[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (683)

Search Parameters:
Keywords = altitude pattern

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 5160 KiB  
Article
Methods for Evaluating Tibial Accelerations and Spatiotemporal Gait Parameters during Unsupervised Outdoor Movement
by Amy Silder, Ethan J. Wong, Brian Green, Nicole H. McCloughan and Matthew C. Hoch
Sensors 2024, 24(20), 6667; https://doi.org/10.3390/s24206667 - 16 Oct 2024
Viewed by 263
Abstract
The purpose of this paper is to introduce a method of measuring spatiotemporal gait patterns, tibial accelerations, and heart rate that are matched with high resolution geographical terrain features using publicly available data. These methods were demonstrated using data from 218 Marines, who [...] Read more.
The purpose of this paper is to introduce a method of measuring spatiotemporal gait patterns, tibial accelerations, and heart rate that are matched with high resolution geographical terrain features using publicly available data. These methods were demonstrated using data from 218 Marines, who completed loaded outdoor ruck hikes between 5–20 km over varying terrain. Each participant was instrumented with two inertial measurement units (IMUs) and a GPS watch. Custom code synchronized accelerometer and positional data without a priori sensor synchronization, calibrated orientation of the IMUs in the tibial reference frame, detected and separated only periods of walking or running, and computed acceleration and spatiotemporal outcomes. GPS positional data were georeferenced with geographic information system (GIS) maps to extract terrain features such as slope, altitude, and surface conditions. This paper reveals the ease at which similar data can be gathered among relatively large groups of people with minimal setup and automated data processing. The methods described here can be adapted to other populations and similar ground-based activities such as skiing or trail running. Full article
(This article belongs to the Special Issue Sensor Technologies and Their Applications in Biomechanics)
Show Figures

Figure 1

Figure 1
<p>Participants hiked on primarily fanglomerate surface (hardtop shown above) while carrying a loaded ruck and weapon. (1) Automated calibration of the IMU relative to the tibia assumed primary motion of the leg was flexion-extension. The extent to which this is true varied among participants, as is demonstrated by Marines pictured on the left and right of the picture. (2) Participants often stopped to rest during their hike, as is demonstrated by the Marine pictured center. Periods of rest were excluded from analysis, by removing movements less than 0.45 m/s. Picture courtesy of Amy Silder.</p>
Full article ">Figure 2
<p>(<b>a</b>) A high resolution publicly available Light Detection and Ranging (LIDaR) scan [<a href="#B31-sensors-24-06667" class="html-bibr">31</a>] of the research area was used to calculate (<b>b</b>) slope in degrees and (<b>c</b>) aspect in positive degrees measured from clockwise North. (<b>d</b>) This was combined with the azimuth between successive time points to estimate gradient. The curved line in the figure represents 2453 azimuths for one participant. Arrows indicate the direction of travel and were subsampled for visualization purposes.</p>
Full article ">Figure 3
<p>Speed of one representative participant as a function of distance travelled. Walking was defined as movement 0.45–1.79 m/s [<a href="#B29-sensors-24-06667" class="html-bibr">29</a>] and running as movement 1.80–5.36 m/s: (<b>a</b>) Taken directly from the GPS watch data; (<b>b</b>) Estimated as the distance travelled between two successive time points, divided by the time elapsed between these two successive time points.</p>
Full article ">Figure 4
<p>(<b>a</b>) Elevation gain and (<b>b</b>) the percent of the distance travelled that was spent walking (0.45–1.79 m/s [<a href="#B29-sensors-24-06667" class="html-bibr">29</a>]) and running (1.80–5.36 m/s) for the 218 participants in this study.</p>
Full article ">
14 pages, 4919 KiB  
Article
Phylogenetic Relations and High-Altitude Adaptation in Wild Boar (Sus scrofa), Identified Using Genome-Wide Data
by Shiyong Fang, Haoyuan Zhang, Haoyuan Long, Dongjie Zhang, Hongyue Chen, Xiuqin Yang, Hongmei Pan, Xiao Pan, Di Liu and Guangxin E
Animals 2024, 14(20), 2984; https://doi.org/10.3390/ani14202984 - 16 Oct 2024
Viewed by 254
Abstract
The Qinghai–Tibet Plateau (QTP) wild boar is an excellent model for investigating high-altitude adaptation. In this study, we analyzed genome-wide data from 93 wild boars compiled from various studies worldwide, including the QTP, southern and northern regions of China, Europe, Northeast Asia, and [...] Read more.
The Qinghai–Tibet Plateau (QTP) wild boar is an excellent model for investigating high-altitude adaptation. In this study, we analyzed genome-wide data from 93 wild boars compiled from various studies worldwide, including the QTP, southern and northern regions of China, Europe, Northeast Asia, and Southeast Asia, to explore their phylogenetic patterns and high-altitude adaptation based on genome-wide selection signal analysis and run of homozygosity (ROH) estimation. The findings demonstrate the alignment between the phylogenetic associations among wild boars and their geographical location. An ADMIXTURE analysis indicated a relatively close genetic relationship between QTP and southern Chinese wild boars. Analyses of the fixation index and cross-population extended haplotype homozygosity between populations revealed 295 candidate genes (CDGs) associated with high-altitude adaptation, such as TSC2, TELO2, SLC5A1, and SLC5A4. These CDGs were significantly overrepresented in pathways such as the mammalian target of rapamycin signaling and Fanconi anemia pathways. In addition, 39 ROH islands and numerous selective CDGs (e.g., SLC5A1, SLC5A4, and VCP), which are implicated in glucose metabolism and mitochondrial function, were discovered in QTP wild boars. This study not only assessed the phylogenetic history of QTP wild boars but also advanced our comprehension of the genetic mechanisms underlying the adaptation of wild boars to high altitudes. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Phylogenetic analysis and population structure of worldwide wild boars. (<b>A</b>) Genome-wide phylogenetic trees of wild boar populations. Each color represents the wild boar population in a different region, including the SCN wild boar population (SCN), the NCN wild boar population (NCN), the EU wild boar population (EU), the NEA wild boar population (NEA), the QTP wild boar population (QTP), and the SEA wild boar population (SEA). (<b>B</b>) Principal component analysis, based on all available data, divided into six groups by region. (<b>C</b>) Neighbor-net graph of worldwide wild boar populations using the pairwise difference (Fs<sub>T</sub>). (<b>D</b>) Analysis of the population structure of each wild boar population. The <span class="html-italic">K</span> value is the number of assumed ancestral populations, which was 2 to 5. #: The most reliable <span class="html-italic">K</span> value was 4, which had the minimum CV error. (<b>E</b>) Cross-validation error for each <span class="html-italic">K</span> value (<span class="html-italic">K</span> = 1–10).</p>
Full article ">Figure 2
<p>Population LD decay and demographic history inference analysis of wild boar populations. (<b>A</b>) LD decay of wild boar populations, including the southern Chinese wild boar population (SCN), northern Chinese wild boar population (NCN), European wild boar population (EU), Northeast Asian wild boar population (NEA), and Qinghai–Tibet Plateau wild boar population (QTP). (<b>B</b>) Effective population sizes of different wild boar populations, inferred from autosomes.</p>
Full article ">Figure 3
<p>Genome-wide selective signal analysis of worldwide wild boars to identify the high-altitude adaptability-related genes in Qinghai–Tibet Plateau wild boars. (<b>A</b>) Manhattan map of F<sub>ST</sub> between groups. (<b>B</b>) Manhattan map of XP-EHH between groups.</p>
Full article ">Figure 4
<p>ROH proportions among the populations. The width of the bar chart in the figure is 20KB, and green marks the regions of the genome with ROH population frequencies greater than 25%. (<b>A</b>) ROH proportions in the Qinghai–Tibet Plateau wild boar population. (<b>B</b>) ROH proportions in wild boar populations not distributed on the Qinghai–Tibet Plateau, including the northern Chinese wild boar population, southern Chinese wild boar population, European wild boar population, and Northeast Asian wild boar population.</p>
Full article ">
24 pages, 5179 KiB  
Article
Modeling Multi-Factor Coupled Pressure Fluctuations in EMU Trains under Extreme Tunnel Conditions
by Miao Zou, Chunjun Chen and Lu Yang
Appl. Sci. 2024, 14(20), 9444; https://doi.org/10.3390/app14209444 - 16 Oct 2024
Viewed by 244
Abstract
As an electric multiple unit (EMU) train passes through an extreme tunnel characterized by high altitude, steep gradient, and extended lengths, the pressure waves generated by the train–tunnel aerodynamic coupling combine with the baseline pressure variations within the tunnel. This interaction results in [...] Read more.
As an electric multiple unit (EMU) train passes through an extreme tunnel characterized by high altitude, steep gradient, and extended lengths, the pressure waves generated by the train–tunnel aerodynamic coupling combine with the baseline pressure variations within the tunnel. This interaction results in rapid fluctuations and extreme external pressure with higher amplitudes, which are transmitted into the carriage, causing pressure fluctuations that can adversely affect passenger comfort. These waves interact with multiple factors within the carriage, such as air ducts, airtight gaps, carbody deformation, oxygen supply systems, and temperature, creating a highly nonlinear internal pressure transmission system. This study first establishes a single-factor internal pressure fluctuation model. Subsequently, a multi-factor coupled internal pressure fluctuation model is constructed based on the ideal gas polytropic process assumption and the law of mass conservation. The model parameters are corrected and the model’s effectiveness and accuracy are validated using experimental data to predict and summarize the internal pressure variation patterns of the EMU train during dynamic operation in such tunnels, ensuring safe train operation and meeting the pressure comfort requirements of passengers. Finally, to address the challenges of maintaining and regulating multi-physical variable comfort under extreme tunnel conditions, this study investigates the impact of partial oxygen pressure and temperature on pressure fluctuations and comfort. The study finds that higher oxygen pressure and temperature significantly increase internal pressure fluctuation amplitude, with the oxygen supply system contributing 18.11% and temperature 5.74% of total variation. Thus, setting appropriate standards for oxygen supply, temperature, and valve operation is crucial for mitigating internal pressure fluctuations and enhancing safety and comfort. This research provides a theoretical foundation for developing a comprehensive comfort evaluation and regulation system under harsh environments. Full article
Show Figures

Figure 1

Figure 1
<p>The study process diagram of the internal pressure fluctuation modeling.</p>
Full article ">Figure 2
<p>Diagram of external and internal pressure transmission pathways.</p>
Full article ">Figure 3
<p>On-site static airtightness test.</p>
Full article ">Figure 4
<p>Test results: (<b>a</b>) Measured internal and external pressure data; (<b>b</b>) equivalent leakage area under different internal and external pressure differentials.</p>
Full article ">Figure 5
<p>Flowchart of the correction process for the internal–external pressure transmission model.</p>
Full article ">Figure 6
<p>(<b>a</b>) Values of <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mrow> <mi>r</mi> <mi>m</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> under different pressure loss correction coefficients; (<b>b</b>) comparison between measured and simulated internal pressure of the train passing through Tunnel 1.</p>
Full article ">Figure 7
<p>(<b>a</b>) Comparison of one-dimensional numerical simulation internal pressure with model calculation results; (<b>b</b>) Comparison of measured internal pressure and model calculation results as the train passes through Tunnel 2.</p>
Full article ">Figure 8
<p>Comparison of internal pressure fluctuations with and without considering temperature and oxygen supply.</p>
Full article ">Figure 9
<p>Impact of internal oxygen partial pressure standards on internal pressure fluctuations at different valve openings: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>100</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>75</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>50</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>25</mn> <mo>%</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Internal pressure peaks under different internal oxygen partial pressure standards. (<b>a</b>) Maximum positive pressure; (<b>b</b>) Maximum negative pressure; (<b>c</b>) Peak-to-peak value.</p>
Full article ">Figure 11
<p>Maximum internal pressure variation under different oxygen supply standards.</p>
Full article ">Figure 12
<p>Impact of temperature standards on internal pressure fluctuations at different valve openings: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>100</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>75</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>50</mn> <mo>%</mo> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>25</mn> <mo>%</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Internal pressure peaks under different temperature standards. (<b>a</b>) Maximum positive pressure; (<b>b</b>) Maximum negative pressure; (<b>c</b>) Peak-to-peak value.</p>
Full article ">Figure 14
<p>Maximum internal pressure variation under different temperature standards.</p>
Full article ">
25 pages, 17434 KiB  
Article
Using UAV RGB Images for Assessing Tree Species Diversity in Elevation Gradient of Zao Mountains
by Thi Cam Nhung Tran, Maximo Larry Lopez Caceres, Sergi Garcia i Riera, Marco Conciatori, Yoshiki Kuwabara, Ching-Ying Tsou and Yago Diez
Remote Sens. 2024, 16(20), 3831; https://doi.org/10.3390/rs16203831 (registering DOI) - 15 Oct 2024
Viewed by 349
Abstract
Vegetation biodiversity in mountainous regions is controlled by altitudinal gradients and their corresponding microclimate. Higher temperatures, shorter snow cover periods, and high variability in the precipitation regime might lead to changes in vegetation distribution in mountains all over the world. In this study, [...] Read more.
Vegetation biodiversity in mountainous regions is controlled by altitudinal gradients and their corresponding microclimate. Higher temperatures, shorter snow cover periods, and high variability in the precipitation regime might lead to changes in vegetation distribution in mountains all over the world. In this study, we evaluate vegetation distribution along an altitudinal gradient (1334–1667 m.a.s.l.) in the Zao Mountains, northeastern Japan, by means of alpha diversity indices, including species richness, the Shannon index, and the Simpson index. In order to assess vegetation species and their characteristics along the mountain slope selected, fourteen 50 m × 50 m plots were selected at different altitudes and scanned with RGB cameras attached to Unmanned Aerial Vehicles (UAVs). Image analysis revealed the presence of 12 dominant tree and shrub species of which the number of individuals and heights were validated with fieldwork ground truth data. The results showed a significant variability in species richness along the altitudinal gradient. Species richness ranged from 7 to 11 out of a total of 12 species. Notably, species such as Fagus crenata, despite their low individual numbers, dominated the canopy area. In contrast, shrub species like Quercus crispula and Acer tschonoskii had high individual numbers but covered smaller canopy areas. Tree height correlated well with canopy areas, both representing tree size, which has a strong relationship with species diversity indices. Species such as F. crenata, Q. crispula, Cornus controversa, and others have an established range of altitudinal distribution. At high altitudes (1524–1653 m), the average shrubs’ height is less than 4 m, and the presence of Abies mariesii is negligible because of high mortality rates caused by a severe bark beetle attack. These results highlight the complex interactions between species abundance, canopy area, and altitude, providing valuable insights into vegetation distribution in mountainous regions. However, species diversity indices vary slightly and show some unusually low values without a clear pattern. Overall, these indices are higher at lower altitudes, peak at mid-elevations, and decrease at higher elevations in the study area. Vegetation diversity indices did not show a clear downward trend with altitude but depicted a vegetation composition at different altitudes as controlled by their surrounding environment. Finally, UAVs showed their significant potential for conducting large-scale vegetation surveys reliably and in a short time, with low costs and low manpower. Full article
(This article belongs to the Special Issue Biomass Remote Sensing in Forest Landscapes II)
Show Figures

Figure 1

Figure 1
<p>The location of the study area in the Zao Mountains. Site 1 (mixed forest); Site 2 (transition from mix to monoculture forest); Site 3 (monoculture).</p>
Full article ">Figure 2
<p>The orthomosaics were generated using raw RGB photos in Metashape software v2.1.3.</p>
Full article ">Figure 3
<p>The figure shows the 3D model of Site 1 was generated from the DPC.</p>
Full article ">Figure 4
<p>The 3D Models of Plot 4 with 5 directions, facilitating vegetation visualization.</p>
Full article ">Figure 5
<p>The Canopy Height Models (CHMs) were generated using 3D Models with the software Global Mapper v21.1.</p>
Full article ">Figure 6
<p>An example for one of the posters that were used for fieldwork purposes.</p>
Full article ">Figure 7
<p>Fourteen sample plots were set up in the study area regarding the increase in elevation.</p>
Full article ">Figure 8
<p>Workflow in this study.</p>
Full article ">Figure 9
<p>The number of individuals and the canopy area of dominant species in the 14 plots along the altitudinal gradient.</p>
Full article ">Figure 9 Cont.
<p>The number of individuals and the canopy area of dominant species in the 14 plots along the altitudinal gradient.</p>
Full article ">Figure 10
<p>Change in tree species composition at different altitude layers within the study area.</p>
Full article ">Figure 11
<p>Change in alpha-diversity indices in the plots along the altitudinal gradient (1336–1667 m).</p>
Full article ">
20 pages, 12135 KiB  
Article
Southern South American Long-Distance Pollen Dispersal and Its Relationship with Atmospheric Circulation
by Claudio F. Pérez, Ana G. Ulke and María I. Gassmann
Aerobiology 2024, 2(4), 85-104; https://doi.org/10.3390/aerobiology2040007 (registering DOI) - 12 Oct 2024
Viewed by 338
Abstract
This paper addresses the study of synoptic-scale meteorological conditions that favor long-range pollen transport in southern South America combining airborne pollen counts, modeled three-dimensional backward trajectories, and synoptic and surface meteorological data. Alnus pollen transport trajectories indicate origins predominantly in montane forests of [...] Read more.
This paper addresses the study of synoptic-scale meteorological conditions that favor long-range pollen transport in southern South America combining airborne pollen counts, modeled three-dimensional backward trajectories, and synoptic and surface meteorological data. Alnus pollen transport trajectories indicate origins predominantly in montane forests of the Yungas between 1500 and 2800 m altitude. The South American Low-Level Jet is the main meteorological feature that explains 64% of the detected pollen arrival at the target site. Podocarpus and Nothofagus pollen instead are linked primarily to the widespread Subantartic forests in southern Patagonia. Their transport patterns are consistent with previous studies, which show an association with synoptic patterns related to cold front passages carrying pollen in the free atmosphere (27% for Nothofagus and 25% for Podocarpus). These results show the significance of understanding long-distance pollen transport for disciplines such as climate change reconstruction and agriculture, emphasizing the need for further research to refine atmospheric circulation models and refine interpretations of past vegetation and climate dynamics. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Even-hour <span class="html-italic">Alnus</span> trajectories arriving at 1500 m a.s.l. from 14 UTC of 31 August–12 UTC of 1 September 2013 and (<b>b</b>) 14 UTC of 1 September–12 UTC of 2nd September 2013. The Yungas Forest is shaded in green.</p>
Full article ">Figure 2
<p>Mean geopotential height at 1000 hPa (black solid lines) and 500/1000 hPa thickness fields (gray dashed lines) for the <span class="html-italic">Alnus</span> case study (31 August–1 September 2013). The shaded area shows the highest heights of the Andes (above 1500 m a.s.l.).</p>
Full article ">Figure 3
<p>Images of 850 hPa winds (vectors, m s<sup>−1</sup>) and areas satisfying the modified Bonner’s criteria for (<b>a</b>) 06 UTC 31 August and (<b>b</b>) 06 UTC 1 September showing the position of the cold front. Shading indicates wind speeds at 850 hPa greater than 12, 16, and 20 m s<sup>−1</sup>. White contours indicate a 700/850 hPa wind difference greater than 6, 8, and 10 m s<sup>−1</sup>. Dashed line masks altitudes above 1500 m.</p>
Full article ">Figure 4
<p>Images of the 800–750 hPa layer mean flow for 06 UTC 31 August (<b>a</b>) and 06 UTC 1 September showing the position of the cold front (<b>b</b>). The dashed line marks the 1500 m altitude, while the shaded area masks altitudes higher than 3250 m. The color scale shows the horizontal wind intensity (m s<sup>−1</sup>).</p>
Full article ">Figure 5
<p>Vertical cross sections (30.97° S) showing the horizontal wind (vectors, m s<sup>−1</sup>) and omega (lines, Pa s<sup>−1</sup>) by the end of the SALLJ event. The star shows the position of Sunchales. Panels show the situation every 6 h from 30 August to 1 September 2013. The shaded area shows the Andes and Córdoba ranges. The star indicates the position of Sunchales.</p>
Full article ">Figure 6
<p>Vertical cross sections (19° S) showing the horizontal wind (vectors, m s<sup>−1</sup>) and omega (lines, Pa s<sup>−1</sup>) at the latitude where the SALLJ passes over the Yungas. Panels show the situation every 6 h from 29 August to 31 August 2013 when the event started. The shaded area shows the Andes and Brazilian ranges.</p>
Full article ">Figure 7
<p>Even-hour <span class="html-italic">Nothofagus</span> (<b>a</b>) and <span class="html-italic">Podocarpus</span> (<b>b</b>) trajectories arriving at 750 m a.s.l. on 14 UTC of 24 November–12 UTC of 25 November 2012, and 14 UTC 24 October–12 UTC 25 October 2013, respectively. Light-colored lines show trajectories not passing over the pollen source area (see text). Straight lines represent the construction cuts of the Hovmöller diagrams in <a href="#aerobiology-02-00007-f008" class="html-fig">Figure 8</a> and <a href="#aerobiology-02-00007-f009" class="html-fig">Figure 9</a>. The shaded area shows the geographic distribution of the Subantarctic forests.</p>
Full article ">Figure 8
<p>Hovmöller diagram for <span class="html-italic">Nothofagus</span> case study from 15 November to 1 December 2012. The space cut corresponds to the straight line in <a href="#aerobiology-02-00007-f007" class="html-fig">Figure 7</a>a. Lines show the 700 hPa geopotential height (gpm) and the shaded areas show 700 hPa omega (Pa s<sup>−1</sup>). The lower panel shows the associated topography and the vertical line represents the geographical location of Sunchales.</p>
Full article ">Figure 9
<p>Hovmöller diagram for <span class="html-italic">Podocarpus</span> case study from 15 October to 1 November 2013. The space cut corresponds to the straight line in <a href="#aerobiology-02-00007-f007" class="html-fig">Figure 7</a>b. Lines show the 700 hPa geopotential height (gpm), and shaded areas show 700 hPa omega (Pa s<sup>−1</sup>). The lower panel shows the associated topography and the vertical line represents the geographical location of Sunchales.</p>
Full article ">Figure A1
<p>Cartoons describing the transient synoptic patterns (see <a href="#aerobiology-02-00007-t001" class="html-table">Table 1</a>, <a href="#aerobiology-02-00007-t002" class="html-table">Table 2</a> and <a href="#aerobiology-02-00007-t003" class="html-table">Table 3</a>) recognized for <span class="html-italic">Alnus</span>, <span class="html-italic">Nothofagus</span>, and <span class="html-italic">Podocarpus</span> pollen arrival at Sunchales. The red star shows the city’s location. (<b>a</b>) leading-edge trough, (<b>b</b>) trough–eastern high, (<b>c</b>) low–eastern high, (<b>d</b>) weak high, (<b>e</b>) eastern high, (<b>f</b>) weak low, (<b>g</b>) ridge, (<b>h</b>) trough, (<b>i</b>) post-frontal, (<b>j</b>) low, (<b>k</b>) high.</p>
Full article ">
20 pages, 4375 KiB  
Article
Differentiating Growth Patterns in Winter Wheat Cultivars via Unmanned Aerial Vehicle Imaging
by Asparuh I. Atanasov, Hristo P. Stoyanov and Atanas Z. Atanasov
AgriEngineering 2024, 6(4), 3652-3671; https://doi.org/10.3390/agriengineering6040208 - 7 Oct 2024
Viewed by 448
Abstract
Wheat is one of the most widely grown cereal crops, serving as a key factor in sustaining the nutritional and food balance in numerous countries. The use of non-contact methods for wheat monitoring allows for the rapid diagnosis of vegetation density, crop growth, [...] Read more.
Wheat is one of the most widely grown cereal crops, serving as a key factor in sustaining the nutritional and food balance in numerous countries. The use of non-contact methods for wheat monitoring allows for the rapid diagnosis of vegetation density, crop growth, and the presence of weeds and diseases in the investigated fields. This study aims to assess the potential for differentiating growth patterns in winter wheat cultivars by examining them with two unmanned aerial vehicles (UAVs), the Mavic 2 Pro and Phantom 4 Pro, equipped with a multispectral camera from the MAPIR™ brand. Based on an experimental study conducted in the Southern Dobruja region (Bulgaria), vegetation reflectance indices, such as the Normalized-Difference Vegetation Index (NDVI), Soil-Adjusted Vegetation Index (SAVI), and Enhanced Vegetation Index 2 (EVI2), were generated, and a database was created to track their changing trends. The obtained results showed that the values of the NDVI, EVI2, and SAVI can be used to predict the productive potential of wheat, but only after accounting for the meteorological conditions of the respective growing season. The proposed methodology provides accurate results in small areas, with a resolution of 0.40 cm/pixel when flying at an altitude of 12 m and 2.3 cm/pixel when flying at an altitude of 100 m. The achieved precision in small and ultra-small agricultural areas, at a width of 1.2 m, will help wheat breeders conduct precise diagnostics of individual wheat varieties. Full article
(This article belongs to the Special Issue Computer Vision for Agriculture and Smart Farming)
Show Figures

Figure 1

Figure 1
<p>UAV DJI Mavic 2 Pro, MAPIR Survey3W Camera (<b>a</b>). DJI Phantom 4 Pro, MAPIR Survey3W Camera (<b>b</b>).</p>
Full article ">Figure 2
<p>RGB image of the field (<b>a</b>), RGN image of the field (<b>b</b>), and generated NDVI (<b>c</b>).</p>
Full article ">Figure 3
<p>Spectral characteristics and numerical value of colors extracted from one plot (<b>a</b>) and spectral characteristics of the MAPIR camera (<b>b</b>).</p>
Full article ">Figure 4
<p>The development of winter wheat during the vegetation stage for the 2021–2022 season.</p>
Full article ">Figure 5
<p>Air temperature and precipitation for the 2021–2022 growing season [<a href="#B36-agriengineering-06-00208" class="html-bibr">36</a>].</p>
Full article ">Figure 6
<p>Air temperature and precipitation for the 2022–2023 growing season [<a href="#B36-agriengineering-06-00208" class="html-bibr">36</a>].</p>
Full article ">Figure 7
<p>Trends of NDVI change by variety for the 2021/2022 growing season.</p>
Full article ">Figure 8
<p>Trends of changes in the EVI2 (<b>a</b>) and SAVI (<b>b</b>) by variety for 2021–2022 throughout the growing season.</p>
Full article ">Figure 9
<p>Trends of NDVI change by variety 2021–2022.</p>
Full article ">Figure 10
<p>Trends of changes in the EVI2 (<b>a</b>) and SAVI (<b>b</b>) by variety for 2022–2023 throughout the growing season.</p>
Full article ">
20 pages, 4810 KiB  
Article
Understanding Spatio-Temporal Hydrological Dynamics Using SWAT: A Case Study in the Pativilca Basin
by Yenica Pachac-Huerta, Waldo Lavado-Casimiro, Melania Zapana and Robinson Peña
Hydrology 2024, 11(10), 165; https://doi.org/10.3390/hydrology11100165 - 4 Oct 2024
Viewed by 645
Abstract
This study investigates the hydrological dynamics of the Pativilca Basin in the Southern Hemisphere using the SWAT (Soil and Water Assessment Tool) model. Seventy-seven watersheds across a mountainous region were analyzed using elevation data, land cover, soil type, and gridded meteorological products (RAIN4PE [...] Read more.
This study investigates the hydrological dynamics of the Pativilca Basin in the Southern Hemisphere using the SWAT (Soil and Water Assessment Tool) model. Seventy-seven watersheds across a mountainous region were analyzed using elevation data, land cover, soil type, and gridded meteorological products (RAIN4PE and PISCO) for hydrological simulations. Watershed delineation, aided by a Digital Elevation Model, enabled the identification of critical drainage points and the definition of Hydrological Response Units (HRUs). The model calibration and validation, performed using the SWAT-CUP with the SUFI-2 algorithm, achieved Nash–Sutcliffe Efficiency (NSE) values of 0.69 and 0.72, respectively. Cluster analysis categorized the watersheds into six distinct groups with unique hydrological and climatic characteristics. The results showed significant spatial variability in the precipitation and temperature, with pronounced seasonality influencing the daily flow patterns. The higher-altitude watersheds exhibited greater soil water storage and more effective aquifer recharge, whereas the lower-altitude watersheds, despite receiving less precipitation, displayed higher flows due to runoff from the upstream areas. These findings emphasize the importance of incorporating seasonality and spatial variability into water resource planning in mountainous regions and demonstrate the SWAT model’s effectiveness in predicting hydrological responses in the Pativilca Basin, laying the groundwork for future research in mountain hydrology. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical map of the Pativilca River Basin (<b>a</b>) study area in Peru; (<b>b</b>) study area in Ancash and Lima regions; (<b>c</b>) study area with elevation and rivers in the basin.</p>
Full article ">Figure 2
<p>Spatial distribution of slope, land cover, and type soil in the Pativilca Basin. (<b>a</b>) Shows how the slope changes, with steeper areas mostly up in the upper part of the basin; (<b>b</b>) maps out the land cover, including vegetation, farms, and urban spots; and (<b>c</b>) highlights the soil types, showing how they affect water retention and erosion throughout the basin.</p>
Full article ">Figure 3
<p>Methodological flowchart.</p>
Full article ">Figure 4
<p>Cluster dendrogram for the regionalization of catchments in the Pativilca Basin. The dendrogram delineates six distinct catchment groups (A–F), represented by color-coded branches. Each group’s representative catchment is highlighted in pink. The vertical axis reflects the degree of dissimilarity between the catchments, with greater heights indicating higher dissimilarity. This regionalization was achieved using hierarchical clustering based on Euclidean distances, facilitating the identification of hydrologically similar catchment groups for further analysis.</p>
Full article ">Figure 5
<p>Regionalization of watersheds in the Pativilca Basin and selection of representative watersheds.</p>
Full article ">Figure 6
<p>Seasonal variations in precipitation, maximum, and minimum temperatures in the Pativilca Basin regions. The first column (blue bars) represents monthly precipitation, while red bars indicate maximum temperatures and orange bars depict minimum temperatures. The groups are arranged vertically from top to bottom, starting with Group A at the uppermost position and concluding with Group F at the lowest. These graphs highlight the temporal distribution and variability in key climatic variables across different seasons, enabling the assessment of seasonal trends and their impact on hydrological processes in the basin.</p>
Full article ">Figure 7
<p>Calibration and validation at the Cahua hydrometric station.</p>
Full article ">Figure 8
<p>Spatial distribution of hydrological components in the Pativilca Basin. The hydrological components include (<b>a</b>) flow out daily mean (<span class="html-italic">Q</span>) and annual precipitation (<span class="html-italic">R<sub>d</sub></span>), (<b>b</b>) evapotranspiration (ET), (<b>c</b>) percolation (<span class="html-italic">W<sub>seep</sub></span>), (<b>d</b>) groundwater contribution to streamflow (<span class="html-italic">Q<sub>gw</sub></span>), (<b>e</b>) average daily soil water storage (SW), and (<b>f</b>) water yield (<span class="html-italic">W<sub>YLD</sub></span>). Each map illustrates the spatial variability across the basin, highlighting the hydrological dynamics. The representative watersheds are bordered in red, indicating their respective groups at the center. Group boundaries are depicted with black dotted lines, enhancing the differentiation between zones. These visual elements allow for a detailed analysis of the distribution and influence of key hydrological processes across the basin’s distinct regions.</p>
Full article ">Figure 9
<p>Temporal distribution of streamflow in the Pativilca Basin. Daily streamflows from 1981 to 2015 show a clear seasonal pattern, with peak flows during the wet season (January to March) and lows in the dry season (June to September). Flow variation is driven by altitude, storage capacity, and watershed connectivity, with lower watersheds redistributing water from upstream areas.</p>
Full article ">Figure 9 Cont.
<p>Temporal distribution of streamflow in the Pativilca Basin. Daily streamflows from 1981 to 2015 show a clear seasonal pattern, with peak flows during the wet season (January to March) and lows in the dry season (June to September). Flow variation is driven by altitude, storage capacity, and watershed connectivity, with lower watersheds redistributing water from upstream areas.</p>
Full article ">
31 pages, 41260 KiB  
Article
Remote Sensing Evaluation and Monitoring of Spatial and Temporal Changes in Ecological Environmental Quality in Coal Mining-Intensive Cities
by Qiqi Huo, Xiaoqian Cheng, Weibing Du, Hao Zhang and Ruimei Han
Appl. Sci. 2024, 14(19), 8814; https://doi.org/10.3390/app14198814 - 30 Sep 2024
Viewed by 654
Abstract
In coal-dependent urban economies, the dichotomy between resource exploitation and ecological conservation presents a pronounced challenge. Traditional remote sensing ecological assessments often overlook the interplay between mining activities and urban environmental dynamics. To address this gap, researchers developed an innovative Resource-Based City Ecological [...] Read more.
In coal-dependent urban economies, the dichotomy between resource exploitation and ecological conservation presents a pronounced challenge. Traditional remote sensing ecological assessments often overlook the interplay between mining activities and urban environmental dynamics. To address this gap, researchers developed an innovative Resource-Based City Ecological Index (RCEI), anchored in a Pressure–State–Response (PSR) framework and synthesized from six discrete ecological indicators. Utilizing geodetic remote sensing data, the RCEI facilitated a comprehensive spatiotemporal analysis of Jincheng City’s ecological quality from 1990 to 2022. The findings corroborated the RCEI’s efficacy in providing a nuanced portrayal of the ecological state within mining regions. (1) Jincheng City’s ecological quality predominantly sustained a mudhopper-tier status, exhibiting an overarching trend of amelioration throughout the study period. (2) Disparities in ecological landscape quality were pronounced at the county level, with Moran’s Index exceeding 0.9, signifying a clustered ecological quality pattern. High–high (H–H) zones were prevalent in areas of elevated altitude and dense vegetation, whereas low–low (L–L) zones were prevalent in urban and mining sectors. (3) Further, a buffer zone analysis of two coal mines, differing in their mining chronology, geographical positioning, and operational status, elucidated the ecological impact exerted over a 32-year trajectory. These insights furnish a robust scientific and technical foundation for resource-centric cities to fortify ecological safeguarding and to advance sustainable development stratagems. Full article
(This article belongs to the Section Ecology Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>Geographical location of Jincheng City.</p>
Full article ">Figure 2
<p>The overall workflow of this study.</p>
Full article ">Figure 3
<p>Trends in the RCEI and RSEI eigenvalues and contributions. (<b>a</b>) Comparison of trends in RCEI and RSEI eigenvalues; (<b>b</b>) Comparison of changes in RCEI and RSEI contributions.</p>
Full article ">Figure 4
<p>Results of the principal component analysis of indicators from 1990 to 2022.</p>
Full article ">Figure 5
<p>A boxplot of the RCEI in Jincheng City from 1990 to 2022.</p>
Full article ">Figure 6
<p>Characteristics of the spatial distribution of the ecological environmental quality level in Jincheng City in 1990–2022.</p>
Full article ">Figure 7
<p>Annual trends of the average RCEI values in county-level regions of Jincheng City.</p>
Full article ">Figure 8
<p>Changing trend in the RCEI for the horizontal area in Jincheng City.</p>
Full article ">Figure 9
<p>Images of the RCEI change monitoring in Jincheng.</p>
Full article ">Figure 10
<p>Land use mapping of Jincheng City in 1990–2022. (<b>a</b>) City land use classification. (<b>b</b>) Land use changes.</p>
Full article ">Figure 11
<p>Positions of different RCEI sample loci in an OLI image, as well as in an RCEI image and OLI image (RGB) for different RCEI levels. (<b>a</b>) Location of samples with different RCEI levels in the RCEI image, (<b>b</b>) Location of samples with different RCEI levels in OLI images, (<b>c</b>) Images of national parks with excellent RCEI levels, (<b>d</b>) OLI images of the Sihe mine with poor RCEI levels, (<b>e</b>) OLI images of the YiCheng mine with poor RCEI levels, (<b>f</b>) OLI images of the Gu mine located in urban areas with poor RCEI levels, (<b>g</b>) Images of national parks with good RCEI levels located on the urban fringe, (<b>h</b>) OLI images of national wetlands with excellent RCEI levels.</p>
Full article ">Figure 12
<p>3D scatterplot of the WET, NDVI, EVI, NDBSI, LST, ICDI and RCEI at sampling points: (<b>a</b>) 3D spatial relationship between the RCEI, NDVI, and EVI; (<b>b</b>) 3D spatial relations between the RCEI, NDBSI, and WET; (<b>c</b>) 3D spatial relations between the RCEI, ICDI, and LST.</p>
Full article ">Figure 13
<p>Correlation analysis between the RCEI and six ecological indicators from 1990 to 2022.</p>
Full article ">Figure 14
<p>Scatterplot of the RCEI Moran’s I in Jincheng from 1990 to 2022.</p>
Full article ">Figure 15
<p>LISA clustering of the RCEI in Jincheng City from 1990 to 2022.</p>
Full article ">Figure 16
<p>Areas producing a million tons of coal and their mean RCEI values with distance. (<b>a</b>) Location of Jincheng City’s Sihe Mine Area and Fenghuangshan Mine Area; (<b>b</b>,<b>c</b>) Raster plots showing the mean RCEI values of the Sihe Mine Area and Fenghuangshan Mine Area with distance; (<b>d</b>,<b>e</b>) Line plots depicting the mean RCEI values of the Fenghuangshan Mine Area and Sihe Mine Area with buffer zones from 1990 to 2022.</p>
Full article ">
16 pages, 1616 KiB  
Article
Species Richness and Similarity of New Zealand Mayfly Communities (Ephemeroptera) Decline with Increasing Latitude and Altitude
by Stephen R. Pohe, Michael J. Winterbourn and Jon S. Harding
Insects 2024, 15(10), 757; https://doi.org/10.3390/insects15100757 - 29 Sep 2024
Viewed by 493
Abstract
The distribution of species in relation to latitude and altitude is of fundamental interest to ecologists and is expected to attain increasing importance as the Earth’s climate continues to change. Species diversity is commonly greater at lower than higher latitudes on a global [...] Read more.
The distribution of species in relation to latitude and altitude is of fundamental interest to ecologists and is expected to attain increasing importance as the Earth’s climate continues to change. Species diversity is commonly greater at lower than higher latitudes on a global scale, and the similarity of communities frequently decreases with distance. Nevertheless, reasons for such patterns are not well understood. We investigated species richness and changes in community composition of mayflies (Ephemeroptera) over 13 degrees of latitude at 81 locations throughout New Zealand by light-trapping and the benthic sampling of streams. Mayflies were also sampled along an altitudinal gradient on a prominent inactive volcano in the east of North Island. Sampled streams were predominantly in the native forest, at a wide range of altitudes from sea level to c. 1000 m a. s. l. A total of 47 of the 59 described New Zealand mayflies were recorded during the study, along with five undescribed morphospecies. Species richness declined and the degree of dissimilarity (beta diversity) of mayfly communities increased significantly from north to south but less strongly with increasing altitude. Our results suggest that the southward decline in species richness has historical origins with the north of the country having acted as a major refuge and region of speciation during the Pleistocene. The increasing dissimilarity of the northern and southern communities may reflect an increasingly harsh climate, variable amounts of subsequent southward dispersal of northern species and, in the South Island, the presence of species which may have evolved in the newly uplifted mountains during the Miocene–Pliocene. Full article
(This article belongs to the Section Insect Ecology, Diversity and Conservation)
Show Figures

Figure 1

Figure 1
<p>The 81 locations (243 sites) within New Zealand where sampling of mayfly communities was undertaken. Latitudinal lines delineate the zones used in analyses. The inset shows the locations and elevations of the seven sites used in the Mt Taranaki altitude study. The 600 m site was also sampled in the latitudinal survey. The black line in the inset indicates the boundary of Egmont National Park.</p>
Full article ">Figure 2
<p>NMS ordination based on presence–absence data for mayfly species recorded at the 81 sampling locations. Stress = 15.3. The different colored symbols represent North Island (blue), South Island (red) and Stewart Island (grey) locations. Note: Stewart Island lies off the southern coast of South Island.</p>
Full article ">Figure 3
<p>Mayfly species richness at all locations in relation to latitude. Richness declined significantly from north to south (<span class="html-italic">r</span><sup>2</sup> = 0.67; <span class="html-italic">p</span> &lt; 0.001; <span class="html-italic">n</span> = 81).</p>
Full article ">Figure 4
<p>Mayfly species richness in relation to altitude. Richness declined significantly as altitude increased (<span class="html-italic">r</span><sup>2</sup> = 0.13; <span class="html-italic">p</span> &lt; 0.01; <span class="html-italic">n</span> = 81).</p>
Full article ">
23 pages, 10753 KiB  
Project Report
Environmental Factors Drive the Biogeographic Pattern of Hippophae rhamnoides Root Endophytic Fungal Diversity in the Arid Regions of Northwest China
by Siyu Guo, Guisheng Ye, Wenjie Liu, Ruoqi Liu, Zhehao Liu and Yuhua Ma
J. Fungi 2024, 10(10), 679; https://doi.org/10.3390/jof10100679 - 29 Sep 2024
Viewed by 390
Abstract
Hippophae rhamnoides subsp. sinensis Rousi (Abbrev. H. rhamnoides) stands as a vital botanical asset in ameliorating the ecological landscape of the arid regions in Northwest China, where its rhizospheric microorganisms serve as linchpins in its growth and developmental dynamics. This study aimed [...] Read more.
Hippophae rhamnoides subsp. sinensis Rousi (Abbrev. H. rhamnoides) stands as a vital botanical asset in ameliorating the ecological landscape of the arid regions in Northwest China, where its rhizospheric microorganisms serve as linchpins in its growth and developmental dynamics. This study aimed to explore the community structure characteristics and origin differences of root endophytic fungi in H. rhamnoides. Samples were collected from 25 areas where H. rhamnoides is naturally distributed along an altitude gradient in the northwest region. Then, endophytic fungi from different regions were analyzed by using high-throughput sequencing technology to compare the structural characteristics of endophytic fungi and examine their association with environmental factors. FUNGuild was employed to analyze the community structure and functions of endophytic fungi, and the results showed that each region had its own dominant endophytic fungal flora, demonstrating the differences in origin of endophytic fungi, and the specific endophytic flora acquired from the original soil in the growing season of H. rhamnoides will help us construct the microecological community structure. Furthermore, the study identified and assessed the diversity of fungi, elucidating the species structure and highlighting dominant species. The RDA analysis revealed that available phosphorus (AP), available potassium (AK), and total nitrogen (TN) exhibit significant correlations with the composition and diversity of root-associated fungi. In conclusion, the fungal community structure is similar within the same region, while significant differences exist in the taxonomic structure and biodiversity among different regions. These findings shed light on the intricate interplay and mechanisms governing the ecological restoration of H. rhamnoides, offering a valuable framework for advancing green ecology initiatives and harnessing the potential of root-associated microorganisms in this species. Full article
(This article belongs to the Section Fungal Evolution, Biodiversity and Systematics)
Show Figures

Figure 1

Figure 1
<p>Map of the 25 field sites for fungal communities in the root rhizosphere and endosphere of <span class="html-italic">H. rhamnoides</span> in the arid regions of Northwest China.</p>
Full article ">Figure 2
<p>Rarefaction curve analysis for all samples. The horizontal axis represents the sequencing depth, while the vertical axis represents the corresponding alpha diversity index. When the curve levels off, it indicates that the sequencing depth has reached a reasonable point, and additional data will not significantly impact the alpha diversity index.</p>
Full article ">Figure 3
<p>Venn diagrams of endophytic fungus species in <span class="html-italic">H. rhamnoides</span> roots at different altitudes at the ASV level. (<b>a</b>) Low altitude; (<b>b</b>) mid-altitude; (<b>c</b>) high altitude. The colors represent different plots, and the numbers represent ASV counts.</p>
Full article ">Figure 4
<p>Shannon index, observed features, dominance, Chao1 index, pielou-e, and Simpson index inter-group difference box plot. The horizontal axis of the box plot represents the groups, while the vertical axis represents the corresponding alpha diversity index values. Different colors represent different plots.</p>
Full article ">Figure 5
<p>The relationship between endophytic fungus alpha diversity (Shannon and Chao1 indices) and geographical factors (longitude, latitude, and altitude). The different colored circles each represent samples from different sites.</p>
Full article ">Figure 6
<p>Weighted and unweighted Unifrac distance box plots. Beta diversity analysis reflects the composition of biological communities between different samples. Different colors represent different plots.</p>
Full article ">Figure 7
<p>NMDS distribution of fungi in rhizomes of <span class="html-italic">H. rhamnoides</span> at different altitudes. (<b>a</b>) Low altitude; (<b>b</b>) mid-altitude; (<b>c</b>) high altitude. NMDS analysis represents samples as points in a multidimensional space, where the degree of difference between samples is reflected by the distance between points. This analysis illustrates both inter-group and intra-group variations among the samples.</p>
Full article ">Figure 8
<p>Weighted Unifrac distance 2D PCoA diagrams. The horizontal axis represents one principal component, while the vertical axis represents another principal component. The percentage indicates the contribution of each principal component to the variation among samples. Each point in the plot represents a sample, with samples from the same group denoted by the same color.</p>
Full article ">Figure 9
<p>Representative sequences of the top 100 genera. (The colors of branches and sectors indicate their corresponding phylum, while the stacked column chart outside the sector ring represents the abundance distribution information of the genus in different samples).</p>
Full article ">Figure 10
<p>UPGMA clustering tree based on weighted Unifrac distance. On the left is the UPGMA clustering tree structure, and on the right is the relative abundance distribution of species at the phylum level for each sample.</p>
Full article ">Figure 11
<p>Environmental factors of rhizosphere soil of <span class="html-italic">H. rhamnoide</span> and db-RDA analysis. The axes represent major variation components, with arrows indicating the direction and strength of environmental variables. Points represent samples or species, and the proximity to an arrow suggests a stronger association with that variable. Longer arrows indicate a greater influence of the environmental variable.</p>
Full article ">Figure 12
<p>Mantel test correlation heat map of fungi (family, genus, species, and ASV) and environmental factors (AMT, AMP, salinity, altitude, latitude, longitude, AP, OM, AK, TP, HN, TK, TN, and pH) in roots. The colors in the heat map represent the strength of the correlation. The significance levels are as follows: <span class="html-italic">p</span> &lt; 0.05, one asterisk (*); <span class="html-italic">p</span> &lt; 0.01, two asterisks (**); <span class="html-italic">p</span> &lt; 0.001, three asterisks (***); <span class="html-italic">p</span> &lt; 0.0001, four asterisks (****).</p>
Full article ">Figure 13
<p>The relationship between soil physicochemical (TN, HN, OM, and pH) and geographical factors (AMT, AMP, altitude, latitude, and longitude). The different colored circles each represent samples from different sites.</p>
Full article ">Figure 14
<p>Relative abundance plot. The horizontal axis represents the sample names; the vertical axis indicates the relative abundance. “Others” represents the sum of relative abundances for all functional information not included among the ten features shown in the figure. (<b>a</b>) Relative abundance of different trophic types; (<b>b</b>) relative abundance bar diagram of ecological function groups.</p>
Full article ">Figure 15
<p>Functional annotations and their abundance information. The heat map displays the correlation between distance matrices, with each cell representing the correlation coefficient between two matrices. The colors indicate the strength and direction of the correlation. The significance levels are as follows: <span class="html-italic">p</span> &lt; 0.05, one asterisk (*); <span class="html-italic">p</span> &lt; 0.01, two asterisks (**); <span class="html-italic">p</span> &lt; 0.001, three asterisks (***); <span class="html-italic">p</span> &lt; 0.0001, four asterisks (****).</p>
Full article ">Figure 16
<p>Molecular network diagram of intra-root fungi. The size of each genus represents its average relative abundance. Nodes of the same color represent the same phylum, and the thickness of the edges between nodes is proportional to the absolute value of the species interaction correlation coefficient. Red edges represent positive correlations between genera.</p>
Full article ">
17 pages, 12703 KiB  
Article
Historical Landscape: A Methodological Proposal to Analyse the Settlements of Monasteries in the Birth of Portugal
by Isabel Vaz de Freitas, Hélder Silva Lopes and Helena Albuquerque
Religions 2024, 15(10), 1158; https://doi.org/10.3390/rel15101158 - 25 Sep 2024
Viewed by 681
Abstract
This study aims to understand and characterise the landscape of monasteries in early medieval Portugal using a methodology to better comprehend the factors influencing monastery construction. The research focuses on variables such as altitude, slope, aspect, hydrology, geomorphology, and topographic prominence. Using Geographic [...] Read more.
This study aims to understand and characterise the landscape of monasteries in early medieval Portugal using a methodology to better comprehend the factors influencing monastery construction. The research focuses on variables such as altitude, slope, aspect, hydrology, geomorphology, and topographic prominence. Using Geographic Information Systems (GIS) for detailed spatial analysis, the study reveals that monasteries were typically located in areas with slight elevations, gentle slopes, and proximity to watercourses, reflecting considerations about resource exploitations, access, and population development. The analysis shows no significant differences in construction preferences among different religious orders, indicating a general adaptability to the local environment rather than distinct criteria for each order. Despite the broad trends, individual orders exhibited some variability in their specific site selections, such as altitude and slope preferences. The findings highlight the importance of integrating historical and environmental data to understand settlement patterns, providing valuable insights into the strategic considerations behind monastery locations. Future research could expand on these findings by incorporating socio-economic impacts, enhancing our understanding of medieval monastic landscapes. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The distribution of the religious orders in mainland Portugal. (<b>b</b>) The reconquest in the period of D. Afonso Henriques. Source: Authors’ own elaborations based on <a href="#B3-religions-15-01158" class="html-bibr">Barroca</a> (<a href="#B3-religions-15-01158" class="html-bibr">2003, p. 49</a>).</p>
Full article ">Figure 2
<p>The distribution of the religious orders between important rivers in mainland Portugal. Source: Authors’ own elaborations based on compiled data.</p>
Full article ">Figure 3
<p>Some of the variables used in modelling. (<b>A</b>) Topographic position index (800 m threshold); (<b>B</b>) slope (in degrees); (<b>C</b>) aspect; and (<b>D</b>) distance from rivers. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 4
<p>Distance between monasteries. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 5
<p>Relationship between religious orders and slope. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 6
<p>Relationship between religious orders and slope. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 7
<p>Relationship between religious orders and elevation. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 8
<p>Characterisation of the elevation (in m). Source: Own elaboration based on compiled data.</p>
Full article ">Figure 9
<p>Relationship between religious orders and distance to rivers. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 10
<p>Characterisation of the distance to rivers (m) in monasteries of different religious orders. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 11
<p>Relationship between religious orders and aspects. Source: Own elaboration based on compiled data.</p>
Full article ">Figure 12
<p>Characterisation of the aspect in monasteries of different religious orders. Source: Own elaboration based on compiled data.</p>
Full article ">
17 pages, 6828 KiB  
Article
Relationships between Precipitation and Elevation in the Southeastern Tibetan Plateau during the Active Phase of the Indian Monsoon
by Lun Luo, Yanggang Zhao, Yanghai Duan, Zeng Dan, Sunil Acharya, Gesang Jimi, Pan Bai, Jie Yan, Liang Chen, Bin Yang and Tianli Xu
Water 2024, 16(18), 2700; https://doi.org/10.3390/w16182700 - 23 Sep 2024
Viewed by 406
Abstract
The precipitation gradient (PG) is a crucial parameter for watershed hydrological models. Analysis of daily precipitation and elevation data from 30 stations in the southeastern Tibetan Plateau (SETP) during the active phase of the Indian monsoon reveals distinct patterns. Below 3000 m, precipitation [...] Read more.
The precipitation gradient (PG) is a crucial parameter for watershed hydrological models. Analysis of daily precipitation and elevation data from 30 stations in the southeastern Tibetan Plateau (SETP) during the active phase of the Indian monsoon reveals distinct patterns. Below 3000 m, precipitation generally decreases with increasing altitude. Between 3000 and 4000 m, precipitation patterns are more complex; in western regions, precipitation increases with elevation, whereas in eastern regions, it decreases. Above 4000 m, up to the highest observation point of 4841 m, precipitation continues to decrease with elevation, with a more pronounced decline beyond a critical height. In the SETP, PGs for LYR and NYR are positive, at 11.3 ± 2.7 mm/100 m and 17.3 ± 3.8 mm/100 m, respectively. Conversely, PLZB exhibits a negative PG of −22.3 ± 4.2 mm/100 m. The Yarlung Zangbo River (YLZBR) water vapor channel plays a significant role in these PGs, with the direction and flux of water vapor potentially influencing both the direction and magnitude of the PG. Additional factors such as precipitation intensity, the number of precipitation days, precipitation frequency, and station selection also significantly impact the PG. Notable correlations between elevation and variables such as the number of precipitation days, non-precipitation days, and precipitation intensity. The precipitation intensity gradients (PIGs) are 0.06 ± 0.02 mm/d/100 m, 0.11 ± 0.04 mm/d/100 m, and −0.18 ± 0.04 mm/d/100 m for the three catchments, respectively. Future research should incorporate remote sensing data and expand site networks, particularly in regions above 5000 m, to enhance the accuracy of precipitation–elevation relationship assessments, providing more reliable data for water resource simulation and disaster warning. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of precipitation observation stations in the SETP.</p>
Full article ">Figure 2
<p>Elevation distribution of the three catchment stations.</p>
Full article ">Figure 3
<p>Cumulative precipitation of different rainfall levels during the active phase of the Indian monsoon at various stations in the SETP.</p>
Full article ">Figure 4
<p>Average precipitation across different precipitation intensities in each river catchment, with line segments representing the range of standard deviations.</p>
Full article ">Figure 5
<p>The relationship between precipitation and elevation across different rain levels for the three study catchments (*: <span class="html-italic">p</span> ≤ 0.05,**: <span class="html-italic">p</span> ≤ 0.01 ). (<b>a</b>–<b>d</b>) represent nongraded precipitation, light precipitation, moderate precipitation, and heavy precipitation, respectively.</p>
Full article ">Figure 6
<p>Regression of rainfall days (<b>a</b>) and no rainfall days (<b>b</b>) with elevation in the river catchments of the SETP (**: <span class="html-italic">p</span> ≤ 0.01 ).</p>
Full article ">Figure 7
<p>The relationship between precipitation intensity and elevation across different rain levels for the three study catchments (*: <span class="html-italic">p</span> ≤ 0.05,**: <span class="html-italic">p</span> ≤ 0.01 ). (<b>a</b>–<b>d</b>) represent nongraded precipitation, light precipitation, moderate precipitation, and heavy precipitation, respectively.</p>
Full article ">Figure 8
<p>Distribution of cumulative precipitation frequency at each station within the catchment: (<b>a</b>) LYR, (<b>b</b>) NYR, and (<b>c</b>) PLZB.</p>
Full article ">Figure 9
<p>The correlation matrix (R<sup>2</sup>) of daily precipitation at each station in the SETP. The blue box represents LYR, the purple box represents NYR, and the green box represents PLZB.</p>
Full article ">
23 pages, 21344 KiB  
Article
Vertical Structure of Heavy Rainfall Events in Brazil
by Eliana Cristine Gatti, Izabelly Carvalho da Costa and Daniel Vila
Meteorology 2024, 3(3), 310-332; https://doi.org/10.3390/meteorology3030016 - 23 Sep 2024
Viewed by 376
Abstract
Intense rainfall events frequently occur in Brazil, often leading to rapid flooding. Despite their recurrence, there is a notable lack of sub-daily studies in the country. This research aims to assess patterns related to the structure and microphysics of clouds driving intense rainfall [...] Read more.
Intense rainfall events frequently occur in Brazil, often leading to rapid flooding. Despite their recurrence, there is a notable lack of sub-daily studies in the country. This research aims to assess patterns related to the structure and microphysics of clouds driving intense rainfall in Brazil, resulting in high accumulation within 1 h. Employing a 40 mm/h threshold and validation criteria, 83 events were selected for study, observed by both single and dual-polarization radars. Contoured Frequency by Altitude Diagrams (CFADs) of reflectivity, Vertical Integrated Liquid (VIL), and Vertical Integrated Ice (VII) are employed to scrutinize the vertical cloud characteristics in each region. To address limitations arising from the absence of polarimetric coverage in some events, one case study focusing on polarimetric variables is included. The results reveal that the generating system (synoptic or mesoscale) of intense rain events significantly influences the rainfall pattern, mainly in the South, Southeast, and Midwest regions. Regional CFADs unveil primary convective columns with 40–50 dBZ reflectivity, extending to approximately 6 km. The microphysical analysis highlights the rapid structural intensification, challenging the event predictability and the issuance of timely, specific warnings. Full article
(This article belongs to the Special Issue Early Career Scientists' (ECS) Contributions to Meteorology (2024))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of CEMADEN, DECEA, and SIPAM radars in the Brazilian territory, with INMET stations selected for the study.</p>
Full article ">Figure 2
<p>Examples of case validations. For each time analysis, an area 5 × 5 km was created centered on the same station coordinate, and the time of the pixel with the highest value (PMAX) is recorded.</p>
Full article ">Figure 3
<p>Example of the tracking carried out for a case that occurred in the municipality of Feira de Santana-BA, which is covered by the Salvador radar. The colors indicate the shapefile extracted at each time step of the storm.</p>
Full article ">Figure 4
<p>Procedure carried out to construct CFADs.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>e</b>) North region, Northeast region, Midwest region, Southeast region and South region. VIL calculation for the analyzed time instants. The PMAX is the reference period for the highest reflectivity value over the station location during the event. The colors indicate the median values of the VIL values, with shades of blue referring to higher medians (higher VIL values) and shades of brown to lower median values (lower VIL values). The red dots are the outliars.</p>
Full article ">Figure 6
<p>Similar to <a href="#meteorology-03-00016-f005" class="html-fig">Figure 5</a> but for cloud-integrated ice content (VII). (<b>a</b>–<b>d</b>) North region, Northeast region, Midwest region, Southeast region. Note that the y-axis in panel (<b>e</b>) differs from the other panels.</p>
Full article ">Figure 7
<p>CFAD of the Northern region of Brazil created from an area 5 × 5 km (25 km<sup>2</sup>), centered on the pixel with the highest VIL for 15 intense rain events selected in the region. n = 375 refers to the number of vertical profiles used in generating the CFAD. As 25 vertical profiles were extracted for each event (due to the size of the area) and 15 cases were studied in this region, there were a total of 375 vertical profiles in analyzing the events as a whole.</p>
Full article ">Figure 8
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Northeast region of Brazil. In total, 200 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 9
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Midwest region of Brazil. In total, 350 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 10
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the Southeast region of Brazil. In total, 725 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 11
<p>Similar to <a href="#meteorology-03-00016-f007" class="html-fig">Figure 7</a> but for the South region of Brazil. In total, 425 vertical profiles were used, referring to 08 selected events.</p>
Full article ">Figure 12
<p>Standard deviations of reflectivity values as a function of height for each instant analyzed in the creation of the CFADs. The vertical line represents the 75th percentile (P75) of the entire deviation dataset. The horizontal line represents the height at which the deviation values are above P75. The colors represent the deviations for each instant and height separated by regions.</p>
Full article ">Figure 13
<p>Location of the Santa Tereza (blue) and Três Marias (green) radars and the INMET automatic stations (red) used in the study. The black dots represent the position of the weather radars.</p>
Full article ">Figure 14
<p>Life cycle of the water (VIL) and ice (VII) contents integrated in the cloud in the highest-intensity pixel (VIL and VII) of each analyzed event (<b>a</b>–<b>g</b>). The dashed vertical line indicates the PMAX instant.</p>
Full article ">Figure 15
<p>CFAD frequency diagram of the reflectivity variable using a 25 km<sup>2</sup> sample centered on the maximum VIL value for each instant analyzed. The CFAD was built from the 7 cases studied, and therefore, with 175 vertical profiles. The PMAX is the reference period in which the maximum reflectivity value on the rain gauge was observed within the hour of recording the accumulated rainfall. The y-axis refers to height in km and the x-axis to reflectivity intervals in dBZ.</p>
Full article ">Figure 16
<p>Similar to <a href="#meteorology-03-00016-f015" class="html-fig">Figure 15</a> but for the <math display="inline"><semantics> <msub> <mi>Z</mi> <mrow> <mi>D</mi> <mi>R</mi> </mrow> </msub> </semantics></math> variable.</p>
Full article ">
26 pages, 3009 KiB  
Article
Phenotypic Diversity and Seed Germination of Elaeagnus angustifolia L. in Relation to the Geographical Environment in Gansu Province, China
by Kaiqiang Zhang, Zhu Zhu, Rongrong Shi, Ningrui Shi, Qing Tian and Xuemei Lu
Agronomy 2024, 14(9), 2165; https://doi.org/10.3390/agronomy14092165 - 22 Sep 2024
Viewed by 527
Abstract
Elaeagnus angustifolia L. is a highly adaptable urban ornamental plant, playing a key role in dry land and saline-alkali protective forests. The diverse geographical and climatic conditions in Gansu Province have resulted in variations in its distribution and growth. This study assesses the [...] Read more.
Elaeagnus angustifolia L. is a highly adaptable urban ornamental plant, playing a key role in dry land and saline-alkali protective forests. The diverse geographical and climatic conditions in Gansu Province have resulted in variations in its distribution and growth. This study assesses the phenotypic diversity of fruits and seeds, and the seed germination characteristics of 82 E. angustifolia plants from nine populations in Gansu Province, exploring their relationship with geographical and climatic factors. We measured 12 phenotypic traits and five germination indices. This study included germination tests under standard conditions, statistical analysis of phenotypic differences, and Pearson and Spearman correlation analyses to examine relationships between traits and geo-climatic factors. Principal component and cluster analyses were also performed to identify key traits and classify populations. The findings were as follows: (1) Significant differences were observed in phenotypic traits and germination characteristics among populations. Single fruit weight showed the highest variation (27.56%), while seed transverse diameter had the lowest (8.76%). The Lanzhou population exhibited the greatest variability (14.27%), while Linze had the lowest (6.29%). (2) A gradient change pattern in traits was observed, primarily influenced by longitude and a combination of geographical and climatic factors. Seed germination was positively correlated with altitude, annual precipitation, and relative humidity, but negatively affected by latitude and traits such as fruit weight. (3) Principal component analysis identified germination rate, germination index, seed shape index, and fruit shape index as primary factors, contributing 27.4%, 20.6%, and 19.9% to the variation, respectively. Cluster analysis grouped the 82 plants into four clusters, not strictly based on geographical distance, suggesting influence from factors such as genotype or environmental conditions. In conclusion, this study lays a foundation for understanding the genetic mechanisms behind the phenotypic diversity and germination characteristics of E. angustifolia. It offers insights into how geo-climatic factors influence these traits, providing valuable information for the species’ conservation, cultivation, and management. Full article
(This article belongs to the Section Horticultural and Floricultural Crops)
Show Figures

Figure 1

Figure 1
<p>The key traits of <span class="html-italic">E. angustifolia</span> fruits and seeds. (<b>a</b>) A color chart with four stages: yellow, orange-yellow, orange-red, and red (Pantone, Inc., Carlstadt, NJ, USA). (<b>b</b>) Measurements of fruit and seed traits, including fruit longitudinal diameter (FLD), fruit transverse diameter (FTD), seed longitudinal diameter (SLD), and seed transverse diameter (STD). FLD and SLD measure the longest distance from top to bottom, while FTD and STD measure the widest distance across.</p>
Full article ">Figure 2
<p>A heatmap of correlations between the phenotypic traits of <span class="html-italic">E. angustifolia</span> fruits and seeds. (<b>a</b>) Displays the correlations between 12 phenotypic traits and (<b>b</b>) shows the correlations between these traits and geo−climatic factors. AL: Altitude, E: Longitude, N: Latitude, AMT: Annual Mean Temperature, AP: Annual Precipitation, AMRH: Annual Mean Relative Humidity. Phenotypic traits include fruit longitudinal diameter (FLD), fruit transverse diameter (FTD), fruit shape index (FSI), fruit weight (FW), fruit stalk length (FRS), seed longitudinal diameter (SLD), seed transverse diameter (STD), seed shape index (SSI), thousand-−grain weight (TGW), fruit shape (FS), fruit color (FRC), and speckles on fruit (SF). Color intensity indicates correlation strength, with blue for negative and red for positive correlations. Significant correlations are marked with * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>A heatmap of the correlations of the phenotypic traits of <span class="html-italic">E. angustifolia</span> fruits and seeds, seed germination indicators, and geographical−climatic factors. (<b>a</b>) The correlation between seed germination indicators (e.g., TG, GP, GI, VI, and GT<sub>50</sub>) and geographic environmental factors (e.g., altitude, latitude, longitude, annual mean temperature, annual precipitation, and annual mean relative humidity). (<b>b</b>) The correlation between phenotypic traits (e.g., fruit shape, fruit color, fruit weight, etc.) and seed germination indicators. The intensity of the color represents the strength of the correlation, with red indicating a positive correlation and blue indicating a negative correlation. Significant correlations are marked with * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>The principal component analysis results of the phenotypic traits of <span class="html-italic">E. angustifolia</span> fruits and seeds in Gansu Province. (<b>a</b>) The loadings and projections of the top three principal components based on the principal component analysis of the phenotypic traits of <span class="html-italic">E. angustifolia</span> fruits and seeds. (<b>b</b>) A scatter plot of the distribution of 82 <span class="html-italic">E. angustifolia</span> samples based on PC1 and PC2.</p>
Full article ">Figure 5
<p>The cluster analysis results based on squared Euclidean distance using Ward’s method. The dendrogram represents the hierarchical clustering of <span class="html-italic">E. angustifolia</span> samples, with different colors (A, B, C, and D) indicating distinct clusters formed by Ward linkage.</p>
Full article ">
24 pages, 8315 KiB  
Article
Spatiotemporal Changes in Vegetation Cover during the Growing Season and Its Implications for Chinese Grain for Green Program in the Luo River Basin
by Xuning Qiao, Jing Zhang, Liang Liu, Jinyuan Zhang and Tongqian Zhao
Forests 2024, 15(9), 1649; https://doi.org/10.3390/f15091649 - 19 Sep 2024
Viewed by 535
Abstract
The Grain for Green Program (GFGP) plays a critical role in enhancing watershed vegetation cover. Analyzing changes in vegetation cover provides significant practical value in guiding ecological conservation and restoration in vulnerable regions. This study utilizes MOD13Q1 NDVI data to construct the Kernel [...] Read more.
The Grain for Green Program (GFGP) plays a critical role in enhancing watershed vegetation cover. Analyzing changes in vegetation cover provides significant practical value in guiding ecological conservation and restoration in vulnerable regions. This study utilizes MOD13Q1 NDVI data to construct the Kernel Normalized Difference Vegetation Index (kNDVI) and analyzes the spatiotemporal evolution and future trends of vegetation cover from 2000 to 2020, covering key periods of the GFGP. The study innovatively combines the optimal parameter geographic detector with constraint lines to comprehensively reveal the nonlinear constraints, intensities, and critical thresholds imposed by various driving factors on the kNDVI. The results indicate that the following: (1) The vegetation cover of the Luo River Basin increased significantly between 2000 and 2020, with a noticeable increase in the percentage of high-quality vegetation. Spatially, the vegetation cover followed a pattern of being “high in the southwest and low in the northeast”, with 73.69% of the region displaying improved vegetation conditions. Future vegetation degradation is predicted to threaten 59.40% of the region, showing a continuous or future declining trend. (2) The primary driving factors for changes in the vegetation cover are evapotranspiration, elevation, population density, and geomorphology type, with temperatures and GDP being secondary factors. Dual-factor enhancement or nonlinear enhancement was observed in interactions among the factors, with evapotranspiration and population density having the largest interaction (q = 0.76). (3) The effects of driving factors on vegetation exhibited various patterns, with thresholds existing for the hump-shaped and concave-waved types. The stability of the kNDVI in 40.23% of the areas showed moderate to high fluctuations, with the most significant fluctuations observed in low-altitude and high-temperature areas, as well as those impacted by dense human activities. (4) By overlaying the kNDVI classifications on the GFGP areas, priority reforestation areas totaling 68.27 km2 were identified. The findings can help decisionmakers optimize the next phase of the GFGP and in effective regional ecological management. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>The study area in the Luo River Basin.</p>
Full article ">Figure 2
<p>The flow chart of the study.</p>
Full article ">Figure 3
<p>Interannual variation; (<b>a</b>) Annual annual mean variation trend in the kNDVI; (<b>b</b>) the results of the Mann–Kendall change–point test.</p>
Full article ">Figure 4
<p>Statistics on the multiyear average kNDVI in the Luo River Basin: (<b>a</b>) average kNDVI spatial distribution; (<b>b</b>) proportion of different kNDVI levels under various geomorphological types.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>d</b>) Characteristics of the changes in the kNDVI during the critical period; (<b>e</b>) variations in the percentage of the various classes of the kNDVI from 2000 to 2020.</p>
Full article ">Figure 6
<p>Significance analysis of the kNDVI trends and future sustainability. (<b>a</b>) the trend of Sen–MK changed significantly; (<b>b</b>) Hurst index; (<b>c</b>) future change trend.</p>
Full article ">Figure 7
<p>The explanatory power of various driving factors on the spatial pattern of the kNDVI in 2020: (<b>a</b>) results of the detection of the driving factors of the kNDVI; (<b>b</b>) results of the detection of interactions among the driving factors. Note * indicates nonlinear enhancement, and other dual-factor enhancement.</p>
Full article ">Figure 8
<p>The constraint relationship between the continuous driving factor indicators and kNDVI. (<b>a</b>) Constraint line of DEM; (<b>b</b>) Constraint line of Slope; (<b>c</b>) Constraint line of PRE; (<b>d</b>) Constraint line of TEM; (<b>e</b>) Constraint line of PET; (<b>f</b>) Constraint line of POP; (<b>g</b>) Constraint line of GDP; (<b>h</b>) Constraint line of NL; (<b>i</b>) Constraint line of NHD; (<b>j</b>) Constraint line of PHD; (<b>k</b>) Constraint line of RD; (<b>l</b>) Constraint line of WD.</p>
Full article ">Figure 9
<p>Spatial distribution pattern of the CV of the annual mean kNDVI in the Luo River Basin.</p>
Full article ">Figure 10
<p>Proportion of the kNDVI stability classes in key impact factor subdivisions.</p>
Full article ">Figure 11
<p>The kNDVI identification of prioritized GFGP areas.</p>
Full article ">
Back to TopTop