figure 8
<p>Spatial consistency comparison across different altitudes. Note: Numbers on the top are the number of datasets showing agreement, full agreement corresponds to a value of 15 (all products indicate the same land cover types), while lowest agreement corresponds to a value of 2 (only 2 products show the same land cover type).</p>
Full article ">Figure 9
<p>Spatial consistency comparison across different land cover types. Note: Numbers on the right indicate spatial consistency of land cover types across 15 datasets: 1 means only one dataset classifies the area as this type, while 15 means all datasets classify it as this type.</p>
Full article ">Figure 10
<p>Accuracy assessment of 15 products. (<b>a</b>) OA tested by four validation sample sets across 15 land cover products; (<b>b</b>,<b>c</b>) box plots of PA and UA; (<b>d</b>–<b>l</b>) F-scores of different land cover products examined by four validation sample sets. Note: Abbreviations used are as follows: 1 (GLC2000), 2 (GLCSHARE), 3 (GLCNMO), 4 (MCD12Q1), 5 (GlobeCover), 6 (CCI-LC), 7 (CGLS-LC100), 8 (Globeland30), 9 (CNLUCC), 10 (FROM-GLC30), 11 (GLC-FCS30), 12 (AGLC), 13 (CLCD), 14 (ESRI), 15 (Worldcover), FR (Forest), SL (Shrubland), GL (Grassland), WL (Wetland), BL (Built-up Land), CL (Cropland), BLD (Bareland), WT (Water), and SI (Snow/Ice).</p>
Full article ">Figure 11
<p>Confusion matrix analysis of 15 land cover products. Note: Abbreviations used are FR (Forest), SL (Shrubland), GL (Grassland), WL (Wetland), BL (Built-up Land), CL (Cropland), BLD (Bareland), WT (Water), and SI (Snow/Ice).</p>
Full article ">Figure 12
<p>Spatial consistency and accuracy assessment of 15 land cover products in northern and southern region. Note: Numbers on the bottom of (<b>a</b>) are the number of datasets showing agreement; full agreement corresponds to a value of 15 (all products indicate the same land cover types), while lowest agreement corresponds to a value of 2 (only 2 products show the same land cover type). Abbreviations used in (<b>b</b>,<b>c</b>) are 1 (GLC2000), 2 (GLCSHARE), 3 (GLCNMO), 4 (MCD12Q1), 5 (GlobeCover), 6 (CCI-LC), 7 (CGLS-LC100), 8 (Globeland30), 9 (CNLUCC), 10 (FROM-GLC30), 11 (GLC-FCS30), 12 (AGLC), 13 (CLCD), 14 (ESRI), 15 (Worldcover).</p>
Full article ">Figure 13
<p>Land cover product comparison in northern region of study area.</p>
Full article ">Figure 14
<p>Land cover product comparison in southern region of study area.</p>
Full article ">
24 pages, 32875 KiB  
Article
Integrating Sequential Backward Selection (SBS) and CatBoost for Snow Avalanche Susceptibility Mapping at Catchment Scale
by Sinem Cetinkaya and Sultan Kocaman
ISPRS Int. J. Geo-Inf. 2024, 13(9), 312; https://doi.org/10.3390/ijgi13090312 - 29 Aug 2024
Viewed by 645
Abstract
Snow avalanche susceptibility (AS) mapping is a crucial step in predicting and mitigating avalanche risks in mountainous regions. The conditioning factors used in AS modeling are diverse, and the optimal set of factors depends on the environmental and geological characteristics of the region. [...] Read more.
Snow avalanche susceptibility (AS) mapping is a crucial step in predicting and mitigating avalanche risks in mountainous regions. The conditioning factors used in AS modeling are diverse, and the optimal set of factors depends on the environmental and geological characteristics of the region. Using a sub-optimal set of input features with a data-driven machine learning (ML) method can lead to challenges like dealing with high-dimensional data, overfitting, and reduced model generalization. This study implemented a robust framework involving the Sequential Backward Selection (SBS) algorithm and a decision-tree based ML model, CatBoost, for the automatic selection of predictive variables for AS mapping. A comprehensive inventory of a large avalanche period, previously derived from satellite images, was used for the investigations in three distinct catchment areas in the Swiss Alps. The integrated SBS-CatBoost approach achieved very high classification accuracies between 94% and 97% for the three catchments. In addition, the Shapley additive explanations (SHAP) method was employed to analyze the contributions of each feature to avalanche occurrences. The proposed methodology revealed the benefits of integrating advanced feature selection algorithms with ML techniques for AS assessment. We aimed to contribute to avalanche hazard knowledge by assessing the impact of each feature in model learning. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic overview of the proposed framework.</p>
Full article ">Figure 2
<p>Engelberger Aa, Meienreuss, and Göschenerreuss catchments (sub-catchments of the Reuss River Basin).</p>
Full article ">Figure 3
<p>Altitude maps of Engelberger Aa, Meienreuss, and Göschenerreuss catchments.</p>
Full article ">Figure 4
<p>LULC classification maps for Engelberger Aa, Meienreuss, and Göschenenreuss.</p>
Full article ">Figure 5
<p>Flowchart of iterative feature elimination using CatBoost with GridSearchCV.</p>
Full article ">Figure 6
<p>(<b>a</b>) The AS map of the Engelberger Aa catchment using the optimal feature set and (<b>b</b>) SHAP summary plot.</p>
Full article ">Figure 7
<p>(<b>a</b>) The AS map of the Meienreuss catchment using the optimal feature set and (<b>b</b>) SHAP summary plot.</p>
Full article ">Figure 8
<p>(<b>a</b>) The AS map of the Göschenerreuss catchment using the optimal feature set and (<b>b</b>) SHAP summary plot.</p>
Full article ">Figure A1
<p>The AS maps for Engelberger Aa, Meienreuss, and Göschenerreuss catchments, based on models with varying numbers of features (16, 15, 14, 13, and 12).</p>
Full article ">Figure A2
<p>The AS maps for Engelberger Aa, Meienreuss, and Göschenerreuss catchments, based on models with varying numbers of features (11, 10, 9, 8, and 7).</p>
Full article ">Figure A3
<p>The AS maps for Engelberger Aa, Meienreuss, and Göschenerreuss catchments, based on models with varying numbers of features (6, 5, 4, 3, and 2).</p>
Full article ">
21 pages, 15343 KiB  
Article
River Ice Mapping from Landsat-8 OLI Top of Atmosphere Reflectance Data by Addressing Atmospheric Influences with Random Forest: A Case Study on the Han River in South Korea
by Hyangsun Han, Taewook Kim and Seohyeon Kim
Remote Sens. 2024, 16(17), 3187; https://doi.org/10.3390/rs16173187 - 29 Aug 2024
Viewed by 478
Abstract
Accurate river ice mapping is crucial for predicting and managing floods caused by ice jams and for the safe operation of hydropower and water resource facilities. Although satellite multispectral images are widely used for river ice mapping, atmospheric contamination limits their effectiveness. This [...] Read more.
Accurate river ice mapping is crucial for predicting and managing floods caused by ice jams and for the safe operation of hydropower and water resource facilities. Although satellite multispectral images are widely used for river ice mapping, atmospheric contamination limits their effectiveness. This study developed river ice mapping models for the Han River in South Korea using atmospherically uncorrected Landsat-8 Operational Land Imager (OLI) multispectral reflectance data, addressing atmospheric influences with a Random Forest (RF) classification approach. The RF-based river ice mapping models were developed by implementing various combinations of input variables, incorporating the Landsat-8 multispectral top-of-atmosphere (TOA) reflectance, normalized difference indices for snow, water, and bare ice, and atmospheric factors such as aerosol optical depth, water vapor content, and ozone concentration from the Moderate Resolution Imaging Spectroradiometer observations, as well as surface elevation from the GLO-30 digital elevation model. The RF model developed using all variables achieved excellent performance in the classification of snow-covered ice, snow-free ice, and water, with an overall accuracy and kappa coefficient exceeding 98.4% and 0.98 for test samples, and higher than 83.7% and 0.75 when compared against reference river ice maps generated by manually interpreting the Landsat-8 images under various atmospheric conditions. The RF-based river ice mapping model for the atmospherically corrected Landsat-8 multispectral surface reflectance was also developed, but it showed very low performance under atmospheric conditions heavily contaminated by aerosol and water vapor. Aerosol optical depth and water vapor content were identified as the most important variables. This study demonstrates that multispectral reflectance data, despite atmospheric contamination, can be effectively used for river ice monitoring by applying machine learning with atmospheric auxiliary data to mitigate atmospheric effects. Full article
Show Figures

Figure 1

Figure 1
<p>Topography of the Han River region in South Korea. The green and yellow rectangles on the left topographic map represent the imaging coverage of Landsat-8 path/row 115/34 and 116/34, respectively. The yellow rectangle on the right map indicates the region of Paldang Lake, which serves as the test site for the river ice mapping models proposed in this study.</p>
Full article ">Figure 2
<p>An example of manual extraction of samples for snow-covered ice, snow-free ice, and water-based on visual investigation of the Landsat-8 RGB true color composite image at the path/row 116/34, obtained on 1 January 2018. The image area corresponds to the yellow box on the right image in <a href="#remotesensing-16-03187-f001" class="html-fig">Figure 1</a>, and the red lines indicate the water boundary.</p>
Full article ">Figure 3
<p>Accuracy evaluation metrics of the RF-based river ice mapping model developed using different variable selection schemes for the test samples from the Landast-8 OLI dataset on (<b>a</b>) 28 January 2022 (medium WV and low AOD), (<b>b</b>) 24 February 2017 (high WV and low ADO), and (<b>c</b>) 15 February 2017 (high WV and high AOD).</p>
Full article ">Figure 4
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 28 January 2022 (low AOD conditions, medium WV content) and the corresponding river ice map generated by (<b>b</b>) manual interpretation and (<b>c</b>–<b>h</b>) the RF models based on variable selection schemes 1 to 6 around the Paldang Lake. (<b>i</b>) Accuracy assessment metrics of the RF model-derived river ice maps calculated based on (<b>b</b>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 24 February 2017 (low AOD conditions, high WV content) and the corresponding river ice map generated by (<b>b</b>) manual interpretation and (<b>c</b>–<b>h</b>) the RF models based on variable selection schemes 1 to 6 around the Paldang Lake. (<b>i</b>) Accuracy assessment metrics of the RF model-derived river ice maps calculated based on (<b>b</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 15 February 2017 (high AOD conditions, high WV content) and the corresponding river ice map generated by (<b>b</b>) manual interpretation and (<b>c</b>–<b>h</b>) the RF models based on variable selection schemes 1 to 6 around the Paldang Lake. (<b>i</b>) Accuracy assessment metrics of the RF model-derived river ice maps calculated based on (<b>b</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 28 January 2022 and (<b>b</b>) the corresponding river ice map generated by the RF model based on variable selection scheme 5. Red polygons in (<b>a</b>) represent the streams of the Han River.</p>
Full article ">Figure 8
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 24 February 2017 and (<b>b</b>) the corresponding river ice map generated by the RF model based on variable selection scheme 5. The red polygons in (<b>a</b>) represent the streams of the Han River.</p>
Full article ">Figure 9
<p>(<b>a</b>) Landsat-8 OLI RGB true-color composite image on 15 February 2017 and (<b>b</b>) the corresponding river ice map generated by the RF model based on variable selection scheme 5. The red polygons in (<b>a</b>) represent the streams of the Han River.</p>
Full article ">Figure 10
<p>(<b>a</b>) Sentinel-2 MSI RGB true-color composite image on 12 February 2017 around the Paldang Lake and (<b>b</b>) the corresponding river ice map generated by manual interpretation.</p>
Full article ">Figure 11
<p>Mean decrease in accuracy of the RF-based river ice mapping model using the input variable selection scheme 5.</p>
Full article ">
31 pages, 19050 KiB  
Article
An Ensemble Machine Learning Approach for Sea Ice Monitoring Using CFOSAT/SCAT Data
by Yanping Luo, Yang Liu, Chuanyang Huang and Fangcheng Han
Remote Sens. 2024, 16(17), 3148; https://doi.org/10.3390/rs16173148 - 26 Aug 2024
Viewed by 692
Abstract
Sea ice is a crucial component of the global climate system. The China–French Ocean Satellite Scatterometer (CFOSAT/SCAT, CSCAT) employs an innovative rotating fan beam system. This study applied principal component analysis (PCA) to extract classification features and developed an ensemble machine learning approach [...] Read more.
Sea ice is a crucial component of the global climate system. The China–French Ocean Satellite Scatterometer (CFOSAT/SCAT, CSCAT) employs an innovative rotating fan beam system. This study applied principal component analysis (PCA) to extract classification features and developed an ensemble machine learning approach for sea ice detection. PCA identified key features from CSCAT’s backscatter information, representing outer and sweet swath observations. The ensemble model’s performances (OA and Kappa) for the Northern and Southern Hemispheres were 0.930, 0.899, and 0.844, 0.747, respectively. CSCAT achieved an accuracy of over 0.9 for close ice and open water but less than 0.3 for open ice, with misclassification of open ice as closed ice. The sea ice extent discrepancy between CSCAT and the National Snow and Ice Data Center (NSIDC) was −0.06 ± 0.36 million km2 in the Northern Hemisphere and −0.03 ± 0.48 million km2 in the Southern Hemisphere. CSCAT’s sea ice closely matched synthetic aperture radar (SAR) imagery, indicating effective sea ice and open water differentiation. CSCAT accurately distinguished sea ice from open water but struggled with open ice classification, with misclassifications in the Arctic’s Greenland Sea and Hudson Bay, and the Antarctic’s sea ice–water boundary. Full article
(This article belongs to the Special Issue Satellite Remote Sensing for Ocean and Coastal Environment Monitoring)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Observation geometry of CSCAT adapted from Zhang et al. [<a href="#B42-remotesensing-16-03148" class="html-bibr">42</a>]. (<b>b</b>) Incidence and azimuth angles versus the cross-track wind vector cell (WVC) number for a row at a latitude of ~43°S from orbit observed on 1 January 2019 at 07:56:26, showcasing WVC views in color and <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> using symbolic circles and forks, respectively. (<b>c</b>) The average number of views at WVC across the swath.</p>
Full article ">Figure 2
<p>Workflow of this study.</p>
Full article ">Figure 3
<p>Location map over (<b>a</b>) the Northern Hemisphere and (<b>b</b>) the Southern Hemisphere for the regions (marked in yellow colors) used in sample selection overlaid on the CAFF Boundary [<a href="#B58-remotesensing-16-03148" class="html-bibr">58</a>], Antarctic Circumpolar Current (<a href="https://data.aad.gov.au/dataset/4892/download" target="_blank">https://data.aad.gov.au/dataset/4892/download</a>, accessed on 20 March 2023) and sea ice median extent [<a href="#B59-remotesensing-16-03148" class="html-bibr">59</a>].</p>
Full article ">Figure 4
<p>Model structure of the soft voting ensemble learning and training process.</p>
Full article ">Figure 5
<p>Pearson’s correlation coefficients in the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere and related principal component analysis (PCA) bioplots of CSCAT backscatter observations over the (<b>c</b>) Northern and (<b>d</b>) Southern Hemispheres on 10 January 2019.</p>
Full article ">Figure 6
<p>Spatial distribution of the first four (out of eight) principal components of <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mo>/</mo> <mi>H</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> polarization in the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere on 10 January 2019, respectively.</p>
Full article ">Figure 7
<p>Time series of cumulative variance of the eigenvalues for principal components in the (<b>a</b>) Northern and (<b>b</b>) Southern Hemispheres between 2019 and 2022.</p>
Full article ">Figure 8
<p>Time series of <math display="inline"><semantics> <mrow> <msubsup> <mi>σ</mi> <mrow> <mi>V</mi> <mi>V</mi> <mo>,</mo> <mi>P</mi> <mi>C</mi> <mn>1</mn> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> with different period lengths in the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere for close ice, open ice, and open water.</p>
Full article ">Figure 9
<p>(<b>a</b>) Feature importance for single models on 10 January 2019 in the Northern Hemisphere (<b>left</b>) and the Southern Hemisphere (<b>right</b>). (<b>b</b>) Statistical results of 10-fold cross-validation F1 scores for different machine learning models from 1 January 2019 to 31 December 2022. (<b>c</b>) Time series of 10-fold cross-validation F1 scores for different machine learning models from 1 January 2019 to 31 December 2022.</p>
Full article ">Figure 10
<p>The time series of the evaluation parameters for (1) overall, (2) close ice, (3) open ice, and (4) open water in the sea ice monitoring ensemble training model in the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere from 1 January 2019 to 31 December 2022, respectively.</p>
Full article ">Figure 11
<p>Daily error analysis for (1) close ice, (2) open ice, and (3) open water in the sea ice monitoring ensemble training model in the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere from 1 January 2019 to 31 December 2022, respectively.</p>
Full article ">Figure 12
<p>Sea ice detection in the (<b>a</b>) Northern Hemisphere on 10 December 2019 and (<b>b</b>) Southern Hemisphere on 10 June 2019 derived from the Dt, Gnb, Knn, Log, Rfc, and ensemble models, respectively.</p>
Full article ">Figure 13
<p>Daily sea ice extent from 2019 to 2022 in the (<b>a1</b>) Northern Hemisphere and (<b>a2</b>) Southern Hemisphere for CSCAT, OSISAF (30% SIC), and NSIDC (30% SIC). Daily sea ice extent difference from 2019 to 2022 in the (<b>b1</b>) Northern Hemisphere and (<b>b2</b>) Southern Hemisphere for CSCAT vs. NSIDC and OSISAF vs. NSIDC. Monthly sea ice extent from 2019 to 2022 over the (<b>c1</b>) Northern Hemisphere and (<b>c2</b>) Southern Hemisphere for CSCAT, OSISAF (30% SIC), and NSIDC (30% SIC). Scatter plot of sea ice extent between CSCAT and NSIDC over the (<b>d1</b>) Northern Hemisphere and (<b>d2</b>) Southern Hemisphere. The pairs are colored by month, and the blue line represents a trend line fitted to the data.</p>
Full article ">Figure 14
<p>Sea ice mapping in the (<b>a</b>) Northern Hemisphere on 18 June 2019 and (<b>b</b>) Southern Hemisphere on 18 June 2019 derived from CSCAT, ASCAT, NSIDC sea ice edge (SIE), and NSIDC sea ice concentration (SIC), respectively.</p>
Full article ">Figure 15
<p>Daily consistency compared to NSIDC for (1) close ice, (2) open ice, and (3) open water over the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere from 1 January 2019 to 31 December 2022, respectively.</p>
Full article ">Figure 16
<p>Monthly mode statistics for CSCAT over the (<b>a</b>) Northern Hemisphere and (<b>b</b>) Southern Hemisphere and for ASCAT over the (<b>c</b>) Northern Hemisphere and (<b>d</b>) Southern Hemisphere, showing sea ice cover differences compared to NSIDC.</p>
Full article ">Figure 17
<p>Comparative analysis of sea ice detection and high-resolution synthetic aperture radar (SAR) images. Comparison between CSCAT-derived sea ice detection results and Sentinel-1 SAR images in the Northern Hemisphere taken on (<b>a</b>) 18 June 2019 and (<b>b</b>) 8 March 2019 and in the Southern Hemisphere on (<b>c</b>) 19 June 2019. The thick red line represents the CSCAT-derived sea ice detection results.</p>
Full article ">
24 pages, 15151 KiB  
Article
Polar Sea Ice Monitoring Using HY-2B Satellite Scatterometer and Scanning Microwave Radiometer Measurements
by Tao Zeng, Lijian Shi, Yingni Shi, Dunwang Lu and Qimao Wang
Remote Sens. 2024, 16(13), 2486; https://doi.org/10.3390/rs16132486 - 6 Jul 2024
Viewed by 944
Abstract
The Ku band microwave scatterometer (SCA) and scanning microwave radiometer (SMR) onboard HaiYang-2B (HY-2B) can simultaneously supply active and passive microwave observations over the polar region. In this paper, a polar ice water discrimination model and Arctic sea-ice-type classification model based on the [...] Read more.
The Ku band microwave scatterometer (SCA) and scanning microwave radiometer (SMR) onboard HaiYang-2B (HY-2B) can simultaneously supply active and passive microwave observations over the polar region. In this paper, a polar ice water discrimination model and Arctic sea-ice-type classification model based on the support vector machine (SVM) method were established and used to produce a daily sea ice extent dataset from 2019 to 2021 with data from SCA and SMR. First, suitable scattering and radiation parameters are chosen as input data for the discriminant model. Then, the sea ice extent was obtained based on the monthly ice water discrimination model, and finally, the ice over the Arctic was classified into multiyear ice (MYI) and first-year ice (FYI). The 3-year ice extent and MYI extent products were consistent with the similar results of the National Snow and Ice Data Center (NSIDC) and Ocean and Sea Ice Satellite Application Facility (OSISAF). Using the OSISAF similar product as validation data, the overall accuracies (OAs) of ice/water discrimination and FYI/MYI discrimination are 99% and 97%, respectively. Compared with the high spatial resolution classification results of the Moderate Resolution Imaging Spectroradiometer (MODIS) and SAR, the OAs of ice/water discrimination and FYI/MYI discrimination are 96% and 86%, respectively. In conclusion, the SAC and SMR of HY-2B have been verified for monitoring polar sea ice, and the sea ice extent and sea-ice-type products are promising for integration into long-term sea ice records. Full article
(This article belongs to the Special Issue Recent Advances in Sea Ice Research Using Satellite Data)
Show Figures

Figure 1

Figure 1
<p>SCA and SMR images over the Arctic region on 26 February 2019. (<b>a</b>) <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> observed using SCA, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">T</mi> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>37</mn> <mi mathvariant="normal">H</mi> </mrow> </msub> </mrow> </semantics></math> observed using SMR, land is shown in grey.</p>
Full article ">Figure 2
<p>Time series of daily histograms of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> (<b>a</b>,<b>c</b>) and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">T</mi> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>37</mn> <mi mathvariant="normal">H</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>,<b>d</b>) over different regions during 2019, x-axis is the day number and y-axis is the value of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">T</mi> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>37</mn> <mi mathvariant="normal">H</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>a</b>,<b>b</b>) are histograms over the hole Arctic region, (<b>c</b>,<b>d</b>) are histograms over the Arctic ice covered area. Each daily histogram in the time series is normalized by maximum observation count, concatenated together and rendered according to the color bar. (<b>a</b>) The small peak at −8 dB in the white box corresponds to MYI and the peak between −15 dB and −20 dB in the red box corresponds to FYI and OW. (<b>b</b>) The peak at approximately 230 K in the white box corresponds to FYI, while the MYI and OW values are between 130 K and 160 K. (<b>c</b>) The small peak near −7 db in the white box corresponds to MYI and the peak between −17 and −19 in the red box corresponds to FYI. (<b>d</b>) The peak near 230 K in the white box corresponds to FYI.</p>
Full article ">Figure 2 Cont.
<p>Time series of daily histograms of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> (<b>a</b>,<b>c</b>) and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">T</mi> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>37</mn> <mi mathvariant="normal">H</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>,<b>d</b>) over different regions during 2019, x-axis is the day number and y-axis is the value of <math display="inline"><semantics> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">T</mi> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>37</mn> <mi mathvariant="normal">H</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>a</b>,<b>b</b>) are histograms over the hole Arctic region, (<b>c</b>,<b>d</b>) are histograms over the Arctic ice covered area. Each daily histogram in the time series is normalized by maximum observation count, concatenated together and rendered according to the color bar. (<b>a</b>) The small peak at −8 dB in the white box corresponds to MYI and the peak between −15 dB and −20 dB in the red box corresponds to FYI and OW. (<b>b</b>) The peak at approximately 230 K in the white box corresponds to FYI, while the MYI and OW values are between 130 K and 160 K. (<b>c</b>) The small peak near −7 db in the white box corresponds to MYI and the peak between −17 and −19 in the red box corresponds to FYI. (<b>d</b>) The peak near 230 K in the white box corresponds to FYI.</p>
Full article ">Figure 3
<p>The correlation matrix of SCA’s five parameters (<b>a</b>) and SMR’s seven parameters (<b>b</b>) over the ocean area on 5 December 2019.</p>
Full article ">Figure 4
<p>The analytical results of the classification distance of SCA’s five parameters (<b>a</b>) and SMR’s seven parameters (<b>b</b>) in 2019 over the Arctic.</p>
Full article ">Figure 5
<p>Different sea ice extent results on 5 March 2019. (<b>a</b>) Result with 5 SCA parameters. (<b>b</b>) Result with 3 SCA parameters. (<b>c</b>) Result with parameters of <math display="inline"><semantics> <mrow> <mfenced close="]" open="["> <mrow> <msubsup> <mi mathvariant="sans-serif">σ</mi> <mrow> <mi>HH</mi> </mrow> <mn>0</mn> </msubsup> <mo>,</mo> <mrow> <mtext> </mtext> <mi>Ratio</mi> </mrow> <mo>,</mo> <msub> <mrow> <mrow> <mtext> </mtext> <mi>STD</mi> </mrow> </mrow> <mi mathvariant="normal">H</mi> </msub> <mo>,</mo> <msub> <mrow> <mrow> <mtext> </mtext> <mi mathvariant="normal">T</mi> </mrow> </mrow> <mrow> <mi mathvariant="normal">b</mi> <mo>,</mo> <mn>18.7</mn> <mi mathvariant="normal">V</mi> </mrow> </msub> <mo>,</mo> <mrow> <mtext> </mtext> <mi>PR</mi> </mrow> </mrow> </mfenced> </mrow> </semantics></math>. (<b>d</b>) Product of OSISAF. The results of (<b>a</b>,<b>b</b>) have some incorrect identifications of ice and water over the area within the red frame.</p>
Full article ">Figure 6
<p>Variation in overall accuracy (dotted line) and Kappa coefficient (solid line) for the Arctic (<b>top</b>) and Antarctic (<b>bottom</b>) from 2019 to 2021: The blue line represents the result of 5 SCA parameters and the red line represents the result of selected parameters of SCA and SMR.</p>
Full article ">Figure 7
<p>Time series of Arctic and Antarctic daily sea ice extents from 2019 to 2021 based on different data.</p>
Full article ">Figure 8
<p>Time series of IIEE between HY-2B and other products over the Arctic and Antarctic from 2019 to 2021.</p>
Full article ">Figure 9
<p>Distribution of Arctic sea ice types on the 15th of each month from January to April (<b>the first row</b>) and from October to December (<b>the second row</b>) in 2019.</p>
Full article ">Figure 10
<p>Time series of Arctic MYI extent derived from HY-2B and OSISAF from 2019 to 2021.</p>
Full article ">Figure 11
<p>Time series of assessment parameters of the ice water discrimination results in the Arctic from 2019 to 2021 for (<b>a</b>) OA and Kappa coefficient, (<b>b</b>) UA of open water and ice, and (<b>c</b>) PA of open water and ice.</p>
Full article ">Figure 12
<p>Time series of assessment parameters of the ice water discrimination results in the Antarctic from 2019 to 2021 for (<b>a</b>) OA and Kappa coefficient, (<b>b</b>) UA of open water and ice, respectively, and (<b>c</b>) PA of open water and ice.</p>
Full article ">Figure 12 Cont.
<p>Time series of assessment parameters of the ice water discrimination results in the Antarctic from 2019 to 2021 for (<b>a</b>) OA and Kappa coefficient, (<b>b</b>) UA of open water and ice, respectively, and (<b>c</b>) PA of open water and ice.</p>
Full article ">Figure 13
<p>Time series of assessment parameters of the ice-type discrimination results in the Arctic from 2019 to 2021 for (<b>a</b>) OA and Kappa coefficient, (<b>b</b>) UA of OW, FYI and MYI, and (<b>c</b>) PA of OW, FYI and MYI.</p>
Full article ">Figure 13 Cont.
<p>Time series of assessment parameters of the ice-type discrimination results in the Arctic from 2019 to 2021 for (<b>a</b>) OA and Kappa coefficient, (<b>b</b>) UA of OW, FYI and MYI, and (<b>c</b>) PA of OW, FYI and MYI.</p>
Full article ">Figure 14
<p>MODIS image and sea ice distribution near Canadian islands on 1 July 2019. (<b>a</b>) MODIS image, the coastline is shown in red. (<b>b</b>) Sea ice extent with a spatial resolution of 25 km based on MODIS image. (<b>c</b>) Sea ice extent result obtained using HY-2B in this paper.</p>
Full article ">Figure 15
<p>MODIS image and sea ice distribution over Ross Sea on 2 January 2020. (<b>a</b>) MODIS image, the coastline is shown in red. (<b>b</b>) Sea ice extent with a spatial resolution of 25 km based on MODIS image. (<b>c</b>) Sea ice extent result obtained using HY-2B in this paper.</p>
Full article ">Figure 16
<p>SAR image and FYI and MYI distribution near Canadian Archipelago on 18 January 2019. (<b>a</b>) SAR image, the coastline is shown in red. (<b>b</b>) Sea ice type with a spatial resolution of 25 km based on the SAR image. (<b>c</b>) Sea ice type result obtained using HY-2B in this paper.</p>
Full article ">Figure 17
<p>SAR image and FYI and MYI distribution near Canadian Archipelago on 23 December 2019. (<b>a</b>) SAR image, the coastline is shown in red. (<b>b</b>) Sea ice type with a spatial resolution of 25 km based on the SAR image. (<b>c</b>) Sea ice type result obtained using HY-2B in this paper.</p>
Full article ">
16 pages, 16286 KiB  
Article
AMELX Mutations and Genotype–Phenotype Correlation in X-Linked Amelogenesis Imperfecta
by Shih-Kai Wang, Hong Zhang, Hua-Chieh Lin, Yin-Lin Wang, Shu-Chun Lin, Figen Seymen, Mine Koruyucu, James P. Simmer and Jan C.-C. Hu
Int. J. Mol. Sci. 2024, 25(11), 6132; https://doi.org/10.3390/ijms25116132 - 1 Jun 2024
Viewed by 990
Abstract
AMELX mutations cause X-linked amelogenesis imperfecta (AI), known as AI types IE, IIB, and IIC in Witkop’s classification, characterized by hypoplastic (reduced thickness) and/or hypomaturation (reduced hardness) enamel defects. In this study, we conducted whole exome analyses to unravel the disease-causing mutations for [...] Read more.
AMELX mutations cause X-linked amelogenesis imperfecta (AI), known as AI types IE, IIB, and IIC in Witkop’s classification, characterized by hypoplastic (reduced thickness) and/or hypomaturation (reduced hardness) enamel defects. In this study, we conducted whole exome analyses to unravel the disease-causing mutations for six AI families. Splicing assays, immunoblotting, and quantitative RT-PCR were conducted to investigate the molecular and cellular effects of the mutations. Four AMELX pathogenic variants (NM_182680.1:c.2T>C; c.29T>C; c.77del; c.145-1G>A) and a whole gene deletion (NG_012494.2:g.307534_403773del) were identified. The affected individuals exhibited enamel malformations, ranging from thin, poorly mineralized enamel with a “snow-capped” appearance to severe hypoplastic defects with minimal enamel. The c.145-1G>A mutation caused a -1 frameshift (NP_001133.1:p.Val35Cysfs*5). Overexpression of c.2T>C and c.29T>C AMELX demonstrated that mutant amelogenin proteins failed to be secreted, causing elevated endoplasmic reticulum stress and potential cell apoptosis. This study reveals a genotype–phenotype relationship for AMELX-associated AI: While amorphic mutations, including large deletions and 5′ truncations, of AMELX cause hypoplastic-hypomaturation enamel with snow-capped teeth (AI types IIB and IIC) due to a complete loss of gene function, neomorphic variants, including signal peptide defects and 3′ truncations, lead to severe hypoplastic/aplastic enamel (AI type IE) probably caused by “toxic” cellular effects of the mutant proteins. Full article
(This article belongs to the Special Issue Molecular Metabolism of Ameloblasts in Tooth Development)
Show Figures

Figure 1

Figure 1
<p>Family 1 with <span class="html-italic">AMELX</span> c.2T&gt;C mutation. (<b>A</b>) The family pedigree suggests a dominant pattern of disease inheritance. The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The proband (III:3) at age 35 had permanent teeth that appeared microdontic and spaced due to severe hypoplastic/aplastic enamel. Tooth surfaces were generally smooth, and dental attrition evident. (<b>C</b>) The bitewing radiograph of the proband confirmed the clinical finding of virtually no enamel covering the teeth. (<b>D</b>) The DNA sequencing chromatograms identified a single nucleotide transition (g.375912A&gt;G) that abolished the translation initiation codon of <span class="html-italic">AMELX</span> (c.2T&gt;C, p.Met1?). While the proband was heterozygous for the mutation, her two sons (IV:2, IV:3) were both hemizygotes.</p>
Full article ">Figure 2
<p>Family 2 with <span class="html-italic">AMELX</span> c.29T&gt;C mutation. (<b>A</b>) The pedigree shows a nuclear family in which the proband (II:1, age 10) and her father (I:1) were both affected. The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The proband had a mixed dentition that exhibited generalized enamel hypoplasia. The surface of the teeth was not particularly rough but presented with black stains. Her maxillary and mandibular dental arches were both markedly narrow. The panoramic radiograph revealed that all her teeth, including unerupted ones, had an extremely thin enamel layer. (<b>C</b>) The DNA sequencing chromatograms indicate that the proband and her father were heterozygous and hemizygous, respectively, for the <span class="html-italic">AMELX</span> missense mutation (g.375885A&gt;G, c.29T&gt;C, p.Leu10Pro). The unaffected mother (I:2) and younger brother (II:2) did not carry the defect.</p>
Full article ">Figure 3
<p>Family 3 with <span class="html-italic">AMELX</span> c.77del mutation. (<b>A</b>) The pedigree indicates that the proband (II:4) was the only affected individual in the family, although the phenotypes of other members could not be confirmed. The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The photographs of the proband at age 20 showed dental crowns with a yellowish-white appearance and lacking normal translucency. The chalky-white discoloration involved more than the incisal or occlusal half of the tooth crowns without a temporal distribution. Dental attrition was evident, particularly over incisal edges and cusp tips. His panorex showed generally thin enamel with reduced radiographic contrast with dentin on all teeth. (<b>C</b>) The DNA sequencing chromatogram shows a single nucleotide deletion (g.373903del, c.77del) that caused a frameshift and premature termination (p.Pro26Leufs*23) of the proband’s AMELX protein. The wild-type (WT) chromatogram was generated from an unrelated healthy individual.</p>
Full article ">Figure 4
<p>Family 4 with <span class="html-italic">AMELX</span> c.77del mutation. (<b>A</b>) The pedigree shows a nuclear family in which the proband (II:1) was affected by enamel defects but was otherwise healthy. The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The photographs of the proband revealed a dental phenotype of snow-capped teeth similar to that of the Family 3 proband but with more severity. Consistently, the panoramic radiograph indicated a combined hypoplastic and hypomaturation malformation. (<b>C</b>) The DNA sequencing chromatogram exhibits the same single nucleotide deletion (g.373903del, c.77del) found in Family 3. While the father (I:1) carried the wild-type <span class="html-italic">AMELX</span> allele, the proband (II:1) and the mother (I:2) were hemizygous and heterozygous for the mutation, respectively.</p>
Full article ">Figure 5
<p>Family 5 with <span class="html-italic">AMELX</span> c.145-1G&gt;A mutation. (<b>A</b>) The pedigree shows a nuclear family in which all the children were affected and inherited enamel malformations from the mother (I:2). The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The proband (II:2, age 10) was at the mixed dentition stage and had enamel defects similar to those of the Family 3 proband. Clinically, the teeth were yellow-white discolored, which resembled snow-capped teeth and suggested a thin and hypomineralized enamel layer. The radiographs confirmed that the enamel was both hypoplastic and hypomature. (<b>C</b>) The DNA sequencing chromatogram from the mother showed a G-to-A transition at the splice acceptor site of Intron 4 (g.372467C&gt;T, c.145-1G&gt;A) in one of her <span class="html-italic">AMELX</span> genes. While all the (male) children were hemizygous for this variant, the father (I:1) was not.</p>
Full article ">Figure 6
<p>Family 6 with a 96240-bp deletion involving <span class="html-italic">AMELX</span>. (<b>A</b>) The family pedigree exhibits a consanguineous family in with the two affected individuals were from a second-cousin marriage. The blue arrow indicates the proband, and the asterisks annotate subjects recruited for genetic testing. (<b>B</b>) The proband’s (V:3, age 14) teeth appeared rough and chalky white with a yellowish hue. Enamel chipping was evident on maxillary central incisors. Radiographically, the enamel was generally thin and of reduced radiopacity. (<b>C</b>) The proband’s brother (V:1, age 20) had a similar enamel phenotype to that of the proband, constituting hypoplastic and hypomaturation AI. (<b>D</b>) The DNA sequencing chromatogram from the proband shows a 96240-bp deletion within the Intron 1 of <span class="html-italic">ARHGAP6</span> (NG_012494.2:g.307534_403773del) that removes the whole <span class="html-italic">AMELX</span> gene, NG_012494.2 (NM_182681.1):c.-26548_*62560del. While the two affected males both carried this deletion and had no <span class="html-italic">AMELX</span> gene, their mother (IV:12) was heterozygous to the mutation and had no overt enamel defects.</p>
Full article ">Figure 7
<p>Molecular characterization of <span class="html-italic">AMELX</span> mutations. (<b>A</b>) Agarose gel electrophoresis exhibited nearly comparable RT-PCR amplicons of 732 and 731 bps from the wild type (WT) and the c.145-1G&gt;A (Mut) minigenes, respectively. The mutant transcript used a new splice acceptor site one nucleotide downstream of the original one, resulting in loss of the first nucleotide of Exon 5 during splicing, which shifted the reading frame and generated a stop codon within the same exon. (<b>B</b>) Alignment of the amino acid sequence of human AMELX (NP_001133.1) signal peptide with those of its vertebral orthologs from pig (NP_999071.1), mouse (NP_001075447.1), platypus (XP_001515115.2), anole (XP_003228746.1), Xenopus (NP_001107153.1), lungfish (XP_043928489.1), and coelacanth (XP_005998289.1) as well as other human P/Q-rich SCPPs, including ENAM (NP_114095.2), AMBN (NP_057603.1), AMTN (NP_997722.1), ODAM (NP_060325.3), and SCPPPQ1 (NP_001392157.1). The Leu10 is in bold and highlighted in yellow. The green underline indicates the h-region of the signal peptide. (<b>C</b>) Immunoblotting revealed that the mutant amelogenin proteins, p.Met1? and p.Leu10Pro, could only be detected in cell lysates, while the wild-type AMELX was mainly found in the culture medium. (<b>D</b>) Anti-phosphoserine (α-pSer) immunoblotting of immunoprecipitated AMELXs only showed positive signals in the wild type. Key: WT, wild type; p.M1?, p.Met1?; p.L10P, p.Leu10Pro; EV, empty vector; rM179, recombinant mouse amelogenin protein; IP, immunoprecipitation.</p>
Full article ">Figure 8
<p>Gene expression under <span class="html-italic">AMELX</span> overexpression. Quantitative reverse transcription PCRs (qRT-PCRs) were performed for <span class="html-italic">AMELX</span> and ER stress related genes, including <span class="html-italic">HSPA5</span>, <span class="html-italic">DDIT3</span>, <span class="html-italic">XBP1s</span>, and <span class="html-italic">ATF6</span>. Expression of <span class="html-italic">TNFRSF10B</span>, an apoptotic gene activated by unmitigated ER stress, was also assessed. Key: WT, wild type; p.M1?, p.Met1?; p.L10P, p.Leu10Pro; EV, empty vector; **, <span class="html-italic">p</span>-value &lt; 0.05; ***, <span class="html-italic">p</span>-value &lt; 0.01.</p>
Full article ">
14 pages, 4681 KiB  
Article
Monitoring Snow Cover in Typical Forested Areas Using a Multi-Spectral Feature Fusion Approach
by Yunlong Wang and Jianshun Wang
Atmosphere 2024, 15(4), 513; https://doi.org/10.3390/atmos15040513 - 22 Apr 2024
Cited by 2 | Viewed by 1144
Abstract
Accurate snow cover monitoring is greatly significant for research on the hydrology model and regional climate variation, especially in Northeast China where forests cover almost forty percent of the total area. However, effectively monitoring snow cover under the forest canopy is still challenging [...] Read more.
Accurate snow cover monitoring is greatly significant for research on the hydrology model and regional climate variation, especially in Northeast China where forests cover almost forty percent of the total area. However, effectively monitoring snow cover under the forest canopy is still challenging with either in situ or remote sensing observations. The global SNOWMAP algorithm pertinent to the fixed normalized difference snow index (NDSI) threshold is, therefore, no longer applicable in a typical forested region of Northeast China. In order to achieve the goal of improving the accuracy of monitoring snow cover in areas with forest, utilizing MOD09GA and MOD13A1 products, a new approach of snow mapping was developed in this study, and it exploits the fusion and coupling of spectral features by integrating and analyzing the normalized difference forest snow index (NDFSI), the normalized difference vegetation index (NDVI), and the NDSI index. Then, Landsat 8 OLI images of high resolution were used to evaluate snow cover mapping precision. The experimental results indicated that the NDFSI index combined with the NDVI index showed great potential for extracting the snow cover distribution in forested regions. Compared with the snow distribution obtained from the Landsat 8 images, the average bias and FAR (false alarm ratio) values of snow cover mapping obtained by this algorithm were 1.23 and 13.54%, which were reduced by 1.98 and 29.36%, respectively. The overall accuracy of 81.31% was reached, which is improved by 20.19%. Thus, the snow classification scheme combining multiple spectral features from MODIS data works effectively in improving the precision of automatic snow cover mapping in typical forested areas of Northeast China, which provides essential support and significant perspectives for the next step of establishing a runoff model and rationally regulating forest water resources. Full article
(This article belongs to the Special Issue Precipitation Monitoring and Databases)
Show Figures

Figure 1

Figure 1
<p>Land cover type distribution in Northeast China through Landsat 8 OLI images.</p>
Full article ">Figure 2
<p>The distribution of NDFSI-NDVI scatter plots in the snow-free and snow-covered forest areas.</p>
Full article ">Figure 3
<p>The optimal threshold distribution for distinguishing between snow-free and snow-covered forest areas.</p>
Full article ">Figure 4
<p>Comparison of snow cover estimated based on the Landsat 8 OLI (<b>a</b>–<b>d</b>), cloud-free snow cover distribution (<b>e</b>–<b>h</b>), and the algorithm in this study (<b>i</b>–<b>l</b>).</p>
Full article ">
21 pages, 14485 KiB  
Article
Analysis of Dual-Polarimetric Radar Observations of Precipitation Phase during Snowstorm Events in Jiangsu Province, China
by Lei Wang, Yi Wang, Mei Liu, Wei Chen and Chiqin Li
Atmosphere 2024, 15(3), 321; https://doi.org/10.3390/atmos15030321 - 4 Mar 2024
Viewed by 912
Abstract
Based on ground observed data, S-band dual-polarization radar data, and ERA-5 reanalysis data, the statistical characteristics of polarimetric parameters and the application of melting layer (ML) and hydrometeor classification (HCL) products during eight snowstorm events in Jiangsu Province from 2020 to 2022 were [...] Read more.
Based on ground observed data, S-band dual-polarization radar data, and ERA-5 reanalysis data, the statistical characteristics of polarimetric parameters and the application of melting layer (ML) and hydrometeor classification (HCL) products during eight snowstorm events in Jiangsu Province from 2020 to 2022 were investigated. A heavy snowstorm that went through different phases of rain, sleet, and pure snow and that occurred on 29 December 2020 was also analyzed as a typical example. The results showed the following: During the phase transition between rain and snow in the Jiangsu region, the basic reflectivity factor ZH ≥ 27 dBZ, the zero-order lag correlation coefficient CC ≤ 0.93, and the differential reflectivity ZDR ≥ 1.0 dB were important indicators for judging the melting layer while the specific differential phase KDP changed slightly. The snowstorm event was well observed and recorded by the Yancheng dual-polarimetric radar, whose low value area of CC coincided mostly with the melting layer. The ML products and HCL products based on fuzzy-logic hydrometeor classification algorithms can help identify the melting layer and the properties of precipitation particles. ML products are more reliable when the melting layer is high and can better show the trends of melting layer decline. They can certainly serve as a reference for detecting and judging precipitation phase changes in winter in Jiangsu Province. Full article
(This article belongs to the Special Issue Data Assimilation for Predicting Hurricane, Typhoon and Storm)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geographical distribution of the studied area in East China region. (<b>b</b>) Location of Jiangsu Province and distribution of observation stations (intervals between radial circles are 50 km and 230 km).</p>
Full article ">Figure 2
<p>Sketch map of melting layer (ML) products of dual-polarization radar.</p>
Full article ">Figure 3
<p>Statistical analysis of (<b>a</b>) horizontal basic reflectivity factor (Z<sub>H</sub>), differential reflectivity (Z<sub>DR</sub>), (<b>b</b>) zero-order lag correlation coefficient (CC), and specific differential phase (K<sub>DP</sub>) of the melting layer during the rain and snow transition from the nearest radar. Values at the tops/bottoms of the bars represent the minimum/maximum values of the corresponding radar parameters.</p>
Full article ">Figure 4
<p>(<b>a</b>) Distribution of maximum snow depth (unit: cm) during the period from 00 LST 28 December to 00 LST 30 December 2020, (<b>b</b>) phase of precipitation particles at 08 LST on 29 December 2020 from ground-based observation, and (<b>c</b>) variations in temperature (unit: °C) at ground level and in upper air at Sheyann station from 20 LST on 27 December to 20 LST on 30 December 2020 in Jiangsu Province.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>e</b>) PPI charts of horizontal basic reflectivity factor (Z<sub>H</sub>), (<b>b</b>,<b>f</b>) zero-order lag correlation coefficient (CC), (<b>c</b>,<b>g</b>) differential reflectivity (Z<sub>DR</sub>), and (<b>d</b>,<b>h</b>) specific differential phase (K<sub>DP</sub>) at 1.5° elevation angle from Yancheng dual-polarization Doppler radar at 7:58 LST and 9:29 LST on 29 December 2020. (Box A and B show the areas with high reflectivity factor values. AB and CD line are marked to show the vertical profile of the reflectivity factor.)</p>
Full article ">Figure 6
<p>(<b>a</b>) Reflectivity factor, (<b>b</b>) CC, and (<b>c</b>) ZDR along the AB line in <a href="#atmosphere-15-00321-f005" class="html-fig">Figure 5</a>e. (<b>d</b>) Reflectivity factor, (<b>e</b>) CC, and (<b>f</b>) ZDR along the CD line in <a href="#atmosphere-15-00321-f005" class="html-fig">Figure 5</a>e at 9:29 LST on 29 December 2020.</p>
Full article ">Figure 7
<p>Superposition of melting layer (ML) product (contours) and CC (color-filled areas) at 08:26 LST at (<b>a</b>) 1.5° elevation angle and (<b>b</b>) 2.4° elevation angle from Yancheng dual-polarization radar and PPI charts of ML at 1.5° elevation angle at (<b>c</b>) 09:29 LST and (<b>d</b>) 15:01 LST on 29 December 2020.</p>
Full article ">Figure 8
<p>PPI charts of HCL at 1.5° elevation angle from Yancheng dual-polarization radar at (<b>a</b>) 07:58 LST and (<b>b</b>) 09:29 LST on 29 December 2020. (Box A and B show the areas with high reflectivity factor values.)</p>
Full article ">
21 pages, 4278 KiB  
Article
Performance of the Thies Clima 3D Stereo Disdrometer: Evaluation during Rain and Snow Events
by Sabina Angeloni, Elisa Adirosi, Alessandro Bracci, Mario Montopoli and Luca Baldini
Sensors 2024, 24(5), 1562; https://doi.org/10.3390/s24051562 - 28 Feb 2024
Viewed by 1008
Abstract
Imaging disdrometers are widely used in field campaigns to provide information on the shape of hydrometeors, together with the diameter and the fall velocity, which can be used to derive information on the shape–size relations of hydrometeors. However, due to their higher price [...] Read more.
Imaging disdrometers are widely used in field campaigns to provide information on the shape of hydrometeors, together with the diameter and the fall velocity, which can be used to derive information on the shape–size relations of hydrometeors. However, due to their higher price compared to laser disdrometers, their use is limited to scientific research purposes. The 3D stereo (3DS) is a commercial imaging disdrometer recently made available by Thies Clima and on which there are currently no scientific studies in the literature. The most innovative feature of the 3DS is its ability in capturing images of the particles passing through the measurement volume, crucial to provide an accurate classification of hydrometeors based on information about their shape, especially in the case of solid precipitation. In this paper. the performance of the new device is analyzed by comparing 3DS with the Laser Precipitation Monitor (LPM) from the same manufacturer, which is a known laser disdrometer used in many research works. The data used in this paper were obtained from measurements of the two instruments carried out at the Casale Calore site in L’Aquila during the CORE-LAQ (Combined Observations of Radar Experiments in L’Aquila) campaign. The objective of the comparison analysis is to analyze the differences between the two disdrometers in terms of hydrometeor classification, number and falling speed of particles, precipitation intensity, and total cumulative precipitation on an event basis. As regards the classification of precipitation, the two instruments are in excellent agreement in identifying rain and snow; greater differences are observed in the case of particles in mixed phase (rain and snow) or frozen phase (hail). Due to the different measurement area of the two disdrometers, the 3DS generally detects more particles than the LPM. The performance differences also depend on the size of the hydrometeors and are more significant in the case of small particles, i.e., D < 1 mm. In the case of rain events, the two instruments are in agreement with respect to the terminal velocity in still air predicted by the Gunn and Kinzer model for drops with a diameter of less than 3 mm, while, for larger particles, terminal velocity is underestimated by both the disdrometers. The agreement between the two instruments in terms of total cumulative precipitation per event is very good. Regarding the 3DS ability to capture images of hydrometeors, the raw data provide, each minute, from one to four images of single particles and information on their size and type. Their number and coarse resolution make them suitable to support only qualitative analysis of the shape of precipitating particles. Full article
(This article belongs to the Special Issue Atmospheric Precipitation Sensors)
Show Figures

Figure 1

Figure 1
<p>Deployment of LPM and 3DS at the Casale Calore site with other instruments used in the CORE-LAQ campaign (adapted from [<a href="#B27-sensors-24-01562" class="html-bibr">27</a>], courtesy of the authors).</p>
Full article ">Figure 2
<p>Barplot of the comparison between the classifications of the two disdrometers in percentage. Each bar relates to a 3DS class.</p>
Full article ">Figure 3
<p>Spectrograms of (<b>a</b>,<b>c</b>) rainy minutes; (<b>b</b>,<b>d</b>) snowy minutes. Top line refers to 3DS, the bottom line to LPM. Colors represent the number of particles detected for each bin. Black curves in rain spectrograms represent the reference fall velocity provided by Formula (<a href="#FD3-sensors-24-01562" class="html-disp-formula">3</a>).</p>
Full article ">Figure 4
<p>Statistics on terminal fall velocity with respect to diameter class (median value, first and third quartiles, and maximum and minimum) for (<b>a</b>,<b>d</b>) rain events; (<b>b</b>,<b>e</b>) rain events with filter; (<b>c</b>,<b>f</b>) snow events. Top line refers to 3DS, the bottom line to LPM. Black curves represent the theoretical speeds provided by Formula (<a href="#FD3-sensors-24-01562" class="html-disp-formula">3</a>) for raindrops, LH for snow particles.</p>
Full article ">Figure 5
<p>Scatterplot between the number of particles detected by the LPM (<span class="html-italic">x</span>-axis) and by the 3DS (<span class="html-italic">y</span>-axis) during rain events for (<b>a</b>,<b>e</b>) all raindrops; (<b>b</b>,<b>f</b>) small raindrops (D &lt; 0.5 mm); (<b>c</b>,<b>g</b>) medium raindrops (0.5 mm &lt; D &lt; 1.5 mm); (<b>d</b>,<b>h</b>) large raindrops (D &gt; 1.5 mm). Top panels refer to raw measurements, the bottom panels to filtered measurements.</p>
Full article ">Figure 6
<p>Barplot of the number of particles (<span class="html-italic">N</span>) detected by LPM and 3DS during rain events as a function of (<b>a</b>) diameter; (<b>b</b>) terminal fall velocity. Black lines mark the filtered values.</p>
Full article ">Figure 7
<p>Scatterplot of rain rate detected by LPM (<span class="html-italic">x</span>-axis) and by 3DS (<span class="html-italic">y</span>-axis) during the common minutes of rain events from 13 December 2022 to 19 July 2023.</p>
Full article ">Figure 8
<p>Scatterplot between event-based rainfall accumulation via LPM (<span class="html-italic">x</span>-axis) and via 3DS (<span class="html-italic">y</span>-axis).</p>
Full article ">Figure 9
<p>Images of particles obtained by 3DS disdrometer during snow events, classified as ‘snow’. The diameter (in mm) is provided at the bottom of each image.</p>
Full article ">Figure 10
<p>Sequence of images referring to 2023-01-20 19:00, classified as ‘rain with snow’ by the 3DS.</p>
Full article ">Figure 11
<p>Images of particles obtained by 3DS disdrometer during hail minutes, classified as ‘hail’. The diameter (in mm) is provided at the bottom of each image.</p>
Full article ">Figure 12
<p>Images of particles obtained by 3DS disdrometer during rain events, classified as ‘rain’. The diameter (in mm) is provided at the bottom of each image.</p>
Full article ">Figure 13
<p>Images of particles obtained by 3DS disdrometer during minutes classified as ‘rain/drizzle with snow’. The diameter (in mm) is provided at the bottom of each image.</p>
Full article ">
10 pages, 246 KiB  
Review
Visual Snow Syndrome in Patient with Migraine: Case Report and Literature Review
by Justyna Chojdak-Łukasiewicz and Edyta Dziadkowiak
J. Clin. Med. 2024, 13(5), 1373; https://doi.org/10.3390/jcm13051373 - 28 Feb 2024
Cited by 1 | Viewed by 2273
Abstract
Visual snow syndrome (VSS) is a rarely diagnosed neurological phenomenon. It is a visual disorder characterised by the presence of numerous white, black, or translucent dots in the visual field, resembling the ‘snow’ of an analogue TV set experiencing reception interference. According to [...] Read more.
Visual snow syndrome (VSS) is a rarely diagnosed neurological phenomenon. It is a visual disorder characterised by the presence of numerous white, black, or translucent dots in the visual field, resembling the ‘snow’ of an analogue TV set experiencing reception interference. According to The International Classification of Headache Disorders, 3rd edition, visual snow is defined as a pattern of continuous small dots across the visual field lasting >3 months and accompanied by at least two of the following four additional symptoms: palinopsia, increased entoptic phenomena, photophobia, and nyctalopia. These complaints are not consistent with a typical migraine with visual aura and cannot be better explained by another disorder. The authors present the case of a 39-year-old woman who was diagnosed with VSS. The symptoms appeared after a migraine attack and had not alleviated. The patient reported a sensation of constant ‘TV screen snow’. A neurological examination found no signs of focal damage to the nervous system. The results of the ophthalmological examination, MRI of the brain with contrast, MRI of the eye sockets, and EEG were normal. VSS is a phenomenon that is still not fully understood, different from migraine aura and associated with a number of additional symptoms. VSS is very difficult to treat. In this case, a lot of drugs were used without improvement. Further research must be conducted to determine the best treatment options for these patients. Full article
(This article belongs to the Section Clinical Neurology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
Back to TopTop