[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (184)

Search Parameters:
Keywords = altitude dependence modelling

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 18893 KiB  
Article
High-Precision Tea Plantation Mapping with Multi-Source Remote Sensing and Deep Learning
by Yicheng Zhou, Lingbo Yang, Lin Yuan, Xin Li, Yihu Mao, Jiancong Dong, Zhenyu Lin and Xianfeng Zhou
Agronomy 2024, 14(12), 2986; https://doi.org/10.3390/agronomy14122986 - 15 Dec 2024
Viewed by 471
Abstract
Accurate mapping of tea plantations is crucial for agricultural management and economic planning, yet it poses a significant challenge due to the complex and variable nature of tea cultivation landscapes. This study presents a high-precision approach to mapping tea plantations in Anji County, [...] Read more.
Accurate mapping of tea plantations is crucial for agricultural management and economic planning, yet it poses a significant challenge due to the complex and variable nature of tea cultivation landscapes. This study presents a high-precision approach to mapping tea plantations in Anji County, Zhejiang Province, China, utilizing multi-source remote sensing data and advanced deep learning models. We employed a combination of Sentinel-2 optical imagery, Sentinel-1 synthetic aperture radar imagery, and digital elevation models to capture the rich spatial, spectral, and temporal characteristics of tea plantations. Three deep learning models, namely U-Net, SE-UNet, and Swin-UNet, were constructed and trained for the semantic segmentation of tea plantations. Cross-validation and point-based accuracy assessment methods were used to evaluate the performance of the models. The results demonstrated that the Swin-UNet model, a transformer-based approach capturing long-range dependencies and global context for superior feature extraction, outperformed the others, achieving an overall accuracy of 0.993 and an F1-score of 0.977 when using multi-temporal Sentinel-2 data. The integration of Sentinel-1 data with optical data slightly improved the classification accuracy, particularly in areas affected by cloud cover, highlighting the complementary nature of Sentinel-1 imagery for all-weather monitoring. The study also analyzed the influence of terrain factors, such as elevation, slope, and aspect, on the accuracy of tea plantation mapping. It was found that tea plantations at higher altitudes or on north-facing slopes exhibited higher classification accuracy, and that accuracy improves with increasing slope, likely due to simpler land cover types and tea’s preference for shade. The findings of this research not only provide valuable insights into the precision mapping of tea plantations but also contribute to the broader application of deep learning in remote sensing for agricultural monitoring. Full article
(This article belongs to the Special Issue Remote Sensing in Smart Agriculture)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographical location and elevation of the study area. The background is the digital elevation model (DEM) of the Anji area derived from NASA’s Shuttle Radar Topography Mission (SRTM). (<b>a</b>) Zhejiang Province; (<b>b</b>) Huzhou City; (<b>c</b>) Anji County.</p>
Full article ">Figure 2
<p>Anji Sentinel-2 satellite imagery (<b>a</b>) and its zoomed-in section (<b>b</b>). The images were acquired on 3 June 2019, using bands B8 (near-infrared), B4 (red), and B3 (green) for false-color composite, representing red, green, and blue, respectively.</p>
Full article ">Figure 3
<p>The acquisition dates of Sentinel-2 (S2) and Sentinel-1 (S1) satellite images in the study area, as well as the time series NDVI curves of tea plantations (<b>c</b>). The time series NDVI curves of tea plantations were derived from Sentinel-2 images of tea gardens within the study area that were not obscured by clouds in 2019. Blue area represents the period of NDVI decrease and recovery caused by intensive pruning of Anji white tea plants. (<b>a</b>,<b>b</b>) were taken on-site at tea gardens during the period indicated by the arrows.</p>
Full article ">Figure 4
<p>Anji Sentinel-1 satellite imagery (<b>a</b>) and its zoomed-in section (<b>b</b>). The image was synthesized through false-color composite using Sentinel-1 VV imagery from 1 May, 1 June, and 1 October 2019, as red, green, and blue channels, respectively.</p>
Full article ">Figure 5
<p>The drone imagery (<b>a</b>) and on-site photographs (<b>b</b>–<b>d</b>) of the tea garden in Anji. (<b>b</b>) illustrates the tea garden in its pre-pruning state, whereas (<b>c</b>,<b>d</b>) depict the post-pruning condition of the tea garden.</p>
Full article ">Figure 6
<p>Overall technical flowchart.</p>
Full article ">Figure 7
<p>Sample points and data distribution.</p>
Full article ">Figure 8
<p>Diagram of the UNet model structure used in this study.</p>
Full article ">Figure 9
<p>Diagram of the SE module structure.</p>
Full article ">Figure 10
<p>Diagram of the SE-UNet model structure.</p>
Full article ">Figure 11
<p>The diagram of the Swin transformer block module.</p>
Full article ">Figure 12
<p>The structural diagram of the Swin-UNet model.</p>
Full article ">Figure 13
<p>Tea classification results based on time-series Sentinel-1 dataset. (<b>a</b>) Classification results based on the UNet model. (<b>b</b>) Classification results based on the SE-UNet model. (<b>c</b>) Classification results based on the SWIN-UNet model. (<b>d</b>) Local classification result of SWIN-UNet model.</p>
Full article ">Figure 14
<p>Tea classification results based on time-series Sentinel-2 dataset. (<b>a</b>) Classification results based on the UNet model. (<b>b</b>) Classification results based on the SE-UNet model. (<b>c</b>) Classification results based on the SWIN-UNet model. (<b>d</b>) Local classification result of SWIN-UNet model.</p>
Full article ">Figure 15
<p>Magnified comparative illustration of tea garden classification results based on various deep learning methods and time-series Sentinel-2 imagery. The blue areas in the figure indicate regions where tea gardens were erroneously omitted. The yellow areas indicate regions where non-tea garden pixels were misclassified as tea garden pixels.</p>
Full article ">Figure 16
<p>Tea classification results based on Sentinel-1 + Sentinel-2 dataset and deep learning models.</p>
Full article ">Figure 17
<p>Tea plantation mapping results in cloud-affected areas based on different image combinations. (<b>a</b>) UAV image, (<b>b</b>) Sentinel-2 image, (<b>c</b>) Sentinel-1 derived tea plantation map, (<b>d</b>) Sentinel-2 derived tea plantation map and (<b>e</b>) tea plantation mapping result based on combined Sentinel-1 and Sentinel-2 data.</p>
Full article ">Figure 18
<p>Relationship between distribution of Anji tea gardens and altitudes, aspects, and slopes. (<b>a</b>) Percentage of tea gardens distributed at different altitudes; (<b>b</b>) percentage of tea gardens distributed at different aspects; (<b>c</b>) percentage of tea gardens distributed at different slopes.</p>
Full article ">Figure 19
<p>Tea plantation accuracy under different terrain conditions. (<b>a</b>–<b>c</b>) show the impact of elevation on the F1-score, precision, and recall of tea plantation classification, respectively. (<b>d</b>–<b>f</b>) depict the effect of aspect on the F1-score, precision, and recall of tea plantation classification. (<b>g</b>–<b>i</b>) illustrate the influence of slope on the F1-score, precision, and recall of tea plantation classification.</p>
Full article ">
28 pages, 18069 KiB  
Article
An AI-Based Deep Learning with K-Mean Approach for Enhancing Altitude Estimation Accuracy in Unmanned Aerial Vehicles
by Prot Piyakawanich and Pattarapong Phasukkit
Drones 2024, 8(12), 718; https://doi.org/10.3390/drones8120718 - 29 Nov 2024
Viewed by 497
Abstract
In the rapidly evolving domain of Unmanned Aerial Vehicles (UAVs), precise altitude estimation remains a significant challenge, particularly for lightweight UAVs. This research presents an innovative approach to enhance altitude estimation accuracy for UAVs weighing under 2 kg without cameras, utilizing advanced AI [...] Read more.
In the rapidly evolving domain of Unmanned Aerial Vehicles (UAVs), precise altitude estimation remains a significant challenge, particularly for lightweight UAVs. This research presents an innovative approach to enhance altitude estimation accuracy for UAVs weighing under 2 kg without cameras, utilizing advanced AI Deep Learning algorithms. The primary novelty of this study lies in its unique integration of unsupervised and supervised learning techniques. By synergistically combining K-Means Clustering with a multiple-input deep learning regression-based model (DL-KMA), we have achieved substantial improvements in altitude estimation accuracy. This methodology represents a significant advancement over conventional approaches in UAV technology. Our experimental design involved comprehensive field data collection across two distinct altitude environments, employing a high-precision Digital Laser Distance Meter as the reference standard (Class II). This rigorous approach facilitated a thorough evaluation of our model’s performance across varied terrains, ensuring robust and reliable results. The outcomes of our study are particularly noteworthy, with the model demonstrating remarkably low Mean Squared Error (MSE) values across all data clusters, ranging from 0.011 to 0.072. These results not only indicate significant improvements over traditional methods, but also establish a new benchmark in UAVs altitude estimation accuracy. A key innovation in our approach is the elimination of costly additional hardware such as Light Detection and Ranging (LiDAR), offering a cost-effective, software-based solution. This advancement has broad implications, enhancing the accessibility of advanced UAVs technology and expanding its potential applications across diverse sectors including precision agriculture, urban planning, and emergency response. This research represents a significant contribution to the integration of AI and UAVs technology, potentially unlocking new possibilities in UAVs applications. By enhancing the capabilities of lightweight UAVs, we are not merely improving a technical aspect, but revolutionizing the potential applications of UAVs across industries. Our work sets the stage for safer, more reliable, and precise UAVs operations, marking a pivotal moment in the evolution of aerial technology in an increasingly UAV-dependent world. Full article
Show Figures

Figure 1

Figure 1
<p>A method for measuring the altitude of a drone using a digital laser distance meter.</p>
Full article ">Figure 2
<p>Hardware Components and System Architecture of the Experimental UAV Platform.</p>
Full article ">Figure 3
<p>Experimental Setup for UAV Altitude Data Collection and Validation Using Digital Laser Measurement.</p>
Full article ">Figure 4
<p>UAV Flight Data Extraction and Validation Process Using Video-synchronized Log Analysis.</p>
Full article ">Figure 5
<p>The image presents a schematic representation of a high-precision fixed distance range measurement experiment, likely designed for unmanned aerial vehicle (UAV) altitude calibration.</p>
Full article ">Figure 6
<p>The DL-KMA predicts altitude estimation accuracy in each cluster of Unmanned Aerial Vehicles (UAVs).</p>
Full article ">Figure 7
<p>An advanced deep learning pipeline for enhancing UAV alti-tude precision.</p>
Full article ">Figure 8
<p>A system for precise drone altitude control using DL-KMA.</p>
Full article ">Figure 9
<p>Diagram Illustrating Area Configuration and UAV Deployment Setup.</p>
Full article ">Figure 10
<p>Example Raw Data from UAV MAVLink logs binary format and converts to CSV files.</p>
Full article ">Figure 11
<p>Data Collection and Preprocessing Workflow for UAV Flight Log Analysis with VDO Integration.</p>
Full article ">Figure 12
<p>Elbow method analysis of UAV telemetry data demonstrating optimal K=4 clusters based on WCSS (Within-Cluster Sum of Squares) metric and corresponding flight parameters.</p>
Full article ">Figure 13
<p>Model Performance Metrics for DL-KMA in altitude estimation.</p>
Full article ">Figure 13 Cont.
<p>Model Performance Metrics for DL-KMA in altitude estimation.</p>
Full article ">
20 pages, 20078 KiB  
Article
Pollutant Dispersion Dynamics Under Horizontal Wind Shear Conditions: Insights from Bidimensional Traffic Flow Models
by Anis Chaari, Waleed Mouhali, Nacer Sellila, Mohammed Louaked and Houari Mechkour
Fluids 2024, 9(11), 265; https://doi.org/10.3390/fluids9110265 - 14 Nov 2024
Viewed by 560
Abstract
Meteorological factors, specifically wind direction and magnitude, influence the dispersion of atmospheric pollutants due to road traffic by affecting their spatial and temporal distribution. In this study, we are interested in the effect of the evolution of horizontal wind components, i.e., in the [...] Read more.
Meteorological factors, specifically wind direction and magnitude, influence the dispersion of atmospheric pollutants due to road traffic by affecting their spatial and temporal distribution. In this study, we are interested in the effect of the evolution of horizontal wind components, i.e., in the plane perpendicular to the altitude axis. A two-dimensional numerical model for solving the coupled traffic flow/pollution problem, whose pollutants are generated by vehicles, is developed. The numerical solution of this model is computed via an algorithm combining the characteristics method for temporal discretization with the finite-element method for spatial discretization. The numerical model is validated through a sensitivity study on the diffusion coefficient of road traffic and its impact on traffic density. The distribution of pollutant concentration, computed based on a source generated by traffic density, is presented for a single direction and different magnitudes of the wind velocity (stationary, Gaussian, linearly increasing and decreasing, sudden change over time), taking into account the stretching and tilting of plumes and patterns. The temporal evolution of pollutant concentration at various relevant locations in the domain is studied for two wind velocities (stationary and sudden change). Three regimes were observed for transport pollution depending on time and velocity: nonlinear growth, saturation, and decrease. Full article
(This article belongs to the Section Mathematical and Computational Fluid Mechanics)
Show Figures

Figure 1

Figure 1
<p>Road network <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math>.</p>
Full article ">Figure 2
<p>Region <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mo>Ω</mo> <mi>p</mi> </msub> </mstyle> </semantics></math> including the road network <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math>.</p>
Full article ">Figure 3
<p>Googl maps image of a portion of “Périphérique de Paris”.</p>
Full article ">Figure 4
<p>Mesh of <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math>.</p>
Full article ">Figure 5
<p>Zoom on the mesh of <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mo>Ω</mo> <mi>p</mi> </msub> </mstyle> </semantics></math>.</p>
Full article ">Figure 6
<p>Density distribution <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>ρ</mi> </mstyle> </semantics></math> on the road network <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math> for diffusion coefficient <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mi>k</mi> <mo>=</mo> <mn>1</mn> </mrow> </mstyle> </semantics></math> km<sup>2</sup>/h at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 7
<p>Density distribution <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>ρ</mi> </mstyle> </semantics></math> on the road network <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math> for diffusion coefficient <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mi>k</mi> <mo>=</mo> <mn>1.5</mn> </mrow> </mstyle> </semantics></math> km<sup>2</sup>/h at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 8
<p>Density distribution <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>ρ</mi> </mstyle> </semantics></math> on the road network <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mo>Ω</mo> </mstyle> </semantics></math> for diffusion coefficient <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mi>k</mi> <mo>=</mo> <mn>3</mn> </mrow> </mstyle> </semantics></math> km<sup>2</sup>/h at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 9
<p>Concentration distribution for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>−</mo> <mi>π</mi> <mo>/</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> direction wind with velocity magnitude <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mtext> </mtext> <mo>=</mo> <mn>2</mn> </mrow> </mstyle> </semantics></math> km/h at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 10
<p>Concentration distribution for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>−</mo> <mi>π</mi> <mo>/</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> direction for the Gaussian model wind with velocity magnitude at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 11
<p>Concentration distribution for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>−</mo> <mi>π</mi> <mo>/</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> direction wind with increasing step function of velocity magnitude <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mo>=</mo> <mi>g</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mstyle> </semantics></math> at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 12
<p>Concentration distribution for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>−</mo> <mi>π</mi> <mo>/</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> direction wind with decreasing step function of velocity magnitude <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mo>=</mo> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mstyle> </semantics></math> at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 13
<p>Concentration distribution for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>−</mo> <mi>π</mi> <mo>/</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> direction wind with decreasing chaotic step function of velocity magnitude <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mo>=</mo> <mi>i</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mstyle> </semantics></math> at different times <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math>.</p>
Full article ">Figure 14
<p>Position of points <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mi>A</mi> <mi>i</mi> </msub> </mstyle> </semantics></math>, <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mi>B</mi> <mi>i</mi> </msub> </mstyle> </semantics></math>, <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mi>C</mi> <mi>i</mi> </msub> </mstyle> </semantics></math> and <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mi>D</mi> <mi>i</mi> </msub> </mstyle> </semantics></math> for <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> </mrow> </mstyle> </semantics></math> in <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mo>Ω</mo> <mi>p</mi> </msub> </mstyle> </semantics></math>.</p>
Full article ">Figure 15
<p>Evolution of the CO concentration <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>ϕ</mi> </mstyle> </semantics></math> as a function of time <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mi>t</mi> </mstyle> </semantics></math> at the different points on <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <msub> <mo>Ω</mo> <mi>p</mi> </msub> </mstyle> </semantics></math> with velocity magnitude <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mo>=</mo> <mn>2</mn> </mrow> </mstyle> </semantics></math> km/h on the left (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mrow> <mo>∥</mo> <mi mathvariant="bold">u</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi>t</mi> <mo>)</mo> <mo>∥</mo> <mo>=</mo> <mi>i</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mstyle> </semantics></math> on the right (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>).</p>
Full article ">
40 pages, 79804 KiB  
Article
The GRAZ Method—Determination of Urban Surface Temperatures from Aerial Thermography Based on a Three-Dimensional Sampling Algorithm
by Daniel Rüdisser, Thomas Posch and Wolfgang Sulzer
Remote Sens. 2024, 16(21), 3949; https://doi.org/10.3390/rs16213949 - 23 Oct 2024
Viewed by 830
Abstract
A novel method to derive surface temperatures from aerial thermography is proposed. Its theoretical foundation, details regarding the implementation, relevant sensitivities, and its application on a day and night survey are presented here. The method differs from existing approaches particularly in two aspects: [...] Read more.
A novel method to derive surface temperatures from aerial thermography is proposed. Its theoretical foundation, details regarding the implementation, relevant sensitivities, and its application on a day and night survey are presented here. The method differs from existing approaches particularly in two aspects: first, a three-dimensional sampling approach is used to determine the reflected thermal radiation component. Different surface classes based on hyperspectral classification with specific properties regarding the reflection and emission of thermal radiation are considered in this sampling process. Second, the method relies on a detailed, altitude-dependent, directionally and spectrally resolved modelling of the atmospheric radiation transfer and considers the spectral sensitivity of the sensor used. In order to accurately consider atmospheric influences, the atmosphere is modelled as a function of altitude regarding temperature, pressure and greenhouse gas concentrations. The atmospheric profiles are generated specifically for the time of the survey based on measurements, meteorological forecasts and generic models. The method was initially developed for application in urban contexts, as it is able to capture the pronounced three-dimensional character of such environments. However, due to the detailed consideration of elevation and atmospheric conditions, the method is also valuable for the analysis of rural areas. The included case studies covering two thermographic surveys of city area of Graz during daytime and nighttime demonstrate the capabilities and feasibility of the method. In relation to the detected brightness temperatures apparent to the sensor, the determined surface temperatures vary considerably and generally cover an increased temperature range. The two processed surface temperature maps of the city area of Graz are finally used to validate the method based on available temperature recordings. Full article
(This article belongs to the Section Urban Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Overview of the main components considered in the method.</p>
Full article ">Figure 2
<p>Atmospheric integration paths (N = 10).</p>
Full article ">Figure 3
<p>Example of spectral radiances upwelling (LU), downwelling at different angles (LD1-LD10: <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> </mrow> </semantics></math> [87.1°, 81.4°, 75.5°, 69.5°, 63.3°, 56.6°, 49.5°, 41.4°, 31.8°, 18.2°]); and spectral sensitivity of sensor S (summer day; clear sky; elevation: 350 m MSL; sensor altitude 1675 m MSL).</p>
Full article ">Figure 4
<p>Elevation dependency of upwelling (LU) and downwelling radiance at different angles (LD1-LD10); sensor integrated values; exemplary for a summer day with clear sky; and sensor altitude 1675 m MSL (true altitude above mean sea level).</p>
Full article ">Figure 5
<p>Elevation dependency of atmospheric transmittance for different blackbody radiation temperatures; sensor integrated values; exemplary for a summer day with clear sky; and sensor altitude 1675 m MSL.</p>
Full article ">Figure 6
<p>Angular dependence of reflectivity for thermal radiation of different brightness temperatures for common window glass [<a href="#B17-remotesensing-16-03949" class="html-bibr">17</a>].</p>
Full article ">Figure 7
<p>Polar plot representation of directional emissivity and radiant intensity for common window glass vs. Lambert model [<a href="#B17-remotesensing-16-03949" class="html-bibr">17</a>].</p>
Full article ">Figure 8
<p>Applied reflectivities for thermal radiation of land cover classes.</p>
Full article ">Figure 9
<p>Proposed “backside foliage approach” for the radiometric determination of near air temperatures.</p>
Full article ">Figure 10
<p>03 UTC survey log–log representation of relevant atmospheric gas concentrations, pressure and air temperature (semi-log).</p>
Full article ">Figure 11
<p>UTC survey, log–log representation of relevant atmospheric gas concentrations, pressure and air temperature (semi-log).</p>
Full article ">Figure 12
<p>Absolute humidity vertical profiles for both surveys (solid graphs) and cumulative (air column) humidity values (dashed graphs).</p>
Full article ">Figure 13
<p>Removal of overhead wiring; laser scanner-based DSM before (<b>left</b>) and after (<b>right</b>) processing.</p>
Full article ">Figure 14
<p>Determined slope angles in the east-west (<b>left</b>) and north-south direction (<b>right</b>) based on [<a href="#B15-remotesensing-16-03949" class="html-bibr">15</a>].</p>
Full article ">Figure 15
<p>Sampled view factors for the city center of Graz; top-left: for urban (built-up) environment; top right: for vegetation; bottom-left: for topmost sky segment D10 (0°–25.8°); bottom-right: for sky segment D4 (72.5°–66.4°); and depiction based on [<a href="#B15-remotesensing-16-03949" class="html-bibr">15</a>].</p>
Full article ">Figure 16
<p>Heat survey “12 UTC”: apparent sensor temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>T</mi> </mrow> <mo>^</mo> </mover> </mrow> <mrow> <mi>s</mi> <mi>e</mi> <mi>n</mi> <mi>s</mi> <mi>o</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> (<b>left</b>) vs. determined surface temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>T</mi> </mrow> <mo>^</mo> </mover> </mrow> <mrow> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> (<b>right</b>).</p>
Full article ">Figure 17
<p>Night survey “03 UTC”: apparent sensor temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>T</mi> </mrow> <mo>^</mo> </mover> </mrow> <mrow> <mi>s</mi> <mi>e</mi> <mi>n</mi> <mi>s</mi> <mi>o</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> (<b>left</b>) vs. determined surface temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mover accent="true"> <mrow> <mi>T</mi> </mrow> <mo>^</mo> </mover> </mrow> <mrow> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> (<b>right</b>).</p>
Full article ">Figure 18
<p>Location of water temperature sensor in Mur River. (<b>left</b>): orthophoto, (<b>right</b>): processed thermal map.</p>
Full article ">Figure 19
<p>Contact temperature measurement “Lendplatz”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 20
<p>Contact temperature measurement “Färberplatz”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 21
<p>Determination of near-air temperature “Schlossberg”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 22
<p>Determination of near-air temperature “Oeverseepark”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 23
<p>Determination of near-air temperature “Lustbuehel”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 24
<p>Determination of near-air temperature “Plabutsch”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 25
<p>Determination of near-air temperature “Mariatrost”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 26
<p>Determination of near-air temperature “Graz university 2 m”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 27
<p>Determination of near-air temperature “Graz university 5 cm”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 28
<p>Determination of near-air temperature “Graz Strassgang 2 m”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 29
<p>Determination of near-air temperature “Graz Strassgang 5 cm”—(<b>left</b>): orthophoto of meas. Location; (<b>center</b>): measurement path in processed thermal map; and (<b>right</b>): evaluation of temperatures along measurement path vs. reference (white arrow indicates measurement path).</p>
Full article ">Figure 30
<p>Solar irradiance at time of survey flight (yellow) vs. deviation of determined radiometric water temperature and contact-measured water temperature (red).</p>
Full article ">Figure 31
<p>Exemplary evaluation for both surveys for different elevations and emissivity values. Red graphs: determined surface temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> for the 12 UTC survey, blue graphs: determined surface temperatures <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> for the 03 UTC survey. Input parameters/assumptions: unobstructed horizon, sensed temperature <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>s</mi> <mi>e</mi> <mi>n</mi> <mi>s</mi> <mi>o</mi> <mi>r</mi> </mrow> </msub> <mo>=</mo> <mn>20</mn> <mo> </mo> <mo>°</mo> <mi mathvariant="normal">C</mi> </mrow> </semantics></math>, and emissivity ε = [0.8, 0.9, 1.0].</p>
Full article ">Figure 32
<p>Orthophoto “solar thermal field” (evaluation path indicated as arrow).</p>
Full article ">Figure 33
<p>Determination of surface temperatures “solar thermal field”—survey 12 UTC. (<b>left</b>): detected temperatures; (<b>right</b>): determined temperatures—evaluation path indicated as arrow.</p>
Full article ">Figure 34
<p>Temperatures along indicated path “solar thermal field”.</p>
Full article ">Figure 35
<p>Orthophoto “city center: Mur River and Schlossberg” (evaluation path indicated as arrow).</p>
Full article ">Figure 36
<p>Determination of surface temperatures “city center: Mur River and Schlossberg”—survey 12 UTC. (<b>left</b>): detected temperatures; (<b>right</b>): determined temperatures—evaluation path indicated as arrow.</p>
Full article ">Figure 37
<p>Determined and detected surface temperatures along indicated path “city center: Mur River and Schlossberg”.</p>
Full article ">Figure 38
<p>Orthophoto “Hauenstein” (evaluation path indicated as arrow).</p>
Full article ">Figure 39
<p>Determination of surface temperatures “Hauenstein—radiation inversion”—survey 03 UTC. (<b>left</b>): detected temperatures; (<b>right</b>): determined temperatures—evaluation path indicated as arrow.</p>
Full article ">Figure 40
<p>Determined and detected surface temperatures along indicated path “Hauenstein—radiation inversion”.</p>
Full article ">
22 pages, 5579 KiB  
Article
Experimental Study on LTE Mobile Network Performance Parameters for Controlled Drone Flights
by Janis Braunfelds, Gints Jakovels, Ints Murans, Anna Litvinenko, Ugis Senkans, Rudolfs Rumba, Andis Onzuls, Guntis Valters, Elina Lidere and Evija Plone
Sensors 2024, 24(20), 6615; https://doi.org/10.3390/s24206615 - 14 Oct 2024
Viewed by 1071
Abstract
This paper analyzes the quantitative quality parameters of a mobile communication network in a controlled drone logistic use-case scenario. Based on the analysis of standards and recommendations, the values of key performance indicators (KPIs) are set. As the main network-impacting parameters, reference signal [...] Read more.
This paper analyzes the quantitative quality parameters of a mobile communication network in a controlled drone logistic use-case scenario. Based on the analysis of standards and recommendations, the values of key performance indicators (KPIs) are set. As the main network-impacting parameters, reference signal received power (RSRP), reference signal received quality (RSRQ), and signal to interference and noise ratio (SINR) were selected. Uplink (UL), downlink (DL), and ping parameters were chosen as the secondary ones, as they indicate the quality of the link depending on primary parameters. The analysis is based on experimental measurements performed using a Latvian mobile operator’s “LMT” JSC infrastructure in a real-life scenario. To evaluate the altitude impact on the selected network parameters, the measurements were performed using a drone as transport for the following altitude values: 40, 60, 90, and 110 m. Network parameter measurements were implemented in automatic mode, allowing switching between LTE4–LTE2 standards, providing the opportunity for more complex analysis. Based on the analysis made, the recommendations for the future mobile networks employed in controlled drone flights should correspond to the following KPI and their values: −100 dBm for RSRP, −16 dB for RSRQ, −5 dB for SINR, 4096 kbps for downlink, 4096 kbps for uplink, and 50 ms for ping. Lastly, recommendations for a network coverage digital twin (DT) model with integrated KPIs are also provided. Full article
(This article belongs to the Section Navigation and Positioning)
Show Figures

Figure 1

Figure 1
<p>Potential drone flight trajectory and interfering antenna placement.</p>
Full article ">Figure 2
<p>Map of measurement locations in Latvia.</p>
Full article ">Figure 3
<p>The drone’s flight path (thick yellow line).</p>
Full article ">Figure 4
<p>DJI M30 and OnePlus NORD BE2029 mobile phone are prepared for measurements and ready to take off.</p>
Full article ">Figure 5
<p>The conjunction between RSRQ and RSRP at RSRQ range (−16 to −18 dB) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m and (<b>D</b>) 110 m heights.</p>
Full article ">Figure 6
<p>The conjunction between RSRQ and SINR at RSRQ range (−16 to −18 dB) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m, and (<b>D</b>) 110 m height.</p>
Full article ">Figure 7
<p>RSRQ average value over 50 m distance during the whole flights at 40, 60, 90 and 110 m heights.</p>
Full article ">Figure 8
<p>The conjunction between RSRP and RSRQ at RSRP range (−100 to −105 dBm) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m and (<b>D</b>) 110 m height.</p>
Full article ">Figure 9
<p>The conjunction between RSRP and SINR at RSRP range (−100 to −105 dB) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m, and (<b>D</b>) 110 m height.</p>
Full article ">Figure 10
<p>RSRP average value over 50 m distance during the whole flights at 40, 60, 90 and 110 m height.</p>
Full article ">Figure 11
<p>The conjunction between SINR and RSRP in SINR range (−5 to −7.5 dB) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m, and (<b>D</b>) 110 m height.</p>
Full article ">Figure 12
<p>The conjunction between SINR and RSRQ in SINR range (−5 to −7.5 dB) at (<b>A</b>) 40 m, (<b>B</b>) 60 m, (<b>C</b>) 90 m, and (<b>D</b>) 110 m height.</p>
Full article ">Figure 13
<p>SINR average value over 50 m distance during the whole flight at 40, 60, 90 and 110 m height.</p>
Full article ">Figure 14
<p>DT model of mobile network coverage and UAV flight.</p>
Full article ">
17 pages, 68906 KiB  
Article
STFM: Accurate Spatio-Temporal Fusion Model for Weather Forecasting
by Jun Liu, Li Wu, Tao Zhang, Jianqiang Huang, Xiaoying Wang and Fang Tian
Atmosphere 2024, 15(10), 1176; https://doi.org/10.3390/atmos15101176 - 30 Sep 2024
Viewed by 749
Abstract
Meteorological prediction is crucial for various sectors, including agriculture, navigation, daily life, disaster prevention, and scientific research. However, traditional numerical weather prediction (NWP) models are constrained by their high computational resource requirements, while the accuracy of deep learning models remains suboptimal. In response [...] Read more.
Meteorological prediction is crucial for various sectors, including agriculture, navigation, daily life, disaster prevention, and scientific research. However, traditional numerical weather prediction (NWP) models are constrained by their high computational resource requirements, while the accuracy of deep learning models remains suboptimal. In response to these challenges, we propose a novel deep learning-based model, the Spatiotemporal Fusion Model (STFM), designed to enhance the accuracy of meteorological predictions. Our model leverages Fifth-Generation ECMWF Reanalysis (ERA5) data and introduces two key components: a spatiotemporal encoder module and a spatiotemporal fusion module. The spatiotemporal encoder integrates the strengths of convolutional neural networks (CNNs) and recurrent neural networks (RNNs), effectively capturing both spatial and temporal dependencies. Meanwhile, the spatiotemporal fusion module employs a dual attention mechanism, decomposing spatial attention into global static attention and channel dynamic attention. This approach ensures comprehensive extraction of spatial features from meteorological data. The combination of these modules significantly improves prediction performance. Experimental results demonstrate that STFM excels in extracting spatiotemporal features from reanalysis data, yielding predictions that closely align with observed values. In comparative studies, STFM outperformed other models, achieving a 7% improvement in ground and high-altitude temperature predictions, a 5% enhancement in the prediction of the u/v components of 10 m wind speed, and an increase in the accuracy of potential height and relative humidity predictions by 3% and 1%, respectively. This enhanced performance highlights STFM’s potential to advance the accuracy and reliability of meteorological forecasting. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>Overview of the study area, including elevation changes, rivers, and urban distribution in the area.</p>
Full article ">Figure 2
<p>The overall structure of STFM and the data flow in encoder. STFM consists of a spatiotemporal encoder, a spatiotemporal fusion module (STFU), and a decoder.</p>
Full article ">Figure 3
<p>The overall structure of the spatiotemporal fusion unit (STFU) and spatiotemporal attention unit. STFU consists of a spatiotemporal attention unit and a temporal convolution.</p>
Full article ">Figure 4
<p>Visualization results of our STFM and TAU on the 850 hPa temperature dataset.</p>
Full article ">Figure 5
<p>Visualization results of our STFM and TAU on the 2 m temperature dataset.</p>
Full article ">Figure 6
<p>Visualization results of our STFM and TAU on the 10 m wind speed u-component dataset.</p>
Full article ">Figure 7
<p>Visualization results of our STFM and TAU on the 10 m wind speed v-component dataset.</p>
Full article ">Figure 8
<p>Visualization results of our STFM and UniFormer on the 500 hPa potential dataset.</p>
Full article ">Figure 9
<p>Visualization results of our STFM and SimVP v2 on the 500 hPa relative humidity dataset.</p>
Full article ">
20 pages, 1730 KiB  
Article
Time-Efficient Neural-Network-Based Dynamic Area Optimization Algorithm for High-Altitude Platform Station Mobile Communications
by Wataru Takabatake, Yohei Shibata, Kenji Hoshino and Tomoaki Ohtsuki
Future Internet 2024, 16(9), 332; https://doi.org/10.3390/fi16090332 - 11 Sep 2024
Viewed by 772
Abstract
There is a growing interest in high-altitude platform stations (HAPSs) as potential telecommunication infrastructures in the stratosphere, providing direct communication services to ground-based smartphones. Enhanced coverage and capacity can be realized in HAPSs by adopting multicell configurations. To improve the communication quality, previous [...] Read more.
There is a growing interest in high-altitude platform stations (HAPSs) as potential telecommunication infrastructures in the stratosphere, providing direct communication services to ground-based smartphones. Enhanced coverage and capacity can be realized in HAPSs by adopting multicell configurations. To improve the communication quality, previous studies have investigated methods based on search algorithms, such as genetic algorithms (GAs), which dynamically optimize antenna parameters. However, these methods face hurdles in swiftly adapting to sudden distribution shifts from natural disasters or major events due to their high computational requirements. Moreover, they do not utilize the previous optimization results, which require calculations each time. This study introduces a novel optimization approach based on a neural network (NN) model that is trained on GA solutions. The simple model is easy to implement and allows for instantaneous adaptation to unexpected distribution changes. However, the NN faces the difficulty of capturing the dependencies among neighboring cells. To address the problem, a classifier chain (CC), which chains multiple classifiers to learn output relationships, is integrated into the NN. However, the performance of the CC depends on the output sequence. Therefore, we employ an ensemble approach to integrate the CCs with different sequences and select the best solution. The results of simulations based on distributions in Japan indicate that the proposed method achieves a total throughput whose cumulative distribution function (CDF) is close to that obtained by the GA solutions. In addition, the results show that the proposed method is more time-efficient than GA in terms of the total time required to optimize each user distribution. Full article
(This article belongs to the Special Issue Moving towards 6G Wireless Technologies—Volume II)
Show Figures

Figure 1

Figure 1
<p>Concept diagram of the HAPS mobile communication infrastructure.</p>
Full article ">Figure 2
<p>Beam parameters for a cell.</p>
Full article ">Figure 3
<p>SINR Heatmap based on the beams obtained by genetic algorithm (GA).</p>
Full article ">Figure 4
<p>Neural network (NN) learning the solutions generated by GA.</p>
Full article ">Figure 5
<p>Challenge encountered in area optimization using individually trained NNs.</p>
Full article ">Figure 6
<p>Example of model configuration using CCs (cells #1, #2, and #3 in sequence).</p>
Full article ">Figure 7
<p>Schematic of the ensemble approach (Three-cell horizontal beam directions).</p>
Full article ">Figure 8
<p>The workflow of NN-based dynamic area optimization from input to output.</p>
Full article ">Figure 9
<p>Example of user distributions. (<b>a</b>) Dense area around Tokyo. (<b>b</b>) Sparse area around Akita.</p>
Full article ">Figure 10
<p>SINR heatmaps based on the estimated beams for three-cell configuration (NN, NN+CC, and GA).</p>
Full article ">Figure 11
<p>Constraint violation ratios for three- and six-cell configurations (NN, NN+CC, NN+CC+Ensemble).</p>
Full article ">Figure 12
<p>Cumulative distribution curves of the total throughput for the (<b>a</b>) three- and (<b>b</b>) six-cell configurations.</p>
Full article ">Figure 13
<p>Cumulative distribution curves of the errors for the beam directions, averaged for each cell. (<b>a</b>) Three-cell configuration. (<b>b</b>) Six-cell configuration.</p>
Full article ">Figure 14
<p>Relationship between the total throughput median and the number of extracted models for the ensemble approach.</p>
Full article ">Figure 15
<p>Relationship between the total throughput median and the training data amount.</p>
Full article ">Figure 16
<p>Relationship between the total time taken for optimizing each user distribution and the number of optimized distributions.</p>
Full article ">
33 pages, 11504 KiB  
Article
Perpendicular Electrical Conductivity in the Topside Ionosphere Derived from Swarm Measurements
by Fabio Giannattasio, Alessio Pignalberi, Roberta Tozzi, Paola De Michelis, Simone Mestici, Giuseppe Consolini, Igino Coco and Michael Pezzopane
Remote Sens. 2024, 16(17), 3129; https://doi.org/10.3390/rs16173129 - 24 Aug 2024
Viewed by 928
Abstract
The study of the physical properties of the topside ionosphere is fundamental to investigating the energy balance of the ionosphere and developing accurate models to predict relevant phenomena, which are often at the root of Space Weather effects in the near-Earth environment. One [...] Read more.
The study of the physical properties of the topside ionosphere is fundamental to investigating the energy balance of the ionosphere and developing accurate models to predict relevant phenomena, which are often at the root of Space Weather effects in the near-Earth environment. One of the most important physical parameters characterising the ionospheric medium is electrical conductivity, which is crucial for the onset and amplification of ionospheric currents and for calculating the power density dissipated by such currents. We characterise, for the first time, electrical conductivity in the direction perpendicular to the geomagnetic field, namely Pedersen and Hall conductivities, in the topside ionosphere at an altitude of about 450 km. For this purpose, we use eight years of in situ simultaneous measurements of electron density, electron temperature and geomagnetic field strength acquired by the Swarm A satellite. We present global statistical maps of perpendicular electrical conductivity and study their variations depending on magnetic latitude and local time, seasons, and solar activity. Our findings indicate that the most prominent features of perpendicular electrical conductivity are located at low latitudes and are probably driven by the complex dynamics of the Equatorial Ionisation Anomaly. At higher latitudes, perpendicular conductivity is a few orders of magnitude lower than that at low latitudes. Nevertheless, conductivity features are modulated by solar activity and seasonal variations at all latitudes. Full article
(This article belongs to the Section Satellite Missions for Earth and Planetary Exploration)
Show Figures

Figure 1

Figure 1
<p>Climatological map in QD-MLT coordinates of electrical conductivity parallel to the geomagnetic field, <math display="inline"><semantics> <msub> <mi>σ</mi> <mrow> <mo>|</mo> <mo>|</mo> </mrow> </msub> </semantics></math>. Values within each bin are saturated below 3.5 × 10<sup>11</sup> s<sup>−1</sup> and above 7.5 × 10<sup>11</sup> s<sup>−1</sup>.</p>
Full article ">Figure 2
<p>Climatological maps in QD-MLT coordinates of the following. <b>Top panel</b>: Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>. Values within each bin are saturated above 10<sup>6</sup> s<sup>−1</sup> and below 10<sup>2</sup> s<sup>−1</sup>. <b>Bottom panel</b>: Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>. Values within each bin are saturated above 10<sup>4</sup> s<sup>−1</sup> and below 10<sup>−2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Climatological maps in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, in a polar stereographic projection representation in the Northern (<b>left panel</b>) and Southern (<b>right panel</b>) Hemispheres. Values within each bin are saturated below 0.5 × 10<sup>3</sup> s<sup>−1</sup> and above 1 × 10<sup>3</sup> s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Climatological maps in QD-MLT coordinates of Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, in polar stereographic projection representations in the Northern (<b>left panel</b>) and Southern (<b>right panel</b>) Hemisphere. Values within each bin are saturated below 0 and above 2 s<sup>−1</sup>.</p>
Full article ">Figure 5
<p>Seasonal maps in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, around (from top to bottom) the December solstice, the June solstice, the March equinox, and the September equinox. Seasons were selected as specified in the text. Values within each bin are saturated above 10<sup>6</sup> s<sup>−1</sup> and below 10<sup>2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 5 Cont.
<p>Seasonal maps in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, around (from top to bottom) the December solstice, the June solstice, the March equinox, and the September equinox. Seasons were selected as specified in the text. Values within each bin are saturated above 10<sup>6</sup> s<sup>−1</sup> and below 10<sup>2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Seasonal maps in QD-MLT coordinates the Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, around (from top to bottom) the December solstice, the June solstice, the March equinox, and the September equinox. Seasons were selected as specified in the text. Values within each bin are saturated above 10<sup>4</sup> s<sup>−1</sup> and below 10<sup>−2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 6 Cont.
<p>Seasonal maps in QD-MLT coordinates the Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, around (from top to bottom) the December solstice, the June solstice, the March equinox, and the September equinox. Seasons were selected as specified in the text. Values within each bin are saturated above 10<sup>4</sup> s<sup>−1</sup> and below 10<sup>−2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Seasonal maps (from top to bottom: winter, summer, spring, autumn) in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, in polar stereographic projection representation in the Northern (left column) and Southern (right column) Hemispheres.</p>
Full article ">Figure 7 Cont.
<p>Seasonal maps (from top to bottom: winter, summer, spring, autumn) in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, in polar stereographic projection representation in the Northern (left column) and Southern (right column) Hemispheres.</p>
Full article ">Figure 8
<p>Seasonal maps (from top to bottom: winter, summer, spring, autumn) in QD-MLT coordinates of Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, in polar stereographic projection representation in the Northern (left column) and Southern (right column) Hemispheres.</p>
Full article ">Figure 8 Cont.
<p>Seasonal maps (from top to bottom: winter, summer, spring, autumn) in QD-MLT coordinates of Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, in polar stereographic projection representation in the Northern (left column) and Southern (right column) Hemispheres.</p>
Full article ">Figure 9
<p>Daily solar radio flux at 10.7 cm (F10.7, blue solid line) as a proxy of solar activity from 1 April 2014 to 31 March 2022. Flux is expressed in solar flux units (sfu): 1 sfu = <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>19</mn> </mrow> </msup> </semantics></math> erg·s<sup>−1</sup>·cm<sup>−2</sup>·Hz<sup>−1</sup> in cgs units. The shaded area in orange between the red dashed lines marks a two-year period of moderate solar activity. The shaded area in green between the black dashed lines marks a two-year period of low solar activity.</p>
Full article ">Figure 10
<p>Maps in QD-MLT coordinates of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, during the first (<b>top panel</b>) and third (<b>bottom panel</b>) biennia as selected in <a href="#remotesensing-16-03129-f009" class="html-fig">Figure 9</a>. Values within each bin are saturated above 10<sup>6</sup> s<sup>−1</sup> and below 10<sup>2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 11
<p>Maps in QD-MLT coordinates of Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, during the first (<b>top panel</b>) and third (<b>bottom panel</b>) biennia as selected in <a href="#remotesensing-16-03129-f009" class="html-fig">Figure 9</a>. Values within each bin are saturated above 10<sup>4</sup> s<sup>−1</sup> and below 10<sup>−2</sup> s<sup>−1</sup>.</p>
Full article ">Figure 12
<p>Maps in QD-MLT coordinates and stereographic projection of Pedersen electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>P</mi> </msub> </semantics></math>, in the Northern (top row) and Southern (bottom row) Hemispheres during the first (left column) and third (right column) biennia, respectively.</p>
Full article ">Figure 13
<p>Maps in QD-MLT coordinates and stereographic projection of Hall electrical conductivity, <math display="inline"><semantics> <msub> <mi>σ</mi> <mi>H</mi> </msub> </semantics></math>, in the Northern (top row) and Southern (bottom row) Hemispheres during the first (left column) and third (right column) biennia, respectively.</p>
Full article ">
23 pages, 3244 KiB  
Article
Assessment of Hygroscopic Behavior of Arctic Aerosol by Contemporary Lidar and Radiosonde Observations
by Nele Eggers, Sandra Graßl and Christoph Ritter
Remote Sens. 2024, 16(16), 3087; https://doi.org/10.3390/rs16163087 - 21 Aug 2024
Viewed by 754
Abstract
This study presents the hygroscopic properties of aerosols from the Arctic free troposphere by means of contemporary lidar and radiosonde observations only. It investigates the period from the Arctic Haze in spring towards the summer season in 2021. Therefore, a one-parameter growth curve [...] Read more.
This study presents the hygroscopic properties of aerosols from the Arctic free troposphere by means of contemporary lidar and radiosonde observations only. It investigates the period from the Arctic Haze in spring towards the summer season in 2021. Therefore, a one-parameter growth curve model is applied to lidar data from the Koldewey Aerosol Raman Lidar (AWIPEV in Ny-Ålesund, Svalbard) and simultaneous radiosonde measurements. Hygroscopic growth depends on different factors like aerosol diameter and chemical composition. To detangle this dependency, three trends in hygroscopicity are additionally investigated by classifying the aerosol first by its dry color ratio, and then by its season and altitude. Generally, we found a complex altitude dependence with the least hygroscopic particles in the middle of the troposphere. The most hygroscopic aerosol is located in the upper free troposphere. A hypothesis based on prior lifting of the particles is given. The expected trend with aerosol diameter is not observed, which draws attention to the complex dependence of hygroscopic growth on geographical region and altitude, and to the development of backscatter with the aerosol size itself. In a seasonal overview, two different modes of stronger or weaker hygroscopic particles are additionally observed. Furthermore, two special days are discussed using the Mie theory. They show, on the one hand, the complexity of analyzing hygroscopic growth by means of lidar data, but on the other hand, they demonstrate that it is in fact measurable with this approach. For these two case studies, we calculated that the aerosol effective radius increased from 0.16μm (dry) to 0.18μm (wet) and from 0.28μm to 0.32μm for the second case. Full article
Show Figures

Figure 1

Figure 1
<p>The daily median of the backscatter (<b>a</b>), the color ratio (<b>b</b>) and the aerosol depolarization (<b>c</b>) was calculated within altitudes between <math display="inline"><semantics> <mrow> <mn>0.7</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>10.0</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>, and for the lidar ratio (<b>d</b>) from <math display="inline"><semantics> <mrow> <mn>0.7</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>. After a decrease in April, the backscatter takes its maximum in May. An unusual second increase in July is observed. The lidar ratio is enhanced throughout the whole season and is maximal in May and June. The color ratio continuously increases, and the depolarization decreases. Three estimated seasons are indicated by dotted lines.</p>
Full article ">Figure 2
<p>The daily median of the backscatter (<b>a</b>) and color ratio (<b>b</b>) are illustrated for four different height intervals: 0.7–2.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>, 2.5–4.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>, 4.5–6.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> and 6.5–10.0 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>. Overall, the backscatter is the highest in the lowest height interval. However, the seasonal development, i.e., the transition from spring to summer, is most pronounced between <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>6.5</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>. The color ratio increases towards summer. The strongest gradient in time is visible below <math display="inline"><semantics> <mrow> <mn>2.5</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The seasonal development of the relative humidity is illustrated in (<b>a</b>). The median was determined between <math display="inline"><semantics> <mrow> <mn>0.7</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>10.0</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>. Dotted lines indicate the three seasons—Haze, summer season and forest fire-impacted season. Figure (<b>b</b>) shows the vertical distribution of data points from the whole season between <math display="inline"><semantics> <mrow> <mn>0.7</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>10.0</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> that provided a relative humidity smaller than 40%. On average, the relative humidity decreases with altitude.</p>
Full article ">Figure 4
<p>The vertical distribution of the color ratio (<b>a</b>) and relative humidity (<b>b</b>) over the troposphere are shown. Values between <math display="inline"><semantics> <mrow> <mn>40</mn> <mo>%</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>60</mn> <mo>%</mo> </mrow> </semantics></math>, as well as the smallest color ratio values, occur most often between <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>4</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>. Note, as no direct comparison of radiosonde and lidar data was performed here, not only the simultaneous data are illustrated, which enhances the data basis.</p>
Full article ">Figure 5
<p>The backscatter development between April and July 2021 with regard to the relative humidity over water is shown in (<b>a</b>). The median backscatter of each percentage of relative humidity is additionally illustrated. In general, the aerosol demonstrates a hygroscopic growth between 40% and 67% relative humidity. Beginning at <math display="inline"><semantics> <mrow> <mn>67</mn> <mo>%</mo> </mrow> </semantics></math> relative humidity, a more irregular behavior dominates. The growth curve is fitted onto the normalized median backscatter between <math display="inline"><semantics> <mrow> <mn>41</mn> <mo>%</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>67</mn> <mo>%</mo> </mrow> </semantics></math> relative humidity in (<b>b</b>). The fitting parameter <math display="inline"><semantics> <mi>γ</mi> </semantics></math> amounts to <math display="inline"><semantics> <mrow> <mn>0.23</mn> </mrow> </semantics></math> with an R<sup>2</sup> of 0.43.</p>
Full article ">Figure 6
<p>The median backscatter of the subdivided dataset is illustrated in scatter plots (<b>a</b>–<b>c</b>) along with relative humidity. The intervals of the subdivision were the following: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>3</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mspace width="3.33333pt"/> <mo>≥</mo> <mspace width="3.33333pt"/> <mn>3.0</mn> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mspace width="3.33333pt"/> <mo>≤</mo> <mspace width="3.33333pt"/> <mn>1.2</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="3.33333pt"/> <mo>≤</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>1.2</mn> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> <mo>≤</mo> <mn>1.7</mn> </mrow> </semantics></math>. It resulted in an increasing aerosol diameter. The growth curve was calculated for each dataset. Hygroscopic growth was the strongest for (<b>b</b>) and the weakest for (<b>a</b>).</p>
Full article ">Figure 7
<p>The growth curve was fitted onto the median backscatter above 40% relative humidity over water. The data were taken from the seasonally classified dataset. It was subdivided into Arctic Haze (<b>a</b>), summer (<b>b</b>) and the season with forest fire impacts (<b>c</b>). Modes of higher and one of lower hygroscopicity were visible during Arctic Haze and summer. The high modes almost coincided, whereas the lower mode of the summer season was still stronger than during Arctic Haze and the forest fire-impacted season.</p>
Full article ">Figure 8
<p>The growth curve was fitted onto the data of the different height intervals 0.7–2.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>a</b>), 2.5–4.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>b</b>), 4.5–6.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>c</b>) and 6.5–10.0 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>d</b>). Except for the uppermost height interval, no clear growth trend was observed. Especially within 2.5–6.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>, random trends seemed to dominate.</p>
Full article ">Figure 9
<p>The development of the color ratio (<b>a</b>) and of the depolarization and lidar ratio (<b>b</b>) with relative humidity on 23 May between <math display="inline"><semantics> <mrow> <mn>2.28</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>3.28</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> is illustrated. <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>532</mn> <mn>355</mn> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>1064</mn> <mn>532</mn> </msubsup> </semantics></math> developed contrarily. While the lidar ratio at <math display="inline"><semantics> <mrow> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> </mrow> </semantics></math> was constantly low, it had a maximum at <math display="inline"><semantics> <mrow> <mn>30</mn> <mo>%</mo> </mrow> </semantics></math> relative humidity for <math display="inline"><semantics> <mrow> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> </mrow> </semantics></math>. In <a href="#sec5dot1-remotesensing-16-03087" class="html-sec">Section 5.1</a> the aerosol radius of this case study is shown to have increased from <math display="inline"><semantics> <mrow> <mn>0.16</mn> <mspace width="0.166667em"/> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mn>0.18</mn> <mspace width="0.166667em"/> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>The backscatter profiles at 10:52:31 and the relative humidity profiles at 11:00:00 between 1550–1900 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> (<b>a</b>), 3000–3800 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> (<b>b</b>) and 6350–6650 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> (<b>c</b>) on 29 April are illustrated. These cases demonstrate the difficulties when analyzing hygroscopic growth with combined radiosonde and lidar data.</p>
Full article ">Figure 11
<p>The color ratio development of <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>532</mn> <mn>355</mn> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>1064</mn> <mn>532</mn> </msubsup> </semantics></math> with regard to relative humidity is displayed between 3000–3800 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> (<b>a</b>) and 6350–6650 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> (<b>b</b>). The development of the two color ratios seems chaotic. No hygroscopic growth is visible.</p>
Full article ">Figure 12
<p>The profiles of color ratio <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>1064</mn> <mn>532</mn> </msubsup> </semantics></math> and relative humidity for the lowest layer, 1550–1900 <math display="inline"><semantics> <mrow> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, are illustrated (<b>a</b>). No strict correlation is seen. In addition, the development of <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>532</mn> <mn>355</mn> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>1064</mn> <mn>532</mn> </msubsup> </semantics></math> with relative humidity is shown (<b>b</b>). In particular, <math display="inline"><semantics> <msubsup> <mi>CR</mi> <mn>532</mn> <mn>355</mn> </msubsup> </semantics></math> stays almost constant.</p>
Full article ">Figure 13
<p>Dependence of the color ratios on the effective radius of aerosol according to Mie theory for a log-normal distribution of geometric standard deviation (<math display="inline"><semantics> <mi>σ</mi> </semantics></math>) of 1.1 and a complex index of refraction of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1.5</mn> <mspace width="0.166667em"/> <mo>+</mo> <mspace width="0.166667em"/> <mn>0.01</mn> <mi>i</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Dependence of the aerosol backscatter at the three colors of <math display="inline"><semantics> <mrow> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> </mrow> </semantics></math> as a function of the effective radius of the aerosol according to Mie theory for a log-normal distribution of geometric standard deviation (<math display="inline"><semantics> <mi>σ</mi> </semantics></math>) of 1.1 and a complex index of refraction of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1.5</mn> <mspace width="0.166667em"/> <mo>+</mo> <mspace width="0.166667em"/> <mn>0.01</mn> <mi>i</mi> </mrow> </semantics></math>. The values on the y-axis are in arbitrary units as the concentration of aerosol is different from case to case.</p>
Full article ">Figure A1
<p>The median extinction over all available data points (without cloud influence) was calculated and is displayed for each height step. The standard deviation for each height step is illustrated as a filled area. It demonstrates the strongly increasing noise in extinction and thus the lidar ratio with altitude. For the previous analysis in this paper, the height intervals 0.7–2.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>, 2.5–4.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math>, 4.5–6.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> and 6.5–10.0 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> were often utilized. In particular, the lidar ratio was only used within the lowest height interval due to noise. To emphasize this strengthening in noise, the mean standard deviation within this height interval is denoted in the figure. It rose by a magnitude of about 2.</p>
Full article ">Figure A2
<p>The vertical distribution of the relative humidity over ice is shown. Significant supersaturation (&gt;110%) occurs only above a <math display="inline"><semantics> <mrow> <mn>4</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> altitude.</p>
Full article ">Figure A3
<p>The backscatter and radiosonde data from April to July 2021 were subdivided according to the profiles’ median dry color ratios. The intervals were the following: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>3</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mspace width="3.33333pt"/> <mo>≥</mo> <mspace width="3.33333pt"/> <mn>3.0</mn> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mspace width="3.33333pt"/> <mo>≤</mo> <mspace width="3.33333pt"/> <mn>1.2</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="3.33333pt"/> <mo>≤</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>355</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <mn>3</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>1.2</mn> <mspace width="3.33333pt"/> <mo>&lt;</mo> <mspace width="3.33333pt"/> <msub> <mi>CR</mi> <mi>dry</mi> </msub> <mrow> <mo>(</mo> <mn>532</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>,</mo> <mo> </mo> <mn>1064</mn> <mspace width="0.166667em"/> <mi>nm</mi> <mo>)</mo> </mrow> <mo>≤</mo> <mn>1.7</mn> </mrow> </semantics></math>. The median backscatter was calculated for each percentage of relative humidity. Overall, the backscatter still rises with humidity, as expected.</p>
Full article ">Figure A4
<p>The lidar and radiosonde data were subdivided into the three seasons, Arctic Haze (<b>a</b>), summer (<b>b</b>) and season with forest fire impacts (<b>c</b>). This separation was based on <a href="#sec3-remotesensing-16-03087" class="html-sec">Section 3</a>. To visualize an average growth behavior of the season, the median backscatter was calculated for each percentage of relative humidity. Note that backscatter developments of individual time steps may have a great impact on the overall trend.</p>
Full article ">Figure A5
<p>The dataset was subdivided by altitude. The backscatter and median backscatter between 0.7–2.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>a</b>), 2.5–4.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>b</b>), 4.5–6.5 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>c</b>) and 6.5–10.0 <math display="inline"><semantics> <mrow> <mi>km</mi> </mrow> </semantics></math> (<b>d</b>) are illustrated.</p>
Full article ">Figure A6
<p>The tropospheric profiles from backscatter, relative humidity (<b>a</b>) and temperature (<b>b</b>) are illustrated. A strong gradient in relative humidity is visible from <math display="inline"><semantics> <mrow> <mn>2.28</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mn>3.28</mn> <mspace width="0.166667em"/> <mi>km</mi> </mrow> </semantics></math>. A focused analysis of this interval was performed in <a href="#sec4dot4-remotesensing-16-03087" class="html-sec">Section 4.4</a>.</p>
Full article ">
11 pages, 1242 KiB  
Article
Mesospheric Ozone Depletion during 2004–2024 as a Function of Solar Proton Events Intensity
by Grigoriy Doronin, Irina Mironova, Nikita Bobrov and Eugene Rozanov
Atmosphere 2024, 15(8), 944; https://doi.org/10.3390/atmos15080944 - 6 Aug 2024
Viewed by 1795
Abstract
Solar proton events (SPEs) affect the Earth’s atmosphere, causing additional ionization in the high-latitude mesosphere and stratosphere. Ionization rates from such solar proton events maximize in the stratosphere, but the formation of ozone-depleting nitrogen and hydrogen oxides begins at mesospheric altitudes. The destruction [...] Read more.
Solar proton events (SPEs) affect the Earth’s atmosphere, causing additional ionization in the high-latitude mesosphere and stratosphere. Ionization rates from such solar proton events maximize in the stratosphere, but the formation of ozone-depleting nitrogen and hydrogen oxides begins at mesospheric altitudes. The destruction of mesospheric ozone is associated with protons with energies of about 10 MeV and higher and will strongly depend on the intensity of the flux of these particles. Most studies investigating the impact of SPEs on the characteristics of the middle atmosphere have been based on either simulations or reanalysis datasets, and some studies have used satellite observations to validate model results. We study the impact of SPEs on cold-season ozone loss in both the northern and southern hemispheres using Aura MLS mesospheric ozone measurements over the 2004 to 2024 period. Here, we show how strongly SPEs can deplete polar mesospheric ozone in different hemispheres and attempt to evaluate this dependence on the intensity of solar proton events. We found that moderate SPEs consisting of protons with an energy of more than 10 MeV and a flux intensity of more than 100 pfu destroy mesospheric ozone in the northern hemisphere up to 47% and in the southern hemisphere up to 33%. For both hemispheres, the peak of winter ozone loss was observed at about 76 km. In the northern hemisphere, maximum winter ozone loss was observed on the second day after a solar proton event, but in the southern hemisphere, winter ozone depletion was already detected on the first day. In the southern hemisphere, mesospheric ozone concentrations return to pre-event levels on the ninth day after a solar proton event, but in the northern hemisphere, even on the tenth day after a solar proton event, the mesospheric ozone layer may not be fully recovered. The strong SPEs with a proton flux intensity of more than 1000 pfu lead to a maximum winter ozone loss of up to 85% in the northern hemisphere, and in the southern hemisphere winter, ozone loss reaches 73%. Full article
(This article belongs to the Special Issue Cosmic Rays, Ozone Depletion and Climate Change)
Show Figures

Figure 1

Figure 1
<p>Results of superposed epoch analysis of Aura MLS ozone altitudinal profiles over 60–80 NH before and after SPEs, which are summarized in <a href="#atmosphere-15-00944-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Results of superposed epoch analysis of Aura MLS ozone altitudinal profiles over 60–80 SH before and after SPEs, which are summarized in <a href="#atmosphere-15-00944-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>Northern hemisphere ozone depletion (in %) after SPEs compared to the average ozone concentration observed before SPEs, which are summarized in <a href="#atmosphere-15-00944-t001" class="html-table">Table 1</a>. The ozone profile for each day is obtained using superposed epoch analysis of Aura MLS ozone altitudinal profiles for 60–80 NH after moderate SPEs with a proton flux intensity of more than 100 pfu.</p>
Full article ">Figure 4
<p>Northern hemisphere ozone depletion (in %) after SPEs compared to the average ozone concentration observed before SPEs, which are summarized in <a href="#atmosphere-15-00944-t001" class="html-table">Table 1</a>. The ozone profile for each day is obtained using superposed epoch analysis of Aura MLS ozone altitudinal profiles for 60–80 NH after strong SPEs with a proton flux intensity of more than 1000 pfu.</p>
Full article ">Figure 5
<p>Southern hemisphere ozone depletion (in %) after SPEs compared to the average ozone concentration observed before SPEs, which are summarized in <a href="#atmosphere-15-00944-t002" class="html-table">Table 2</a>. Each day ozone profile—results superposed epoch analysis of Aura MLS ozone altitudinal profiles for 60–80 SH after solar proton events. Moderate SPEs—with a proton flux intensity of more than 100 pfu.</p>
Full article ">Figure 6
<p>Southern hemisphere ozone depletion (in %) after SPEs compared to the average ozone concentration observed before SPEs, which are summarized in <a href="#atmosphere-15-00944-t002" class="html-table">Table 2</a>. Each day ozone profile—results superposed epoch analysis of Aura MLS ozone altitudinal profiles for 60–80 SH after SPE. Strong solar proton events—with a proton flux intensity of more than 1000 pfu.</p>
Full article ">
17 pages, 6261 KiB  
Article
Impact of Oil Viscosity on Emissions and Fuel Efficiency at High Altitudes: A Response Surface Methodology Analysis
by Milton Garcia Tobar, Oscar Cabrera Ojeda and Fredy Crespo Montaño
Lubricants 2024, 12(8), 277; https://doi.org/10.3390/lubricants12080277 - 3 Aug 2024
Viewed by 1575
Abstract
This study investigates the effect of oil viscosity on pollutant emissions and fuel consumption of an internal combustion engine (ICE) at high altitudes using a response surface methodology (RSM). A Chevrolet Corsa Evolution 1.5 SOHC gasoline engine was used in Cuenca, Ecuador (2560 [...] Read more.
This study investigates the effect of oil viscosity on pollutant emissions and fuel consumption of an internal combustion engine (ICE) at high altitudes using a response surface methodology (RSM). A Chevrolet Corsa Evolution 1.5 SOHC gasoline engine was used in Cuenca, Ecuador (2560 m above sea level), testing three lubricating oils with kinematic viscosities of 9.66, 14.08, and 18.5 mm2/s, measured at a temperature of 100 °C under various engine speeds and loads. Key findings include the following: hydrocarbon (HC) emissions were minimized from 150.22 ppm at the maximum load to 7.25 ppm with low viscosity and load; carbon dioxide (CO2) emissions peaked at 15.2% vol with high viscosity and load; carbon monoxide (CO) ranged from 0.04% to 3.74% depending on viscosity and load; nitrogen oxides (NOx) were significantly influenced by viscosity, RPM, and load, indicating a need for model refinement; and fuel consumption was significantly affected by load and viscosity. RSM-based optimization identified optimal operational conditions with a viscosity of 13 mm2/s, 1473 rpm, and a load of 78%, resulting in 52.35 ppm of HC, 13.97% vol of CO2, 1.2% vol of CO, 0 ppm of NOx, and a fuel consumption of 6.66 L/h. These conditions demonstrate the ability to adjust operational variables to maximize fuel efficiency and minimize emissions. This study underscores the critical role of optimizing lubricant viscosity and operational conditions to mitigate environmental impact and enhance engine performance in high-altitude environments. Full article
(This article belongs to the Special Issue Recent Advances in Automotive Powertrain Lubrication)
Show Figures

Figure 1

Figure 1
<p>Experimental unit.</p>
Full article ">Figure 2
<p>Schematic of the experimental setup. (<b>a</b>) Dynamometer. (<b>b</b>) Dynamometer control unit. (<b>c</b>) Dynamometer’s sensor console. (<b>d</b>) Fuel flow meter. (<b>e</b>) Gas emissions analyzer. (<b>f</b>) Expert design software (version 12).</p>
Full article ">Figure 3
<p>(<b>a</b>) Contour plot and (<b>b</b>) response surface for HC.</p>
Full article ">Figure 4
<p>(<b>a</b>) Contour plot and (<b>b</b>) response surface for CO<sub>2</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Contour plot and (<b>b</b>) Response surface for CO.</p>
Full article ">Figure 6
<p>(<b>a</b>) Contour plot and (<b>b</b>) Response surface for NOx.</p>
Full article ">Figure 7
<p>(<b>a</b>) Contour plot and (<b>b</b>) response surface of consumption.</p>
Full article ">Figure 8
<p>Identified optimum point.</p>
Full article ">Figure 9
<p>Contour plot of desirability.</p>
Full article ">
10 pages, 5049 KiB  
Communication
Kriging Interpolation for Constructing Database of the Atmospheric Refractivity in Korea
by Doyoung Jang, Nammoon Kim and Hosung Choo
Remote Sens. 2024, 16(13), 2379; https://doi.org/10.3390/rs16132379 - 28 Jun 2024
Viewed by 770
Abstract
This paper presents a Kriging interpolation method for constructing a database of atmospheric refractivity in Korea. To collect as much data as possible for the interpolation, meteorological data from 120 regions of Korea, including both land and sea areas, are examined. Then, the [...] Read more.
This paper presents a Kriging interpolation method for constructing a database of atmospheric refractivity in Korea. To collect as much data as possible for the interpolation, meteorological data from 120 regions of Korea, including both land and sea areas, are examined. Then, the normalized atmospheric refractivity Nn is calculated for an altitude of 0 m, because the refractivity tends to vary depending on the altitude of the observation site. In addition, the optimal variogram model to obtain the spatial correlation required for the Kriging method is investigated. The estimation accuracy of the Kriging interpolation is compared with that of the inverse distance weighting (IDW) and the bi-linear methods. The average and maximum estimation errors when using the Kriging method are 0.24 and 1.32, respectively. The result demonstrates that the Kriging method is more suitable for the interpolation of atmospheric refractivity in Korea than the conventional methods. Full article
Show Figures

Figure 1

Figure 1
<p>Kriging interpolation.</p>
Full article ">Figure 2
<p>Example of a variogram cloud that includes all the dissimilarities (γ*) for the data set.</p>
Full article ">Figure 3
<p>Experimental and theoretical variograms.</p>
Full article ">Figure 4
<p>Locations of the observatories in Korea.</p>
Full article ">Figure 5
<p>Altitude of the meteorological observatories in Korea.</p>
Full article ">Figure 6
<p>Histograms of the <span class="html-italic">N</span> and <span class="html-italic">N<sub>n</sub></span> for the atmospheric measurement data.</p>
Full article ">Figure 7
<p>Average and maximum estimation errors of the Kriging interpolation according to the variogram models.</p>
Full article ">Figure 8
<p>Comparison of the interpolation results for 1 January 2022: (<b>a</b>) Bi-linear; (<b>b</b>) IDW; (<b>c</b>) Kriging.</p>
Full article ">Figure 9
<p>Kriging interpolation results for the four-season data: (<b>a</b>) winter; (<b>b</b>) spring; (<b>c</b>) summer; (<b>d</b>) fall.</p>
Full article ">Figure 10
<p>Kriging interpolation results for data on January 1 of each year from 2014 to 2024.</p>
Full article ">
18 pages, 3802 KiB  
Article
Influence of Natural Tropical Oscillations on Ozone Content and Meridional Circulation in the Boreal Winter Stratosphere
by Tatiana Ermakova, Andrey Koval, Kseniia Didenko, Olga Aniskina and Arina Okulicheva
Atmosphere 2024, 15(6), 717; https://doi.org/10.3390/atmos15060717 - 15 Jun 2024
Viewed by 861
Abstract
The dependence of ozone content in the polar stratosphere upon different phases of the quasi-biennial oscillation (QBO) of the zonal wind and the El Niño–Southern Oscillation (ENSO) during winter was studied. The monthly (from November to January) mean residual meridional circulation (RMC) was [...] Read more.
The dependence of ozone content in the polar stratosphere upon different phases of the quasi-biennial oscillation (QBO) of the zonal wind and the El Niño–Southern Oscillation (ENSO) during winter was studied. The monthly (from November to January) mean residual meridional circulation (RMC) was calculated for four different combinations of the main phases of ENSO and QBO using MERRA2 reanalysis data. It has been demonstrated that the QBO phase manifests itself in different vertical distributions of ozone in the equatorial stratosphere, as well as in strengthening/weakening of the secondary meridional circulation in the tropics. The enhancement of the RMC from the tropical to the polar stratosphere is stronger at altitudes where ozone is higher in the tropics under El Niño conditions. The RMC modification and intensification are observed from ozone-depleted areas under La Niña conditions. A “cumulative” effect is observed by February under La Niña conditions and the easterly QBO, which is expressed in the lowest ozone content in the polar stratosphere. The numerical experiments carried out using the Middle and Upper Atmosphere Model (MUAM) confirmed tendencies in changes in the meridional transport detected from the reanalysis data for different combinations of QBO and ENSO. Full article
(This article belongs to the Special Issue Ozone Evolution in the Past and Future (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Mean zonal ozone mixing ratio (MLS, (<b>a</b>,<b>c</b>,<b>e</b>)) (MERRA2, (<b>b</b>,<b>d</b>,<b>f</b>)), averaged for November 2009 (<b>a</b>,<b>b</b>), December 2007 (<b>c</b>,<b>d</b>), and January 2011 (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 2
<p>Mean ozone mixing ratio (kg/kg 10<sup>6</sup>) for three winter seasons under La Niña-EQBO (left column) and the difference between La Niña-EQBO and other combinations: La Niña-WQBO (second column, El Niño-EQBO (third column), and El Niño-WQBO (right column) in November (<b>m</b>–<b>p</b>), in December (<b>i</b>–<b>l</b>), in January (<b>e</b>–<b>h</b>), and in February (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 3
<p>Mean zonal ozone mixing ratio during La Niña-EQBO (<b>a</b>) and the difference between La Niña-EQBO and other combinations (<b>b</b>–<b>d</b>)—shading, mean RMC (<b>a</b>), and its increments (<b>b</b>–<b>d</b>)—arrows, mean zonal temperature (<b>a</b>), and its difference (<b>b</b>–<b>d</b>)—contours in November. Dotted areas reveal statistically significant differences in RMC and temperature at 95%. The vertical component is multiplied by 100 (in sm/s).</p>
Full article ">Figure 4
<p>Mean zonal ozone mixing ratio during La Niña-EQBO (<b>a</b>) and the difference between La Niña-EQBO and other combinations (<b>b</b>–<b>d</b>)—shading, mean RMC (<b>a</b>), and its increments (<b>b</b>–<b>d</b>)—arrows, mean zonal temperature (<b>a</b>), and its difference (<b>b</b>–<b>d</b>)—contours in December. Dotted areas reveal statistically significant differences in RMC and temperature at 95%. The vertical component is multiplied by 100 (in sm/s).</p>
Full article ">Figure 5
<p>Mean zonal ozone mixing ratio during La Niña-EQBO (<b>a</b>) and the difference between La Niña-EQBO and other combinations (<b>b</b>–<b>d</b>)—shading, mean RMC (<b>a</b>), and its increments (<b>b</b>–<b>d</b>)—arrows, mean zonal temperature (<b>a</b>), and its difference (<b>b</b>–<b>d</b>)—contours in January. Dotted areas reveal statistically significant differences in RMC and temperature at 95%. The vertical component is multiplied by 100 (in sm/s).</p>
Full article ">Figure 6
<p>Mean zonal temperature during La Niña-EQBO (<b>a</b>) and the difference between La Niña-EQBO and other combinations (<b>b</b>–<b>d</b>)—shading, mean RMC (<b>a</b>), and its increments (<b>b</b>–<b>d</b>)—arrows, simulated with the MUAM model for January. Dotted areas reveal statistically significant differences in RMC and temperature at 95%.</p>
Full article ">
25 pages, 14328 KiB  
Article
Advanced Computer Vision Methods for Tracking Wild Birds from Drone Footage
by Dimitris Mpouziotas, Petros Karvelis and Chrysostomos Stylios
Drones 2024, 8(6), 259; https://doi.org/10.3390/drones8060259 - 12 Jun 2024
Cited by 1 | Viewed by 2100
Abstract
Wildlife conservationists have historically depended on manual methods for the identification and tracking of avian species, to monitor population dynamics and discern potential threats. Nonetheless, many of these techniques present inherent challenges and time constraints. With the advancement in computer vision techniques, automated [...] Read more.
Wildlife conservationists have historically depended on manual methods for the identification and tracking of avian species, to monitor population dynamics and discern potential threats. Nonetheless, many of these techniques present inherent challenges and time constraints. With the advancement in computer vision techniques, automated bird detection and recognition have become possible. This study aimed to further advance the task of detecting wild birds using computer vision methods with drone footage, as well as entirely automating the process of detection and tracking. However, detecting objects from drone footage presents a significant challenge, due to the elevated altitudes, as well as the dynamic movement of both the drone and the birds. In this study, we developed and introduce a state-of-the-art model titled ORACLE (optimized rigorous advanced cutting-edge model for leveraging protection to ecosystems). ORACLE aims to facilitate robust communication across multiple models, with the goal of data retrieval, rigorously using various computer vision techniques such as object detection and multi-object tracking (MOT). The results of ORACLE’s vision models were evaluated at 91.89% mAP at 50% IoU. Full article
Show Figures

Figure 1

Figure 1
<p>An image depicting wildlife surveillance using drones over an islet, recorded at 40 m altitude.</p>
Full article ">Figure 2
<p>Drone surveillance guidelines and altitudes during nesting season.</p>
Full article ">Figure 3
<p>Original drone image image scaled from 3840 × 2160 pixels down to 640 × 640 pixels.</p>
Full article ">Figure 4
<p>The process flow of the ORACLE model.</p>
Full article ">Figure 5
<p>Layer one pre-processing: input frame example with a visible object to detect. Image frames located at Koronisia, Balla (Location at Koronisia, Nisida Balla [<a href="https://www.google.com/maps/search/?api=1&amp;query=39.05285415470534,20.854573719986032" target="_blank">GPS Pin</a>]), 22 June 2023. (<b>a</b>) Original drone image 3840 × 2160 pixels (width by height). (<b>b</b>) target object.</p>
Full article ">Figure 6
<p>Layer two object detection output: target object depicts two detections. The green grid showcases the processed tiles. (<b>a</b>) Original drone image 3840 × 2160 pixels (width by height). (<b>b</b>) Target object layer two output.</p>
Full article ">Figure 7
<p>Layer three: post-detection processing layer: detection merging.</p>
Full article ">Figure 8
<p>Layer four post-detection processing: detection merging.</p>
Full article ">Figure 9
<p>ORACLE fifth layer output in four different time-frames in a video.</p>
Full article ">Figure 9 Cont.
<p>ORACLE fifth layer output in four different time-frames in a video.</p>
Full article ">Figure 10
<p>Application of the image tiling technique in a high-resolution image (4 k). Tiling the image resulted in a total of 24 images sized at 640 × 640 pixels. (<b>a</b>) High-resolution image 3840 × 2160 (4 k) ready to be tiled. (<b>b</b>) A composition of all the tiles of an image, created by our tiling algorithm, including the padded space.</p>
Full article ">Figure 11
<p>The candidate detection closest to the tiles for merging.</p>
Full article ">Figure 12
<p>Descriptive image of the detection merging technique.</p>
Full article ">Figure 13
<p>The effect of Image tiling in object detection in large-scale images. The green line represents the split between the two tiles, whilst the red boxes represent the detections. (<b>a</b>) Target object split in two detections due to image tiling. (<b>b</b>) Successfully merged target object after being split by the two tiles.</p>
Full article ">Figure 14
<p>Training batch of our dataset containing images augmented using the mosaic algorithm.</p>
Full article ">Figure 15
<p>Estimation count of the objects in the annotated video per life span of the tracks, using three different models YOLOv7x (Normal), YOLOv7x (Augmented), YOLOv8x (Augmented), and YOLOv8x-p2 (Augmented). (The results differed from the previously retrieved results due to improvements in the code, <a href="#drones-08-00259-t011" class="html-table">Table 11</a>).</p>
Full article ">Figure 16
<p>A visualization example of our five-layered model displaying advanced object detection, tracking, object count estimation, and track cisualization.</p>
Full article ">
17 pages, 2789 KiB  
Article
ERNet: A Rapid Road Crack Detection Method Using Low-Altitude UAV Remote Sensing Images
by Zexian Duan, Jiahang Liu, Xinpeng Ling, Jinlong Zhang and Zhiheng Liu
Remote Sens. 2024, 16(10), 1741; https://doi.org/10.3390/rs16101741 - 14 May 2024
Cited by 2 | Viewed by 1303
Abstract
The rapid and accurate detection of road cracks is of great significance for road health monitoring, but currently, this work is mainly completed through manual site surveys. Low-altitude UAV remote sensing can provide images with a centimeter-level or even subcentimeter-level ground resolution, which [...] Read more.
The rapid and accurate detection of road cracks is of great significance for road health monitoring, but currently, this work is mainly completed through manual site surveys. Low-altitude UAV remote sensing can provide images with a centimeter-level or even subcentimeter-level ground resolution, which provides a new, efficient, and economical approach for rapid crack detection. Nevertheless, crack detection networks face challenges such as edge blurring and misidentification due to the heterogeneity of road cracks and the complexity of the background. To address these issues, we proposed a real-time edge reconstruction crack detection network (ERNet) that adopted multi-level information aggregation to reconstruct crack edges and improve the accuracy of segmentation between the target and the background. To capture global dependencies across spatial and channel levels, we proposed an efficient bilateral decomposed convolutional attention module (BDAM) that combined depth-separable convolution and dilated convolution to capture global dependencies across the spatial and channel levels. To enhance the accuracy of crack detection, we used a coordinate-based fusion module that integrated spatial, semantic, and edge reconstruction information. In addition, we proposed an automatic measurement of crack information for extracting the crack trunk and its corresponding length and width. The experimental results demonstrated that our network achieved the best balance between accuracy and inference speed compared to six established models. Full article
(This article belongs to the Section Urban Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A structural overview of our proposed network for crack detection.</p>
Full article ">Figure 2
<p>Illustration of the STDC module. (<b>a</b>) STDC module with a stride of 1. (<b>b</b>) STDC module with a stride of 2.</p>
Full article ">Figure 2 Cont.
<p>Illustration of the STDC module. (<b>a</b>) STDC module with a stride of 1. (<b>b</b>) STDC module with a stride of 2.</p>
Full article ">Figure 3
<p>Illustration of the bilateral decomposed convolutional attention module. The BDAM was used to refine the corresponding combined features of the decoding stage. Among them, depth-wise separable convolution and dilated convolution were used to capture the global content, and an attention vector was used for guidance.</p>
Full article ">Figure 4
<p>Illustration of the edge reconstruction module. (<b>a</b>) Edge reconstruction module. (<b>b</b>) Gate convolution.</p>
Full article ">Figure 5
<p>Illustration of the feature fusion module based on coordinates.</p>
Full article ">Figure 6
<p>Diagram of crack skeleton backbone extraction. (<b>a</b>) Crack prediction map. (<b>b</b>) Zhang–Suen thinning algorithm. (<b>c</b>) Crack backbone extraction.</p>
Full article ">Figure 7
<p>Result of the crack based on the distance transformation method. (<b>a</b>) DTM values of the crack prediction map. (<b>b</b>) Result of weighting the crack skeleton with DTM values.</p>
Full article ">Figure 8
<p>A visualization of the different semantic segmentation detection results of the compared methods on Crack500.</p>
Full article ">Figure 9
<p>A visualization of the different semantic segmentation detection results of the compared methods on DeepCrack.</p>
Full article ">Figure 10
<p>A visualization of the different semantic segmentation detection results of the compared methods on the UAV remote sensing dataset.</p>
Full article ">Figure 11
<p>A visualization of different mechanisms. (<b>a</b>) Image; (<b>b</b>) without the BDAM; (<b>c</b>) without the ERM; and (<b>d</b>) with the BDAM and ERM.</p>
Full article ">Figure 12
<p>A visualization of the edge reconstruction results.</p>
Full article ">
Back to TopTop